You are on page 1of 13

Leading Edge

Review

Understanding the Intersections


between Metabolism and Cancer Biology
Matthew G. Vander Heiden1,2,* and Ralph J. DeBerardinis3,*
1The Koch Institute for Integrative Cancer Research and Department of Biology, Massachusetts Institute of Technology, Cambridge,

MA 02139, USA
2Dana–Farber Cancer Institute, Boston, MA 02115, USA
3Children’s Medical Center Research Institute, Department of Pediatrics, and Eugene McDermott Center for Human Growth and

Development, The University of Texas Southwestern Medical Center, Dallas, TX 75390


*Correspondence: mvh@mit.edu (M.G.V.H.), ralph.deberardinis@utsouthwestern.edu (R.J.D.)
http://dx.doi.org/10.1016/j.cell.2016.12.039

Transformed cells adapt metabolism to support tumor initiation and progression. Specific meta-
bolic activities can participate directly in the process of transformation or support the biological
processes that enable tumor growth. Exploiting cancer metabolism for clinical benefit requires
defining the pathways that are limiting for cancer progression and understanding the context spec-
ificity of metabolic preferences and liabilities in malignant cells. Progress toward answering these
questions is providing new insight into cancer biology and can guide the more effective targeting of
metabolism to help patients.

Introduction to oxidize pyruvate (Curi et al., 1988), consistent with preferen-


Changes in cell metabolism can contribute to transformation and tial conversion of glucose to lactate without loss of respira-
tumor progression. Metabolic phenotypes can also be exploited tion. This distinction is important because intermediates
to image tumors, provide prognostic information, and treat generated by the tricarboxylic acid (TCA) cycle are precursors
cancer. Thus, understanding cancer metabolism has implica- for lipids, amino acids, and nucleotides. These precursors com-
tions for understanding basic cancer pathophysiology and for plement precursor metabolites from glycolysis and other path-
clinical oncology. ways and are necessary to support proliferation (DeBerardinis
et al., 2007).
Cell-Autonomous Reprogramming of Cancer Fuels besides glucose also contribute to core metabolic
Metabolism functions of cancer cells: energy formation, biomass assimila-
Many recent cancer metabolism studies have focused on the tion, and redox control. Glutamine is a prominent example
cell-autonomous effects of cancer-promoting mutations. This (Altman et al., 2016); however, recent work has revealed that
has uncovered new principles in metabolic regulation and in a diversity of nutrients and pathways support these func-
crosstalk between signaling and metabolic networks. No tions. The expanding metabolic repertoire of cancer cells
pathway has received more attention than aerobic glycolysis has been reviewed extensively, with acetate and other fatty
(the Warburg effect) (Vander Heiden et al., 2009). This phenom- acids, lactate, branched chain amino acids, serine, and
enon involves the propensity for proliferating cells, including glycine representing some of the nutrients that are needed to
cancer cells, to take up glucose and secrete the carbon as fuel different cancers (DeBerardinis and Chandel, 2016; White,
lactate even when oxygen is present. Principles governing glyco- 2013).
lytic regulation in cancer cells have been extensively reviewed The growing list of cancer fuels contrasts with prevailing views
(Cairns and Mak, 2016). from just a few years ago, which suggested that oncogenic
Warburg interpreted tumor lactate secretion as an indication signaling imposed a rigid dependence on specific nutrients (Van-
that oxidative metabolism (i.e., respiration) was damaged. How- der Heiden et al., 2009). The current picture is more logical when
ever, numerous studies, including Warburg’s original work, fail one considers that cancer cells must compete for fuels in a
to demonstrate defective respiration as a general feature of crowded, nutrient-limited tissue environment. The ability to use
malignant cells (Koppenol et al., 2011). Instead, respiration and a panoply of fuels is advantageous, with some cancer cells
other mitochondrial activities are required for tumor growth even relying on autophagic degradation or scavenging of macro-
(DeBerardinis and Chandel, 2016). Furthermore, in non-trans- molecules (Commisso et al., 2013; Guo et al., 2016; Palm
formed cells the Warburg effect is a reversible phenomenon et al., 2015).
tethered to proliferation, indicating that it reflects proliferation- This Review builds on the increasingly sophisticated under-
associated changes in metabolism rather than a unique feature standing of cancer metabolism derived from work over the last
of malignancy. decade. Our goal is to convey emerging paradigms and ques-
Proliferating cells tend to express glucose transporters and tions in order to guide future research so that altered metabolism
glycolytic enzymes out of proportion to the machinery required can be exploited to improve patient care.

Cell 168, February 9, 2017 ª 2017 Elsevier Inc. 657


2009). D-2HG accumulates to high levels in IDH mutant tumors
and interferes with the function of several aKG-dependent diox-
ygenases, including the prolyl hydroxylases that target HIF-a
subunits for degradation and epigenetic enzymes that regulate
the methylation status of histones and DNA (Losman and Kaelin,
2013). This interferes with expression of genes required for
differentiation.
SDH and FH catalyze sequential reactions in the TCA cycle
(Figure 2). Both enzymes act as tumor suppressors in familial
cancer syndromes where patients inherit one loss-of-function
allele and lose expression of the other allele in tumors (Baysal
et al., 2000; Tomlinson et al., 2002). Consequently, tumors
accumulate high levels of succinate and/or fumarate. Like
D-2HG, these dicarboxylic acids interfere with dioxygenase
function (Sciacovelli et al., 2016; Selak et al., 2005; Xiao
et al., 2012).
In mouse models, mutations in these metabolic enzymes do
not transform cells on their own but cooperate with other muta-
tions to promote neoplasia. Loss of FH in the kidney leads to cyst
formation, and hyperproliferation results when p21 is also
deleted (Zheng et al., 2015). Idh2 mutations synergize with onco-
Figure 1. Classification of Reprogrammed Metabolic Activities genic Kras to produce intrahepatic cholangiocarcinoma (Saha
Some enzyme mutations result in perturbed metabolic activities or metabolite et al., 2014). IDH mutations increase the number of hematopoi-
levels that contribute to cancer initiation (transforming activities). Mutations in
etic progenitor cells (Sasaki et al., 2012) and cooperate with
oncogenes and tumor-suppressor genes transform cells, in part, by activating
proliferative signaling and/or by inducing broad changes in gene expression increased HoxA9 and Meis1a expression or with mutant Flt3 or
that favor cell proliferation, both of which induce metabolic reprogramming. Nras to cause AML (Chen et al., 2013; Kats et al., 2014).
Some metabolic alterations enable cancer progression, whereas others are An important concept emerged from these studies. In all
neutral and not required for cancer cell proliferation or survival. Enzyme mu-
tations resulting in transforming activities may also induce metabolic network cases, an accumulated metabolite inhibits enzymes using aKG
changes that are enabling or neutral for tumor progression. as a co-substrate to promote tumorigenesis (Kaelin and
McKnight, 2013). Thus, a unifying feature of D-2HG, succinate,
and fumarate as ‘‘oncometabolites’’ is causing non-metabolic
Functional Classification of Reprogrammed Metabolic effects that promote transformation in receptive contexts. Inter-
Activities in Cancer Cells estingly, modestly elevated levels of the L-enantiomer of 2HG
Not all reprogrammed metabolic activities contribute equally to might also contribute to malignancy (Shim et al., 2014). Given
cancer. With many metabolic activities under oncogenic control, the importance of dioxygenases in regulating epigenetics, fluctu-
categorizing them based on whether they are transforming, ations in metabolite levels may also play functional roles in
enabling, or neutral can clarify the role of each activity in cancer normal development. Consistent with this idea, extreme abnor-
biology and predict how it might be exploited in basic research malities in metabolite levels observed in monogenic metabolic
and clinical oncology (Figure 1). diseases, such as deficiency in the dehydrogenases converting
Transforming Activities L- and D-2HG to aKG, are associated with developmental abnor-
These activities directly contribute to cell transformation, and malities (Erez and DeBerardinis, 2015).
blocking them might prevent tumorigenesis in susceptible pa- Enabling Activities
tients or antagonize disease progression. At present, only a These activities are altered in cancer cells but are not involved in
handful of metabolic activities can be considered as transform- transformation. They carry out conventional metabolic tasks
ing based on genetic evidence. These include metabolic such as supporting energetics, generating macromolecules,
perturbations caused by enzyme mutations that either arise and maintaining redox state and are required for tumor progres-
somatically in a recurrent fashion and/or are inherited in the sion (Figure 1). In many cases they are effectors of oncogenes
germlines of patients with heritable cancer predisposition syn- and tumor-suppressor genes. For example, oncogene expres-
dromes. Three examples of these types of mutations have sion is sufficient to enhance glucose uptake in fibroblasts (Flier
been intensely studied: mutations in the genes encoding isoci- et al., 1987). Transcriptional targets of c-MYC include metabolite
trate dehydrogenases-1 and -2 (IDH1, IDH2); mutations in transporters and enzymes required for glucose metabolism, glu-
components of the succinate dehydrogenase (SDH) complex; taminolysis, and biosynthesis, and inhibiting these activities sup-
and mutations in fumarate hydratase (FH) (Losman and Kae- presses growth of c-MYC-driven tumors (Altman et al., 2016).
lin, 2013). Oncogenic KRAS also regulates nutrient acquisition, macromo-
Somatic mutations in IDH1 and IDH2 occur in several tumor lecular synthesis, and redox homeostasis, and inhibiting these
types (Losman and Kaelin, 2013). These monoallelic mutations pathways suppresses oncogenic KRAS-driven tumor growth
generate an enzyme with neomorphic ability to convert a-keto- (White, 2013). Activation of mTORC1, an essential component
glutarate (aKG) to (D)-2-hydroxyglutarate (D-2HG) (Dang et al., of growth factor signaling networks, enables protein, lipid, and

658 Cell 168, February 9, 2017


Figure 2. Nutrient Availability and the Meta-
bolic Network Both Influence Metabolic
Phenotypes
Both nutrient availability and metabolic network
configuration affect how cells use metabolism to
produce ATP, generate macromolecules, and
regulate redox state. Key reactions in central
carbon metabolism are shown, including how the
TCA cycle and the electron transport chain are
involved in purine and pyrimidine synthesis. Some
of the reactions catalyzed by enzymes and me-
tabolites discussed in this Review are also shown
for reference. PKM2, pyruvate kinase M2; LDH,
lactate dehydrogenase; PDH, pyruvate dehydro-
genase; PDHK, pyruvate dehydrogenase kinase;
PC, pyruvate carboxylase; IDH, isocitrate dehy-
drogenase; SDH, succinate dehydrogenase; FH,
fumarate hydratase; GLS, glutaminase; DHODH,
dihydroorotate dehydrogenase; Ox PPP, oxidative
pentose phosphate pathway; Non-Ox PPP, non-
oxidative pentose phosphate pathway; aKG,
a-ketoglutarate; OAA, oxaloacetate; ROS, reactive
oxygen species.

the glycolytic enzyme pyruvate kinase


that is expressed in most cancers
(Figure 2) and is regulated by oncogenic
signaling (Israelsen and Vander Heiden,
2015). PKM2 expression is not required
for growth of breast tumors, liver tumors,
leukemia, or xenograft tumors from
various human cancer cell lines in mice
(Dayton et al., 2016; Israelsen and Vander
Heiden, 2015). However, studies of PKM2
still provide insight into cancer metabolic
regulation. PKM2 loss promotes cancer
in some mouse models, and character-
izing these models has argued that how
glucose is metabolized can influence tu-
mor progression (Dayton et al., 2016; Isra-
elsen et al., 2013). In fact, activating
PKM2 can slow growth of some tumors,
but the ability of tumors to lose pyruvate
nucleic acid synthesis through a variety of mechanisms (Howell kinase expression argues against PKM2 being a good therapeu-
et al., 2013). Genetic suppression of metabolic activities may tic target for most cancers (Israelsen and Vander Heiden, 2015).
also enable tumor growth. For example, somatic deletion of
genes involved in arginine synthesis allows aspartate to be What Products of Metabolism are Limiting for
diverted toward nucleotide synthesis, thereby supporting cell Proliferation?
proliferation (Rabinovich et al., 2015). Most existing and investi- One approach to identify enabling activities is to determine
gational metabolic therapies are directed against enabling activ- which aspects of metabolism are limiting for cancer cell prolifer-
ities (Table 1). ation. Targeting activities that supply limiting materials for prolif-
Neutral Activities eration is therapeutically attractive, especially if the pathways
Many metabolic features of cancer cells are dispensable for tu- used are less important in normal proliferative tissues. Although
mor growth (Figure 1). In a given context, these activities are several metabolic products have been proposed as critical out-
predicted to be poor therapeutic targets. Fluctuating nutrient puts of cancer metabolism, which are rate limiting for prolifera-
access may cause activities to be required in some contexts tion remains controversial.
and dispensable in others. Thus, confidently classifying an activ- Is ATP Limiting for Proliferating Cells?
ity as neutral is challenging and requires definitive proof that loss ATP plays a critical role in all cells to enable otherwise unfavor-
of the activity does not impair tumor progression. able processes. Aerobic glycolysis has been historically inter-
Pyruvate kinase M2 (PKM2) provides an example of an preted as a shift from oxidative phosphorylation toward
enzyme dispensable for tumor growth. PKM2 is an isoform of fermentation to generate ATP (Gatenby and Gillies, 2004; Vander

Cell 168, February 9, 2017 659


Table 1. Drugs Targeting Metabolic Activities to Treat Malignancies or Suppress Immune Cell Proliferation
DRUG TARGET(S) EXAMPLE INDICATIONS
Nucleotide Metabolism
Methotrexate dihydrofolate reductase (DHFR) lymphoma
breast cancer
choriocarcinoma
osteosarcoma
Pemetrexed DHFR lung cancer
glycinamide ribonucleotide formyl- mesothelioma
transferase (GARFT)
thymidylate synthase
5-Fluorouracil thymidylate synthase colon cancer
pancreatic cancer
breast cancer
Gemcitabine ribonucleotide reductase pancreatic, lung cancer
Cytarabine DNA synthesis leukemia
Fludarabine lymphoma
Hydroxyurea ribonucleotide reductase acute myeloid leukemia (AML)
Leflunomide DHODH rheumatoid arthritis
psoriatic arthritis
Azathioprine DNA synthesis immunosuppression for organ transplant

Amino Acid Metabolism


L-Asparginase asparagine depletion acute lymphoblastic leukemia (ALL)

Select Drugs in Trials


AG120, AG221, AG881, IDH305, IDH1 and/or IDH2 clinical trials for AML, MDS, and solid tumors
BAY1436032, FT-2102 with mutant IDH1 or IDH2
Epacadostat indoleamine 2,3-dioxygenase-1 (IDO1) clinical trials for chronic lymphocytic leukemia
(CLL), ovarian cancer, melanoma, other cancers
CB-839 glutaminase clinical trials for AML, ALL, solid tumors
TVB-2640 Fatty-acid synthase clinical trials for solid tumors
PEG-BCT-100 arginine depletion clinical trials for hepatocellular carcinoma and
(ADI-PEG20) other hematological and solid tumors including
AEB-1102 those with ASS1 deficiency
The cell target(s) for several drugs as well as some approved indications are shown. Some drugs targeting metabolism in clinical trials are also shown.

Heiden et al., 2009). This route of ATP production yields less ATP sufficiently high to limit electron transport, and substrates are
per mole of glucose but allows fast ATP generation even in low available in excess of the demand for ATP synthesis. Finally,
oxygen (Koppenol et al., 2011). Energy is derived from ATP hy- increasing ATP consumption can promote proliferation (Fang
drolysis to ADP (or AMP) and thus depends on the ATP/ADP et al., 2010), providing further evidence that the ability to
(or ATP/AMP) ratio. Therefore, cells must continuously oxidize generate ATP is greater than what is needed by prolifer-
nutrients to regenerate ATP to maintain homeostasis. Although ating cells.
many proliferative processes also require ATP hydrolysis, the Is NADPH Limiting for Proliferating Cells?
additional ATP needed for proliferation is small relative to the re- Cells store energy as reduced carbon in carbohydrates and
quirements for homeostasis (Vander Heiden et al., 2009). Thus, lipids, and the anabolic reactions used to synthesize these mate-
although producing ATP is critical for survival, it may not be rials requires electrons from NADPH (Figure 3), leading to the hy-
limiting for proliferation. pothesis that NADPH may be limiting for proliferation (Vander
Little evidence supports ATP limitation as a reason cells use Heiden et al., 2009). However, the fact that some cancers,
aerobic glycolysis. First, many proliferating cells use aerobic such as clear cell renal cell carcinoma, accumulate lipids (Qiu
glycolysis regardless of nutrient and oxygen availability (Vander et al., 2015) argues that these cancers have sufficient NADPH
Heiden et al., 2009). Second, cancer cells retain the capacity to produce lipids in excess of those needed for cell membranes.
to increase fermentation when respiration is inhibited and also Most macromolecular biosynthesis occurs in the cytosol where
can increase respiration when mitochondria are uncoupled several enzymes can produce NADPH, including malic enzyme 1,
(Andrzejewski et al., 2014; Birsoy et al., 2014). The existence of IDH1, and folate metabolism; however, the oxidative pentose
‘‘spare respiratory capacity’’ argues that the ATP/ADP ratio is phosphate pathway (oxPPP) is a major source of cytosolic

660 Cell 168, February 9, 2017


Figure 3. Role of Redox Cofactors in Energy
Generation and Biosynthesis
Cells rely on nutrient oxidation to generate ATP
(energy), either through glycolysis or via NADH
generation to fuel oxidative phosphorylation.
Generating biomass can involve either nutrient
reduction or nutrient oxidation. Continued nutrient
oxidation requires cycling of NADH back to NAD+,
which necessitates transfer of electrons to an
electron acceptor such as oxygen.

of TCA cycle carbon in biomass (Hosios


et al., 2016). Aspartate, asparagine, pro-
line, and glutamate can all be produced
from TCA cycle intermediates, and cells
synthesize these amino acids from gluta-
mine even when they are present in the
environment (Hosios et al., 2016). Indeed,
synthesis of both aspartate (Birsoy et al.,
2015; Sullivan et al., 2015) and proline
(Loayza-Puch et al., 2016) can be limiting
under some conditions.
Nucleotide Synthesis Can Be
Limiting for Proliferation
Supplying nucleotide bases can rescue
proliferation of glutamine-deprived cells
NAPDH (Fan et al., 2014; Lewis et al., 2014) (Figure 2). Despite and cells with high pyruvate kinase activity, suggesting that
this, most ribose is produced via the non-oxidative pentose nucleotide base production is important downstream of major
phosphate pathway, which does not produce NAPDH (Boros nutrient pathways (Gaglio et al., 2009; Lunt et al., 2015).
et al., 2000; Ying et al., 2012). A major regulator of oxPPP flux Acquiring deoxyribonucleotide bases is limiting in other cancer
is the NADP+/NADPH ratio, and exposing cells to oxidative contexts (Aird et al., 2013), and nucleotide synthesis is the
stress increases oxPPP flux (Kuehne et al., 2015), arguing that target of several chemotherapeutics (Table 1), arguing that
cells have excess capacity to produce NADPH even when acquiring nucleotides might be a metabolic bottleneck for
ROS levels increase. Thus, although NADPH is important to diverse cancers. Why might nucleotide base synthesis be
help cancer cells cope with ROS, generating enough cytosolic limiting on a biochemical level? One important factor is that
NADPH for competing metabolic pathways may not be limiting unlike lipids and proteins, nucleotides may not be provided in
for proliferation. This assertion is supported by both classic liter- sufficient quantities or proportions by the tumor microenviron-
ature (Reitzer et al., 1980) and the fact that reduced oxPPP flux ment, and synthesis of these complex molecules requires the
resulting from G6PD deficiency has no impact on human cancer integrated metabolism of non-essential amino acids, ribose,
risk (Cocco, 1987). and one-carbon donors.
Products of the TCA Cycle Can Be Limiting for The DHODH step of pyrimidine biosynthesis is directly
Proliferation coupled to the mitochondrial electron transport chain and oxy-
Clues to which aspects of metabolism are most limiting for pro- gen consumption (Figure 2) and is important for both AML and
liferation come from studies where proliferation arrest from melanoma progression (Sykes et al., 2016; White et al., 2011).
metabolic disruption is rescued by re-supplying metabolites. Whereas oxygen is abundant in standard tissue culture, oxygen
A classic example is glutamine metabolism for TCA cycle ana- levels in even the best perfused tissues are 9% or less (Bertout
plerosis (DeBerardinis et al., 2007; Yuneva et al., 2007). Most et al., 2008), a value closer to the oxygen levels used to study
cultured cells require glutamine, and providing cells with aKG hypoxia responses in culture. Oxygen delivery can be more
or other TCA cycle intermediates can rescue proliferation in limiting than glucose for normal tissue physiology, as evidenced
low-glutamine conditions (Yuneva et al., 2007). Activating pyru- by lactic acid build-up in muscle during exercise, by increased
vate carboxylase to allow anaplerosis from glucose can also hematocrit enhancing the performance of elite athletes, and by
enable proliferation in low-glutamine and contexts where gluta- the fact that angiogenesis is regulated by oxygen levels rather
mine is not used (Figure 2) (Cheng et al., 2011; Davidson et al., than by nutrients even though tissue delivery of both relies on
2016b; Sellers et al., 2015). the vasculature. Oxygen can also be limiting for tumor growth.
The TCA cycle carbon requirement suggests that a product of Individuals living at high altitude have lower cancer incidence
this cycle can be limiting for proliferation. TCA cycle intermedi- (Burtscher, 2013), and increasing oxygen delivery to tissues
ates are precursors for several biosynthetic intermediates, but with erythropoietin promotes cancer progression (Szenajch
production of amino acids accounts for the dominant disposition et al., 2010).

Cell 168, February 9, 2017 661


NAD+/NADH ratio could also explain the use of fermentation
by many proliferating cells across organisms and conditions.
Cancer cells have a very low NAD+/NADH ratio (Hung et al.,
2011), and changes in NAD+/NADH ratio correlate better with
tumor growth rate than changes in ATP/ADP ratio (Gui et al.,
2016). Regenerating NAD+ via orthogonal pathways can also in-
crease cell proliferation when respiration is impaired (Birsoy
et al., 2015; Sullivan et al., 2015; Titov et al., 2016), and disposal
of excess electrons can limit proliferation in prokaryotic systems
(Dietrich et al., 2013).
A low NAD+/NADH ratio also could explain increased ROS in
cancer. ROS levels might increase because cells are deficient
in electron donors to detoxify ROS. However, electron acceptor
deficiency leading to inefficient mitochondrial electron transport
chain function will also lead to increased ROS (Figure 2). The
Figure 4. Intrinsic and Extrinsic Influences on Cancer Cell Metabolic
latter explanation is consistent with angiogenesis and oxygen
Reprogramming delivery being a limitation for tumor growth (Gatenby and Gillies,
Reprogrammed metabolic pathways, including pathways involved in bio- 2004).
energetics, anabolism, and redox homeostasis, are common features of tumor
tissue. The metabolic phenotype of cancer cells is the cumulative result of a
variety of processes both intrinsic and extrinsic to the malignant cell. An in- What Determines How Different Tumors Use
tegrated understanding of the relative contributions of all processes, in the Metabolism?
context of intact tumors, will be necessary to exploit metabolic reprogramming Pathways downstream of oncogenes and tumor suppressors
in the clinic. Both intrinsically and extrinsically regulated pathways provide
regulate cancer cell metabolism. Genomic alterations can also
opportunities for clinical translation, including new targets for therapy and for
non-invasive imaging to detect and monitor cancer. result in copy-number gains and losses of genes encoding meta-
bolic enzymes, and this may induce vulnerabilities (Li et al., 2014;
Locasale et al., 2011; Possemato et al., 2011). However, the
Cancer cells require oxidative metabolism to form tumors extent to which metabolic preferences are hard-wired by the tu-
(Davidson et al., 2016b; Weinberg et al., 2010). Oxygen con- mor genotype is less clear because many non-genetic factors
sumption is coupled to ATP synthesis, but maintaining an also influence tumor metabolism (Figure 2). As in all tissues, tu-
adequate ATP/ADP ratio may not be why respiration is limiting mor metabolism is dictated by a variety of intrinsic and extrinsic
for proliferation. Respiration also regenerates NAD+ for oxidation factors (Figure 4). We need to understand how these factors are
reactions (Figure 3), which are required for aspartate synthesis integrated to create metabolic dependencies.
(Birsoy et al., 2015; Sullivan et al., 2015), providing another link The Environment Can Affect Cancer Cell Metabolism
between oxygen consumption and nucleotide synthesis. Nucle- Cancer cells in culture have a different metabolic phenotype than
otide base carbon is more oxidized than the carbon in most nu- tumors. Whereas many cancer cell lines quantitatively convert
trients. Aspartate is a key oxidized precursor for both purines glucose to lactate, glucose oxidation is prevalent in tumors
and pyrimidines (Figure 2), blood aspartate levels are low, and (Davidson et al., 2016b; Hensley et al., 2016; Maher et al.,
most cells are unable to take up aspartate from the environment 2012; Marin-Valencia et al., 2012). Cultured lung cancer cells
(Birsoy et al., 2015). Thus, stoichiometric reduction of oxygen or use glutamine to supply TCA cycle carbon, whereas lung tumors
another molecule is required for aspartate and nucleotide base in mice prefer glucose as a TCA cycle fuel (Davidson et al.,
synthesis (Sullivan et al., 2015). The ability to produce aspartate 2016b; Hensley et al., 2016; Sellers et al., 2015). These differ-
can be limiting for some tumors (Gui et al., 2016), and respiration ences translate into altered vulnerabilities, as lung cancer cell
also supports production of folate species for purine synthesis lines require glutaminase for proliferation whereas tumors
(Bao et al., 2016; Meiser et al., 2016). Taken together, these derived from these same cells do not; the converse is true for en-
studies argue that access to oxygen or alternative electron ac- zymes involved in glucose oxidation (Davidson et al., 2016b).
ceptors limits nucleotide synthesis for some cancer cells. The environment can also affect the efficacy of drugs targeting
Consequences of Electron Acceptor Limitation in metabolism. Metformin and other biguanides are mitochondrial
Cancer complex I inhibitors that slow tumor growth by preventing com-
The notion that chemical disposal of excess electrons to synthe- plex I-mediated NAD+ regeneration (Wheaton et al., 2014). Thus,
size nucleotides is limiting for proliferation is attractive because it alternative NAD+ regeneration pathways decrease complex I
explains several cancer metabolism phenotypes. First, it pro- dependence and promote metformin resistance (Gui et al.,
vides an explanation for the propensity of cancer cells to pro- 2016). Lipid depletion potentiates the effect of acetyl-coA
duce lactate. NAD+ regeneration from NADH requires electron carboxylase inhibitors in culture, but the same drugs impair
disposal (Figure 3). Conversion of pyruvate to lactate is driven lung tumor growth in vivo despite the presumed availability of
by the NAD+/NADH ratio, and decreasing this ratio in prolifer- fatty acids (Svensson et al., 2016). Tumors vary in the fraction
ating cells because of increased nucleotide production may be of actively proliferating cells, which influences metabolic depen-
sufficient to drive increased lactate production. All proliferating dency (Coloff et al., 2016). Both genetic and drug screens have
cells must replicate DNA, so the accompanying reduction in identified context-specific metabolic liabilities, implying that

662 Cell 168, February 9, 2017


the tumor environment can influence sensitivity to metabolic in- Sullivan et al., 2015) (Figure 2). However, in lung tumors where
hibitors (Possemato et al., 2011; Wenzel et al., 2014). oxygen is less abundant, aspartate is produced from glucose
Cell Lineage Can Also Affect Cancer Metabolism via a series of reactions with fewer oxidation steps (Davidson
Most cytotoxic chemotherapies, including those that target et al., 2016b; Hensley et al., 2016; Sellers et al., 2015). Pancreatic
nucleotide metabolism, are only effective against select cancers. tumors are particularly oxygen-limited and use yet another strat-
A genetic predictor of response is lacking for most of these egy, scavenging amino acids from extracellular protein through
drugs, with sensitive cancer types defined instead by the cancer KRAS-dependent macropinocytosis and circumventing any
tissue of origin. This suggests that some metabolic depen- need to synthesize aspartate from other nutrients (Commisso
dencies might be defined by tumor lineage. Indeed, acute et al., 2013; Davidson et al., 2016a; Kamphorst et al., 2015).
lymphoblastic leukemia (ALL) cells are auxotrophic for aspara- Interactions with Benign Cells Can Affect Cancer Cell
gine and are sensitive to asparagine depletion by L-asparagi- Metabolism
nase. DHODH activity is a lineage-specific requirement of Tumors consist of a complex milieu of malignant and non-malig-
myeloid cells that impacts differentiation (Sykes et al., 2016). nant cell types with distinct metabolic preferences. Specific nu-
MYC induces distinct metabolic phenotypes in mouse liver and trients might be available to cancer cells because they are
lung tumors, with liver tumors displaying enhanced glutamine produced in a given tissue, and differential consumption of nutri-
catabolism and lung tumors retaining the ability to synthesize ents by cancer and non-cancer cells might further affect metab-
glutamine from glucose (Yuneva et al., 2012). Mouse lung and olite levels. In the brain, astrocytes metabolize glucose and
pancreatic tumors initiated by the same driver mutations secrete lactate to be used by neurons as a source of energy (Bit-
also exhibit differences in amino acid metabolism with lung tu- tar et al., 1996). A similar symbiotic relationship may exist within
mors relying on uptake of free amino acids from the blood tumors, where some cells ferment glucose to lactate, which is
(Mayers et al., 2016), whereas pancreatic tumors obtain then used as a respiratory substrate for other cells (Sonveaux
amino acids from extracellular protein (Davidson et al., 2016a). et al., 2008). Alanine produced by pancreatic stellate cells can
These differences in amino acid metabolism translate into a dif- promote pancreatic cancer cell proliferation (Sousa et al.,
ferential dependency of lung and pancreatic tumors on 2016), and bone marrow stromal cells provide cysteine to pro-
branched-chain amino acid nitrogen for nucleotide synthesis mote survival of chronic lymphocytic leukemia cells (Zhang
(Mayers et al., 2016). et al., 2012). Similar relationships likely exist in all cancers, but
An explanation for how lineage influences tumor metabolism is dissecting how cell populations share nutrients within tumors re-
suggested by metabolic gene-expression studies. The meta- mains a challenge.
bolic network of tumors is more similar to the normal tissue
from where the cancer arose than it is to the metabolic network Roles for Metabolic Reprogramming during Cancer
of tumors arising from another tissue (Hu et al., 2013), and normal Progression
tissues wire metabolism differently to support their unique func- Most cancer-related deaths occur after therapy resistance has
tions. Mechanistically, tissue identity is established through emerged and/or after tumor cells have spread from the primary
epigenetic regulation of gene expression, and despite altered site. Whether specific metabolic activities acquired during tumor
gene expression in cancer, many aspects of tissue identity are progression promote therapy resistance or metastasis is being
maintained (Gaude and Frezza, 2016). Metabolism can also influ- investigated.
ence epigenetic state, but the fact that tumors retain metabolic Role of Metabolism in Metastasis
features from their parental normal tissue argues that cancers Epithelial-mesenchymal transition (EMT) is a transdifferentiation
adapt existing tissue metabolism to support abnormal prolifera- program associated with metastasis and chemotherapeutic
tion rather than converging on a universal proliferation program. resistance (Ye and Weinberg, 2015). Enforcing an EMT-like
This distinction is critical because it suggests that drugs target- phenotype in lung cancer cells reduces lipogenesis and in-
ing proliferative metabolism may not be effective against all creases respiration (Jiang et al., 2015). Silencing fatty-acid
tumors. synthase in these cells induces EMT and enhances metastatic
Tumor lineage also modifies the penetrance of transforming seeding in mice. Similarly, in a breast cancer model cellular inva-
mutations. The restricted tumor spectrum in patients with familial sion, migration, and metastasis correlate with enhanced respira-
cancer syndromes caused by SDH and FH mutations indicates tion and mitochondrial biogenesis driven by the transcriptional
an exquisite context dependence for these events to cause co-activator PGC1a (LeBleu et al., 2014). Silencing PGC1a
malignancy. Why specific mutations cause lineage-restricted reduces metastasis without altering tumor growth at the primary
cancers is not understood, but one might speculate that some site. In this model, PGC1a is required for cancer cell invasion but
metabolic alterations are only tolerated in the context of a spe- not proliferation, perhaps explaining its role in metastasis and
cific metabolic network. dispensability for primary tumor growth.
Tissue-specific differences in nutrient availability impose Several studies have examined metabolic transitions accom-
further constraints on metabolism. How cancer cells generate panying detachment from extracellular matrix, an early step in
aspartate in different contexts illustrates how such constraints metastasis. In breast epithelial cells, loss of matrix attachment
can affect production of a key metabolite for nucleotide synthe- results in oxidative stress and cell death (Schafer et al., 2009).
sis. Cells in culture where oxygen is in excess produce aspartate Treating detached cells with antioxidants or enhancing NAPDH
from glutamine via the TCA cycle, a pathway that involves multi- production by the oxPPP promotes survival. In lung cancer cells,
ple oxidation steps (Birsoy et al., 2015; DeBerardinis et al., 2007; detachment suppresses pathways required to maximize cell

Cell 168, February 9, 2017 663


growth and induces shuttling of cytosolic-reducing equivalents Can Cancer Metabolism Be Exploited to Improve
into the mitochondria to combat ROS (Jiang et al., 2016). Specif- Therapy?
ically, reductive carboxylation of cytosolic aKG is induced to pro- To target metabolism for therapy, limiting metabolic processes
duce citrate, which then enters the mitochondria to produce must be identified and understood sufficiently to target the pro-
NADPH. This pathway is dispensable in monolayer culture but cess safely and select responsive patients. Using the classifica-
required in the anchorage-independent state. tions described above, transforming and/or enabling activities
Additional bottlenecks can prevent metastasis after cells enter must be identified with an adequate therapeutic index. Clinical
the circulation. Circulating melanoma cells and metastases have experience with cytotoxic chemotherapy highlights the chal-
elevated ROS compared to primary tumors (Piskounova et al., lenges that will likely confront new metabolic therapies. Many
2015). Highly metastatic tumors undergo reversible metabolic chemotherapies inhibit nucleotide metabolism (Table 1). These
changes allowing them to increase mitochondrial NADPH and drugs are highly effective in some tumors, but efficacy is not
combat ROS stress. Inhibiting this response reduces metastasis, solely determined by proliferative index. Tumor cells acquire mu-
whereas systemic treatment with antioxidants enhances metas- tations in cell-cycle checkpoint genes that contribute to thera-
tasis. Exacerbating oxidative stress potentiates some chemo- peutic window, but the differential sensitivity of different cancers
therapeutic responses (Harris et al., 2015; Raj et al., 2011) and to these drugs is incompletely understood. Evidence suggests
enhances response to radiation therapy (Robbins and Zhao, that tumors use diverse mechanisms to maintain nucleotide
2004). Altogether, these studies underscore the importance of pools, and this may contribute to efficacy. However, a better un-
managing ROS during tumor progression and suggest that this derstanding of differential sensitivity to metabolism-targeted
might be exploited therapeutically. chemotherapy may help in using these drugs more effectively
Metabolism and Therapy Resistance and in suggesting new targets.
Therapy-resistant tumors have altered metabolic phenotypes Recent Challenges Associated with Targeting Cancer
relative to treatment-naive tumors, with enhanced reliance on Metabolism
mitochondrial metabolism in the resistant cancers. Genetic inac- Whether studies of cell-autonomous metabolic reprogramming
tivation of oncogenic KRASG12D in pancreatic cancer leads to tu- will identify the best therapeutic targets is unproven. Success
mor regression; however, a small number of cancer cells survive targeting glycolysis has been limited. Exposing patients to high
in a dormant state and regenerate tumors if KRASG12D is reacti- levels of 2-deoxyglucose generated tumor responses but
vated (Viale et al., 2014). Compared to cells with constitutive caused hypoglycemia symptoms, and more tolerable doses in
KRASG12D expression, the surviving cells have enhanced respi- modern trials have been disappointing (Vander Heiden, 2011).
ration, reduced glycolysis, and an impaired ability to increase Attempts to target the Warburg effect directly have also been
glycolysis when respiration is inhibited. This renders surviving met with limited success. Dependence on the enzyme lactate
cells susceptible to respiration inhibitors, and treating mice dehydrogenase (LDH) was demonstrated both genetically and
with an ATP synthase inhibitor during tumor regression delays pharmacologically (Fantin et al., 2006; Le et al., 2010; Shim
relapse after KRASG12D reactivation. et al., 1997), leading to development of LDHA inhibitors (Bou-
Mitochondria also contribute to melanoma therapy resistance. dreau et al., 2016), but none progressed to clinical trials, sug-
Higher PGC1a expression portends poor outcome in human gesting either unacceptable toxicity, insufficient drug exposure,
melanoma (Vazquez et al., 2013). In melanoma cells, PGC1a or a lack of LDHA dependence in human tumors. Lactate trans-
expression is regulated by the melanocyte-lineage transcription port inhibition is currently being tested as an alternative
factor MITF and is required to maintain high levels of respiration, approach. Activating pyruvate dehydrogenase (PDH) by target-
low ROS, and resistance to the oxidizing agent piperlongumine. ing its inhibitory kinase (PDHK) has been proposed as a thera-
Silencing PGC1a synergizes with piperlongumine to reduce mel- peutic strategy. The PDHK inhibitor dichloroacetate (DCA) elicits
anoma xenograft growth. In human melanoma with mutant metabolic effects in human tumors (Michelakis et al., 2010), but
BRAF, gene-expression patterns associated with mitochondrial minimal efficacy data have been reported. Recent data indicate
biogenesis predict reduced survival, and tumors that relapse that PDH is active and required by some tumors (Davidson et al.,
after MAPK inhibitor treatment have increased mitochondrial 2016b; Hensley et al., 2016), tempering hope that further acti-
biogenesis gene expression (Zhang et al., 2016). Inhibiting mito- vating the enzyme will be therapeutically useful.
chondrial biogenesis using a mitochondrial-targeted Hsp90 in- One success in targeting cancer metabolism is the reversal of
hibitor decreases MAPK inhibitor resistance in xenografts. 2HG-mediated reprogramming in IDH mutant tumors (Losman
In autochthonous models of breast and lung cancer, anti- and Kaelin, 2013). Drugs targeting mutant IDH can elicit re-
angiogenic kinase inhibitors enhance AMPK signaling and shift sponses in patients with hematological malignancies but have
metabolism toward a more oxidative phenotype. Treated cells been less effective in glioma models (Tateishi et al., 2015). The
are more sensitive to inhibitors of oxidative metabolism, and mechanism of 2HG-mediated transformation may differ in
the same drugs synergize with anti-angiogenics to suppress tu- different cancers; however, liabilities caused by high 2HG might
mor growth (Navarro et al., 2016). Mitochondrial inhibition with be exploited as an alternative approach to treat these cancers.
metformin can kill chemotherapy-resistant breast tumor stem Identifying Tumors Susceptible to Metabolic Therapy
cells (Janzer et al., 2014). Taken together, these studies demon- Transforming mutations in metabolic enzymes present clinical
strate the importance of mitochondrial function in enabling ther- opportunities where patients can be selected based on tumor
apeutic resistance and suggest that mitochondrial inhibitors may genetics. However, most metabolic changes are not driven by
help curtail cancer progression. metabolic enzyme mutations, and even in the case of IDH

664 Cell 168, February 9, 2017


mutations, a durable state of mutant enzyme dependence may Cancer cells compete with T cells for glucose in tumors, and
not always be present (Tateishi et al., 2015). Thus, how to select restricting T cell glucose metabolism causes lymphocyte
patients for drugs targeting metabolic enzymes remains a critical exhaustion (Chang et al., 2015; Ho et al., 2015). Thus, therapies
therapeutic question. that decrease glucose use by cancer cells could make glucose
Based on successes developing protein kinase inhibitors, one available for T cells and enhance immune-effector functions to
approach has been to screen cell-line panels to identify geneti- further limit tumor growth. However, if the same therapies inhibit
cally susceptible cancers and then to validate these hypotheses T cell glucose metabolism, they may limit anti-tumor immunity.
in animal models. However, differential nutrient use in culture Other nutrient levels can affect both cancer and immune cells.
and in tumors suggests that the use of cell lines to identify sen- High arginine levels promote enhanced T cell survival and anti-
sitive cancers could identify activities that are neutral in vivo. tumor activity (Geiger et al., 2016), and therapies that deplete tu-
Conversely, enabling pathways in vivo may be less important mor arginine promote immune-suppressor cell accumulation
in culture. A requirement for glutamine metabolism has been (Fletcher et al., 2015). Lactate can blunt T and NK cell function
observed in many cancer types (Altman et al., 2016), but both tis- (Brand et al., 2016). Kynurenine, a breakdown product of trypto-
sue of origin and genetics contribute to this phenotype (Yuneva phan, can modulate both the innate and adaptive immune
et al., 2012). The auxotrophy of some cancers for individual system and has been implicated in cancer-associated immuno-
amino acids is also influenced by a combination of genetic and suppression (Platten et al., 2015), and indoleamine-2,3-dioxyge-
non-genetic factors. Auxotrophy for arginine or asparagine can nase (IDO) inhibitors that decrease tryptophan metabolism can
be explained in some cases by decreased expression of synthe- help break immune tolerance and potentiate chemotherapy
sis enzymes; however, treatment of ALL with asparaginase is (Muller et al., 2005). Increased MTA levels in MTAP null tumors
effective even when asparagine synthase is expressed, suggest- can also suppress T cell activation (Henrich et al., 2016) and
ing that additional factors underlie the need for exogenous suggest that some metabolic interventions might enhance anti-
asparagine (Krall et al., 2016; Stams et al., 2003). Increased cancer immune responses.
biguanide sensitivity has been attributed to various mutations The relationship between tumors and other organ systems is
(Buzzai et al., 2007; Shackelford et al., 2013), but tissue environ- not limited to the immune system. Endothelial cells also undergo
ment also affects response to these drugs (Gui et al., 2016). factor-induced metabolic reprogramming, and metabolic thera-
Selecting patients for chemotherapy based on tumor type pies can limit endothelial cell proliferation and angiogenesis (De
argues that cell lineage can be an effective tool to stratify pa- Bock et al., 2013). Cancer also causes alterations in whole-body
tients and argues that in addition to genetics, tumor type and metabolism that may influence tumor-nutrient availability. Modu-
location should be considered when developing metabolism-tar- lating the amino acid composition of the diet can slow cancer
geted drugs. growth (Maddocks et al., 2013), and investigations into how
Genetically Defined Metabolic Targets diet affects tumor growth remain an underexplored area.
Sensitivity to some enabling metabolic targets is determined by Clinical Utility of Neutral Metabolic Activities
genetics, including passenger deletion of enzymes that prevent By definition, neutral metabolic activities are not therapeutic tar-
metabolic compensation in tumors. For example, most glioblas- gets, but some may still present useful opportunities in clinical
toma cells express two isoforms of the glycolytic enzyme oncology, particularly as predictive biomarkers or for tumor im-
enolase encoded by the genes ENO1 and ENO2. ENO1 resides aging. PKM2 expression in tumors is the basis for a positron
on chromosome 1p36 at a tumor-suppressor locus that is homo- emission tomography (PET) imaging agent. N,N-diarylsulfona-
zygously deleted in 1%–5% of glioblastomas (Muller et al., mide (DASA) compounds bind to PKM2, and an 11C-labeled
2012). This reduces metabolic redundancy and creates a thera- analog of DASA-23 is taken up by PKM2-expressing orthotopic
peutic opportunity to target ENO2. The gene encoding the gliomas in mice, providing imaging contrast with the normal brain
enzyme methylthioadenosine phosphorylase (MTAP) can be (Witney et al., 2015). Other investigational PET agents exploit the
deleted with the p16/CDKN2A tumor-suppressor gene. The re- ability of some tumors to take up and retain glutamine (Venneti
sulting accumulation of MTA reduces PRMT5 methyltransferase et al., 2015). Conversion of 13C-labeled pyruvate to lactate can
activity, creating another therapeutic opportunity (Kryukov et al., be imaged in patients by hyperpolarized MRI (Nelson et al.,
2016; Marjon et al., 2016; Mavrakis et al., 2016). Gene-expres- 2013). Some targetable activities are also the basis of new imag-
sion changes that create auxotrophies for specific amino acids ing techniques, with the detection of 2HG by magnetic reso-
can also be targeted, with L-asparaginase providing an example nance spectroscopy a prominent recent example (Andronesi
of success. Drugs that deplete arginine are being tested in pa- et al., 2012; Choi et al., 2012). Finally, the heterogeneity of meta-
tients with ASS1-deficient tumors (Phillips et al., 2013). Limiting bolic phenotypes among human cancer patients may provide
serine in the diet can be efficacious in some p53 mutant cancers biomarkers that correlate with disease progression (Hakimi
(Maddocks et al., 2013). et al., 2016).
Non-Cell-Autonomous Targeting of Tumor Metabolism
Understanding the interactions between cancer cell and immune Future Perspectives
cell metabolism will be critical for combining metabolism-tar- Analysis of cell-intrinsic metabolic preferences imposed by the
geted therapies with immunotherapies. Nutrient availability and oncogenotype has been informative to uncover new paradigms
metabolism affect T cell and macrophage differentiation, raising in proliferating cell metabolic regulation. However, the next
the possibility that targeting tumor cell metabolism may either phase of cancer metabolism research will need to address
promote or prevent anti-cancer immune responses. increasingly complex questions about how intrinsic and extrinsic

Cell 168, February 9, 2017 665


influences integrate to create exploitable metabolic phenotypes Boros, L.G., Torday, J.S., Lim, S., Bassilian, S., Cascante, M., and Lee, W.N.
in cancer. This will require consideration of the metabolic (2000). Transforming growth factor beta2 promotes glucose carbon incorpora-
tion into nucleic acid ribose through the nonoxidative pentose cycle in lung
preferences hard-wired into cancer cells by tissue of origin, inter-
epithelial carcinoma cells. Cancer Res. 60, 1183–1185.
actions between benign and malignant cells within the microen-
Boudreau, A., Purkey, H.E., Hitz, A., Robarge, K., Peterson, D., Labadie, S.,
vironment, and influences of the diet and microbiome on the
Kwong, M., Hong, R., Gao, M., Del Nagro, C., et al. (2016). Metabolic plasticity
host. Cell culture systems must be improved to better reflect underpins innate and acquired resistance to LDHA inhibition. Nat. Chem. Biol.
the metabolic limitations in tumors, and these studies will be pro- 12, 779–786.
pelled by improvements in quantitative assessment of metabolic Brand, A., Singer, K., Koehl, G.E., Kolitzus, M., Schoenhammer, G., Thiel, A.,
fluxes in different contexts. Ultimately this will enable matching Matos, C., Bruss, C., Klobuch, S., Peter, K., et al. (2016). LDHA-associated lac-
of the right therapies with the right patients to exploit altered tic acid production blunts tumor immunosurveillance by T and NK cells. Cell
metabolism and improve cancer outcomes. Metab. 24, 657–671.
Burtscher, M. (2013). Effects of living at higher altitudes on mortality: a narra-
ACKNOWLEDGMENTS tive review. Aging Dis. 5, 274–280.
Buzzai, M., Jones, R.G., Amaravadi, R.K., Lum, J.J., DeBerardinis, R.J., Zhao,
We regret that due to space limitations we were unable to cite many excellent
F., Viollet, B., and Thompson, C.B. (2007). Systemic treatment with the antidi-
studies that shaped our modern understanding of cancer metabolism. We
abetic drug metformin selectively impairs p53-deficient tumor cell growth.
thank members of the R.J.D. and M.G.V.H. laboratories for insight and exten-
Cancer Res. 67, 6745–6752.
sive critique of the manuscript and Brooke Bevis for help with figures. M.G.V.H.
is supported by the NCI (CA168653, CA201276), the Ludwig Center at MIT, Cairns, R.A., and Mak, T.W. (2016). The current state of cancer metabolism.
SU2C, the Lustgarten Foundation, and the Howard Hughes Medical Institute Nat. Rev. Cancer 16, 613–614.
(Faculty Scholars Award). R.J.D. is supported by grants from the NCI Chang, C.H., Qiu, J., O’Sullivan, D., Buck, M.D., Noguchi, T., Curtis, J.D.,
(CA157996), Cancer Prevention and Research Institute of Texas (RP160089), Chen, Q., Gindin, M., Gubin, M.M., van der Windt, G.J., et al. (2015). Metabolic
V Foundation (Translational Research Award), Robert A. Welch Foundation competition in the tumor microenvironment is a driver of cancer progression.
(I-1733), and the Howard Hughes Medical Institute (Faculty Scholars Award). Cell 162, 1229–1241.
M.G.V.H. is on the scientific advisory board of Agios Pharmaceuticals and Chen, C., Liu, Y., Lu, C., Cross, J.R., Morris, J.P., 4th, Shroff, A.S., Ward, P.S.,
Aeglea Biotherapeutics, and R.J.D. is on the scientific advisory board of Agios Bradner, J.E., Thompson, C., and Lowe, S.W. (2013). Cancer-associated IDH2
Pharmaceuticals. mutants drive an acute myeloid leukemia that is susceptible to Brd4 inhibition.
Genes Dev. 27, 1974–1985.
REFERENCES Cheng, T., Sudderth, J., Yang, C., Mullen, A.R., Jin, E.S., Matés, J.M., and
DeBerardinis, R.J. (2011). Pyruvate carboxylase is required for glutamine-in-
Aird, K.M., Zhang, G., Li, H., Tu, Z., Bitler, B.G., Garipov, A., Wu, H., Wei, Z., dependent growth of tumor cells. Proc. Natl. Acad. Sci. USA 108, 8674–8679.
Wagner, S.N., Herlyn, M., and Zhang, R. (2013). Suppression of nucleotide
Choi, C., Ganji, S.K., DeBerardinis, R.J., Hatanpaa, K.J., Rakheja, D., Kovacs,
metabolism underlies the establishment and maintenance of oncogene-
Z., Yang, X.L., Mashimo, T., Raisanen, J.M., Marin-Valencia, I., et al. (2012).
induced senescence. Cell Rep. 3, 1252–1265.
2-hydroxyglutarate detection by magnetic resonance spectroscopy in IDH-
Altman, B.J., Stine, Z.E., and Dang, C.V. (2016). From Krebs to clinic: gluta-
mutated patients with gliomas. Nat. Med. 18, 624–629.
mine metabolism to cancer therapy. Nat. Rev. Cancer 16, 619–634.
Cocco, P. (1987). Does G6PD deficiency protect against cancer? A critical re-
Andronesi, O.C., Kim, G.S., Gerstner, E., Batchelor, T., Tzika, A.A., Fantin,
view. J. Epidemiol. Community Health 41, 89–93.
V.R., Vander Heiden, M.G., and Sorensen, A.G. (2012). Detection of 2-hy-
droxyglutarate in IDH-mutated glioma patients by in vivo spectral-editing Coloff, J.L., Murphy, J.P., Braun, C.R., Harris, I.S., Shelton, L.M., Kami, K.,
and 2D correlation magnetic resonance spectroscopy. Sci. Transl. Med. 4, Gygi, S.P., Selfors, L.M., and Brugge, J.S. (2016). Differential glutamate meta-
116ra4. bolism in proliferating and quiescent mammary epithelial cells. Cell Metab. 23,
867–880.
Andrzejewski, S., Gravel, S.P., Pollak, M., and St-Pierre, J. (2014). Metformin
directly acts on mitochondria to alter cellular bioenergetics. Cancer Metab. Commisso, C., Davidson, S.M., Soydaner-Azeloglu, R.G., Parker, S.J., Kam-
2, 12. phorst, J.J., Hackett, S., Grabocka, E., Nofal, M., Drebin, J.A., Thompson,
C.B., et al. (2013). Macropinocytosis of protein is an amino acid supply route
Bao, X.R., Ong, S.E., Goldberger, O., Peng, J., Sharma, R., Thompson, D.A.,
in Ras-transformed cells. Nature 497, 633–637.
Vafai, S.B., Cox, A.G., Marutani, E., Ichinose, F., et al. (2016). Mitochondrial
dysfunction remodels one-carbon metabolism in human cells. eLife 5, 5. Curi, R., Newsholme, P., and Newsholme, E.A. (1988). Metabolism of pyruvate
Baysal, B.E., Ferrell, R.E., Willett-Brozick, J.E., Lawrence, E.C., Myssiorek, D., by isolated rat mesenteric lymphocytes, lymphocyte mitochondria and iso-
Bosch, A., van der Mey, A., Taschner, P.E., Rubinstein, W.S., Myers, E.N., et al. lated mouse macrophages. Biochem. J. 250, 383–388.
(2000). Mutations in SDHD, a mitochondrial complex II gene, in hereditary par- Dang, L., White, D.W., Gross, S., Bennett, B.D., Bittinger, M.A., Driggers, E.M.,
aganglioma. Science 287, 848–851. Fantin, V.R., Jang, H.G., Jin, S., Keenan, M.C., et al. (2009). Cancer-associated
Bertout, J.A., Patel, S.A., and Simon, M.C. (2008). The impact of O2 availability IDH1 mutations produce 2-hydroxyglutarate. Nature 462, 739–744.
on human cancer. Nat. Rev. Cancer 8, 967–975. Davidson, S.M., Jonas, O., Keibler, M.A., Hou, H.W., Luengo, A., Mayers, J.R.,
Birsoy, K., Possemato, R., Lorbeer, F.K., Bayraktar, E.C., Thiru, P., Yucel, B., Wyckoff, J., Del Rosario, A.M., Whitman, M., Chin, C.R., et al. (2016a). Direct
Wang, T., Chen, W.W., Clish, C.B., and Sabatini, D.M. (2014). Metabolic deter- evidence for cancer-cell-autonomous extracellular protein catabolism in
minants of cancer cell sensitivity to glucose limitation and biguanides. Nature pancreatic tumors. Nat. Med. Published online December 26, 2016. http://
508, 108–112. dx.doi.org/10.1038/nm.4256.
Birsoy, K., Wang, T., Chen, W.W., Freinkman, E., Abu-Remaileh, M., and Sa- Davidson, S.M., Papagiannakopoulos, T., Olenchock, B.A., Heyman, J.E., Kei-
batini, D.M. (2015). An essential role of the mitochondrial electron transport bler, M.A., Luengo, A., Bauer, M.R., Jha, A.K., O’Brien, J.P., Pierce, K.A., et al.
chain in cell proliferation is to enable aspartate synthesis. Cell 162, 540–551. (2016b). Environment impacts the metabolic dependencies of Ras-driven non-
Bittar, P.G., Charnay, Y., Pellerin, L., Bouras, C., and Magistretti, P.J. (1996). small cell lung cancer. Cell Metab. 23, 517–528.
Selective distribution of lactate dehydrogenase isoenzymes in neurons and as- Dayton, T.L., Gocheva, V., Miller, K.M., Israelsen, W.J., Bhutkar, A., Clish,
trocytes of human brain. J. Cereb. Blood Flow Metab. 16, 1079–1089. C.B., Davidson, S.M., Luengo, A., Bronson, R.T., Jacks, T., and Vander

666 Cell 168, February 9, 2017


Heiden, M.G. (2016). Germline loss of PKM2 promotes metabolic distress and Henrich, F.C., Singer, K., Poller, K., Bernhardt, L., Strobl, C.D., Limm, K., Ritter,
hepatocellular carcinoma. Genes Dev. 30, 1020–1033. A.P., Gottfried, E., Völkl, S., Jacobs, B., et al. (2016). Suppressive effects of tu-
De Bock, K., Georgiadou, M., Schoors, S., Kuchnio, A., Wong, B.W., Can- mor cell-derived 50 -deoxy-50 -methylthioadenosine on human T cells. OncoIm-
telmo, A.R., Quaegebeur, A., Ghesquière, B., Cauwenberghs, S., Eelen, G., munology 5, e1184802.
et al. (2013). Role of PFKFB3-driven glycolysis in vessel sprouting. Cell 154, Hensley, C.T., Faubert, B., Yuan, Q., Lev-Cohain, N., Jin, E., Kim, J., Jiang, L.,
651–663. Ko, B., Skelton, R., Loudat, L., et al. (2016). Metabolic heterogeneity in human
DeBerardinis, R.J., and Chandel, N.S. (2016). Fundamentals of cancer meta- lung tumors. Cell 164, 681–694.
bolism. Sci. Adv. 2, e1600200. Ho, P.C., Bihuniak, J.D., Macintyre, A.N., Staron, M., Liu, X., Amezquita, R.,
DeBerardinis, R.J., Mancuso, A., Daikhin, E., Nissim, I., Yudkoff, M., Wehrli, S., Tsui, Y.C., Cui, G., Micevic, G., Perales, J.C., et al. (2015). Phosphoenolpyr-
and Thompson, C.B. (2007). Beyond aerobic glycolysis: transformed cells can uvate is a metabolic checkpoint of anti-tumor T cell responses. Cell 162,
engage in glutamine metabolism that exceeds the requirement for protein and 1217–1228.
nucleotide synthesis. Proc. Natl. Acad. Sci. USA 104, 19345–19350. Hosios, A.M., Hecht, V.C., Danai, L.V., Johnson, M.O., Rathmell, J.C., Stein-
Dietrich, L.E., Okegbe, C., Price-Whelan, A., Sakhtah, H., Hunter, R.C., and hauser, M.L., Manalis, S.R., and Vander Heiden, M.G. (2016). Amino acids
Newman, D.K. (2013). Bacterial community morphogenesis is intimately linked rather than glucose account for the majority of cell mass in proliferating
to the intracellular redox state. J. Bacteriol. 195, 1371–1380. mammalian cells. Dev. Cell 36, 540–549.
Erez, A., and DeBerardinis, R.J. (2015). Metabolic dysregulation in monogenic Howell, J.J., Ricoult, S.J., Ben-Sahra, I., and Manning, B.D. (2013). A growing
disorders and cancer - finding method in madness. Nat. Rev. Cancer 15, role for mTOR in promoting anabolic metabolism. Biochem. Soc. Trans. 41,
440–448. 906–912.
Fan, J., Ye, J., Kamphorst, J.J., Shlomi, T., Thompson, C.B., and Rabinowitz, Hu, J., Locasale, J.W., Bielas, J.H., O’Sullivan, J., Sheahan, K., Cantley, L.C.,
J.D. (2014). Quantitative flux analysis reveals folate-dependent NADPH pro- Vander Heiden, M.G., and Vitkup, D. (2013). Heterogeneity of tumor-induced
duction. Nature 510, 298–302. gene expression changes in the human metabolic network. Nat. Biotechnol.
Fang, M., Shen, Z., Huang, S., Zhao, L., Chen, S., Mak, T.W., and Wang, X. 31, 522–529.
(2010). The ER UDPase ENTPD5 promotes protein N-glycosylation, the War- Hung, Y.P., Albeck, J.G., Tantama, M., and Yellen, G. (2011). Imaging cytosolic
burg effect, and proliferation in the PTEN pathway. Cell 143, 711–724. NADH-NAD(+) redox state with a genetically encoded fluorescent biosensor.
Fantin, V.R., St-Pierre, J., and Leder, P. (2006). Attenuation of LDH-A expres- Cell Metab. 14, 545–554.
sion uncovers a link between glycolysis, mitochondrial physiology, and tumor Israelsen, W.J., Dayton, T.L., Davidson, S.M., Fiske, B.P., Hosios, A.M., Bellin-
maintenance. Cancer Cell 9, 425–434. ger, G., Li, J., Yu, Y., Sasaki, M., Horner, J.W., et al. (2013). PKM2 isoform-spe-
Fletcher, M., Ramirez, M.E., Sierra, R.A., Raber, P., Thevenot, P., Al-Khami, cific deletion reveals a differential requirement for pyruvate kinase in tumor
A.A., Sanchez-Pino, D., Hernandez, C., Wyczechowska, D.D., Ochoa, A.C., cells. Cell 155, 397–409.
and Rodriguez, P.C. (2015). l-Arginine depletion blunts antitumor T-cell re- Israelsen, W.J., and Vander Heiden, M.G. (2015). Pyruvate kinase: Function,
sponses by inducing myeloid-derived suppressor cells. Cancer Res. 75, regulation and role in cancer. Semin. Cell Dev. Biol. 43, 43–51.
275–283.
Janzer, A., German, N.J., Gonzalez-Herrera, K.N., Asara, J.M., Haigis, M.C.,
Flier, J.S., Mueckler, M.M., Usher, P., and Lodish, H.F. (1987). Elevated levels
and Struhl, K. (2014). Metformin and phenformin deplete tricarboxylic acid
of glucose transport and transporter messenger RNA are induced by ras or src
cycle and glycolytic intermediates during cell transformation and NTPs in can-
oncogenes. Science 235, 1492–1495.
cer stem cells. Proc. Natl. Acad. Sci. USA 111, 10574–10579.
Gaglio, D., Soldati, C., Vanoni, M., Alberghina, L., and Chiaradonna, F. (2009).
Jiang, L., Xiao, L., Sugiura, H., Huang, X., Ali, A., Kuro-o, M., Deberardinis,
Glutamine deprivation induces abortive s-phase rescued by deoxyribonucleo-
R.J., and Boothman, D.A. (2015). Metabolic reprogramming during TGFb1-
tides in k-ras transformed fibroblasts. PLoS ONE 4, e4715.
induced epithelial-to-mesenchymal transition. Oncogene 34, 3908–3916.
Gatenby, R.A., and Gillies, R.J. (2004). Why do cancers have high aerobic
Jiang, L., Shestov, A.A., Swain, P., Yang, C., Parker, S.J., Wang, Q.A., Terada,
glycolysis? Nat. Rev. Cancer 4, 891–899.
L.S., Adams, N.D., McCabe, M.T., Pietrak, B., et al. (2016). Reductive carbox-
Gaude, E., and Frezza, C. (2016). Tissue-specific and convergent metabolic ylation supports redox homeostasis during anchorage-independent growth.
transformation of cancer correlates with metastatic potential and patient sur- Nature 532, 255–258.
vival. Nat. Commun. 7, 13041.
Kaelin, W.G., Jr., and McKnight, S.L. (2013). Influence of metabolism on epige-
Geiger, R., Rieckmann, J.C., Wolf, T., Basso, C., Feng, Y., Fuhrer, T., Koga-
netics and disease. Cell 153, 56–69.
deeva, M., Picotti, P., Meissner, F., Mann, M., et al. (2016). L-Arginine modu-
lates T cell metabolism and enhances survival and anti-tumor activity. Cell 167, Kamphorst, J.J., Nofal, M., Commisso, C., Hackett, S.R., Lu, W., Grabocka, E.,
829–842.e13. Vander Heiden, M.G., Miller, G., Drebin, J.A., Bar-Sagi, D., et al. (2015). Human
pancreatic cancer tumors are nutrient poor and tumor cells actively scavenge
Gui, D.Y., Sullivan, L.B., Luengo, A., Hosios, A.M., Bush, L.N., Gitego, N., Da-
extracellular protein. Cancer Res. 75, 544–553.
vidson, S.M., Freinkman, E., Thomas, C.J., and Vander Heiden, M.G. (2016).
Environment dictates dependence on mitochondrial complex I for NAD+ and Kats, L.M., Reschke, M., Taulli, R., Pozdnyakova, O., Burgess, K., Bhargava,
aspartate production and determines cancer cell sensitivity to metformin. P., Straley, K., Karnik, R., Meissner, A., Small, D., et al. (2014). Proto-onco-
Cell Metab. 24, 716–727. genic role of mutant IDH2 in leukemia initiation and maintenance. Cell Stem
Cell 14, 329–341.
Guo, J.Y., Teng, X., Laddha, S.V., Ma, S., Van Nostrand, S.C., Yang, Y., Khor,
S., Chan, C.S., Rabinowitz, J.D., and White, E. (2016). Autophagy provides Koppenol, W.H., Bounds, P.L., and Dang, C.V. (2011). Otto Warburg’s contri-
metabolic substrates to maintain energy charge and nucleotide pools in butions to current concepts of cancer metabolism. Nat. Rev. Cancer 11,
Ras-driven lung cancer cells. Genes Dev. 30, 1704–1717. 325–337.
Hakimi, A.A., Reznik, E., Lee, C.H., Creighton, C.J., Brannon, A.R., Luna, A., Krall, A.S., Xu, S., Graeber, T.G., Braas, D., and Christofk, H.R. (2016). Aspar-
Aksoy, B.A., Liu, E.M., Shen, R., Lee, W., et al. (2016). An integrated metabolic agine promotes cancer cell proliferation through use as an amino acid ex-
atlas of clear cell renal cell carcinoma. Cancer Cell 29, 104–116. change factor. Nat. Commun. 7, 11457.
Harris, I.S., Treloar, A.E., Inoue, S., Sasaki, M., Gorrini, C., Lee, K.C., Yung, Kryukov, G.V., Wilson, F.H., Ruth, J.R., Paulk, J., Tsherniak, A., Marlow, S.E.,
K.Y., Brenner, D., Knobbe-Thomsen, C.B., Cox, M.A., et al. (2015). Glutathione Vazquez, F., Weir, B.A., Fitzgerald, M.E., Tanaka, M., et al. (2016). MTAP dele-
and thioredoxin antioxidant pathways synergize to drive cancer initiation and tion confers enhanced dependency on the PRMT5 arginine methyltransferase
progression. Cancer Cell 27, 211–222. in cancer cells. Science 351, 1214–1218.

Cell 168, February 9, 2017 667


Kuehne, A., Emmert, H., Soehle, J., Winnefeld, M., Fischer, F., Wenck, H., Gal- Muller, A.J., DuHadaway, J.B., Donover, P.S., Sutanto-Ward, E., and Prender-
linat, S., Terstegen, L., Lucius, R., Hildebrand, J., and Zamboni, N. (2015). gast, G.C. (2005). Inhibition of indoleamine 2,3-dioxygenase, an immunoregu-
Acute activation of oxidative pentose phosphate pathway as first-line latory target of the cancer suppression gene Bin1, potentiates cancer
response to oxidative stress in human skin cells. Mol. Cell 59, 359–371. chemotherapy. Nat. Med. 11, 312–319.
Le, A., Cooper, C.R., Gouw, A.M., Dinavahi, R., Maitra, A., Deck, L.M., Royer, Muller, F.L., Colla, S., Aquilanti, E., Manzo, V.E., Genovese, G., Lee, J., Eisen-
R.E., Vander Jagt, D.L., Semenza, G.L., and Dang, C.V. (2010). Inhibition of son, D., Narurkar, R., Deng, P., Nezi, L., et al. (2012). Passenger deletions
lactate dehydrogenase A induces oxidative stress and inhibits tumor progres- generate therapeutic vulnerabilities in cancer. Nature 488, 337–342.
sion. Proc. Natl. Acad. Sci. USA 107, 2037–2042. Navarro, P., Bueno, M.J., Zagorac, I., Mondejar, T., Sanchez, J., Mourón, S.,
LeBleu, V.S., O’Connell, J.T., Gonzalez Herrera, K.N., Wikman, H., Pantel, K., Muñoz, J., Gómez-López, G., Jimenez-Renard, V., Mulero, F., et al. (2016).
Haigis, M.C., de Carvalho, F.M., Damascena, A., Domingos Chinen, L.T., Ro- Targeting tumor mitochondrial metabolism overcomes resistance to antian-
cha, R.M., et al. (2014). PGC-1a mediates mitochondrial biogenesis and oxida- giogenics. Cell Rep. 15, 2705–2718.
tive phosphorylation in cancer cells to promote metastasis. Nat. Cell Biol. 16, Nelson, S.J., Kurhanewicz, J., Vigneron, D.B., Larson, P.E., Harzstark, A.L.,
992–1003, 1–15. Ferrone, M., van Criekinge, M., Chang, J.W., Bok, R., Park, I., et al. (2013).
Lewis, C.A., Parker, S.J., Fiske, B.P., McCloskey, D., Gui, D.Y., Green, C.R., Metabolic imaging of patients with prostate cancer using hyperpolarized
Vokes, N.I., Feist, A.M., Vander Heiden, M.G., and Metallo, C.M. (2014). [1-13C]pyruvate. Sci. Transl. Med. 5, 198ra108.
Tracing compartmentalized NADPH metabolism in the cytosol and mitochon- Palm, W., Park, Y., Wright, K., Pavlova, N.N., Tuveson, D.A., and Thompson,
dria of mammalian cells. Mol. Cell 55, 253–263. C.B. (2015). The utilization of extracellular proteins as nutrients is suppressed
Li, B., Qiu, B., Lee, D.S., Walton, Z.E., Ochocki, J.D., Mathew, L.K., Mancuso, by mTORC1. Cell 162, 259–270.
A., Gade, T.P., Keith, B., Nissim, I., and Simon, M.C. (2014). Fructose-1,6-bi- Phillips, M.M., Sheaff, M.T., and Szlosarek, P.W. (2013). Targeting arginine-
sphosphatase opposes renal carcinoma progression. Nature 513, 251–255. dependent cancers with arginine-degrading enzymes: opportunities and chal-
Loayza-Puch, F., Rooijers, K., Buil, L.C., Zijlstra, J., Oude Vrielink, J.F., Lopes, lenges. Cancer Res. Treat. 45, 251–262.
R., Ugalde, A.P., van Breugel, P., Hofland, I., Wesseling, J., et al. (2016). Piskounova, E., Agathocleous, M., Murphy, M.M., Hu, Z., Huddlestun, S.E.,
Tumour-specific proline vulnerability uncovered by differential ribosome Zhao, Z., Leitch, A.M., Johnson, T.M., DeBerardinis, R.J., and Morrison, S.J.
codon reading. Nature 530, 490–494. (2015). Oxidative stress inhibits distant metastasis by human melanoma cells.
Locasale, J.W., Grassian, A.R., Melman, T., Lyssiotis, C.A., Mattaini, K.R., Nature 527, 186–191.
Bass, A.J., Heffron, G., Metallo, C.M., Muranen, T., Sharfi, H., et al. (2011). Platten, M., von Knebel Doeberitz, N., Oezen, I., Wick, W., and Ochs, K. (2015).
Phosphoglycerate dehydrogenase diverts glycolytic flux and contributes to Cancer immunotherapy by targeting IDO1/TDO and their downstream effec-
oncogenesis. Nat. Genet. 43, 869–874. tors. Front. Immunol. 5, 673.
Losman, J.A., and Kaelin, W.G., Jr. (2013). What a difference a hydroxyl makes: Possemato, R., Marks, K.M., Shaul, Y.D., Pacold, M.E., Kim, D., Birsoy, K., Se-
mutant IDH, (R)-2-hydroxyglutarate, and cancer. Genes Dev. 27, 836–852. thumadhavan, S., Woo, H.K., Jang, H.G., Jha, A.K., et al. (2011). Functional ge-
Lunt, S.Y., Muralidhar, V., Hosios, A.M., Israelsen, W.J., Gui, D.Y., Newhouse, nomics reveal that the serine synthesis pathway is essential in breast cancer.
L., Ogrodzinski, M., Hecht, V., Xu, K., Acevedo, P.N., et al. (2015). Pyruvate ki- Nature 476, 346–350.
nase isoform expression alters nucleotide synthesis to impact cell prolifera- Qiu, B., Ackerman, D., Sanchez, D.J., Li, B., Ochocki, J.D., Grazioli, A., Bo-
tion. Mol. Cell 57, 95–107. brovnikova-Marjon, E., Diehl, J.A., Keith, B., and Simon, M.C. (2015). HIF2a-
Maddocks, O.D., Berkers, C.R., Mason, S.M., Zheng, L., Blyth, K., Gottlieb, E., dependent lipid storage promotes endoplasmic reticulum homeostasis in
and Vousden, K.H. (2013). Serine starvation induces stress and p53-depen- clear-cell renal cell carcinoma. Cancer Discov. 5, 652–667.
dent metabolic remodelling in cancer cells. Nature 493, 542–546. Rabinovich, S., Adler, L., Yizhak, K., Sarver, A., Silberman, A., Agron, S., Stettner,
Maher, E.A., Marin-Valencia, I., Bachoo, R.M., Mashimo, T., Raisanen, J., Ha- N., Sun, Q., Brandis, A., Helbling, D., et al. (2015). Diversion of aspartate in ASS1-
tanpaa, K.J., Jindal, A., Jeffrey, F.M., Choi, C., Madden, C., et al. (2012). Meta- deficient tumours fosters de novo pyrimidine synthesis. Nature 527, 379–383.
bolism of [U-13 C]glucose in human brain tumors in vivo. NMR Biomed. 25, Raj, L., Ide, T., Gurkar, A.U., Foley, M., Schenone, M., Li, X., Tolliday, N.J., Go-
1234–1244. lub, T.R., Carr, S.A., Shamji, A.F., et al. (2011). Selective killing of cancer cells by
Marin-Valencia, I., Yang, C., Mashimo, T., Cho, S., Baek, H., Yang, X.L., Raja- a small molecule targeting the stress response to ROS. Nature 475, 231–234.
gopalan, K.N., Maddie, M., Vemireddy, V., Zhao, Z., et al. (2012). Analysis of tu- Reitzer, L.J., Wice, B.M., and Kennell, D. (1980). The pentose cycle. Control
mor metabolism reveals mitochondrial glucose oxidation in genetically diverse and essential function in HeLa cell nucleic acid synthesis. J. Biol. Chem.
human glioblastomas in the mouse brain in vivo. Cell Metab. 15, 827–837. 255, 5616–5626.
Marjon, K., Cameron, M.J., Quang, P., Clasquin, M.F., Mandley, E., Kunii, K., Robbins, M.E., and Zhao, W. (2004). Chronic oxidative stress and radiation-
McVay, M., Choe, S., Kernytsky, A., Gross, S., et al. (2016). MTAP deletions in induced late normal tissue injury: a review. Int. J. Radiat. Biol. 80, 251–259.
cancer create vulnerability to targeting of the MAT2A/PRMT5/RIOK1 axis. Cell Saha, S.K., Parachoniak, C.A., Ghanta, K.S., Fitamant, J., Ross, K.N., Najem,
Rep. 15, 574–587. M.S., Gurumurthy, S., Akbay, E.A., Sia, D., Cornella, H., et al. (2014). Mutant
Mavrakis, K.J., McDonald, E.R., 3rd, Schlabach, M.R., Billy, E., Hoffman, G.R., IDH inhibits HNF-4a to block hepatocyte differentiation and promote biliary
deWeck, A., Ruddy, D.A., Venkatesan, K., Yu, J., McAllister, G., et al. (2016). cancer. Nature 513, 110–114.
Disordered methionine metabolism in MTAP/CDKN2A-deleted cancers leads Sasaki, M., Knobbe, C.B., Munger, J.C., Lind, E.F., Brenner, D., Brüstle, A.,
to dependence on PRMT5. Science 351, 1208–1213. Harris, I.S., Holmes, R., Wakeham, A., Haight, J., et al. (2012). IDH1(R132H)
Mayers, J.R., Torrence, M.E., Danai, L.V., Papagiannakopoulos, T., Davidson, mutation increases murine haematopoietic progenitors and alters epigenetics.
S.M., Bauer, M.R., Lau, A.N., Ji, B.W., Dixit, P.D., Hosios, A.M., et al. (2016). Nature 488, 656–659.
Tissue of origin dictates branched-chain amino acid metabolism in mutant Schafer, Z.T., Grassian, A.R., Song, L., Jiang, Z., Gerhart-Hines, Z., Irie, H.Y.,
Kras-driven cancers. Science 353, 1161–1165. Gao, S., Puigserver, P., and Brugge, J.S. (2009). Antioxidant and oncogene
Meiser, J., Tumanov, S., Maddocks, O., Labuschagne, C.F., Athineos, D., Van rescue of metabolic defects caused by loss of matrix attachment. Nature
Den Broek, N., Mackay, G.M., Gottlieb, E., Blyth, K., Vousden, K., et al. (2016). 461, 109–113.
Serine one-carbon catabolism with formate overflow. Sci. Adv. 2, e1601273. Sciacovelli, M., Gonçalves, E., Johnson, T.I., Zecchini, V.R., da Costa, A.S.,
Michelakis, E.D., Sutendra, G., Dromparis, P., Webster, L., Haromy, A., Niven, Gaude, E., Drubbel, A.V., Theobald, S.J., Abbo, S.R., Tran, M.G., et al.
E., Maguire, C., Gammer, T.L., Mackey, J.R., Fulton, D., et al. (2010). Metabolic (2016). Fumarate is an epigenetic modifier that elicits epithelial-to-mesen-
modulation of glioblastoma with dichloroacetate. Sci. Transl. Med. 2, 31ra34. chymal transition. Nature 537, 544–547.

668 Cell 168, February 9, 2017


Selak, M.A., Armour, S.M., MacKenzie, E.D., Boulahbel, H., Watson, D.G., Vazquez, F., Lim, J.H., Chim, H., Bhalla, K., Girnun, G., Pierce, K., Clish, C.B.,
Mansfield, K.D., Pan, Y., Simon, M.C., Thompson, C.B., and Gottlieb, E. Granter, S.R., Widlund, H.R., Spiegelman, B.M., and Puigserver, P. (2013).
(2005). Succinate links TCA cycle dysfunction to oncogenesis by inhibiting PGC1a expression defines a subset of human melanoma tumors with
HIF-alpha prolyl hydroxylase. Cancer Cell 7, 77–85. increased mitochondrial capacity and resistance to oxidative stress. Cancer
Sellers, K., Fox, M.P., Bousamra, M., 2nd, Slone, S.P., Higashi, R.M., Miller, Cell 23, 287–301.
D.M., Wang, Y., Yan, J., Yuneva, M.O., Deshpande, R., et al. (2015). Pyruvate Venneti, S., Dunphy, M.P., Zhang, H., Pitter, K.L., Zanzonico, P., Campos, C.,
carboxylase is critical for non-small-cell lung cancer proliferation. J. Clin. Carlin, S.D., La Rocca, G., Lyashchenko, S., Ploessl, K., et al. (2015). Gluta-
Invest. 125, 687–698. mine-based PET imaging facilitates enhanced metabolic evaluation of gliomas
Shackelford, D.B., Abt, E., Gerken, L., Vasquez, D.S., Seki, A., Leblanc, M., in vivo. Sci. Transl. Med. 7, 274ra17.
Wei, L., Fishbein, M.C., Czernin, J., Mischel, P.S., and Shaw, R.J. (2013). Viale, A., Pettazzoni, P., Lyssiotis, C.A., Ying, H., Sánchez, N., Marchesini, M.,
LKB1 inactivation dictates therapeutic response of non-small cell lung cancer Carugo, A., Green, T., Seth, S., Giuliani, V., et al. (2014). Oncogene ablation-
to the metabolism drug phenformin. Cancer Cell 23, 143–158. resistant pancreatic cancer cells depend on mitochondrial function. Nature
Shim, H., Dolde, C., Lewis, B.C., Wu, C.S., Dang, G., Jungmann, R.A., Dalla-Fa- 514, 628–632.
vera, R., and Dang, C.V. (1997). c-Myc transactivation of LDH-A: implications Weinberg, F., Hamanaka, R., Wheaton, W.W., Weinberg, S., Joseph, J., Lo-
for tumor metabolism and growth. Proc. Natl. Acad. Sci. USA 94, 6658–6663. pez, M., Kalyanaraman, B., Mutlu, G.M., Budinger, G.R., and Chandel, N.S.
Shim, E.H., Livi, C.B., Rakheja, D., Tan, J., Benson, D., Parekh, V., Kho, E.Y., (2010). Mitochondrial metabolism and ROS generation are essential for
Ghosh, A.P., Kirkman, R., Velu, S., et al. (2014). L-2-Hydroxyglutarate: an Kras-mediated tumorigenicity. Proc. Natl. Acad. Sci. USA 107, 8788–8793.
epigenetic modifier and putative oncometabolite in renal cancer. Cancer Dis- Wenzel, C., Riefke, B., Gründemann, S., Krebs, A., Christian, S., Prinz, F., Os-
cov. 4, 1290–1298. terland, M., Golfier, S., Räse, S., Ansari, N., et al. (2014). 3D high-content
Sonveaux, P., Végran, F., Schroeder, T., Wergin, M.C., Verrax, J., Rabbani, screening for the identification of compounds that target cells in dormant tu-
Z.N., De Saedeleer, C.J., Kennedy, K.M., Diepart, C., Jordan, B.F., et al. mor spheroid regions. Exp. Cell Res. 323, 131–143.
(2008). Targeting lactate-fueled respiration selectively kills hypoxic tumor cells Wheaton, W.W., Weinberg, S.E., Hamanaka, R.B., Soberanes, S., Sullivan,
in mice. J. Clin. Invest. 118, 3930–3942. L.B., Anso, E., Glasauer, A., Dufour, E., Mutlu, G.M., Budigner, G.S., and Chan-
Sousa, C.M., Biancur, D.E., Wang, X., Halbrook, C.J., Sherman, M.H., Zhang, del, N.S. (2014). Metformin inhibits mitochondrial complex I of cancer cells to
L., Kremer, D., Hwang, R.F., Witkiewicz, A.K., Ying, H., et al. (2016). Pancreatic reduce tumorigenesis. eLife 3, e02242.
stellate cells support tumour metabolism through autophagic alanine secre- White, E. (2013). Exploiting the bad eating habits of Ras-driven cancers. Genes
tion. Nature 536, 479–483. Dev. 27, 2065–2071.
Stams, W.A., den Boer, M.L., Beverloo, H.B., Meijerink, J.P., Stigter, R.L., van White, R.M., Cech, J., Ratanasirintrawoot, S., Lin, C.Y., Rahl, P.B., Burke, C.J.,
Wering, E.R., Janka-Schaub, G.E., Slater, R., and Pieters, R. (2003). Sensitivity Langdon, E., Tomlinson, M.L., Mosher, J., Kaufman, C., et al. (2011). DHODH
to L-asparaginase is not associated with expression levels of asparagine syn- modulates transcriptional elongation in the neural crest and melanoma. Nature
thetase in t(12;21)+ pediatric ALL. Blood 101, 2743–2747. 471, 518–522.
Sullivan, L.B., Gui, D.Y., Hosios, A.M., Bush, L.N., Freinkman, E., and Vander Witney, T.H., James, M.L., Shen, B., Chang, E., Pohling, C., Arksey, N.,
Heiden, M.G. (2015). Supporting aspartate biosynthesis is an essential func- Hoehne, A., Shuhendler, A., Park, J.H., Bodapati, D., et al. (2015). PET imaging
tion of respiration in proliferating cells. Cell 162, 552–563. of tumor glycolysis downstream of hexokinase through noninvasive measure-
Svensson, R.U., Parker, S.J., Eichner, L.J., Kolar, M.J., Wallace, M., Brun, ment of pyruvate kinase M2. Sci. Transl. Med. 7, 310ra169.
S.N., Lombardo, P.S., Van Nostrand, J.L., Hutchins, A., Vera, L., et al. Xiao, M., Yang, H., Xu, W., Ma, S., Lin, H., Zhu, H., Liu, L., Liu, Y., Yang, C., Xu,
(2016). Inhibition of acetyl-CoA carboxylase suppresses fatty acid synthesis Y., et al. (2012). Inhibition of a-KG-dependent histone and DNA demethylases
and tumor growth of non-small-cell lung cancer in preclinical models. Nat. by fumarate and succinate that are accumulated in mutations of FH and SDH
Med. 22, 1108–1119. tumor suppressors. Genes Dev. 26, 1326–1338.
Sykes, D.B., Kfoury, Y.S., Mercier, F.E., Wawer, M.J., Law, J.M., Haynes, Ye, X., and Weinberg, R.A. (2015). Epithelial-mesenchymal plasticity: a central
M.K., Lewis, T.A., Schajnovitz, A., Jain, E., Lee, D., et al. (2016). Inhibition of regulator of cancer progression. Trends Cell Biol. 25, 675–686.
dihydroorotate dehydrogenase overcomes differentiation blockade in acute Ying, H., Kimmelman, A.C., Lyssiotis, C.A., Hua, S., Chu, G.C., Fletcher-San-
myeloid leukemia. Cell 167, 171–186.e15. anikone, E., Locasale, J.W., Son, J., Zhang, H., Coloff, J.L., et al. (2012). Onco-
Szenajch, J., Wcislo, G., Jeong, J.Y., Szczylik, C., and Feldman, L. (2010). The genic Kras maintains pancreatic tumors through regulation of anabolic glucose
role of erythropoietin and its receptor in growth, survival and therapeutic metabolism. Cell 149, 656–670.
response of human tumor cells From clinic to bench - a critical review. Bio- Yuneva, M., Zamboni, N., Oefner, P., Sachidanandam, R., and Lazebnik, Y.
chim. Biophys. Acta 1806, 82–95. (2007). Deficiency in glutamine but not glucose induces MYC-dependent
Tateishi, K., Wakimoto, H., Iafrate, A.J., Tanaka, S., Loebel, F., Lelic, N., Wie- apoptosis in human cells. J. Cell Biol. 178, 93–105.
derschain, D., Bedel, O., Deng, G., Zhang, B., et al. (2015). Extreme vulnera- Yuneva, M.O., Fan, T.W., Allen, T.D., Higashi, R.M., Ferraris, D.V., Tsukamoto,
bility of IDH1 mutant cancers to NAD+ depletion. Cancer Cell 28, 773–784. T., Matés, J.M., Alonso, F.J., Wang, C., Seo, Y., et al. (2012). The metabolic
Titov, D.V., Cracan, V., Goodman, R.P., Peng, J., Grabarek, Z., and Mootha, profile of tumors depends on both the responsible genetic lesion and tissue
V.K. (2016). Complementation of mitochondrial electron transport chain by type. Cell Metab. 15, 157–170.
manipulation of the NAD+/NADH ratio. Science 352, 231–235. Zhang, G., Frederick, D.T., Wu, L., Wei, Z., Krepler, C., Srinivasan, S., Chae, Y.C.,
Tomlinson, I.P., Alam, N.A., Rowan, A.J., Barclay, E., Jaeger, E.E., Kelsell, D., Xu, X., Choi, H., Dimwamwa, E., et al. (2016). Targeting mitochondrial biogenesis
Leigh, I., Gorman, P., Lamlum, H., Rahman, S., et al.; Multiple Leiomyoma to overcome drug resistance to MAPK inhibitors. J. Clin. Invest. 126, 1834–1856.
Consortium (2002). Germline mutations in FH predispose to dominantly in- Zhang, W., Trachootham, D., Liu, J., Chen, G., Pelicano, H., Garcia-Prieto, C.,
herited uterine fibroids, skin leiomyomata and papillary renal cell cancer. Lu, W., Burger, J.A., Croce, C.M., Plunkett, W., et al. (2012). Stromal control of
Nat. Genet. 30, 406–410. cystine metabolism promotes cancer cell survival in chronic lymphocytic
Vander Heiden, M.G. (2011). Targeting cancer metabolism: a therapeutic win- leukaemia. Nat. Cell Biol. 14, 276–286.
dow opens. Nat. Rev. Drug Discov. 10, 671–684. Zheng, L., Cardaci, S., Jerby, L., MacKenzie, E.D., Sciacovelli, M., Johnson,
Vander Heiden, M.G., Cantley, L.C., and Thompson, C.B. (2009). Understand- T.I., Gaude, E., King, A., Leach, J.D., Edrada-Ebel, R., et al. (2015). Fumarate
ing the Warburg effect: the metabolic requirements of cell proliferation. Sci- induces redox-dependent senescence by modifying glutathione metabolism.
ence 324, 1029–1033. Nat. Commun. 6, 6001.

Cell 168, February 9, 2017 669

You might also like