You are on page 1of 16

REVIEWS

Metabolism of immune cells in cancer


Robert D. Leone and Jonathan D. Powell ✉

Abstract | Through the successes of checkpoint blockade and adoptive cellular therapy,
immunotherapy has become an established treatment modality for cancer. Cellular metabolism
has emerged as a critical determinant of the viability and function of both cancer cells and
immune cells. In order to sustain prodigious anabolic needs, tumours employ a specialized
metabolism that differs from untransformed somatic cells. This metabolism leads to a tumour
microenvironment that is commonly acidic, hypoxic and/or depleted of critical nutrients required
by immune cells. In this context, tumour metabolism itself is a checkpoint that can limit
immune-​mediated tumour destruction. Because our understanding of immune cell metabolism
and cancer metabolism has grown significantly in the past decade, we are on the cusp of being
able to unravel the interaction of cancer cell metabolism and immune metabolism in
therapeutically meaningful ways. Although there are metabolic processes that are seemingly
fundamental to both cancer and responding immune cells, metabolic heterogeneity and
plasticity may serve to distinguish the two. As such, understanding the differential metabolic
requirements of the diverse cells that comprise an immune response to cancer offers an
opportunity to selectively regulate immune cell function. Such a nuanced evaluation of cancer
and immune metabolism can uncover metabolic vulnerabilities and therapeutic windows upon
which to intervene for enhanced immunotherapy.

Oxidative phosphorylation
Work over the past several decades has shown that acti­ toxic at certain concentrations. A significant amount
(OXPHOS). A highly efficient vated immune cells employ many metabolic pathways of glucose from the TME is metabolized through aer­
form of cellular respiration attributed to cancer cells1–3 (Fig. 1). This convergence of obic glycolysis, generating high amounts of lactate and
synthesizing ATP from the metabolic adaptations creates a fundamental competi­ H+, thereby lowering the intratumoural pH. That said,
phosphorylation of ADP using
tion for nutrients required by cancer cells and immune it is likely that the balance between lactate-​generating
electrochemical potential
energy generated by the
cells within the tumour microenvironment (TME). glycolysis and oxidative phosphorylation (OXPHOS) is
transfer of electrons from However, we are coming to find fundamental differences dependent on the degree of hypoxia, which can be both
NADH or FADH2 to oxygen between the metabolic programmes of cancer cells and heterogeneous and wide ranging within the TME. It is
through a series of immune cells, as well as between different immune cells. instructive to note that in moderately hypoxic regions,
mitochondrial electron carriers.
Understanding these differences can reveal specific meta­ CO2 derived from mitochondrial respiration is hydrated
bolic vulnerabilities and, consequently, novel targets for by extracellular carbonic anhydrase enzymes, forming
therapeutic approaches aimed at metabolic programming HCO3– and H+. Thus, oxidative metabolism can be a
in order to enhance cancer immunotherapy. significant and often overlooked source of extracellular
Although the ability of cancer cells and tumour tis­ acidification within the TME.
sue to upregulate glycolytic catabolism of glucose to Given the recent establishment of cancer immu­
form lactate, even in oxygen-​replete conditions (aero­ notherapy, including the use of blocking antibodies
bic glycolysis), a process known as the ‘Warburg effect’, against immune checkpoint pathways and adoptive cell
has been considered a hallmark of malignancy, it has therapy with chimeric antigen receptor T cells (CAR
Bloomberg~Kimmel Institute
for Cancer Immunotherapy, become increasingly clear that cancer metabolism is het­ T cells), several recent studies have begun to establish
Sidney Kimmel erogeneous, and that cancer cells can engage in a broad the relationship of tumour-​intrinsic metabolism to
Comprehensive Cancer range of metabolic programmes to meet the demands of successful immunotherapy. For instance, it has been
Research Center, Department growth and proliferation, and that in addition to aerobic reported that increased glycolytic metabolism in mela­
of Oncology, Johns Hopkins
University School of Medicine,
glycolysis, mitochondrial respiration is fundamentally noma cells is associated with resistance to adoptive
Baltimore, MD, USA. important in this regard4–7. Predictably, highly metabol­ T cell therapy and checkpoint blockade8,9. Other stud­
✉e-​mail: poweljo@jhmi.edu ically active cancer cells (Fig. 1) impart profound effects ies have shown that signalling through immune check­
https://doi.org/10.1038/ on the TME, leading to nutrient depletion, hypoxia, point proteins on tumour cells, including PD1 and
s41568-020-0273-​y acidity and the generation of metabolites that can be B7-​H3, was responsible for increased glucose depletion

Nature Reviews | Cancer


Reviews

Tumour microenvironment Necrotic


Tryptophan IDO Kynurenine cancer cell
Cancer cell
Arginine ARG1 Polyamines K+
Treg cell
ATP CD39 Adenosine Necrosis
CD73

Glucose Glycolysis Fructose-6-phosphate 3-phosphoglycerate


MDSC Teff cell
Glucose-6- Hexosamine Cellular
phosphate biosynthesis pathway regulation

REDOX homeostasis Serine–


glycine–
Pentose
one-carbon
phosphate Nucleotide synthesis pathway
Leaky pathway
blood Lipid synthesis
vessel Acetyl-CoA
Pyruvate
Amino acid and
Amino acids protein synthesis Aerobic
glycolysis

Citrate Lactate, H+ Lactate, H+


Dendritic TCA
cell Fibroblast Glutamine cycle Energy
homeostasis
Mitochondrion ROS
Macrophage Oxygen ETC
R-2-HG
Cancer cell

Depleted nutrients Immunosuppressive metabolites

Fig. 1 | Cancer cell metabolism and derangements in the TMe. upregulated in the setting of increased proliferation72. In addition to
Mitochondrial oxidation of nutrients, including glucose, amino acids and supplying the TCA cycle with carbon skeletons that maintain intermediates
fatty acids, through the tricarboxylic acid (TCA) cycle and the electron for biosynthesis of amino acids, nucleic acids and fatty acids (a process
transport chain (ETC) is a highly efficient means of producing energy for known as anaplerosis), glutamine is the primary source of nitrogen used for
quiescent, differentiated cells. However, during periods of increased amino acid and nucleic acid synthesis. These cells also upregulate a broad
proliferation, such as after immune activation or malignant transformation, range of amino acid transporters and maintain tightly controlled redox
cells upregulate an alternative pathway for glucose metabolism, called balance, primarily through NADPH synthesis. Many cells within the tumour
aerobic glycolysis. Although less efficient in generating ATP, aerobic microenvironment (TME) express ectoenzymes, such as indoleamine 2,3-​
glycolysis allows for more rapid metabolism of glucose, efficient disposal of dioxygenase (IDO), arginase 1 (ARG1) and CD73, which deplete nutrients,
excess carbon and regeneration of NAD+ while preserving mitochondrial as well as increase immunosuppressive metabolites, such as kynurenine and
enzymatic activity for anabolic processes242. Glycolytic intermediates are adenosine. Along with a deranged microvasculature, these metabolic
channelled through other essential pathways, such as the pentose adaptations can have profound effects on the metabolic make-​up of
phosphate pathway, the one-​carbon pathway and the hexosamine the TME, leading to depletion of vital nutrients, hypoxia, acidosis and the
biosynthesis pathway. These pathways support cellular processes that are generation of immune-​toxic metabolites as shown. MDSC, myeloid-​derived
critical for highly proliferative cells, such as synthesis of fatty acids and suppressor cell; R-2-​HG, (R)-2-​hydroxyglutarate; ROS, reactive oxygen
nucleic acids. Pathways for the metabolism of glutamine are also species; Teff, effector T; Treg, regulatory T.

within the TME10–12. Interestingly, some immunosup­ in the immune response to cancer; how these metabolic
pressive checkpoint pathways are actually induced as a programmes might be perturbed within the TME; the
Immune checkpoint
pathways direct consequence of tumour acidification13. Further, implications of metabolic derangements in the TME for
Pathways mediated by cell immune checkpoint blockade can dampen glycolysis current immunotherapeutic paradigms; and how meta­
surface proteins on immune of tumour cells, restore glucose in the TME and per­ bolic interventions might be leveraged to enhance the
cells, such as PD1 or CTLA4, mit T cell glycolysis and cytokine production14. Several antitumour immune response.
that serve to suppress the
immune response, which can
recent studies have demonstrated that targeting specific
be activated by ligands within aspects of tumour-​intrinsic metabolism, such as the The TME and immune contexture
the tumour microenvironment hexosamine biosynthesis pathway (HBP) or glutamine Highly active metabolic pathways that are characteristic
or draining lymph nodes. metabolism, could foster an immune response and of cancer cells (Fig. 1) can create profound changes in
sensitize tumours to checkpoint blockade15,16. the composition of nutrients and other small molecules
Chimeric antigen receptor
T cells Because of the emergence of immunotherapy as within the TME. This can have critical effects on the
(CAR T cells). T cells harvested a pillar of oncologic therapy, it is increasingly vital to immune response. The high metabolic activity of cancer
from a patient’s blood and understand as much as possible about the metabolic cells and disorganized vasculature within the TME can
genetically modified to express interdependence of infiltrating immune cells and can­ contribute to a nutrient-​depleted and hypoxic microen­
a special receptor that can
recognize and respond to
cer. This Review aims to discuss the following funda­ vironment, establishing metabolic competition between
specific, predefined molecular mental questions: which metabolic programmes are cancer cells and infiltrating immune cells14,17,18. Indeed,
targets on tumour cells. critical for the function of specific cell subsets involved the glucose uptake and effector function of antitumour

www.nature.com/nrc
Reviews

Table 1 | Functional and metabolic phenotypes of immune cells within the TMe expression, which encodes the rate-​limiting enzyme in
the glycolytic pathway18. Metabolic programmes active
Cell type Function Metabolic phenotype within cells of the TME can also lead to the generation
Immune activation or inflammatory of toxic concentrations of certain metabolites. Elevated
NK cell MHC-​independent Glycolysis and OXPHOS levels of adenosine, kynurenine, ornithine, reactive oxy­
cytotoxicity: gen species (ROS) and potassium, as well as increased
Perforin, granzymes acidosis, have all been reported in the TME, and each
can have profound effects in suppressing the antitumour
FASL, TRAIL
immune response.
IFNγ, TNF The immune contexture of the TME comprises
Inflammatory TAM MHC-​independent Glycolysis and PPP a range of distinct cell types 19 (Table 1) . Effector
cytotoxicity: cells perform functions aimed at cell killing and can
TNF, IL-1β arise from either the innate (non-​specific) or adap­
Oxidative burst tive (antigen-​specific) arms of the immune system.
Antigen presentation Antitumour effector cells arising from the adaptive
system include CD4+ and CD8+ Teff cells, which orches­
DC DAMP processing Glycolysis
trate and carry out antigen-​specific killing of cancer
Teff cell activation cells, respectively. CD8+ Teff cells are critically important
Antigen presentation in direct tumour cell killing through the induction of
apoptosis and cytokine secretion. CD4+ T cells comprise
Teff cell Antigen-​specific cytotoxicity: Highly glycolytic and numerous subsets. Some of these subsets, the most well
OXPHOS studied of which is the T helper 1 (TH1) subset, can also
Perforin, granzymes
Amino acid metabolism
provide significant antitumour activity. These antitu­
FASL mour CD4+ T cells, collectively termed conventional
(arginine, tryptophan, serine,
IFNγ, TNF leucine, glutamine, cysteine) CD4+ (CD4+conv) T cells, are distinct from immunosup­
PPP pressive, pro-​tumorigenic CD4+ T cells known as regu­
latory T (Treg) cells. Although CD4+conv cells may engage
Tmem cell Maintain long-​lived response OXPHOS
in direct tumour cell killing, they primarily contribute
to antitumour immunity through cytokine secretion and
assisting in CD8+ T cell activation. Antitumour CD4+conv
T cells share significant metabolic characteristics with
Immunosuppression CD8+ Teff cells. Although less well understood in terms
of antitumour immunity, B cells may also perform effec­
MDSC IL-10, TGFβ Glycolysis and OXPHOS
tor roles in the TME20. Importantly, as part of the adap­
Amino acid depletion tive immune system, T cells and B cells can give rise to
Polyamines, kynurenine memory cell populations, which can persist long after
the resolution of an infection or tumour response. CD8+
Immunosuppressive IL-10 OXPHOS, HBP memory T (Tmem) cells are a crucial aspect of long-​term
TAM tumour control. Innate cells, such as natural killer (NK)
Amino acid depletion
cells and inflammatory macrophages, perform critical
Polyamines, kynurenine antitumour effector functions as well. There are also
VEGF immunosuppressive cell populations within the TME,
including CD4+FOXP3+ Treg cells, myeloid-​derived sup­
Treg cell IL-2 sequestration: OXPHOS pressor cells (MDSCs), anti-​inflammatory macrophages
Dampen APC co-​stimulation and some B cell populations20. Through various mech­
IL-10, TGFβ anisms, including cytokine secretion and metabolic
Adenosine derangements, these cells can dampen or eliminate the
effectiveness of antitumour effector cell populations.
APC, antigen-​presenting cell; DAMP, damage-​associated molecular pattern; DC, dendritic
cell; FASL, fas ligand; HBP, hexosamine biosynthesis pathway; IFNγ, interferon-​γ; MDSC, Lastly, antigen-​presenting cells, such as intratumoural
myeloid-​derived suppressor cell; MHC, major histocompatibility complex; NK, natural killer; dendritic cells (DCs), have been shown to perform
OXPHOS, oxidative phosphorylation; PPP, pentose phosphate pathway; TAM, tumour-​associated essential roles in maintaining active adaptive immune
macrophage; TGFβ, transforming growth factor-β; Teff, effector T; TME, tumour microenvironment;
Tmem, memory T; TNF, tumour necrosis factor; TRAIL, TNF-​related apoptosis-​inducing ligand; response within the TME 21,22. Numerous excellent
Treg, regulatory T; VEGF, vascular endothelial growth factor. reviews can be referred to for more detailed discussions
of tumour immunology and immunotherapy19,23–25.
CD4+ T cells has been shown to be inversely propor­
tional to glycolytic activity of cancer cells in mouse The metabolism of the antitumour response
models18, and glucose availability in the TME allows Glucose metabolism of antitumour effector T cells.
Hexosamine biosynthesis for improved cytokine expression from antitumour CD4+conv and CD8+ Teff cells form the critical effector
pathway CD8+ T cells14. Furthermore, transcriptomic analyses compartment of the antitumour response. When naive
(HBP). A branch of glycolysis
that generates building blocks
of patients with melanoma from The Cancer Genome CD4+ and CD8+ T cells, which are non-​proliferative,
used for glycosylation of Atlas revealed that effector T (Teff ) cell genes, such as recognize their cognate antigen in the context of
proteins and lipids. CD40lg and IFNG, are inversely correlated with HK2 co-​stimulatory signalling, they become proliferative and

Nature Reviews | Cancer


Reviews

enact metabolic features to support immense growth26–28. Teff cells are highly proliferative and upregulate specific
Although many early investigations highlighted the glycolytic programmes, including aerobic glycolysis,
upregulation of aerobic glycolysis as a hallmark of T cell PPP, HBP and TCA cycle support, to allow massive cell
activation, it is now clear that upregulated tricarboxylic division and effector functions.
acid (TCA) cycle metabolism and OXPHOS are also a
critical aspect of CD4+conv and CD8+ T cell activation. T cells and glucose restriction in the TME. Glucose lim­
Although TCA cycle metabolism is upregulated within itation within the TME can markedly affect the T cell
24 h post activation, upregulated aerobic glycolysis response. For example, low-​glucose conditions (0.1 mM)
appears to be a more rapid event, occurring within 6 h suppressed the generation of the glycolytic intermedi­
after activation27–32. ate phosphoenolpyruvate (PEP) in T cells, which dis­
The transcriptional activity of MYC and hypoxia rupted calcium-​dependent NFAT signalling in vitro18.
inducible factor 1 (HIF-1) are both upregulated in Compared with control, decreasing the glucose con­
response to T cell activation and promote metabolic centration in growth media has been shown to suppress
reprogramming26,29,33,34. Notably, although HIF-1 is well the extracellular acidification rate (a measure of aerobic
known to regulate metabolism in response to hypoxia, its glycolysis), augment the oxygen consumption rate (a
activity is also induced in response to T cell activation in measure of OXPHOS), attenuate mTOR signalling and
the absence of hypoxia. MYC and HIF-1 transcriptional suppress the effector function of both CD4+ and CD8+
activity leads to upregulation of genes encoding enzymes Teff cells46–48. Reduced mTOR complex 1 (mTORC1)
that promote glycolysis, such as pyruvate kinase (PKM1), signalling interfered with Teff cell differentiation and,
hexokinase 2 (HK2) and GLUT1 (refs29,34,35). Pathways in the case of CD4+ T cells, specifically favoured the
emanating from proximal metabolites in the glycolytic development of immunosuppressive, pro-​tumorigenic
pathway are also integral components of T cell activa­ Treg cells49. Interestingly, in CD8+ T cells, mTOR block­
tion and function (Fig. 1). The pentose phosphate pathway ade with rapamycin favoured differentiation of long-​
(PPP) metabolizes glucose-6-​phosphate to generate lived Tmem cells, which may play an important role in
NADPH and ribose-5-​phosphate36. Glucose shuttling sustaining antitumour responses49–51. Decreasing glu­
into the PPP is significantly increased upon CD4+ T cell cose availability in culture suppressed production of the
activation29. The PPP is the primary cellular source for critical effector molecules interferon-​γ (INFγ), IL-17
NADPH, which is required for fatty acid and plasma and granzyme B in Teff cells compared with control
membrane synthesis in newly activated CD8+ T cells37. growth media47,48,52,53. In activated CD4+ T cells cultured
NADPH is also critical for REDOX homeostasis in pro­ in glucose-​free media containing the alternative sugar
liferating mammalian cells38–40. ROS levels in activated fuel galactose (which suppresses aerobic glycolysis), the
T cells need to be finely regulated. Although dysregulated glycolytic enzyme GAPDH assumed a moonlighting
ROS levels can be toxic39,41,42, ROS play an important role role, binding the 3′ untranslated region of Ifng mRNA
in Teff cell activation, having been shown to promote and suppressing its translation and Teff cell function31.
nuclear factor of activated T cells (NFAT)-​dependent IL-2 Glucose restriction in media conditioned by primary
expression in CD4+ and CD8+ T cells. Another pathway ovarian cancer cells led to microRNA-​mediated suppres­
originating from early glycolytic reactions, the HBP, is sion of the histone methylase EZH2 (enhancer of zeste
the primary cellular source of glycosylation substrates, homologue 2), leading to decreased NOTCH signalling,
which mediate a variety of effects on a broad range of suppressed cytokine production and decreased viability
proteins, including stability, trafficking and function. of Teff cells54.
Pentose phosphate pathway The HBP relies on metabolism of glucose and glutamine Increasing the glycolytic capacity of mouse sarcoma
(PPP). A metabolic branch of
glycolysis generating NADPH,
and is responsive to their availability. The main substrate cells through either pharmacologic treatment with the
used for fatty acid synthesis produced by the HBP, UDP-​GlcNAc, is critical for effec­ AKT activator 4-​hydroxytamoxifen in co-​culture exper­
and redox homeostasis, and tor CD4+ and CD8+ T cell expansion and function43. iments or overexpression of key glycolytic enzymes (for
5-​carbon sugars used in Lastly, the serine–glycine–one-​carbon pathway allows example, Glut1, Hk2 and Pdk1) in tumour cells followed
nucleotide synthesis.
cells the ability to generate serine, glycine, NADPH and by injection into mice led to suppression of CD8+ T cell
Nuclear factor of activated one-​carbon units for use in the folate cycle. Teff cell pro­ effector function compared with vehicle-​treated tumour
T cells liferation and function were dependent on sufficient cells or empty vector overexpression, respectively14.
(NFAT). A calcium-​dependent serine metabolism in vitro and in vivo44 (Fig. 1). Similarly, compared with wild-​type tumours, implanted
transcription factor activated in
Glucose carbons that are not metabolized to lactate or Hk2-​o verexpressing melanoma cells suppressed
response to T cell receptor
stimulation. Cooperation with
by proximal glycolytic pathways contribute significantly CD4+ T cell antitumour effector function and in vivo
the AP-1 transcription factor to the TCA cycle in Teff cells6 (Fig. 1). In highly prolifer­ responses in mouse models18. Furthermore, expres­
results in a productive immune ative cells, intermediates of the TCA cycle are rapidly sion of glycolysis-​related genes in tumour samples
response and transcription of consumed to serve as building blocks for a broad range from patients with melanoma and non-​small-​cell lung
pro-​inflammatory cytokines,
of biomolecular syntheses, a process called cataplerosis45. cancer was inversely correlated to T cell infiltration8.
such as IL-2 and interferon-​γ.
For example, citrate can be exported to the cytoplasm to Tipping the metabolic balance can also be accomplished
Cataplerosis regenerate acetyl-​CoA for use in lipid and cholesterol through directly manipulating T cell metabolism. For
The loss of metabolic synthesis, both of which are critical for producing mem­ example, overexpression of the glycolytic enzyme
intermediates in a metabolic
branes in proliferative Teff cells. Other TCA cycle inter­ PEP carboxykinase in tumour-​specific CD4+ T cells
pathway (particularly the
tricarboxylic acid cycle)
mediates function as building blocks for biosynthesis improved antitumour responses compared with control
owing to consumption or of, for example, nucleotides and amino acids, which are vector-​transfected T cells in an adoptive T cell model
degradation. in high demand during proliferation. Like cancer cells, using melanoma-​specific T cells18.

www.nature.com/nrc
Reviews

Anaplerosis Mitochondrial respiration is also a critical aspect of in vitro human and mouse studies61. Essential amino
The process of replenishing Teff cell metabolism. Several recent studies have reported acids must be obtained exogenously. For example,
intermediates of the that T cells in patients with cancer (compared with leucine was required for mTORC1 signalling, effector
tricarboxylic acid cycle to healthy controls) and tumour-​infiltrating CD8+ T cells in function and proper differentiation in effector CD8+
support biosynthesis.
tumour-​bearing mice (compared with non-​infiltrating and CD4+conv T cells. Interestingly, deletion of the leucine
CD8+ T cells) displayed decreased mitochondrial mass transporter, Slc7a5, in mouse models caused metabolic
as well as indicators of mitochondrial dysfunction55–57. failure during in vitro activation and cytokine-​directed
Mitochondrial fitness of resting peripheral CD8+ T cells differentiation of CD4+ (TH1, IL-17-​producing TH17)
was impaired in patients with chronic lymphocytic leu­ and CD8+ Teff cells, but had no adverse effect on the dif­
kaemia compared with healthy controls55. Furthermore, ferentiation of Treg cells26,62. Activated T cells also rapidly
the degree of response in these patients to CAR T cell metabolize arginine, and exogenous arginine supple­
therapy was negatively correlated to the degree of mentation leads to improved T cell fitness and increased
mitochondrial impairment of infused CAR T cells55. generation of central Tmem cells63. Serine, tryptophan and
Tumour-​infiltrating CD8+ T cells from patients with cysteine are also vital nutrients for T cell responses and,
renal cell carcinoma showed dysregulated mitochon­ as such, are important mediators of antitumour immune
drial dynamics and function, including elevated levels responses44,64–66. Tryptophan is an essential amino acid
of mitochondrial ROS and hyperpolarization, com­ and its availability within the TME is an important fac­
pared with CD8+ T cells from healthy donors56. Normal tor in determining strength and quality of the T cell
ex vivo activation of these T cells could be rescued response. Human T cell proliferation and activation were
with mitochondrial ROS scavengers or pyruvate sup­ strongly suppressed in tryptophan-​free media compared
plementation. Mitochondrial biogenesis and function with normal growth media66,67. Cancer cells, tumour-​
are particularly deranged in a subset of dysfunctional associated macrophages (TAMs), MDSCs, suppressive
tumour-​infiltrating CD8+ T cells termed exhausted DCs and cancer-​associated fibroblasts can deplete tryp­
T cells (Box 1). As a whole, these studies demonstrate that tophan levels through enzymatic activity of indoleam­
cancer itself can lead to derangements in the metabo­ ine 2,3-​dioxygenase (IDO)68, which can be expressed at
lism of Teff cells, including mitochondrial dynamics, and high levels in these cells within the TME. Underlining
that a reciprocal relationship exists between the degree the importance of this metabolic pathway for tumour
of glycolytic activity of cancer cells and the antitumour growth, IDO expression has been correlated with poor
effector function of infiltrating T cells. outcomes in patients with several cancer types, includ­
ing gastric cancer, colorectal cancer, non-​small-​cell lung
Amino acids and the antitumour T cell response. Like cancer and melanoma69–71.
cancer cells, highly proliferative immune cells, such as In proliferating cells, glutamine provides nitrogen
activated T cells, are reliant on amino acid metabolism for amino acid and nucleic acid synthesis and carbon
to support protein and nucleotide synthesis. As such, to replenish TCA cycle intermediates that are syphoned
amino acid transporters, including SLC7A5 (also known off as building blocks for biosynthesis — a process
as LAT1)58, SLC38A1 (also known as SNAT1), SLC38A2 called anaplerosis (Fig. 1). Cancer cells and some acti­
(also known as SNAT2)59 and SLC1A5 (also known as vated immune cells, such as T cells and macrophages,
ASCT2)60, have been found to be highly upregulated are generally highly glutamine avid59,72. The expression
during T cell activation compared with naive cells in of glutamine transporters SLC1A5 and SLC38A1 and/or
SLC38A2 was significantly upregulated during in vitro
Box 1 | Metabolism of T cell exhaustion stimulation of murine CD4+conv T cells60. Driven by
MYC, glutamine is metabolized by glutaminase (GLS)
CD8+ T cells in the tumour microenvironment (TME) can adopt a state of functional
exhaustion wherein they are poorly proliferative and unable to generate sufficient
to glutamate, which may enter the TCA cycle after
cytotoxicity against target cancer cells. A similar cell subset exists during chronic viral conversion by glutamate dehydrogenase (GLUD1) to
infections, such as lymphocytic choriomeningitis virus (LCMV) clone 13 in mice or α-​ketoglutarate (αKG; also known as 2-​oxoglutarate).
hepatitis C virus in humans243. There are numerous metabolic features that are emerging αKG is subsequently metabolized to succinate and
as characteristic of this set of immune cells, termed CD8+ exhausted T cells. Some of the fumarate in the TCA cycle29. Interestingly, in settings
metabolic characteristics appear to be a consequence of co-​inhibitory signalling, such of glutamine restriction, some cancer cell lines switch
as PD1, which is highly characteristic of CD8+ exhausted T cells. In a chronic LCMV to glucose-​fuelled anaplerosis, wherein pyruvate is con­
mouse model, PD1 signalling inhibited the expression of peroxisome proliferator-​ verted by pyruvate carboxylase to oxaloacetate, which
activated receptor-​γ co-​activator 1α (PGC1α), which disrupted mitochondrial and enters the TCA cycle73. Our group has recently shown
effector function and led to significantly less oxidative capacity compared with T cells
that effector CD8+ T cells are also capable of upregu­
responding to an acute LCMV infection244. Overexpression of PGC1α in adoptively
transferred T cells improved mitochondrial function and restored T cell function. T cells
lating pyruvate carboxylase activity under conditions of
infiltrating tumours showed similar mitochondrial dysfunction and loss of oxidative glutamine blockade in vitro16.
capacity secondary to inhibited PGC1α activity57. Interestingly, antitumour T cells also Although effector function and proliferation in dif­
regained function through overexpression of PGC1α, implying that the activity of a ferentiated CD8+ Teff cells was suppressed by limiting glu­
metabolic programme can, in and of itself, overcome functional T cell exhaustion. PD1 tamine in media59, if glutamine availability was restricted
signalling also suppressed mTOR complex 1 (mTORC1) signalling and glycolytic activity during activation of CD8+ T cells, it altered differenti­
in infiltrating CD8+ T cells in a progressive mouse tumour model14. Given the dependence ation towards a long-​lived, memory phenotype74. This
of T cell function (and loss of function) on metabolic programming, more studies are effect on differentiation was shown to be mediated by
needed to assess the determinants of metabolic dysfunction and associated T cell αKG. αKG and other TCA metabolites, such as succinate
exhaustion within the TME.
and fumarate, can modulate the activity of a wide range

Nature Reviews | Cancer


Reviews

of cellular processes, including epigenetic remodelling dioxygenases (2OGDD) through alterations in these
and the stability of critical transcription factors, such as TCA metabolites has been shown to increase memory
HIF-1α (refs45,75). cell differentiation in CD8+ T cells84,85. Although glu­
cose, glutamine and fatty acids are the primary nutrient
Lipid metabolism and T cells. Activated T cells sources fuelling the TCA cycle, a range of other nutri­
also reprogramme lipid metabolism, upregulating ents, such as amino acids and acetate, can also enter
de novo lipid synthesis and cholesterol uptake, which are the cycle. In particular, acetate metabolism is emerging
critical for membrane synthesis and mediated by the as an important source of acetyl-​CoA in CD8+ T cells
transcription factors sterol regulatory element-​binding and some cancer types16,86–88. In the mitochondria,
protein 1 (SREBP1) and SREBP2, respectively 37,76. acetate can enter the TCA cycle after it is metabolized
Proliferation, metabolic reprogramming and antiviral by acyl-​CoA synthetase short chain family member 1
activity were dramatically suppressed in activated mouse (ACSS1) to form acetyl-​CoA. Alternatively, acetate can
CD8+ T cells lacking SREBP1and SREBP2 functional­ be converted to acetyl-​CoA by ACSS2 in the cytoplasm,
ity. In addition, the cholesterol content in membranes where it can contribute to fatty acid synthesis and acetyl­
during CD8+ T cell activation and expansion in vitro ation reactions important in epigenetic reprogramming
was, in part, regulated by cholesterol esterification and post-​translational modifications. The metabolism of
enzyme acetyl-​CoA acetyltransferase (ACAT1). Acat1 acetate is an important metabolic pathway for promot­
knockout CD8+ T cells showed increased membrane ing the function of memory CD8+ T cells88. Interestingly,
cholesterol and improved T cell receptor clustering blockade of glutamine metabolism during T cell acti­
and signalling, leading to enhanced proliferation, func­ vation increased Tmem cell differentiation and induced
tion and improved tumour killing in adoptive-​transfer acetate metabolism and associated enzymes, including
mouse tumour models76. Pharmacologic inhibition of ACSS1 and ACSS216,85. As quiescent cells, Tmem cells
ACAT1 with avasimide improved the antitumour effect preferentially rely on OXPHOS relative to aerobic gly­
in mice compared with vehicle-​treated control animals. colysis and have significant mitochondrial reserve that
Cholesterol metabolism and antitumour T cell function is required to upregulate OXPHOS further upon antigen
is an evolving story, however. A recent study by Ma et al. activation. Tmem cells can adapt several distinct nutrient
demonstrated that a high cholesterol content in tumours sources to fuel this metabolic programme.
can induce T cell dysfunction by activating the endo­
plasmic reticulum stress response77. As such, although Hypoxia and the antitumour T cell response. Although
cholesterol is important for Teff cell proliferation and tumours are highly heterogeneous, high levels of meta­
metabolism, the benefit of targeting specific aspects bolic activity and associated oxygen consumption, as
of cholesterol metabolism to improve the antitumour well as disorganized, poorly functioning vasculature, can
immune response needs further study. generate hypoxic regions with median oxygen saturation
levels <2% (compared with a median of about 5% in nor­
The metabolism of immunologic memory. Unlike mal tissues)89,90. The effect of hypoxia on Teff cells is not
Teff cells, CD8 + Tmem cells preferentially rely on straightforward. Complicating this area of study is the
OXPHOS78–81. Compared with CD8+ Teff cells, enhanced fact that HIF-1 transcriptional activity is upregulated in
spare respiratory capacity, a parameter indicative of the response to T cell activation in normoxic conditions34, so
ability of cells to upregulate OXPHOS, is also highly it is challenging to understand the effect of hypoxia on
characteristic of Tmem cells78. Initial studies using eto­ further augmenting HIF-1 activity while also evaluating
moxir as an inhibitor of carnitine palmitoyl transferase HIF-1-​independent effects. Early in vitro studies of CD8+
1A (CPT1A), a mitochondrial transporter responsible Teff cell activation, differentiation and function showed
for the import of long-​chain fatty acids destined for fatty that whereas proliferation and the expression of some
acid β-​oxidation (FAO), implicated FAO as the primary cytokines were suppressed in hypoxia, the lytic capac­
fuel for OXPHOS in Tmem cells. However, more recent ity, activation markers and survival were improved91.
work using T cell-​specific Cpt1a knockout models has Subsequent in vivo studies showed that CD4+ and CD8+
called this into question and demonstrated that off-​ splenic T cells were more poorly activated after concana­
target effects of high-​dose etomoxir (200 μM) are likely valin A challenge in mice exposed to subatmospheric
responsible for the earlier findings82. This should not be O2 tension (8%) compared with mice exposed to ambi­
De novo lipid synthesis taken to imply that Tmem cells do not use FAO in support ent O2 tension (20%)92. Other studies showed that in vitro
The cellular biosynthesis of
fatty acids, triglycerides,
of OXPHOS and spare respiratory capacity but, rather, hypoxic exposure causes intracellular accumulation of
cholesterol and other lipids that FAO is not the sole pathway responsible for this the metabolite (S)-2-​hydroxyglutarate (S-2-​HG), which
from carbohydrates or other metabolic phenotype. Indeed, the expression of CPT1A profoundly alters CD8+ T cell activation and differen­
non-​lipid precursors. is consistently upregulated in CD8+ Tmem cells com­ tiation, suppressing cytokine secretion and cytolytic
pared with Teff cells. Furthermore, a CD8+ T cell subset capacity, but, interestingly, augmenting proliferation,
2-​Oxoglutarate-​dependent
dioxygenases known as tissue-​resident memory cells were specifically long-​term survival and antitumour response after
(2OGDD). A family of enzymes dependent on fatty acid binding protein 4 (FABP4) and in vivo transfer in mouse models84. Contrary to previ­
that catalyse the hydroxylation FABP5 to import extracellular fatty acids for FAO and for ous findings demonstrating the necessity of oxidative
of macromolecules, often as a maintenance of a long-​term memory phenotype83. metabolism and oxidative metabolic capacity in form­
prerequisite to demethylation,
reliant on α-​ketoglutarate, Fe2+,
Intermediates of the TCA cycle, such as αKG, succi­ ing long-​lived memory CD8+ T cells, glycolytic activity
ascorbate and oxygen as nate and fumarate, are particularly important in adap­ enforced through constitutive HIF-1α activity (achieved
cofactors. tive memory. Inhibition of 2-​oxoglutarate-​dependent through conditional knockout of the HIF-1 regulator

www.nature.com/nrc
Reviews

Tumour microenvironment
Treg cell Cancer Depleted Teff cell Immunosuppressive
cell nutrients metabolites
• Metabolic derangement • Disrupted activation
↓ Glucose ↓ Glycolysis ↓ TCR signalling and ↑ H+
↑ AMP:ATP ratio NFAT
• Disrupted activation • Suppressed effector
MDSC ↓ NFAT and Ca2+ signalling function Lactate
↓ Tryptophan • Suppressed effector ↓ IFNγ, IL-2, TNFα and
function cytotoxic killing
↓ IFNγ, IL-2, IL-17, Granz B • Altered signalling Adenosine
and cytotoxic killing ↑ Cyclic AMP and PKA
• Altered signalling • Altered differentiation • Kynurenine
Teff cell ↓ mTOR ↓ TH1 cell
↓ Glutamine • Ammonia
↑ AMPK ↑ Treg, Teff cell • Polyamines
• Altered miRNA activity exhaustion and Tscm cell
Leaky
↓ Viability, NOTCH • Altered phenotype
blood
↓ Arginine signalling and EZH2 and mobility R-2-HG
vessel
• Altered differentiation ↑ Anergy
↓ Teff cell ↓ Migration and
Fibroblast ↓ Oxygen ↑ Treg and Tmem cells proliferation ↑ K+
Dendritic
cell Macrophage

Fig. 2 | Metabolic derangements in the TME inhibit T cell function. The metabolic milieu of the tumour microenvironment
(TME) is a reflection of cancer metabolic programmes. Nutrient deprivation, hypoxia and toxic metabolites are conditions
within the TME that confront and influence T cell metabolism and function. The consequences of TME conditions on
immune cell responses can be predicted based on a growing literature of preclinical, translational and clinical studies.
AMPK, AMP kinase; EZH2, enhancer of zeste homologue 2; Granz B, granzyme B; IFNγ, interferon-​γ; MDSC, myeloid-​
derived suppressor cell; miRNA, microRNA; NFAT, nuclear factor of activated T cells; PKA, protein kinase A; R-2-​HG,
(R)-2-​hydroxyglutarate; TCR, T cell receptor; Teff, effector T; TH1, T helper 1; Tmem, memory T; Treg, regulatory T; Tscm, stem cell
memory T; TNF, tumour necrosis factor.

Vhl) actually favoured the formation of long-​lived effec­ CD8+ T cells activated in the presence of elevated lactate
tor memory cells in mouse vaccine models93. Other work and H+ compared with control media105.
has demonstrated that hypoxia induced the expression The accumulation of specific amino acids within
of the ectonucleotidases CD39 and CD73 on various tumours can also suppress the Teff cell response. Probably
cells in the TME94,95. These enzymes break down ATP the most well studied in this regard are the effects of
in the TME to adenosine. Adenosine is a ligand for the tryptophan metabolites, especially kynurenine, which is
A2A and A2B purinergic receptors, which are expressed generated through the activity of IDO1106. Functioning
on a large range of immune cells, and is broadly immu­ as an endogenous ligand for aryl hydrocarbon receptors
nosuppressive, inhibiting effector cell function and pro­ on T cells, kynurenine caused upregulation of the PD1
liferation of Teff cells96–101. Interestingly, supplemental co-​inhibitory pathway on activated CD8+ T cells in vitro
oxygen enhanced the antitumour immune response of compared with vehicle-​treated control107. This upregu­
T cells in mice by downregulating the adenosine signal­ lation of PD1 was also observed on tumour-​infiltrating
ling pathway102. The effect of hypoxia on antitumour CD8+ T cells in mouse models treated with exogenous
T cells is an evolving area of study. Further research will kynurenine compared with vehicle-​treated controls.
clearly benefit the field of immunotherapy, given both Tumour-​infiltrating CD8+ T cells from kynurenine
the prevalence of hypoxic regions in tumours as well treated tumour-​b earing mice produced less IFNγ
as the profound effects hypoxia can have on the adaptive and TNF.
immune response. Cancer cells have also been reported to suppress
T cell activity through release of the oncometabolite
Toxic metabolites. In addition to adenosine, many other (R)-2-​hydroxyglutarate (R-2-​HG). This metabolite
products generated from cancer cell metabolism influ­ can inhibit epigenetic dioxygenase enzymes, such as
ence infiltrating T cells (Figs 1 and 2). Elevated levels histone demethylases, leading to increased methyla­
of extracellular lactate and H+ in the TME can suppress tion and modified transcription. R-2-​HG produced
T cell proliferation, survival, cytotoxicity and cytokine by isocitrate dehydrogenase (IDH)-​mutant human gli­
production in in vitro studies of mouse and human oma was taken up by T cells in in vitro studies. R-2-​HG
CD8+ T cells103,104. The upregulation of the gene encoding interfered with proliferation, T cell receptor signalling,
the key Teff cell transcription factor NFAT was impaired NFAT activity and polyamine biosynthesis in activated
during in vitro activation of mouse CD8+ T cells in the human CD4+ and CD8+ T cells in vitro108. This was cor­
presence of high levels of lactate and H+ compared with roborated by the finding that R-2-​HG released into the
standard growth media104. In vivo mouse studies showed TME in IDH-​mutant glioma-​bearing mice inhibited
that mouse melanoma cells that have Ldha knocked complement-​mediated antitumour response as well as
down produced less lactate and were more responsive to T cell migration, proliferation and cytokine secretion109.
immune-​mediated tumour rejection than empty vector-​ These studies highlight the intricate interplay of can­
transfected control melanoma cells104. MAP kinase sig­ cer metabolites and immune function within the
nalling was also severely impaired in human effector TME (Fig. 2).

Nature Reviews | Cancer


Reviews

High levels of necrosis lead to increased levels of in vitro studies104,125,126. Lastly, elevated adenosine levels
potassium within the TME, which limits T cell effector within the TME can strongly suppress NK cell effector
function110. Mediated by reduced cytoplasmic levels of function and proliferation127.
acetyl-​CoA, this state induced epigenetic remodelling Other innate cells, macrophages and DCs also
of activated T cells, inducing a dysfunctional state of enact specific metabolic programmes upon activation.
Teff cells within the TME111. However, this dysfunctional Although early in vitro work using activation schemes
state was enriched with characteristics of T cell stemness. with specific cytokines classified macrophages into
In accord with the induction of a stem-​like state, ex vivo inflammatory (M1) or immunosuppressive (M2) pheno­
stimulation and expansion of Teff cells in high potassium types, there is poor evidence that these polarized phe­
produced T cells with improved in vivo persistence, notypes play distinct roles in vivo128,129. More recent work
multipotency and capacity for tumour clearance111. The has uncovered a spectrum of macrophage phenotypes
generation of T cell-​suppressive metabolites through characterized by distinct transcriptional states130. That
tumour necrosis and metabolic activity intrinsic to said, macrophages with inflammatory characteristics
the TME forms an important mechanism of tumour can play an important role in antitumour immunity19,
immune evasion. and in this regard it is instructive to examine what has
been established regarding the metabolic programming
Metabolism and the innate effector response. Because of in vitro derived ‘M1’ macrophages. Glucose metab­
NK cells are particularly adept at cell killing during olism is a vital aspect of the inflammatory phenotype
major histocompatibility complex class I (MHC-​I) down­ in macrophages. Upon activation, for example by a
regulation, a common evasion strategy of cancer cells, Toll-​like receptor agonist, these cells showed increased
they form a critical effector component of the innate expression of glycolytic genes, high levels of glucose
response. Metabolically, aerobic glycolysis and OXPHOS uptake, increased lactate production and upregulated
were upregulated after in vitro cytokine stimulation glutamine anaplerosis131. This metabolic reprogram­
(IL-12 and IL-15) of NK cells112. Interestingly, SREBP ming led to increased succinate levels, which increased
transcription factors were required for these cytokine-​ expression of the inflammatory cytokine IL-1β by sta­
induced metabolic changes during in vitro NK cell bilizing HIF-1 (ref.132). Inflammatory macrophages are
stimulation113. Pharmacologic inhibition of SREBP also particularly reliant on the PPP for the generation
activity suppressed metabolic reprogramming, cytokine of NADPH, with 13C-​glucose tracing studies confirming
production and cytotoxicity in vitro and curtailed anti­ increased routing of glucose though this pathway upon
tumour response in an adoptive NK cell mouse model. activating inflammatory phenotypes in culture133,134.
Interestingly, it has been reported that endogenous NADPH is necessary to produce high levels of ROS as
SREBP inhibitors, such as 27-​hydroxycholesterol, can part of an oxidative burst, a key effector mechanism for
be increased within the TME and thus may be a mech­ these cells132,135,136. Arginine is also a critical nutrient
anism of NK cell suppression112,114–118. Lung cancer pro­ in the function of pro-​inflammatory ‘M1’ macrophages
gression in mice and tumour-​associated transforming as they express high levels of inducible nitric oxide syn­
growth factor-​β (TGFβ) are correlated with increased thase (iNOS) compared with alternatively activated or
fructose-1,6-​bisphosphatase (FBP1) expression in ‘M2’ polarized macrophages in in vitro studies137. iNOS
tumour-​associated NK cells119. FBP1 is a key enzyme requires arginine to generate cytotoxic nitric oxide, an
in gluconeogenesis, which, when activated, strongly important pro-​inflammatory mediator of antitumour
suppressed glycolysis in NK cells, leading to dysfunction response138,139.
and diminished viability. Interestingly, pharmacologic Specific nutrient deficits within the TME, particularly
Major histocompatibility
complex
inhibition of FBP1 was sufficient to re-​establish gly­ glucose and arginine, can severely limit the metabolism
(MHC). MHC class I (MHC-​I) is colytic metabolism, as well as cytokine production and and related elaboration of effector programmes in these
expressed on all nucleated cytotoxicity in vitro, and improve antitumour response cells. Glucose limitation not only suppresses glycol­
cells, a molecular complex in adoptive cell therapy mouse models119. These stud­ ysis as a whole but can curtail PPP activity and TCA
presenting intracellular peptide
ies showed that rescuing NK function through FBP1 cycle function, thus limiting the generation of NADPH,
epitopes for CD8+ T cell
receptor recognition. Also inhibition was dependent on restoration of glucose ROS and succinate, all of which can severely limit M1
expressed on antigen-​ metabolism, as blocking glucose metabolism with macrophage function. Supporting this idea, the secre­
presenting cells, allowing initial 2-​deoxyglucose (2-​D G) prevented the rescue caused tion of pro-​inflammatory cytokines by macrophages
antigen-​specific activation of by FBP1 inhibition. 2-​DG by itself also led to NK cell was significantly reduced by glycolysis inhibition
cytotoxic CD8+ T cells. MHC-​II
is highly expressed on
dysfunction119, implying that inhibition of glucose with 2-​DG140.
antigen-​presenting cells for metabolism could have profound effects on NK cell DCs are an important class of antigen-​presenting
presenting antigenic epitopes antitumour response. Other metabolic derangements cells involved in the antitumour response. Intratumoural
for CD4+ T cell receptor within the TME are likely to affect NK cell function as DCs that are capable of antigen cross-​presentation have
recognition and activation.
well. For instance, low arginine levels can impair NK cell specifically emerged as a vital component of this
Antigen cross-​presentation proliferation and IFNγ production120,121, and hypoxia response22. Upon activation, DCs undergo maturation
The ability of antigen-​ can suppress cytolytic activity122–124. Human NK cell-​ allowing antigen processing and presentation to T cells.
presenting cells to process activating receptors, such as NKp46 and NKp30, are This response was coupled to a metabolic switch from
extracellular antigens and suppressed in response to hypoxia or low arginine in OXPHOS to aerobic glycolysis, mediated by HIF-1α
present them to CD8+ T cells
through major histo­
in vitro studies121,124. High lactate levels and associated in response to in vitro LPS activation141, and by the
compatibility complex low pH, as in the TME, also suppressed NK cell cyto­ PI3K–AKT pathway in response to Toll-​like receptor
class 1 presentation. toxicity, cytokine production and NFAT signalling in stimulation in vitro142. This switch to glycolysis and away

www.nature.com/nrc
Reviews

from OXPHOS during DC activation is critical for The unique metabolism of amino acids within
DC survival, production of stimulatory cytokines and the TME can also have a profound effect on Treg cells.
activation of T cells142. Interestingly, pharmacologic acti­ IDO1 activity can strongly promote Treg cell differen­
vation of AMPK, which promotes mitochondrial bio­ tiation in vitro, an effect that appears to be secondary
genesis and oxidative respiration143, was sufficient to to both tryptophan deficiency as well as the generation
block DC maturation in vitro142. Given this critical of downstream metabolites, such as kynurenine160,161.
dependence on aerobic glycolysis, glucose competition Kynurenine has been found to induce the generation of
in the TME may significantly suppress DC activation FOXP3-​expressing Treg cells by functioning as an endog­
and viability and thus limit the ability of DCs to foster enous ligand for aryl hydrocarbon receptors on T cells162.
an effective and persistent T cell response. Interestingly, many of the qualities of the TME that make
it inhospitable for Teff cells are either well tolerated by
The metabolism of cancer immune evasion Treg cells (elevated lactate and H+) or can induce Treg cell
Metabolism of adaptive immune suppression. Immuno­ responses (for example, accumulation of adenosine,
suppressive Treg cells preferentially rely on TCA cycle func­ kynurenine and hypoxia).
tion and mitochondrial respiration144,145. Although initial
studies demonstrating the dependence of Treg cells on FAO Metabolism of innate immune suppression. TAMs
did not account for off-​target effects of etomoxir, other can adopt phenotypes that are highly immunosup­
studies have shown that FAO does support OXPHOS pressive. Although there is a spectrum of TAM phe­
in Treg cells, although not as the sole pathway82,146,147. In notypes as mentioned above, it is useful to examine
contrast to Teff cells, Treg cells showed decreased glucose the metabolic programming of what heretofore had
uptake and expressed lower levels of GLUT1 in vitro144. been established as an ‘M2’ anti-​inflammatory mac­
Interestingly, although glycolysis did not appear to play rophage subset, characteristics of which are clearly evi­
a crucial role in Treg cell differentiation or a long-​lived dent within immunosuppressive TAMs. Like Treg cells,
phenotype, our laboratory has reported that a subset of M2 macrophages upregulate FAO and mitochondrial
highly active Treg cells, termed effector Treg cells, relied on respiration134,163. Although early studies demonstrating
the upregulation of glycolysis for optimal function148. As that FAO is obligatory in M2 macrophages did not take
such, Treg cells appear to be metabolically flexible, which into account the off-​target effects of the CPT1A inhibitor
may allow them to thrive within relatively harsh and heter­ etomoxir164, forced induction of FAO and mitochondrial
ogeneous conditions, such as the TME. To this end, it has biogenesis by overexpression of Pgc1α primes mac­
been reported that the Treg cell-​defining transcription fac­ rophages for an immunosuppressive phenotype and
tor, FOXP3, reprogrammes cellular metabolism through strongly suppresses the production of pro-​inflammatory
suppression of MYC favouring OXPHOS and NAD(H) cytokines163.
oxidation149. In conditions of low glucose and high lac­ M2 macrophages metabolize amino acids in a
tate, such as found in the TME, these adaptations allow distinct manner from inflammatory macrophages,
for a metabolic advantage of these immunosuppressive expressing high levels of arginase 1 (ARG1), which
cells, allowing Treg cells to resist lactate-​induced functional depletes arginine and generates polyamines that are
and proliferative suppression (unlike Teff cells) in vitro149. important mediators of wound healing but also highly
Glucose or glutamine deprivation (leading to reduced immunosuppressive165–168.
intracellular αKG) in media during in vitro skewing Another group of tumour-​associated immunosup­
experiments can alter CD4 differentiation and favour the pressive innate cells, MDSCs, appear to be highly met­
development of Treg cells150,151. abolically active. Both aerobic glycolysis and OXPHOS
Similar to Teff cells, Treg cell response to hypoxia is not were upregulated in tumour-​associated MDSCs com­
entirely clear. Hypoxia has been shown to promote pared with MDSCs in the periphery169. In another study,
cytokine-​mediated recruitment of Treg cells into the granulocytic MDSCs from the spleens of tumour-​bearing
tumour environment152. Other work has demonstrated mice also showed increases in both aerobic glycolysis
that FOXP3 transcript is actually increased in response to and OXPHOS compared with splenic neutrophils from
HIF-1α induction153–155. Also, adoptively transferred the same mice. Interestingly, MDSC expansion in vitro
Treg cell-​specific Hif1 knockout cells failed to migrate and accumulation in the TME in mouse breast cancer
into brain tumours in mouse models compared with models could be attenuated through blockade of glycol­
wild-​type controls, an effect that was also observed ysis with 2-​DG, likely by causing increased ROS levels
in dichloroacetate-​treated Treg cells, in which glycol­ in these cells170.
ysis is inhibited compared with vehicle-​treated control Hypoxic regions within tumours have been asso­
Treg cells156. Interestingly, hypoxia-​responsive adeno­ ciated with the accumulation of macrophages, where
sine signalling through the adenosine receptor A2A on they aid tumour development through the production
Treg cells induced proliferation and significantly stronger of angiogenesis factors, mitogenic factors and cytokines
immunoregulatory activity in mixed lymphocyte cul­ associated with tumour metastasis171–173. Furthermore,
ture experiments157,158. Conversely, several groups have hypoxia can promote the generation of immunosuppres­
reported that hypoxia-​induced HIF-1α can destabilize sive macrophage phenotypes174. Adenosine, which can
Treg cells, with reports demonstrating that hypoxia can be generated as a result of hypoxia, can trigger signalling
promote TH17 CD4+ T cells through direct HIF-1α through A2A and A2B receptors on macrophages, both
interactions with the cell subtype-​defining transcription of which augment the differentiation and functional
factors FOXP3 and RORγt, respectively52,53,159. capabilities of immunosuppressive macrophages as well

Nature Reviews | Cancer


Reviews

Table 2 | Metabolic inhibitors with potential for clinical translation


Pathway Target representative drugs representative refs
phase II/III
clinical trials
Glycolysis GLUT1 WZB117 NA 187

HK 2-​DG NA 50

3-​Bromopyruvate NA 188

PFKFB3 3-(3-​Pyridinyl)-1-(4-​pyridinyl)- NA 189

2-​propen-1-​one (3-​PO), PFK158


GAPDH Dimethyl fumarate (DMF) Phase IIa 190,191

Heptelidic acid NA 192

PDHK1 Dichloroacetate (DCA) Phase II 193,194

LDHA NCI-737, NCI-006 NA 195

FAO CPT1A Etomoxir NA 196

Perhexidine NA 197

CD36 ABT-511 Phase II 198,199

FAS ACC1 Firsocostat NA 200

Cholesterol esterification ACAT Pactimibe NA 201

Electron transport and/or AMPK Metformin Phase IIa 202–205

mitochondrial function
Glutaminolysisb GLS BPTES NA 206

CB-839 Phase IIa 207,208

Glutamine metabolism Glutamine DON NA 16

requiring enzymes
JHU083 NA 209

One-carbon metabolism SHMT2 RZ-2994 NA 44

One-carbon metabolism TS, DHFR, GARFT Pemetrexed Phase II and


a 192,210,211

phase IIIa
Methotrexate Phase IIa 212–214

TS 5-​Fluorouracil Phase IIIa 215–217

Reduction of ROS levels Antioxidant NAC Phase II 218,219

Increase of ROS levels GSH Menadione NA 220

Pentose phosphate pathway G6PD Polydatin NA 221

Hexosamine biosynthetic pathway PGM3 FR054 NA 222

Arginine pathway ARG1 CB-1158 Phase IIa 223–225

Arginine pathway Pegylated arginine ADI-​PEG20 NA 226

deiminase
IDO inhibitors IDO Epacadostat Phase IIa 227–229

Indoximod Phase IIa 227,230,231

Navoximod NA 227

R-2-​HG synthesis Mutant IDH1 FT-2102 Phase IIa 232–234

Adenosine pathway CD73 Oleclumab Phase IIa 98,235–237

BMS-986179 Phase IIa 98,235,238

NZV930, CPI-006 NA 98,235

Adenosine PBF-509 Phase IIa 98,239

receptor A2A
CPI-444 Phase IIa 98,240

AZD4635 Phase IIa 98,241

ACAT, acetyl-​CoA acetyltransferase; ACC1, acetyl-​CoA carboxylase; AMPK, AMP kinase; ARG1, arginase; CPT1A, carnitine palmitoyl
transferase 1; 2-​DG, 2-​deoxyglucose; DHFR, dihydrofolate reductase; DON, 6-​diazo-5-​oxo-​l-​norleucine; GAPDH, glyceraldehyde
3-​phosphate dehydrogenase; GARFT, glycinamide ribonucleotide formyltransferase; GLS, glutaminase; GLUT1, glucose transporter,
type 1; G6PD, glucose-6-​phosphate dehydrogenase; GSH, glutathione; HK, hexokinase; IDH1, isocitrate dehydrogenase; IDO,
indoleamine 2,3-​dioxygenase; LDHA, lactate dehydrogenase A; NA, not applicable; NAC, N-​acetyl cysteine; PDHK1, pyruvate
dehydrogenase kinase; PFKFB3, 6-​phosphofructo-2-​kinase/fructose-2,6-​biphosphatase; PGM3, phosphoglucomutase; R-2-​HG,
(R)-2-hydroxyglutarate; ROS, reactive oxygen species; SHMT2, serine hydroxymethyltransferase, mitochondrial; TS, thymidylate
synthase. aDenotes clinical trial in combination with established immunotherapy agents. bThe enzymatic catabolism of glutamine
to generate tricarboxylic acid cycle intermediates.

www.nature.com/nrc
Reviews

Box 2 | Adoptive T cell therapies enable highly flexible approaches to metabolic therapy through ex vivo culture conditioning or
genetic targeting of metabolic processes
T cells in such regimens can be genetically modified to express chimeric response50,74,111,179 (see the figure). Recent studies demonstrated that
antigen receptors (CARs) that recognize known tumour surface antigens forced expression of peroxisome proliferator-activated receptor-​γ
and trigger T cell receptor signalling in the absence of major histo­ co-​activator 1α (PGC1α), which promotes mito­chondrial biogenesis, in
compatibility complex presentation. T cells can be metabolically adoptively transferred CD8+ T cells resulted in superior intratumoural
conditioned through the use of chemical inhibitors or genetic editing metabolic and effector function57. Several groups have also reported the
during ex vivo expansion, through metabolically engineered media or importance of the co-​stimulatory receptor 4-1BB in positively conditioning
through pharmacologic treatment after T cell reinfusion. Many of the mitochondrial health and biogenesis for robust antitumour immunity245,246.
interventions discussed, including inhibiting glycolysis, glutamine These findings have been validated in the CAR T cell field, wherein the
metabolism with 6-​diazo-5-​oxo-​l-​norleucine (DON), AKT–mTOR addition of the 4-1BB receptor module has enhanced T cell persistence
signalling and potassium supplementation, have been used in ex vivo and increased therapeutic efficacy247–251. 2-​DG, 2-​deoxyglucose;
T cell expansion and led to improved T cell persistence and antitumour TCA, tricarboxylic acid.

Glucose Glutamine
CAR Pharmacologic
T cell metabolic
CAR Glycolysis infusion intervention

Glutamine
2-DG metabolism
TCA DON
cycle • Metabolic
Leukopheresis conditioning
• Activation,
Enhanced transduction
memory and
Rapalog phenotype ↑ K+ expansion
Tumour
CAR
T cell Enhanced
mTORC1 ↑ PGC1α antitumour
function
AKT
inhibitor AKT
VIII

as attenuate the cytokine release of pro-​inflammatory for tumour treatment, not only allowed for increased
macrophages in vitro96,175. Elevated lactic acid in cul­ survival of antitumour T cells but also for improved
ture has been shown to promote the M2 phenotype, cytokine production and cytotoxicity. Similar phe­
increasing ARG1 expression and polyamine-​dependent nomena have been reported in CD8 + Teff cells in
immunosuppression 176 . High glycolytic rates in response to AKT inhibition, glutamine blockade,
triple-​negative breast cancer were shown to promote hypoxia, arginine supplementation and potassium
MDSCs, whereas restricting glycolysis in these cancer supplementation 16,63,74,93,111,179. It may be possible to
cells inhibited granulocyte colony-​stimulating fac­ differentially affect cancer and the immune response.
tor and granulocyte–macrophage colony-​stimulating For example, acetate metabolism can rescue T cell
factor secretion from cancer cells and limited MDSC function in glucose-​restricted CD8+ Teff cells87. In addi­
development177. Interestingly, hypoxia skewed MDSCs tion, our group recently demonstrated the importance
towards an immunosuppressive, M2-​like TAM pheno­ of this pathway in maintaining metabolic homeosta­
type compared with normoxic MDSCs in vitro. This sis in CD8+ T cells undergoing glutamine blockade16.
occurred via HIF-1α mechanisms, as Hif1a knockout These findings could imply a generalizable therapeutic
MDSCs displayed increased tumour growth compared strategy, in that blocking the use of typical metabolic
with wild-​type MDSCs in mouse melanoma tumour fuels, such as glucose or glutamine, may render some
models178. cancers metabolically compromised, but may leave
antitumour T cells metabolically intact and functional
Exploiting differential metabolic plasticity given their ability to use alternative sources, such as ace­
Whereas the activation, proliferation and function of tate. Although specific metabolic interventions may be
Teff cells can be attenuated through inhibiting numerous introduced pharmacologically as adjuncts to checkpoint
metabolic pathways, other attributes, such as long-​term blockade (Table 2), these targets may be particularly
viability or effector function upon restimulation, may applicable to CAR T cell therapy, wherein manipulation
be enhanced. Although inhibiting glycolytic metab­ of metabolic pathways can be precisely defined through
olism with 2-​D G inhibited Teff cell generation, it also genetic means (Box 2). Future studies delineating the
conditioned T cells towards a long-​lived, memory-​like degree of metabolic flexibility possible within a given
phenotype50. Interestingly, blocking glycolysis during immune cell subset and functional capacity are clearly
ex vivo T cell activation and expansion, before reinfusion warranted.

Nature Reviews | Cancer


Reviews

Fatty acids Glucose 2-DG to enhance Glucose


memory phenotype
CD36 Glycolysis Cancer
blockade cell
CD36 Glycolysis
CD73
IDO DON to
Fatty acid enhance Glutamine
oxidation memory metabolism
Fatty acid
oxidation phenotype
blockade TCA TCA
cycle cycle
ETC
ARG1 IDO IDO inhibition CD73 blockade
MDSC Teff cell
A2AR
Metformin
Glucose Glucose
Glutamate
ARG1 blockade Glycolysis transamination Glycolysis
blockade
PGM3 blockade with AOA G6PD blockade
CD73
ARG1
Fatty acid Hexosamine Glutamate Hexosamine
oxidation pathway pathway
Fatty acid
oxidation Adenosine
blockade TCA IDO α-Ketoglutarate TCA A2AR
cycle cycle A2AR blockade
ETC ETC Fatty acid
oxidation
Macrophage blockade
Fatty acid
Treg cell oxidation
Metformin Metformin

Fig. 3 | Potential metabolic targets for enhancing immune response in cancer. Using small molecules, monoclonal
antibodies and genetic editing, metabolic processes can be targeted to either disable cancer and suppressive immune cell
metabolism or, conversely, engage and support effector cell metabolism. Metabolic processes in suppressive immune
populations and cancer cells can be targeted to directly decrease viability, as well as to disable metabolic pathways that
deplete nutrients (for example, arginase 1 (ARG1) and indoleamine 2,3-​dioxygenase (IDO)), lead to toxic metabolites
(for example, lactate and CD73) or induce metabolic control of effector cell populations (for example, mutant IDH1
generation of the oncometabolite (R)-2-​hydroxyglutarate (R-2-HG)). Metabolic interventions may also be able to induce
beneficial changes in effector populations, such as increasing longevity and antigen-​specific immunologic memory. A2AR,
adenosine receptor subtype A2A; AOA, amino-​oxyacetic acid; 2-​DG, 2-​deoxyglucose; DON, 6-​diazo-5-​oxo-​l-​norleucine;
ETC, electron transport chain; G6PD, glucose-6-​phosphate dehydrogenase; MDSC, myeloid-​derived suppressor cell;
PGM3, phosphoglucomutase; TCA, tricarboxylic acid; Teff, effector T; Treg, regulatory T.

Checkpoint blockade and immunometabolism. It is of models inhibited tumour growth and conditioned the
great interest to define both the metabolic consequences TME to be more hospitable for antitumour effector
of checkpoint therapy and the metabolic determinants of cells16. Also, metabolically reprogramming T cells to
response. Checkpoint signalling has been shown to make them more robust, long-​lasting memory cells
regulate metabolism in several studies. For example, might improve their response to checkpoint inhibitors.
PDL1 expression on cancer cells can drive Akt–mTOR This has been heralded by recent clinical trials combin­
activation and glycolysis in cancer cells, increasing glu­ ing the anti-​folate pemetrexed with anti-​PDL1 immune
cose uptake and augmenting competition with T cells checkpoint blockade183. In addition to having direct
for glucose14. CD155-​TIGIT signalling in T cells from antitumour effects, pemetrexed treatment enhanced
human gastric cancer tissue dampened glucose uptake, the metabolic fitness and effector function of antitumour
lactate production and expression of glycolytic enzymes CD8+ T cells, as well as induced immunologic cell death
GLUT1 and HK2180. Conversely, agonism of the co-​ of cancer cells to trigger the immune response.
stimulatory pathway GITR broadly increased T cell
metabolic activity and proliferation compared with Conclusion and perspective
isotype-​treated control T cells181. Lastly, in vitro PD1 and Although much of the foundation of immunometab­
CTLA4 signalling on activated human T cells suppressed olism has been informed by observations of cancer
metabolic pathways, such as aerobic glycolysis, that are metabolism, it is clear that there are distinct differences
associated with T cell activation182. To this end, the pros­ between cancer and immunologic metabolic repro­
pect of combining metabolic inhibitors (Table 2) with gramming. These differences provide opportunities
checkpoint inhibitors holds promise to enhance the effi­ to target metabolism as a means of enhancing the effi­
cacy of checkpoint blockade. Targeting tumour metab­ cacy of immunotherapy (Fig. 3). Such an approach can
olism by inhibiting glutamine metabolism in mouse be achieved through numerous different strategies.

www.nature.com/nrc
Reviews

These include targeting tumour metabolic programmes can be induced and co-​opted to benefit malignant pro­
to inhibit growth and alter the TME, targeting the gression. It has been reported that pancreatic stellate
metabolism of suppressive immune cells to inhibit cells can provide alanine to cancer cells and, thus, fuel
their function and targeting effector cell metabolism proliferation184, and bone marrow stromal cells have
to enhance tumour killing. Likewise, ex vivo pharma­ been reported to provide cysteine to promote survival
cologic or genetic reprogramming of T cell metabolic of chronic lymphocytic leukaemia cells185. In another
pathways prior to adoptive cellular therapy offers an report, ammonia from cancer cell glutamine metabo­
opportunity to dramatically engineer enhanced fea­ lism diffused through the TME and triggered autophagy
tures, which may include longevity or enhanced effector in cancer-​associated fibroblasts, which in turn provided
function (Box 2). protein breakdown products, such as glutamine itself,
Future work should begin to focus on the meta­ to further support cancer cell metabolism186. It will be
bolic interdependence of immune cells and cancer important to understand whether and by what mech­
cells within the TME. In addition to nutrient depletion anism immune-​evading cancers may be co-​opting the
and the generation of metabolites that can suppress metabolic machinery of immune cells and benefitting
the immune response at certain concentrations, can­ from their remarkable metabolic flexibility.
cer cells can engage in metabolic crosstalk with other
cells within the TME, wherein metabolic programmes Published online xx xx xxxx

1. Fox, C. J., Hammerman, P. S. & Thompson, C. B. 16. Leone, R. D. et al. Glutamine blockade induces 30. Gatza, E. et al. Manipulating the bioenergetics of
Fuel feeds function: energy metabolism and the T-​cell divergent metabolic programs to overcome tumor alloreactive T cells causes their selective apoptosis and
response. Nat. Rev. Immunol. 5, 844–852 (2005). immune evasion. Science 366, 1013–1021 (2019). arrests graft-​versus-​host disease. Sci. Transl Med. 3,
This paper reviews the critical determinants of This study demonstrates, using pharmacologic 67ra68 (2011).
metabolic reprogramming that occur during T cell glutamine blockade, the potential of leveraging the 31. Chang, C. H. et al. Posttranscriptional control of
activation, including the roles of co-​stimulatory ability of CD8+ T cells to use alternative metabolic T cell effector function by aerobic glycolysis. Cell 153,
signalling and growth factors, to meet increased pathways, including acetate metabolism and glucose 1239–1251 (2013).
bioenergetic demands required for pathogen anaplerosis, to enhance the antitumour response. This study shows that IFNγ translation in activated
response. 17. Lukey, M. J., Katt, W. P. & Cerione, R. A. Targeting T cells is dependent on the upregulation of aerobic
2. Andrejeva, G. & Rathmell, J. C. Similarities and amino acid metabolism for cancer therapy. Drug glycolysis. The study reports a novel mechanism
distinctions of cancer and immune metabolism in Discov. Today 22, 796–804 (2017). wherein GAPDH, in the absence of glycolysis-​driven
inflammation and tumors. Cell Metab. 26, 49–70 18. Ho, P. C. et al. Phosphoenolpyruvate is a metabolic NAD+, blocks IFNγ translation through a moonlighting
(2017). checkpoint of anti-​tumor T cell responses. Cell 162, role by binding to the 3′ untranslated region of
3. Bauer, D. E. et al. Cytokine stimulation of aerobic 1217–1228 (2015). IFNγ mRNA.
glycolysis in hematopoietic cells exceeds proliferative This study demonstrates the role of glucose 32. Sena, L. A. et al. Mitochondria are required for
demand. FASEB J. 18, 1303–1305 (2004). deprivation within the TME as an novel checkpoint antigen-​specific T cell activation through reactive
4. Kim, J. & DeBerardinis, R. J. Mechanisms and for T cell tumouricidal effector functions. Additionally, oxygen species signaling. Immunity 38, 225–236
implications of metabolic heterogeneity in cancer. the study shows a critical role for the glycolysis (2013).
Cell Metab. 30, 434–446 (2019). metabolite, phosphoenolpyruvate, in sustaining 33. Pollizzi, K. N. & Powell, J. D. Integrating canonical and
5. Weinberg, F. et al. Mitochondrial metabolism and Ca2+-​NFAT signalling in activated antitumour T cells. metabolic signalling programmes in the regulation of
ROS generation are essential for Kras-​mediated 19. Fridman, W. H., Zitvogel, L., Sautes-​Fridman, C. & T cell responses. Nat. Rev. Immunol. 14, 435–446
tumorigenicity. Proc. Natl Acad. Sci. USA 107, Kroemer, G. The immune contexture in cancer prognosis (2014).
8788–8793 (2010). and treatment. Nat. Rev. Clin. Oncol. 14, 717–734 (2017). 34. Finlay, D. K. et al. PDK1 regulation of mTOR and
6. Ma, E. H. et al. Metabolic profiling using stable 20. Tsou, P., Katayama, H., Ostrin, E. J. & Hanash, S. M. hypoxia-​inducible factor 1 integrate metabolism and
isotope tracing reveals distinct patterns of glucose The emerging role of B cells in tumor immunity. migration of CD8+ T cells. J. Exp. Med. 209,
utilization by physiologically activated CD8+ T cells. Cancer Res. 76, 5597–5601 (2016). 2441–2453 (2012).
Immunity 51, 856–870.e5 (2019). 21. Roberts, E. W. et al. Critical role for CD103+/CD141+ 35. Osthus, R. C. et al. Deregulation of glucose
7. Chen, P. H. et al. Metabolic diversity in human dendritic cells bearing CCR7 for tumor antigen transporter 1 and glycolytic gene expression by
non-​small cell lung cancer cells. Mol. Cell 76, trafficking and priming of T cell immunity in c-​Myc. J. Biol. Chem. 275, 21797–21800 (2000).
838–851.e5 (2019). melanoma. Cancer Cell 30, 324–336 (2016). 36. Patra, K. C. & Hay, N. The pentose phosphate pathway
8. Cascone, T. et al. Increased tumor glycolysis 22. Jansen, C. S. et al. An intra-​tumoral niche maintains and cancer. Trends Biochem. Sci. 39, 347–354 (2014).
characterizes immune resistance to adoptive T cell and differentiates stem-​like CD8 T cells. Nature 576, 37. Kidani, Y. et al. Sterol regulatory element-​binding
therapy. Cell Metab. 27, 977–987.e4 (2018). 465–470 (2019). proteins are essential for the metabolic programming
This study identifies elevated tumour glycolysis as 23. Pardoll, D. Cancer and the immune system: basic of effector T cells and adaptive immunity. Nat. Immunol.
a determinant of immune resistance in melanoma concepts and targets for intervention. Semin. Oncol. 14, 489–499 (2013).
in an adoptive cell therapy model. 42, 523–538 (2015). 38. Gorrini, C., Harris, I. S. & Mak, T. W. Modulation of
9. Renner, K. et al. Restricting glycolysis preserves T cell 24. Miller, J. F. & Sadelain, M. The journey from oxidative stress as an anticancer strategy. Nat. Rev.
effector functions and augments checkpoint therapy. discoveries in fundamental immunology to cancer Drug Discov. 12, 931–947 (2013).
Cell Rep. 29, 135–150.e9 (2019). immunotherapy. Cancer Cell 27, 439–449 (2015). 39. Hosios, A. M. & Vander Heiden, M. G. The redox
10. Kleffel, S. et al. Melanoma cell-​intrinsic PD-1 receptor 25. Becht, E., Giraldo, N. A., Dieu-​Nosjean, M. C., requirements of proliferating mammalian cells.
functions promote tumor growth. Cell 162, 1242–1256 Sautes-​Fridman, C. & Fridman, W. H. Cancer immune J. Biol. Chem. 293, 7490–7498 (2018).
(2015). contexture and immunotherapy. Curr. Opin. Immunol. 40. Przybytkowski, E. & Averill-​Bates, D. A. Correlation
11. Nunes-​Xavier, C. E. et al. Decreased expression of 39, 7–13 (2016). between glutathione and stimulation of the pentose
B7-​H3 reduces the glycolytic capacity and sensitizes 26. Patel, C. H., Leone, R. D., Horton, M. R. & Powell, J. D. phosphate cyclein situin Chinese hamster ovary cells
breast cancer cells to AKT/mTOR inhibitors. Targeting metabolism to regulate immune responses exposed to hydrogen peroxide. Arch. Biochem.
Oncotarget 7, 6891–6901 (2016). in autoimmunity and cancer. Nat. Rev. Drug Discov. Biophysics 325, 91–98 (1996).
12. Lim, S. et al. Immunoregulatory protein B7-​H3 18, 669–688 (2019). 41. Sena, L. A. & Chandel, N. S. Physiological roles of
reprograms glucose metabolism in cancer cells by 27. Menk, A. V. et al. Early TCR signaling induces rapid mitochondrial reactive oxygen species. Mol. Cell 48,
ROS-​mediated stabilization of HIF1α. Cancer Res. 76, aerobic glycolysis enabling distinct acute T cell effector 158–167 (2012).
2231 (2016). functions. Cell Rep. 22, 1509–1521 (2018). 42. Mak, T. W. et al. Glutathione primes T cell metabolism
13. Johnston, R. J. et al. VISTA is an acidic pH-​selective 28. Frauwirth, K. A. et al. The CD28 signaling pathway for inflammation. Immunity 46, 675–689 (2017).
ligand for PSGL-1. Nature 574, 565–570 (2019). regulates glucose metabolism. Immunity 16, 43. Swamy, M. et al. Glucose and glutamine fuel protein
14. Chang, C. H. et al. Metabolic competition in the tumor 769–777 (2002). O-​GlcNAcylation to control T cell self-​renewal and
microenvironment is a driver of cancer progression. This study shows that the upregulated glycolytic malignancy. Nat. Immunol. 17, 712–720 (2016).
Cell 162, 1229–1241 (2015). rate in activated T cells is dependent on CD28 This study identifies the regulation of protein
This study demonstrates that glucose consumption co-​stimulation acting through PI3K–AKT signalling. O-​GlcNAcylation through glucose and glutamine
by tumours can restrict the glycolytic capacity and 29. Wang, R. et al. The transcription factor Myc controls metabolism as a key controller of T cell clonal
IFNγ production of T cells, and that this nutrient metabolic reprogramming upon T lymphocyte expansion.
competition can be attenuated through checkpoint activation. Immunity 35, 871–882 (2011). 44. Ma, E. H. et al. Serine is an essential metabolite for
blockade with antibodies against PD1/PDL1 and This study demonstrates that acute metabolic effector T cell expansion. Cell Metab. 25, 345–357
CTLA4. reprogramming in activated T cells, including the (2017).
15. Sharma, N. S. et al. Targeting tumor-​intrinsic hexosamine upregulation of gylycolytic, pentose phosphate This study identifies extracellular serine as a key
biosynthesis sensitizes pancreatic cancer to anti-​PD1 and glutaminolysis pathways, is dependent on the immunometabolite that is required for optimal
therapy. J. Clin. Investigation 130, 451–465 (2020). activity of the MYC transcription factor. T cell expansion, in glucose-​replete conditions.

Nature Reviews | Cancer


Reviews

45. Martínez-Reyes, I. & Chandel, N. S. Mitochondrial TCA 68. Liu, M. et al. Targeting the IDO1 pathway in cancer: tumor growth in mice. J. Clin. Investigation 121,
cycle metabolites control physiology and disease. from bench to bedside. J. Hematol. Oncol. 11, 2371–2382 (2011).
Nat. Commun. 11, 102 (2020). 100–100 (2018). 95. Allard, B., Longhi, M. S., Robson, S. C. & Stagg, J.
46. Blagih, J. et al. The energy sensor AMPK regulates 69. Liu, H. et al. Increased expression of IDO associates The ectonucleotidases CD39 and CD73: novel
T cell metabolic adaptation and effector responses with poor postoperative clinical outcome of patients with checkpoint inhibitor targets. Immunol. Rev. 276,
in vivo. Immunity 42, 41–54 (2015). gastric adenocarcinoma. Sci. Rep. 6, 21319 (2016). 121–144 (2017).
47. Cham, C. M. & Gajewski, T. F. Glucose availability 70. Li, R. et al. IDO inhibits T-​cell function through 96. Ferrante, C. J. et al. The adenosine-​dependent
regulates IFN-​γ production and p70S6 kinase suppressing Vav1 expression and activation. angiogenic switch of macrophages to an M2-​like
activation in CD8+ effector T cells. J. Immunol. 174, Cancer Biol. Ther. 8, 1402–1408 (2009). phenotype is independent of interleukin-4 receptor α
4670–4677 (2005). 71. Godin-​Ethier, J., Hanafi, L. A., Piccirillo, C. A. & (IL-4Rα) signaling. Inflammation 36, 921–931
48. Cham, C. M., Driessens, G., O’Keefe, J. P. & Lapointe, R. Indoleamine 2,3-​dioxygenase expression (2013).
Gajewski, T. F. Glucose deprivation inhibits multiple in human cancers: clinical and immunologic 97. Leone, R. D., Lo, Y. C. & Powell, J. D. A2aR
key gene expression events and effector functions in perspectives. Clin. Cancer Res. 17, 6985–6991 (2011). antagonists: next generation checkpoint blockade for
CD8+ T cells. Eur. J. Immunol. 38, 2438–2450 (2008). 72. Cluntun, A. A., Lukey, M. J., Cerione, R. A. & cancer immunotherapy. Comput. Struct. Biotechnol. J.
49. Delgoffe, G. M. et al. The mTOR kinase differentially Locasale, J. W. Glutamine metabolism in cancer: 13, 265–272 (2015).
regulates effector and regulatory T cell lineage understanding the heterogeneity. Trends Cancer 3, 98. Leone, R. D. & Emens, L. A. Targeting adenosine for
commitment. Immunity 30, 832–844 (2009). 169–180 (2017). cancer immunotherapy. J. Immunother. Cancer 6, 57
50. Sukumar, M. et al. Inhibiting glycolytic metabolism 73. Cheng, T. et al. Pyruvate carboxylase is required (2018).
enhances CD8+ T cell memory and antitumor function. for glutamine-​independent growth of tumor cells. 99. Waickman, A. T. et al. Enhancement of tumor
J. Clin. investigation 123, 4479–4488 (2013). Proc. Natl Acad. Sci. USA 108, 8674–8679 (2011). immunotherapy by deletion of the A2A adenosine
This study shows that inhibiting glycolysis with 74. Nabe, S. et al. Reinforce the antitumor activity of CD8+ receptor. Cancer Immunol. Immunother. 61, 917–926
2-​DG during in vitro activation enhances the CD8+ T cells via glutamine restriction. Cancer Sci. 109, (2012).
T cell memory phenotype and leads to improved 3737–3750 (2018). 100. Ohta, A. et al. A2A adenosine receptor protects
antitumour activity after adoptive transfer. 75. Reid, M. A., Dai, Z. & Locasale, J. W. The impact of tumors from antitumor T cells. Proc. Natl Acad. Sci.
51. Araki, K. et al. mTOR regulates memory CD8 T-​cell cellular metabolism on chromatin dynamics and 103, 13132–13137 (2006).
differentiation. Nature 460, 108–112 (2009). epigenetics. Nat. Cell Biol. 19, 1298–1306 (2017). This study demonstrates that pharmacologic or
52. Shi, L. Z. et al. HIF1α-​dependent glycolytic pathway 76. Yang, W. et al. Potentiating the antitumour response genetic inhibition of adenosine-​A2AR receptor
orchestrates a metabolic checkpoint for the of CD8+ T cells by modulating cholesterol metabolism. signalling on T cells enhanced T cell-​mediated
differentiation of TH17 and Treg cells. J. Exp. Med. 208, Nature 531, 651–655 (2016). tumour regression and destruction of metastases.
1367–1376 (2011). This work demonstrates the potential to enhance 101. Leone, R. D. et al. Inhibition of the adenosine A2a
53. Dang, E. V. et al. Control of TH17/Treg balance by CD8+ T cell effector function, proliferation and receptor modulates expression of T cell coinhibitory
hypoxia-​inducible factor 1. Cell 146, 772–784 antitumour activity through pharmacologic and receptors and improves effector function for enhanced
(2011). genetic blockade of the cholesterol-​esterifying checkpoint blockade and ACT in murine cancer
54. Zhao, E. et al. Cancer mediates effector T cell enzyme ACAT1, which allows for increased models. Cancer Immunol. Immunother. 67, 1271–1284
dysfunction by targeting microRNAs and EZH2 via cholesterol content in plasma membranes, (2018).
glycolysis restriction. Nat. Immunol. 17, 95–103 facilitating T cell receptor clustering. 102. Hatfield, S. M. et al. Immunological mechanisms of
(2016). 77. Ma, X. et al. Cholesterol induces CD8+ T cell the antitumor effects of supplemental oxygenation.
55. van Bruggen, J. A. C. et al. Chronic lymphocytic exhaustion in the tumor microenvironment. Sci. Transl Med. 7, 277ra230 (2015).
leukemia cells impair mitochondrial fitness in CD8+ Cell Metab. 30, 143–156.e5 (2019). This report shows the potential of respiratory
T cells and impede CAR T-​cell efficacy. Blood 134, 78. van der Windt, G. J. et al. Mitochondrial respiratory hyperoxygenation in mice to suppress adenosine-​
44–58 (2019). capacity is a critical regulator of CD8+ T cell memory A2AR signalling in antitumour CD8+ T cells
56. Siska, P. J. et al. Mitochondrial dysregulation and development. Immunity 36, 68–78 (2012). and NK cells, while also weakening Treg cell
glycolytic insufficiency functionally impair CD8 T cells 79. van der Windt, G. J. et al. CD8 memory T cells have immunosuppression, leading to enhanced tumour
infiltrating human renal cell carcinoma. JCI Insight 2, a bioenergetic advantage that underlies their rapid regression and survival.
e93411 (2017). recall ability. Proc. Natl Acad. Sci. USA 110, 103. Fischer, K. et al. Inhibitory effect of tumor cell–derived
57. Scharping, N. E. et al. The tumor microenvironment 14336–14341 (2013). lactic acid on human T cells. Blood 109, 3812–3819
represses T cell mitochondrial biogenesis to drive 80. Pollizzi, K. N. et al. mTORC1 and mTORC2 selectively (2007).
intratumoral T cell metabolic insufficiency and regulate CD8+ T cell differentiation. J. Clin. This study demonstrates the suppressive effect of
dysfunction. Immunity 45, 374–388 (2016). Investigation 125, 2090–2108 (2015). lactic acid on CD8+ Teff cell proliferation, cytokine
58. Nii, T. et al. Molecular events involved in up-​regulating 81. Pollizzi, K. N. et al. Asymmetric inheritance of mTORC1 production and cytotoxicity.
human Na+-​independent neutral amino acid kinase activity during division dictates CD8+ T cell 104. Brand, A. et al. LDHA-​associated lactic acid production
transporter LAT1 during T-​cell activation. Biochem. J. differentiation. Nat. Immunol. 17, 704–711 (2016). blunts tumor immunosurveillance by T and NK cells.
358, 693–704 (2001). 82. Raud, B. et al. Etomoxir actions on regulatory and Cell Metab. 24, 657–671 (2016).
59. Carr, E. L. et al. Glutamine uptake and metabolism memory T cells are independent of Cpt1a-​mediated 105. Mendler, A. N. et al. Tumor lactic acidosis suppresses
are coordinately regulated by ERK/MAPK during fatty acid oxidation. Cell Metab. 28, 504–515.e7 CTL function by inhibition of p38 and JNK/c-​Jun
T lymphocyte activation. J. Immunol. 185, 1037–1044 (2018). activation. Int. J. Cancer 131, 633–640 (2012).
(2010). 83. Pan, Y. et al. Survival of tissue-​resident memory T cells 106. Labadie, B. W., Bao, R. & Luke, J. J. Reimagining IDO
60. Nakaya, M. et al. Inflammatory T cell responses rely requires exogenous lipid uptake and metabolism. Pathway inhibition in cancer immunotherapy via
on amino acid transporter ASCT2 facilitation of Nature 543, 252–256 (2017). downstream focus on the tryptophan–kynurenine–aryl
glutamine uptake and mTORC1 kinase activation. 84. Tyrakis, P. A. et al. S-2-​hydroxyglutarate regulates CD8+ hydrocarbon axis. Clin. Cancer Res. 25, 1462–1471
Immunity 40, 692–705 (2014). T-​lymphocyte fate. Nature 540, 236–241 (2016). (2019).
61. Ren, W. et al. Amino-​acid transporters in T-​cell 85. Chisolm, D. A. et al. CCCTC-​binding factor translates 107. Liu, Y. et al. Tumor-​repopulating cells induce PD-1
activation and differentiation. Cell Death Dis. 8, interleukin 2- and α-​ketoglutarate-​sensitive metabolic expression in CD8+ T cells by transferring kynurenine
e2655 (2017). changes in T cells into context-​dependent gene and AhR activation. Cancer Cell 33, 480–494.e7
62. Sinclair, L. V. et al. Control of amino-​acid transport programs. Immunity 47, 251–267.e7 (2017). (2018).
by antigen receptors coordinates the metabolic 86. Schug, Z. T., Vande Voorde, J. & Gottlieb, E. The 108. Bunse, L. et al. Suppression of antitumor T cell immunity
reprogramming essential for T cell differentiation. metabolic fate of acetate in cancer. Nat. Rev. Cancer by the oncometabolite (R)-2-​hydroxyglutarate. Nat. Med.
Nat. Immunol. 14, 500–508 (2013). 16, 708–717 (2016). 24, 1192–1203 (2018).
63. Geiger, R. et al. l-​Arginine modulates T cell 87. Qiu, J. et al. Acetate promotes T cell effector function This study shows that tumour-​derived R-2-​HG is
metabolism and enhances survival and anti-​tumor during glucose restriction. Cell Rep. 27, 2063–2074. taken up by T cells and interferes with NFAT
activity. Cell 167, 829–842.e813 (2016). e5 (2019). transcriptional activity and polyamine biosynthesis
This study shows that elevating l-​arginine levels 88. Balmer, M. L. et al. Memory CD8+ T cells require to suppress T cell activity. The antitumour immune
during T cell activation promotes long-​lived central increased concentrations of acetate induced by stress response is enhanced by inhibition of enzymatic
memory-​like cells with enhanced antitumour for optimal function. Immunity 44, 1312–1324 (2016). activity of IDH1.
activity in a mouse model. 89. Petrova, V., Annicchiarico-​Petruzzelli, M., Melino, G. 109. Zhang, L. et al. d-2-​Hydroxyglutarate is an intercellular
64. Srivastava, M. K., Sinha, P., Clements, V. K., & Amelio, I. The hypoxic tumour microenvironment. mediator in IDH-​mutant gliomas inhibiting complement
Rodriguez, P. & Ostrand-​Rosenberg, S. Myeloid-derived Oncogenesis 7, 10 (2018). and T cells. Clin. Cancer Res. 24, 5381–5391 (2018).
suppressor cells inhibit T-​cell activation by depleting 90. McKeown, S. R. Defining normoxia, physoxia and 110. Eil, R. et al. Ionic immune suppression within the
cystine and cysteine. Cancer Res. 70, 68–77 (2010). hypoxia in tumours—implications for treatment tumour microenvironment limits T cell effector
65. Munn, D. H. et al. Prevention of allogeneic fetal response. Br. J. Radiol. 87, 20130676 (2014). function. Nature 537, 539–543 (2016).
rejection by tryptophan catabolism. Science 281, 91. Caldwell, C. C. et al. Differential effects of This study shows that necrosis-​related potassium
1191–1193 (1998). physiologically relevant hypoxic conditions on release occurs in human and mouse tumours and
66. Munn, D. H. et al. Inhibition of T cell proliferation by T lymphocyte development and effector functions. that elevated extracellular potassium concentration
macrophage tryptophan catabolism. J. Exp. Med. J. Immunol. 167, 6140–6149 (2001). leads to suppression of Akt–mTOR signalling and
189, 1363–1372 (1999). 92. Ohta, A. et al. In vivo T cell activation in lymphoid T cell function in a PP2A-​dependent manner.
This paper uncovers the ability of macrophages tissues is inhibited in the oxygen-​poor 111. Vodnala, S. K. et al. T cell stemness and dysfunction
to induce cell cycle arrest in Teff cells through the microenvironment. Front. Immunol. 2, 27 (2011). in tumors are triggered by a common mechanism.
IDO-​dependent catabolism of tryptophan. 93. Phan, A. T. et al. Constitutive glycolytic metabolism Science 363, eaau0135 (2019).
67. Munn, D. H. et al. GCN2 kinase in T cells mediates supports CD8+ T cell effector memory differentiation 112. Keppel, M. P., Saucier, N., Mah, A. Y., Vogel, T. P.
proliferative arrest and anergy induction in response during viral infection. Immunity 45, 1024–1037 (2016). & Cooper, M. A. Activation-​specific metabolic
to indoleamine 2,3-​dioxygenase. Immunity 22, 94. Wang, L. et al. CD73 has distinct roles in requirements for NK cell IFN-​γ production. J. Immunol.
633–642 (2005). nonhematopoietic and hematopoietic cells to promote 194, 1954–1962 (2015).

www.nature.com/nrc
Reviews

113. Assmann, N. et al. Srebp-​controlled glucose respiratory inhibition in tumor target cells. J. Exp. 161. Chen, W., Liang, X., Peterson, A. J., Munn, D. H.
metabolism is essential for NK cell functional Med. 169, 1543–1555 (1989). & Blazar, B. R. The indoleamine 2,3-dioxygenase
responses. Nat. Immunol. 18, 1197–1206 (2017). 139. El-​Gayar, S., Thuring-​Nahler, H., Pfeilschifter, J., pathway is essential for human plasmacytoid dendritic
114. Wu, Q. et al. 27-Hydroxycholesterol promotes Rollinghoff, M. & Bogdan, C. Translational control of cell-​induced adaptive T regulatory cell generation.
cell-​autonomous, ER-​positive breast cancer growth. inducible nitric oxide synthase by IL-13 and arginine J. Immunol. 181, 5396–5404 (2008).
Cell Rep. 5, 637–645 (2013). availability in inflammatory macrophages. J. Immunol. 162. Mezrich, J. D. et al. An interaction between kynurenine
115. Rossin, D. et al. Increased production of 171, 4561–4568 (2003). and the aryl hydrocarbon receptor can generate
27-hydroxycholesterol in human colorectal cancer 140. Moon, J. S. et al. mTORC1-​induced HK1-​dependent regulatory T cells. J. Immunol. 185, 3190 (2010).
advanced stage: possible contribution to cancer cell glycolysis regulates NLRP3 inflammasome activation. 163. Vats, D. et al. Oxidative metabolism and PGC-1β
survival and infiltration. Free. Radic. Biol. Med. 136, Cell Rep. 12, 102–115 (2015). attenuate macrophage-​mediated inflammation.
35–44 (2019). 141. Jantsch, J. et al. Hypoxia and hypoxia-​inducible factor- Cell Metab. 4, 13–24 (2006).
116. Li, D., Long, W., Huang, R., Chen, Y. & Xia, M. 1α modulate lipopolysaccharide-​induced dendritic cell 164. Divakaruni, A. S. et al. Etomoxir inhibits macrophage
27-Hydroxycholesterol inhibits sterol regulatory activation and function. J. Immunol. 180, 4697–4705 polarization by disrupting CoA homeostasis. Cell Metab.
element-​binding protein 1 activation and hepatic lipid (2008). 28, 490–503.e7 (2018).
accumulation in mice. Obesity 26, 713–722 (2018). 142. Krawczyk, C. M. et al. Toll-​like receptor-​induced 165. Shearer, J. D., Richards, J. R., Mills, C. D. &
117. Guo, F. et al. Upregulation of 24(R/S),25-​ changes in glycolytic metabolism regulate dendritic Caldwell, M. D. Differential regulation of macrophage
epoxycholesterol and 27-​hydroxycholesterol cell activation. Blood 115, 4742–4749 (2010). arginine metabolism: a proposed role in wound
suppresses the proliferation and migration of gastric 143. Hardie, D. G. AMP-​activated/SNF1 protein kinases: healing. Am. J. Physiol. 272, E181–E190 (1997).
cancer cells. Biochem. Biophys. Res. Commun. 504, conserved guardians of cellular energy. Nat. Rev. Mol. 166. Ye, C. et al. Targeting ornithine decarboxylase by
892–898 (2018). Cell Biol. 8, 774–785 (2007). α-​difluoromethylornithine inhibits tumor growth
118. Sun, J. C., Ma, A. & Lanier, L. L. Cutting edge: 144. Michalek, R. D. et al. Cutting edge: distinct glycolytic by impairing myeloid-​derived suppressor cells.
IL-15-independent NK cell response to mouse and lipid oxidative metabolic programs are essential J. Immunol. 196, 915 (2016).
cytomegalovirus infection. J. Immunol. 183, for effector and regulatory CD4+ T cell subsets. 167. Mills, C. D., Shearer, J., Evans, R. & Caldwell, M. D.
2911–2914 (2009). J. Immunol. 186, 3299–3303 (2011). Macrophage arginine metabolism and the inhibition or
119. Cong, J. et al. Dysfunction of natural killer cells by 145. Gerriets, V. A. et al. Metabolic programming and stimulation of cancer. J. Immunol. 149, 2709–2714
FBP1-​induced inhibition of glycolysis during lung cancer PDHK1 control CD4+ T cell subsets and inflammation. (1992).
progression. Cell Metab. 28, 243–255.e5 (2018). J. Clin. Investigation 125, 194–207 (2015). 168. Hayes, C. S. et al. Polyamine-​blocking therapy reverses
120. Oberlies, J. et al. Regulation of NK cell function by 146. Pacella, I. et al. Fatty acid metabolism complements immunosuppression in the tumor microenvironment.
human granulocyte arginase. J. Immunol. 182, glycolysis in the selective regulatory T cell expansion Cancer Immunol. Res. 2, 274–285 (2014).
5259–5267 (2009). during tumor growth. Proc. Natl Acad. Sci. USA 115, 169. Hossain, F. et al. Inhibition of fatty acid oxidation
121. Lamas, B. et al. Altered functions of natural killer cells E6546–e6555 (2018). modulates immunosuppressive functions of myeloid-​
in response to l-​arginine availability. Cell Immunol. 147. Cluxton, D., Petrasca, A., Moran, B. & Fletcher, J. M. derived suppressor cells and enhances cancer therapies.
280, 182–190 (2012). Differential regulation of human Treg and TH17 cells by Cancer Immunol. Res. 3, 1236–1247 (2015).
122. Sarkar, S. et al. Hypoxia induced impairment of NK fatty acid synthesis and glycolysis. Front. Immunol. 10, 170. Jian, S. L. et al. Glycolysis regulates the expansion
cell cytotoxicity against multiple myeloma can be 115 (2019). of myeloid-​derived suppressor cells in tumor-​bearing
overcome by IL-2 activation of the NK cells. PLoS One 148. Sun, I. H. et al. mTOR complex 1 signaling regulates hosts through prevention of ROS-​mediated apoptosis.
8, e64835 (2013). the generation and function of central and effector Cell Death Dis. 8, e2779 (2017).
123. Loftus, R. M. et al. Amino acid-​dependent cMyc Foxp3+ regulatory T cells. J. Immunol. 201, 481–492 171. Casazza, A. et al. Impeding macrophage entry into
expression is essential for NK cell metabolic and (2018). hypoxic tumor areas by Sema3A/Nrp1 signaling
functional responses in mice. Nat. Commun. 9, 2341 149. Angelin, A. et al. Foxp3 reprograms T cell metabolism blockade inhibits angiogenesis and restores antitumor
(2018). to function in low-​glucose, high-​lactate environments. immunity. Cancer Cell 24, 695–709 (2013).
124. Balsamo, M. et al. Hypoxia downregulates the Cell Metab. 25, 1282–1293.e87 (2017). 172. Murdoch, C. & Lewis, C. E. Macrophage migration
expression of activating receptors involved in This study shows that the Treg cell-​defining and gene expression in response to tumor hypoxia.
NK-​cell-​mediated target cell killing without affecting transcription factor, FOXP3, suppresses MYC Int. J. Cancer 117, 701–708 (2005).
ADCC. Eur. J. Immunol. 43, 2756–2764 (2013). activity and glycolysis, while increasing OXPHOS, 173. Henze, A. T. & Mazzone, M. The impact of hypoxia on
125. Husain, Z., Huang, Y., Seth, P. & Sukhatme, V. P. allowing Treg cells to resist lactate-​mediated tumor-​associated macrophages. J. Clin. Investigation
Tumor-​derived lactate modifies antitumor immune suppression in low-​glucose environments such 126, 3672–3679 (2016).
response: effect on myeloid-​derived suppressor cells as the TME. 174. Tripathi, C. et al. Macrophages are recruited to
and NK cells. J. Immunol. 191, 1486–1495 (2013). 150. Araujo, L., Khim, P., Mkhikian, H., Mortales, C. L. hypoxic tumor areas and acquire a pro-​angiogenic
126. Harmon, C. et al. Lactate-​mediated acidification of & Demetriou, M. Glycolysis and glutaminolysis M2-​polarized phenotype via hypoxic cancer cell
tumor microenvironment induces apoptosis of cooperatively control T cell function by limiting derived cytokines Oncostatin M and Eotaxin.
liver-​resident NK cells in colorectal liver metastasis. metabolite supply to N-​glycosylation. eLife 6, e21330 Oncotarget 5, 5350–5368 (2014).
Cancer Immunol. Res. 7, 335–346 (2019). (2017). 175. Csóka, B. et al. Adenosine promotes alternative
127. Young, A. et al. A2AR adenosine signaling suppresses 151. Klysz, D. et al. Glutamine-​dependent α-​ketoglutarate macrophage activation via A2A and A2B receptors.
natural killer cell maturation in the tumor micro­ production regulates the balance between T helper 1 FASEB J. 26, 376–386 (2012).
environment. Cancer Res. 78, 1003–1016 (2018). cell and regulatory T cell generation. Sci. Signal. 8, 176. Colegio, O. R. et al. Functional polarization of
128. Orecchioni, M., Ghosheh, Y., Pramod, A. B. & Ley, K. ra97 (2015). tumour-​associated macrophages by tumour-​derived
Macrophage polarization: different gene signatures in 152. Facciabene, A. et al. Tumour hypoxia promotes lactic acid. Nature 513, 559–563 (2014).
M1(LPS+) vs. classically and M2(LPS–) vs. alternatively tolerance and angiogenesis via CCL28 and Treg cells. This study demonstrates that lactic acid from
activated macrophages. Front. Immunol. 10, 1084 Nature 475, 226–230 (2011). tumour cells induces an M2-​like phenotype in
(2019). 153. Clambey, E. T. et al. Hypoxia-​inducible factor- macrophages with upregulation of vascular
129. Martinez, F. O. & Gordon, S. The M1 and M2 paradigm 1α-​dependent induction of FoxP3 drives regulatory endothelial growth factor and ARG1.
of macrophage activation: time for reassessment. T-​cell abundance and function during inflammatory 177. Li, W. et al. Aerobic glycolysis controls myeloid-​derived
F1000Prime Rep. 6, 13 (2014). hypoxia of the mucosa. Proc. Natl Acad. Sci. USA 109, suppressor cells and tumor immunity via a specific
130. Aras, S. & Zaidi, M. R. TAMeless traitors: macrophages E2784–E2793 (2012). CEBPB isoform in triple-​negative breast cancer.
in cancer progression and metastasis. Br. J. Cancer 154. Westendorf, A. M. et al. Hypoxia enhances Cell Metab. 28, 87–103.e6 (2018).
117, 1583–1591 (2017). immunosuppression by inhibiting CD4+ effector T cell 178. Corzo, C. A. et al. HIF-1α regulates function and
131. Freemerman, A. J. et al. Metabolic reprogramming of function and promoting Treg activity. Cell Physiol. differentiation of myeloid-​derived suppressor cells in
macrophages: glucose transporter 1 (GLUT1)-​mediated Biochem. 41, 1271–1284 (2017). the tumor microenvironment. J. Exp. Med. 207,
glucose metabolism drives a proinflammatory 155. Ben-​Shoshan, J., Maysel-​Auslender, S., Mor, A., 2439–2453 (2010).
phenotype. J. Biol. Chem. 289, 7884–7896 (2014). Keren, G. & George, J. Hypoxia controls CD4+CD25+ 179. Crompton, J. G. et al. Akt inhibition enhances
132. Tannahill, G. M. et al. Succinate is an inflammatory regulatory T-​cell homeostasis via hypoxia-​inducible expansion of potent tumor-​specific lymphocytes
signal that induces IL-1β through HIF-1α. Nature 496, factor-1α. Eur. J. Immunol. 38, 2412–2418 (2008). with memory cell characteristics. Cancer Res. 75,
238–242 (2013). 156. Miska, J. et al. HIF-1α is a metabolic switch between 296–305 (2015).
133. Haschemi, A. et al. The sedoheptulose kinase CARKL glycolytic-​driven migration and oxidative This work shows that harvested tumour infiltrating
directs macrophage polarization through control of phosphorylation-​driven immunosuppression of Tregs in lymphocytes expanded in the presence of an
glucose metabolism. Cell Metab. 15, 813–826 (2012). glioblastoma. Cell Rep. 27, 226–237.e4 (2019). inhibitor of AKT show enhanced transcriptional,
134. Jha, A. K. et al. Network integration of parallel 157. Ohta, A. & Sitkovsky, M. Extracellular adenosine-​ metabolic and functional properties of Tmem cells,
metabolic and transcriptional data reveals metabolic mediated modulation of regulatory T cells. Front. imbuing these T cells with enhanced in vivo
modules that regulate macrophage polarization. Immunol. 5, 304 (2014). persistence and augmented antitumor activity
Immunity 42, 419–430 (2015). 158. Ohta, A. et al. The development and immunosuppressive in mouse models.
135. Iles, K. E. & Forman, H. J. Macrophage signaling and functions of CD4+CD25+FoxP3+ regulatory T cells are 180. He, W. et al. CD155T/TIGIT signaling regulates CD8+
respiratory burst. Immunol. Res. 26, 95–105 (2002). under influence of the adenosine-​A2A adenosine T-​cell metabolism and promotes tumor progression in
136. Viola, A., Munari, F., Sanchez-​Rodriguez, R., Scolaro, T. receptor pathway. Front. Immunol. 3, 190 (2012). human gastric cancer. Cancer Res. 77, 6375–6388
& Castegna, A. The metabolic signature of macrophage 159. Lee, J. H., Elly, C., Park, Y. & Liu, Y. C. E3 ubiquitin (2017).
responses. Front. Immunol. 10, 1462 (2019). ligase VHL regulates hypoxia-​inducible factor-1α to 181. Sabharwal, S. S. et al. GITR agonism enhances cellular
137. Zajac, E. et al. Angiogenic capacity of M1- and maintain regulatory T cell stability and suppressive metabolism to support CD8+ T-​cell proliferation and
M2-​polarized macrophages is determined by the capacity. Immunity 42, 1062–1074 (2015). effector cytokine production in a mouse tumor model.
levels of TIMP-1 complexed with their secreted 160. Fallarino, F. et al. The combined effects of tryptophan Cancer Immunol. Res. 6, 1199 (2018).
proMMP-9. Blood 122, 4054–4067 (2013). starvation and tryptophan catabolites down-​regulate 182. Parry, R. V. et al. CTLA-4 and PD-1 receptors inhibit
138. Stuehr, D. J. & Nathan, C. F. Nitric oxide. A T cell receptor ζ-​chain and induce a regulatory phenotype T-​cell activation by distinct mechanisms. Mol. Cell Biol.
macrophage product responsible for cytostasis and in naive T cells. J. Immunol. 176, 6752–6761 (2006). 25, 9543–9553 (2005).

Nature Reviews | Cancer


Reviews

183. Schaer, D. A. et al. The folate pathway inhibitor 206. Xiang, Y. et al. Targeted inhibition of tumor-​specific 235. Li, X. et al. Navigating metabolic pathways to enhance
pemetrexed pleiotropically enhances effects of cancer glutaminase diminishes cell-​autonomous tumorigenesis. antitumour immunity and immunotherapy. Nat. Rev.
immunotherapy. Clin. Cancer Res. 25, 7175–7188 J. Clin. Investigation 125, 2293–2306 (2015). Clin. Oncol. 16, 425–441 (2019).
(2019). 207. Gross, M. I. et al. Antitumor activity of the glutaminase 236. US National Library of Medicine. ClinicalTrials.gov
This study demonstrates that pemetrexed inhibitor CB-839 in triple-​negative breast cancer. https://clinicaltrials.gov/ct2/show/NCT03742102
augments antitumour immunity in combination with Mol. Cancer Therapeutics 13, 890–901 (2014). (2020).
anti-​PDL1 checkpoint blockade in mouse models, 208. US National Library of Medicine. ClinicalTrials.gov 237. US National Library of Medicine. ClinicalTrials.gov
in part by enhancing CD8+ T cell metabolic health https://clinicaltrials.gov/ct2/show/NCT04265534 https://clinicaltrials.gov/ct2/show/NCT04262388
through stimulating mitochondrial biogenesis with (2020). (2020).
subsequent increased T cell infiltration and activation. 209. Lee, C. F. et al. Preventing allograft rejection by targeting 238. US National Library of Medicine. ClinicalTrials.gov
184. Sousa, C. M. et al. Pancreatic stellate cells support immune metabolism. Cell Rep. 13, 760–770 (2015). https://clinicaltrials.gov/ct2/show/NCT02754141
tumour metabolism through autophagic alanine 210. Adjei, A. A. Pemetrexed (ALIMTA), A. Novel (2020).
secretion. Nature 536, 479–483 (2016). multitargeted antineoplastic agent. Clin. Cancer Res. 239. US National Library of Medicine. ClinicalTrials.gov
185. Zhang, W. et al. Stromal control of cystine metabolism 10, 4276s (2004). https://clinicaltrials.gov/ct2/show/NCT02403193
promotes cancer cell survival in chronic lymphocytic 211. US National Library of Medicine. ClinicalTrials.gov (2020).
leukaemia. Nat. Cell Biol. 14, 276–286 (2012). https://clinicaltrials.gov/ct2/show/NCT03793179 240. US National Library of Medicine. ClinicalTrials.gov
186. Ko, Y.-H. et al. Glutamine fuels a vicious cycle of (2020). https://clinicaltrials.gov/ct2/show/NCT03337698
autophagy in the tumor stroma and oxidative 212. Chabner, B. A., Myers, C. E., Coleman, C. N. & (2020).
mitochondrial metabolism in epithelial cancer cells: Johns, D. G. The clinical pharmacology of 241. US National Library of Medicine. ClinicalTrials.gov
implications for preventing chemotherapy resistance. antineoplastic agents (first of two parts). N. Engl. https://clinicaltrials.gov/ct2/show/NCT03381274
Cancer Biol. Ther. 12, 1085–1097 (2011). J. Med. 292, 1107–1113 (1975). (2020).
187. Zhang, Z. et al. Differential glucose requirement in skin 213. US National Library of Medicine. ClinicalTrials.gov 242. Vander Heiden, M. G. & DeBerardinis, R. J.
homeostasis and injury identifies a therapeutic target https://clinicaltrials.gov/ct2/show/NCT00989352 Understanding the intersections between metabolism
for psoriasis. Nat. Med. 24, 617–627 (2018). (2020). and cancer biology. Cell 168, 657–669 (2017).
188. Okano, T. et al. 3-Bromopyruvate ameliorate 214. US National Library of Medicine. ClinicalTrials.gov 243. Saeidi, A. et al. T-​cell exhaustion in chronic infections:
autoimmune arthritis by modulating TH17/Treg cell https://clinicaltrials.gov/ct2/show/NCT03643276 reversing the state of exhaustion and reinvigorating
differentiation and suppressing dendritic cell (2020). optimal protective immune responses. Front. Immunol.
activation. Sci. Rep. 7, 42412 (2017). 215. Longley, D. B., Harkin, D. P. & Johnston, P. G. 9, 2569 (2018).
189. Telang, S. et al. Discovery of a PFKFB3 inhibitor for 5-Fluorouracil: mechanisms of action and clinical 244. Bengsch, B. et al. Bioenergetic insufficiencies due to
phase I trial testing that synergizes with the B-​Raf strategies. Nat. Rev. Cancer 3, 330–338 (2003). metabolic alterations regulated by the inhibitory
inhibitor vemurafenib. Cancer Metab. 2, P14–P14 216. US National Library of Medicine. ClinicalTrials.gov receptor PD-1 are an early driver of CD8+ T cell
(2014). https://clinicaltrials.gov/ct2/show/NCT02997228 exhaustion. Immunity 45, 358–373 (2016).
190. Kornberg, M. D. et al. Dimethyl fumarate targets (2020). This paper reports the suppression of glycolysis,
GAPDH and aerobic glycolysis to modulate immunity. 217. US National Library of Medicine. ClinicalTrials.gov mitochondrial dysfunction and mitochondrial
Science 360, 449–453 (2018). https://clinicaltrials.gov/ct2/show/NCT03777813 biogenesis in CD8+ T cells through PD1 signalling
191. US National Library of Medicine. ClinicalTrials.gov (2020). leading to an exhausted T cell phenotype.
https://clinicaltrials.gov/ct2/show/NCT02546440 218. Suwannaroj, S., Lagoo, A., Keisler, D. & McMurray, R. W. 245. Teijeira, A. et al. Mitochondrial morphological
(2020). Antioxidants suppress mortality in the female NZB × and functional reprogramming following CD137
192. Liberti, M. V. et al. A predictive model for selective NZW F1 mouse model of systemic lupus (4-1BB) costimulation. Cancer Immunol. Res. 6,
targeting of the Warburg effect through GAPDH erythematosus (SLE). Lupus 10, 258–265 (2001). 798 (2018).
inhibition with a natural product. Cell Metab. 26, 219. US National Library of Medicine. ClinicalTrials.gov 246. Menk, A. V. et al. 4-1BB costimulation induces T cell
648–659.e8 (2017). https://clinicaltrials.gov/ct2/show/NCT00003346 mitochondrial function and biogenesis enabling cancer
193. Dunbar, E. M. et al. Phase 1 trial of dichloroacetate (2020). immunotherapeutic responses. J. Exp. Med. 215,
(DCA) in adults with recurrent malignant brain tumors. 220. Yang, Z. et al. Restoring oxidant signaling suppresses 1091 (2018).
Invest. N. Drugs 32, 452–464 (2014). proarthritogenic T cell effector functions in rheumatoid 247. Zhao, Z. et al. Structural design of engineered
194. US National Library of Medicine. ClinicalTrials.gov arthritis. Sci. Transl Med. 8, 331ra338 (2016). costimulation determines tumor rejection kinetics and
https://clinicaltrials.gov/ct2/show/NCT01386632 221. Mele, L. et al. A new inhibitor of glucose-6-phosphate persistence of CAR T cells. Cancer Cell 28, 415–428
(2020). dehydrogenase blocks pentose phosphate pathway (2015).
195. Yeung, C. et al. Targeting glycolysis through inhibition and suppresses malignant proliferation and metastasis 248. Milone, M. C. et al. Chimeric receptors containing
of lactate dehydrogenase impairs tumor growth in in vivo. Cell Death Dis. 9, 572 (2018). CD137 signal transduction domains mediate enhanced
preclinical models of Ewing sarcoma. Cancer Res. 79, 222. Ricciardiello, F. et al. Inhibition of the hexosamine survival of T cells and increased antileukemic efficacy
5060–5073 (2019). biosynthetic pathway by targeting PGM3 causes in vivo. Mol. Ther. 17, 1453–1464 (2009).
196. Holubarsch, ChristianJ. F. et al. A double-​blind breast cancer growth arrest and apoptosis. Cell Death 249. Imai, C. et al. Chimeric receptors with 4-1BB signaling
randomized multicentre clinical trial to evaluate Dis. 9, 377 (2018). capacity provoke potent cytotoxicity against acute
the efficacy and safety of two doses of etomoxir in 223. Steggerda, S. M. et al. Inhibition of arginase by lymphoblastic leukemia. Leukemia 18, 676–684
comparison with placebo in patients with moderate CB-1158 blocks myeloid cell-​mediated immune (2004).
congestive heart failure: the ERGO (Etomoxir for the suppression in the tumor microenvironment. 250. Finney, H. M., Akbar, A. N. & Lawson, A. D. Activation
Recovery of Glucose Oxidation) study. Clin. Sci. 113, J. Immunother. Cancer 5, 101 (2017). of resting human primary T cells with chimeric receptors:
205–212 (2007). 224. US National Library of Medicine. ClinicalTrials.gov costimulation from CD28, inducible costimulator,
197. Senanayake, E. L. et al. Multicentre double-​blind https://clinicaltrials.gov/ct2/show/NCT03361228 CD134, and CD137 in series with signals from the
randomized controlled trial of perhexiline as a metabolic (2020). TCRζ chain. J. Immunol. 172, 104–113 (2004).
modulator to augment myocardial protection in patients 225. US National Library of Medicine. ClinicalTrials.gov 251. Davila, M. L. et al. Efficacy and toxicity management
with left ventricular hypertrophy undergoing cardiac https://clinicaltrials.gov/ct2/show/NCT02903914 of 19-28z CAR T cell therapy in B cell acute
surgery. Eur. J. Cardiothorac. Surg. 48, 354–362 (2020). lymphoblastic leukemia. Sci. Transl Med. 6, 224ra225
(2015). 226. Przystal, J. M. et al. Efficacy of arginine depletion (2014).
198. US National Library of Medicine. ClinicalTrials.gov by ADI-​PEG20 in an intracranial model of GBM.
https://clinicaltrials.gov/ct2/show/NCT00602199 Cell Death Dis. 9, 1192 (2018). Author contributions
(2020). 227. Prendergast, G. C., Malachowski, W. P., The authors contributed equally to all aspects of the article.
199. US National Library of Medicine. ClinicalTrials.gov DuHadaway, J. B. & Muller, A. J. Discovery of IDO1
https://clinicaltrials.gov/ct2/show/NCT00061646 inhibitors: from bench to bedside. Cancer Res. 77, Competing interests
(2020). 6795–6811 (2017). J.D.P. is a scientific founder, a paid consultant and has equity
200. Alkhouri, N., Lawitz, E., Noureddin, M., DeFronzo, R. & 228. US National Library of Medicine. ClinicalTrials.gov in Dracen Pharmaceuticals. Technology arising in part from
Shulman, G. I. GS-0976 (Firsocostat): an investigational https://clinicaltrials.gov/ct2/show/NCT03493945 the studies described herein was patented by Johns Hopkins
liver-​directed acetyl-​CoA carboxylase (ACC) inhibitor for (2020). U n i ve rs i t y a n d s u b s e q u e n t l y l i c e n s e d to D ra c e n
the treatment of non-​alcoholic steatohepatitis (NASH). 229. US National Library of Medicine. ClinicalTrials.gov Pharmaceuticals (JHU083 is currently labelled as DRP-083).
Expert. Opin. Investig. Drugs 29, 135–141 (2020). https://clinicaltrials.gov/ct2/show/NCT04231864 R.D.L. and J.D.P. are inventors for pending patent application
201. Meuwese, M. C. et al. ACAT inhibition and progression (2020). no. PCT/US16/44829 submitted by Johns Hopkins University
of carotid atherosclerosis in patients with familial 230. US National Library of Medicine. ClinicalTrials.gov that covers the use of glutamine analogues, such as JHU083
hypercholesterolemia: the CAPTIVATE randomized https://clinicaltrials.gov/ct2/show/NCT04049669 (DRP-083), for cancer immunotherapy. J.D.P has been a paid
trial. Jama 301, 1131–1139 (2009). (2020). consultant for Corvus Pharmaceuticals and has equity in the
202. Scharping, N. E., Menk, A. V., Whetstone, R. D., 231. US National Library of Medicine. ClinicalTrials.gov company.
Zeng, X. & Delgoffe, G. M. Efficacy of PD-1 blockade Is https://clinicaltrials.gov/ct2/show/NCT02073123 (2020).
potentiated by metformin-​induced reduction of tumor 232. Megias-​Vericat, J. E., Ballesta-​Lopez, O., Barragan, E. Peer review information
hypoxia. Cancer immunology Res. 5, 9–16 (2017). & Montesinos, P. IDH1-​mutated relapsed or refractory Nature Reviews Cancer thanks N. Chandel, J. Fan and the
203. US National Library of Medicine. ClinicalTrials.gov AML: current challenges and future prospects. other, anonymous, reviewer(s) for their contribution to
https://clinicaltrials.gov/ct2/show/NCT03048500 Blood Lymphat. Cancer 9, 19–32 (2019). the peer review of this work.
(2020). 233. US National Library of Medicine. ClinicalTrials.gov
204. US National Library of Medicine. ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT03684811 Publisher’s note
https://clinicaltrials.gov/ct2/show/NCT03800602 (2020). (2020). Springer Nature remains neutral with regard to jurisdictional
205. US National Library of Medicine. ClinicalTrials.gov 234. US National Library of Medicine. ClinicalTrials.gov claims in published maps and institutional affiliations.
https://clinicaltrials.gov/ct2/show/NCT03994744 https://clinicaltrials.gov/ct2/show/NCT02719574
(2020). (2020). © Springer Nature Limited 2020

www.nature.com/nrc

You might also like