Anda di halaman 1dari 11

Review

Mitochondria and Hypoxia: Metabolic


Crosstalk in Cell-Fate Decisions
David Bargiela,1,2 Stephen P. Burr,1,2 and Patrick F. Chinnery1,2,*

Alterations in mitochondrial metabolism influence cell differentiation and Highlights


growth. This process is regulated by the activity of 2-oxoglutarate (2OG)- Alignment of transcriptional activity
dependent dioxygenases (2OGDDs) – a diverse superfamily of oxygen-con- with nutrient availability is crucial to
maintain homeostasis in mammalian
suming enzymes – through modulation of the epigenetic landscape and tran- cells. The 2-oxoglutarate-dependent
scriptional responses. Recent reports have described the role of mitochondrial dioxygenase (2OGDD) superfamily
integrates mitochondrial metabolic
metabolites in directing 2OGDD-driven cell-fate switches in stem cells (SCs), signals to coordinate downstream
immune cells, and cancer cells. An understanding of the metabolic mecha- transcriptional responses.
nisms underlying 2OGDD autoregulation is required for therapeutic targeting of
Oxygen and mitochondrial metabolite
this system. We propose a model dependent on oxygen and metabolite avail- concentrations signal via 2OGDDs to
ability and discuss how this integrates 2OGDD metabolic signalling, the hyp- alter DNA/histone methylation and
hypoxic transcriptional activity.
oxic transcriptional response, and fate-determining epigenetic changes.
Targeting 2OGDD activity affects cell
fate in vivo and has relevance in the
Introduction treatment of immune diseases and
Oxygen is essential for almost all animal life on the planet, allowing aerobic metabolism and the cancer.

efficient production of ATP via oxidative phosphorylation (OXPHOS). Consequently, multicel-


lular organisms have evolved a number of mechanisms to rapidly adapt to decreased oxygen
availability (termed hypoxia), prolonging survival in the absence of this critical resource.

Mitochondria serve as metabolic hubs in the cell and are intimately linked with the adaptive
response to hypoxia. A mitochondrial–nuclear link via 2OGDD enzymes allows crosstalk
between mitochondrial metabolism and nuclear transcriptional activity. The 2OGDD enzyme
family includes regulators of hypoxia-inducible factor (HIF), the master transcription factor
orchestrating cellular adaptation to hypoxia, as well as DNA and histone demethylase enzymes
involved in modifying the epigenetic landscape [1]. Recent work has shown that modulation of
2OGDD activity through these pathways is able to shape cell fate, with particular relevance to
immunity and cancer.

In this review we describe the key interactions between mitochondria and the hypoxia
response, focussing on the regulation of transcription by mitochondrial metabolism and the
2OGDD family. We offer a model of metabolic cross-regulation of 2OGDD family members,
1
MRC Mitochondrial Biology Unit,
illustrating cellular mechanisms that may represent promising future therapeutic targets.
Cambridge Biomedical Campus,
Cambridge CB2 0XY, UK
2
Hypoxia, HIFs, and the 2OGDDs Department of Clinical
Neurosciences, Cambridge
Several cellular mechanisms are involved in the adaptation to acute hypoxia, such as activation of ion
Biomedical Campus, University of
channels via gaseous signalling in carotid body glomus cells [2] and the direct upregulation of Cambridge, Cambridge CB2 0QQ, UK
glycolysis by AMPK in cardiac myocytes [3]. However, the major pathway controlling the cellular
response to hypoxia occurs at the transcriptional level and is coordinated through a number of key
*Correspondence:
transcriptions factors, resulting in upregulation of genes involved in metabolism, angiogenesis and,
pfc25@medschl.cam.ac.uk
haematopoiesis [4–6]. To date, the best studied of these transcriptional regulators are the HIFs. (P.F. Chinnery).

Trends in Endocrinology & Metabolism, April 2018, Vol. 29, No. 4 https://doi.org/10.1016/j.tem.2018.02.002 249
© 2018 Elsevier Ltd. All rights reserved.
HIFs are members of the basic helix–loop–helix (bHLH)/Per–Arnt–Sim (PAS) family of transcription
factors and form functional heterodimers comprising one of three known HIFa subunits and the
HIFb/aryl hydrocarbon receptor nuclear translocator (ARNT) subunit [6]. Active HIF heterodimers
formed in the nucleus bind DNA at conserved HIF-responsive elements (HREs) located in the
enhancer and promoter regions of over 200 HIF-target genes [7,8], including erythropoietin (EPO)
[9], vascular endothelial growth factor (VEGF) [10], and glucose transporter 1 (GLUT-1) [11],
enhancing the transcription of these genes and enabling tissues to adapt to hypoxia.

Wang and Semenza first identified that oxygen-dependent control of this transcriptional
response is conferred through the labile HIFa subunits (the major mammalian isoforms being
HIF1a and HIF2a/EPAS), which are stabilised in hypoxia but rapidly turned over under aerobic
conditions [9,12]. The key oxygen-sensing component of this pathway involves post-transla-
tional hydroxylation of two conserved proline residues of the HIFa protein by a family of prolyl
hydroxylase domain-containing proteins (PHDs) [13,14]. Following hydroxylation, the von
Hippel–Lindau (VHL) E3 ligase complex polyubiquitinates both HIF1a and HIF2a in an oxy-
gen-dependent manner [15,16]. This facilitates rapid proteasomal degradation of the HIFa
subunits in normoxia but is abolished on exposure to hypoxia, leading to stabilisation and
activation of HIF signalling [16,17] (Figure 1A).

PHDs are a critical component of the canonical hypoxia response pathway and also form part of a
larger group of 2OGDDs (Box 1). This enzyme family includes the asparaginyl hydroxylase factor
inhibiting HIF (FIH), a regulator of HIF transcriptional activity [18], as well as epigenetic modulators –
the Jumonji C-domain-containing histone demethylases (JHDs) [19,20] and the ten–eleven
translocation (TET) dioxygenase enzymes [21–23], which promote DNA demethylation. As their
name suggests, all 2OGDD members have a functional requirement for 2OG and oxygen, and
additionally utilise ferrous iron and ascorbate as cofactors [24] (Figure 1B, inset). Therefore,
changes in cofactor levels, such as reduced oxygen availability in hypoxic environments, result
in altered activity of these enzymes. Furthermore, competition for a finite cofactor pool may also
represent a level of regulation, integrating information on oxygen, metabolite, and macromolecule
availability with epigenomic and transcriptomic programmes, as discussed below.

Mitochondria and Hypoxia: Crosstalk via 2OGDDs


Mitochondria are metabolic hubs in cells and generate energy in the form of ATP by OXPHOS
[38]. As a major consumer of oxygen, mitochondrial OXPHOS has a significant impact on the
availability of molecular oxygen to regulate the hypoxic response and other 2OGDD-dependent
reactions. Multiple studies have shown that reduced activity of mitochondrial complex I,
through either chemical inhibition or genetic mutation, decreases mitochondrial oxygen con-
sumption, increasing oxygen availability in the cytoplasm, and destabilises both HIF1a and
HIF2a under hypoxic conditions [39–42]. There is evidence that cells may be able to actively
redistribute oxygen under hypoxic conditions and modulate HIF signalling through production
of endogenous ETC inhibitors, such as nitric oxide [43], or altered expression of proteins that
regulate mitochondrial oxygen consumption [44] (Figure 1A). Furthermore, HIF-mediated
transcription of PDK1 and LDHA is able to limit the entry of acetyl-CoA into the oxidative
tricarboxylic acid (TCA) cycle (by blocking the conversion of pyruvate to acetyl-CoA via PDK1
inhibition of PDH) and instead switches cells towards glycolytic metabolism [45–48].

A number of HIF-dependent mechanisms in turn also feed back on mitochondrial function,


highlighting the reciprocal relationship between these two key cellular processes. HIFs have
been shown to directly regulate the expression of a number of ETC subunits [49,50] and the
HIF-responsive noncoding miRNA-210 has been implicated in suppression of the iron–sulphur

250 Trends in Endocrinology & Metabolism, April 2018, Vol. 29, No. 4
(A) Normoxia Hypoxia

HydroxylaƟon
HIF1α
HIF1α
DimerisaƟon
UbiquiƟnaƟon OH
HIF1α
PHD
HIF1α HIF1β
Poly-Ub OH
HIF1α
VHL
HIF1β TranscripƟon
DegradaƟon
HRE

Nucleus
Proteasome

(B) 2OGDDs
O2 + + CO2
O2
O2 s
ink TETs EpigeneƟc
2-OG + Fe(II)
+ Ascorbate
Succinate

2OG remodelling PHDs


TCA cycle
JHDs HIF1α
ETC HIF1β TranscripƟon
HIF O O
ROS HRE
signaling
Mitochondrion Succinate PHDs HIF
N N HIF–OH
Fumarate Nucleus (stable)
(degraded)
2HG
ETC OH

components

Figure 1. 2-Oxoglutarate (2OG)-Dependent Dioxygenases (2OGDDs) Link Mitochondria and Transcriptional Control. Hypoxia-inducible factor 1 alpha
(HIF1a) is stabilised in hypoxic conditions, dimerises with HIF1 beta (HIF1b)/aryl hydrocarbon receptor nuclear translocator (ARNT), and binds to hypoxia response
elements (HREs) to enhance transcription at target genes (A, right). Under normoxic conditions (A, left), HIF1a undergoes prolyl hydroxylase domain-containing protein
(PHD)-mediated hydroxylation and polyubiquitinylation by the von Hippel–Lindau (VHL) E3 ligase complex followed by proteasomal degradation. 2OGDD’s regulation of
transcriptional activity is dependent on oxygen and 2OG and is inhibited by mitochondrial metabolites [succinate, fumarate, 2-hydroxyglutarate (2HG)] and reactive
oxygen species (ROS) (B). 2OGDDs convert 2OG and oxygen (O2) to succinate and carbon dioxide (CO2), respectively (B, inset). Ferrous iron [Fe (II)] and ascorbate serve
as further cofactors. Abbreviations: ETC, electron transport chain; JHDs, Jumonji domain-containing histone demethylase enzymes; Me, methyl; TCA, tricarboxylic
acid; TETs, ten–eleven translocation dioxygenase enzymes.

cluster assembly proteins ISCU1 and 2, which are critical for the function of a number of ETC
components [51,52]. These feedback loops between mitochondria and the HIFs are likely to be
involved in the fine-tuning of both metabolic and transcriptional responses to hypoxia.

Reactive oxygen species (ROS) from the mitochondrial ETC (mROS) were originally thought to
be merely a harmful byproduct of OXPHOS, but it is now clear that they function as signalling
molecules in a wide range of cellular processes. The production of mROS during hypoxia may
represent an important mito-hypoxia signalling mechanism. Chandel et al. first reported that
mROS produced by mitochondrial complex III during hypoxia are able to inhibit PHDs and
stabilise HIFa, suggesting a role for mROS in the activation of HIF under hypoxic conditions
[53]. Later studies supported this mechanism, showing that blockade of mROS production,
either genetically or pharmacologically, resulted in impaired induction of HIF in hypoxia [54–56].
mROS-mediated inhibition of HIF has been proposed to occur via oxidation of the cysteine
residues or Fe(II) cofactor [57] required for PHD2 activity [57–59] or via inhibition of FIH [60].

Trends in Endocrinology & Metabolism, April 2018, Vol. 29, No. 4 251
Box 1. The 2OGDDs
The 2OGDDs are a family of over 60 enzymes responsible for catalysing a number of key oxidative modifications in a
diverse range of target proteins [25]. The active site of the 2OGDDs comprises a characteristic ‘jelly-roll’ motif
comprising two b sheets within which an Fe(II) atom is bound by conserved histidine and aspartate residues [26].
The iron-bound holoenzyme binds 2-OG to form a stable enzyme–Fe(II)–2-OG complex [27], which then associates with
the target protein. In the presence of molecular oxygen, 2-OG is decarboxylated, resulting in the release of CO2,
production of succinate, and oxidation of the target [24]:
RH þ O2 þ C5 H6 O5 ! ROH þ CO2 þ C4 H6 O4 : [I]

In addition to Fe(II) and 2-OG, many of the 2OGDDs also utilise ascorbate as a cofactor, although this may not be
absolutely required [25].

The 2OGDDs catalyse a range of important post-translational protein modifications [28] and this family of enzymes is
responsible for regulating a number of key intracellular pathways in both prokaryotic and eukaryotic cells [29]. Perhaps
the best-studied of the 2OGDD family members in recent years include the HIF PHDs responsible for oxygen sensing
and control of the hypoxic response [30,31], the TET dioxygenases, which hydroxylate 5-methylcytosine and promote
DNA demethylation [32,33], and the Jumonji-C demethylases, which play a crucial role in histone demethylation [34,35].

Since these enzymes require both oxygen and the mitochondrial intermediate metabolite 2OG to function, their activity
can be significantly influenced by the metabolic state of the cell, providing a key link between metabolism and major
intracellular signalling pathways such as the hypoxic response and epigenetic regulation [36,37].

The regulation of OXPHOS by metabolic intermediates and oxygen allows mitochondria to


respond to alterations in nutrient availability, re-establishing homeostasis by redirecting metabolic
flux to compensate. This metabolic reprogramming may represent an important mechanism of
mitochondrial–nuclear signalling whereby the energy status of the cell is kept in sync with the
epigenetic and transcriptional responses required for cell proliferation and differentiation. These
responses in turn influence oxygen supply and usage, tightly regulating ROS, which, although
functioning as key signalling molecules, may also be toxic to the cell when in excess. The position of
2OG at the intersection of catabolic pathways facilitates its role as a rheostat of glucose and fatty
acid metabolism (via TCA cycle activity) as well as glutamine metabolism (by direct production via
glutamate). The availability and flux of 2OG is then able to regulate 2OGDD activity by altering the
ratio of promoting and inhibiting metabolites, discussed further in the following section.

The 2OGDD Shared Cofactor Pool (SCP) Model: Evidence and Implications
Metabolic regulation of transcriptional activity via 2OGDDs is involved in determining cell fate in
development and disease. In this section we briefly discuss another metabolic–transcriptional
process, the removal of epigenetic marks via 2OG-dependent demethylase enzymes, the
activity of which relates to the mitochondrial HIF response through a shared dependency on
2OGDD substrates and cofactors. By integrating these 2OGDD-dependent processes, we
propose a cross-regulation model that allows 2OGDD cofactor levels and activity to be
coordinated across multiple pathways (Figure 2).

As the epigenetic modulators of the 2OGDD family, the JHDs and TET demethylases (TETs) aid
the removal of methylation marks from histones and DNA, respectively, oxidising repressive
methyl groups to a hydroxymethyl form as an initial step in this process [20–22].

2OG serves as a co-substrate for this reaction and is converted to succinate and then fumarate.
Due to their structural similarity to 2OG, succinate and fumarate are able to inhibit enzyme
activity by competing for the 2OG-binding site [61], forming a regulatory loop (Figure 2B).

The same competitive inhibition occurs in the regulation of the hypoxia HIF transcriptional
response. PHDs and FIH require 2OG and other 2OGDD cofactors to regulate the abundance

252 Trends in Endocrinology & Metabolism, April 2018, Vol. 29, No. 4
(A) Mitochondrial metabolism
Mitochondrial metabolites
Forward Reverse Glucose
Glucose
(oxidaƟve) flux
Macromolecular synthesis (reducƟve) flux FaƩy acid
synthesis

Citrate Acetyl-CoA

TCA cycle TCA cycle


2HG
ETC ETC
2OG
Fumarate 2OG O2
ROS
Glutamine Glutamine
Succinate 2OGDD shared
co-factor pool
NucleoƟde/protein
synthesis
Fe2+
CompeƟƟve 2OG CompeƟƟve
inhibiƟon Ascorbate inhibiƟon
(B) EpigeneƟc remodelling (C) HIF acƟvity
O2 O2
2OG
CO2 Proteosomal
Succinate PHD OH
O2 degradaƟon
JHD 2OG
Succinate
hMe
Me CO2 HIF1α
Cytosol
TET/JHD PHD/FIH O2 Nucleus
2OG
acƟvity acƟvity HIF1α/β
FIH

5mC 5hmC
Succinate TranscripƟon
CO2 +
TET Succinate
2OG CompeƟƟve
O2 CO2 HRE
inhibiƟon

Figure 2. The 2-Oxoglutarate (2OG)-Dependent Dioxygenase (2OGDD) Shared Cofactor Pool Model. 2OG is oxidised to form succinate and then
fumarate or reduced to form citrate or 2-hydroxyglutarate (2HG) (A). The direction of tricarboxylic acid (TCA) flux depends on the redox environment and the availability of
downstream metabolites: an increased 2OG:succinate ratio promotes oxidative flux while an increased 2OG:citrate ratio promotes reductive flux. TCA activity relates to
oxidative phosphorylation (OXPHOS) activity via the redox state of electron carriers (e.g., NADH/FADH2) and oxygen availability in the mitochondria. 2OGDDs, involved
in epigenetic remodelling [mediated by the oxidase activity of ten–eleven translocation dioxygenase (TET)/Jumonji domain-containing histone demethylase (JHD)
enzymes] (B), and regulation of the hypoxia/hypoxia-inducible factor (HIF) response (C) both depend on 2OG and cofactor availability for their activity. 2OGDDs
competitively inhibit each other due to a limited supply of 2OGDD cofactors in the cell. Abbreviations: 5hmC, 5-hydroxymethylcytosine; 5mC, 5-methylcytosine; ETC,
electron transport chain; FIH, factor inhibiting HIF; hME, hydroxymethyl; HRE, hypoxia-response element; Me, methyl; PHD, prolyl hydroxylase enzymes; ROS, reactive
oxygen species.

and transcriptional activity of HIF, respectively. Competition for the 2OG-binding site by
succinate or fumarate inhibits the action of PHDs, causing stabilisation of HIF1 under normoxic
conditions, termed pseudohypoxia [62,63]. In tumours with mutations in mitochondrial succi-
nate dehydrogenase (SDH) or fumarate hydratase (FH), succinate or fumarate accumulate
resulting in the inhibition of PHDs and promotion of protumorigenic HIF transcription, including
VEGF-mediated angiogenesis and glycolytic metabolism [63,64]. Likewise, high levels of
succinate and fumarate inhibit the TET and JHD enzymes, leading to hypermethylation of
cell-differentiation genes and promoting malignant growth [65–67]. In both cases inhibition via
succinate or fumarate can be rescued by the addition of cell-permeable 2OG, restoring 2OGDD
activity by outcompeting the inhibitory metabolites [68,69].

Control of the intracellular 2OG:succinate ratio has also been shown to be crucial for the
maintenance of SC pluripotency. Initial studies in mouse embryonic SCs (mESCs) identified
elevated 2OG levels in naïve mESCs resulting in an elevated 2OG:succinate ratio and conse-
quent promotion of histone and DNA demethylation via increased TET/JHD activity [70].
Manipulation of the 2OG:succinate ratio directly impacts mESC cell-fate determination, with

Trends in Endocrinology & Metabolism, April 2018, Vol. 29, No. 4 253
the addition of cell-permeable 2OG or succinate being sufficient to promote mESC pluripotency
or differentiation, respectively [70,71]. The role of 2OG in maintaining stemness has also been
explored in naïve human pluripotent SCs (hPSCs) [72]. However, primed SCs from both mice
and humans have been shown to respond differently, with increased 2OG levels promoting
differentiation rather than maintaining pluripotency despite similar increases in TET and JHD
activity.

2OG can also undergo reductive carboxylation to isocitrate via the enzyme isocitrate dehy-
drogenase (IDH) (Figure 2A). Reductive flux is promoted by increases in the NADH:NAD+ and
2OG:citrate ratios; however, it is dependent on continued oxidation of 2OG to maintain
sufficient levels of NADPH [73]. Hypoxia also promotes reductive carboxylation of 2OG,
providing a crucial carbon source for fatty acid synthesis [74,75]. In cancer cells with mito-
chondrial respiratory chain complex defects, reductive metabolism of 2OG promotes prolifer-
ation [76], and this process may also influence tumour invasiveness by supporting anchorage-
independent growth [7].

Somatic gain-of-functions mutation in IDH, detected in haematological malignancies and solid


tumours, result in the conversion of 2OG to the oncometabolite R-2-hydroxyglutarate (R-2HG)
[77,78]. Structurally similar to 2OG, 2HG is a chiral molecule that exists as two enantiomers, S-
and R-2HG (also known as L- and D-2HG, respectively). Both are produced endogenously in
cells, as evidenced by the presence of specific dehydrogenase enzymes for both the S- and R-
forms [79,80]. However, the biochemical properties of the two enantiomers appear to differ
significantly. R-2HG, but not S-2HG, drives malignant transformation in the context of acute
myeloid leukaemia with IDH mutations [81] despite both 2-HG enantiomers being able to block
TET enzymes [82–84]. This differential response in promoting leukaemogenesis is attributed to
the ability of S-2HG, but not R-2HG, to inhibit PHD enzymes [81]. However, the role of R-2HG in
regulating 2-OGDDs under physiological conditions remains unclear.

S-2HG demonstrates potent inhibition of 2OGDDs both in vitro and in vivo. Metabolomic
profiling of cells in hypoxia revealed significant accumulation of S-2HG [85–87], with carbon-flux
analysis indicating that the source of S-2HG was 2OG derived from glutamine [86]. The
mechanism responsible for the conversion of 2OG to S-2HG appears to depend on the activity
of lactate dehydrogenase A (LDHA) and malate dehydrogenase (MDH) 1 and 2, with all three of
these enzymes potentially contributing to S-2HG production [85,86], although the choice of
enzyme may be cell-type specific [87]. This promiscuous substrate usage by LDHA and MDH1/
2 is favoured at low pH [88,89] and in the context of a low NAD+:NADH ratio [85,90], both of
which occur in hypoxia.

Elevated S-2HG levels have also been seen in normoxic conditions following disruption of the
oxoglutarate dehydrogenase complex (OGDHC) [91] and in fetal mouse haematopoietic SCs
with mitochondrial complex III deficiency [90], suggesting that metabolic defects can create a
permissive redox environment facilitating the accumulation of S-2HG. Tyrakis et al. recently
showed that the activation signal in T cells is sufficient to cause an increase in intracellular S-
2HG levels under normoxic conditions. This effect was amplified when cells were cultured in
hypoxia, resulting in inhibition of TET enzymes and the acquisition of a CD8+ memory-like
phenotype [87]. It is clear that a number of complex mechanisms regulate 2HG concentrations
and enantiomer ratios; further work is required to clarify how these processes may skew
immune responses in autoimmune disease or cancer.

254 Trends in Endocrinology & Metabolism, April 2018, Vol. 29, No. 4
Concluding Remarks and Therapeutic Prospects Outstanding Questions
The involvement of 2OGDDs in pathological processes such as cancer, inflammation, and How do primary mitochondrial disor-
anaemia have made them attractive drug targets. Pharmacological efforts to date have been ders alter 2OGDD’s regulation of cell
fate?
directed towards the dioxygenase enzymes themselves as well as the cofactors that regulate
their activity. How do the actions of the 2-HG enan-
tiomers R-2HG and S-2HG differ in
Small-molecule inhibitors of mutant IDH were developed with the aim of inhibiting 2HG- vivo and what therapeutic implications
does this have in autoimmune disease
mediated inhibition of TET/JHD activity. These compounds showed encouraging results in
and cancer?
vitro using patient-derived leukaemia cells and in vivo in a murine xenograft model, in both
cases reversing aberrant methylation to restore cell differentiation [81,92–94]. Clinical trials in How does 2OGDD-driven HIF metab-
patients with IDH1/2-mutant cancers demonstrated a profile similar to preclinical studies, olism feed back to regulate 2OGDD
showing clinical evidence of restoration of cell differentiation [95,96], and recently enasidenib activity?
(Idhifa, Celgene Corp.) has gained FDA approval as a treatment for patients with relapsed or
What are the molecular mechanisms
refractory AML with IDH2 mutations [97]. Similarly, a number of inhibitors of the PHD enzymes
coordinating 2OGDD–HIF–TET/JHDM
are in Phase II/III clinical trials for the treatment of renal anaemia. These compounds target activity with other metabolic–transcrip-
PHD2 and stimulate endogenous production of EPO via HIF2a-mediated transcription to tional axes (e.g., acetyl-CoA–DNA/his-
promote red blood cell production [98]. However, the activation of other aspects of HIF tone acetylation) in the cell?

transcription, such as angiogenesis and reduction in ROS production, may allow repurposing
of these drugs for other indications; for example, to limit damage from ischaemic injury [99–
102].

Altering oxygen and cofactor concentrations to modulate 2OGDD activity may represent an
alternative therapeutic strategy. Mootha and colleagues recently demonstrated that blockade
of VHL, a member of the hypoxia response pathway downstream of 2OGDDs, may confer
survival benefits in cells with mitochondrial respiratory chain inhibition [103]. They were able to
replicate this therapeutic response in vivo by exposing mice with mitochondrial respiratory
chain complex defects to low oxygen concentrations that limited VHL activity. The use of
moderate hypoxia to treat primary and secondary mitochondrial diseases in humans is an
intriguing possibility, although questions regarding tolerability, off-target effects, and the
duration of hypoxic exposure required for therapeutic effect will need to be ascertained.

Ascorbic acid (AA) (also known as vitamin C), a 2OGDD cofactor, also modulates the hypoxia
response pathway by promoting PHD activity through reduction of the enzyme’s inactive Fe(III)
state to an active Fe(II) state, thus rescuing its catalytic potential [104]. It may act similarly in
promoting TET-dependent 5hmC formation [105–109], although whether it modulates TETs
directly or indirectly via nonspecific reduction of the cellular iron pool has been questioned
[110,111]. Due to the potential antitumorigenic effects of TET-mediated promotion of differen-
tiation or PHD-mediated restriction of the HIF translational response, there has been renewed
interest in the potential use of AA as a cancer treatment [112,113]. High-dose intravenous AA
therapy has been shown to inhibit tumor growth in rodent xenograft models [114–116] and was
well tolerated when administered to pancreatic cancer patients in two small clinical studies
[117,118]. A Phase II trial will assess the effects of AA as an adjunct to chemotherapy for
pancreatic cancer (ClinicalTrials.gov NCT02905578). Regarding hypoxic therapy, further work
will be required to delineate the mechanism driving AA-mediated cytotoxicity, whether this
effect is tissue-specific, and what represents a therapeutic dose of AA. Delivery of AA to the
tumour site, regulation of intracellular AA concentration, and limiting the potential off-target
effects are all challenges that remain to be overcome.

Screening assays will be required to determine the selectivity of new compounds in targeting
2OGDD family members, to avoid the significant potential risk of undesired side-effects by

Trends in Endocrinology & Metabolism, April 2018, Vol. 29, No. 4 255
inhibition of other 2OGDD-dependent processes [119,120]. However, concurrent inhibition or
activation of multiple 2OGDD pathways may be desirable in certain situations. As an example in
the treatment of cancer, combined activation of the PHD and TET enzymes may result in
reduced HIF-mediated angiogenesis and encourage cell differentiation, respectively, with the
overall effect of limiting malignant growth. Further work will be required to understand how to
specifically target malignant cells to maximise the benefits of these types of drug approaches.
Once optimised this approach may be relevant in the treatment of other common multifactorial
diseases with known mitochondrial involvement. Mitochondrial dysfunction may contribute to
pathology in these conditions by altering metabolic signalling through 2OGDD pathways and
thus may be amenable to treatments that redirect metabolic flux.

Finally, there is a need to explore the coordination of the 2OGDD SCP model in the context of
other metabolic–epigenetic signalling pathways, to predict more accurately the consequences
of metabolic programme switches (see Outstanding Questions). For example, the reductive flux
of 2OG to citrate and subsequent conversion to cytoplasmic acetyl-CoA may indirectly affect
histone acetylation. For further insight on the compartmental regulation of acetylation, the
reader is directed towards an excellent recent review [121]. An integrative study of metabolism
in cell-fate decision promises to have far-reaching potential, both in our understanding of cell
development and differentiation and the treatment of these processes when they become
dysfunctional.

Acknowledgements
P.F.C. is a Wellcome Trust Senior Fellow in Clinical Science (101876/Z/13/Z) and a UK National Institute for Health
Research (NIHR) Senior Investigator who receives support from the Medical Research Council Mitochondrial Biology Unit
(MC_UP_1501/2), the Medical Research Council (UK) Centre for Translational Muscle Disease (G0601943), and the NIHR
Biomedical Research Centre based at Cambridge University Hospitals NHS Foundation Trust and the University of
Cambridge. The views expressed are those of the authors and not necessarily those of the NHS, the NIHR, or the
Department of Health.

References
1. Hausinger, R.P. (2015) Biochemical diversity of 2-oxoglutarate- 10. Forsythe, J.A. et al. (1996) Activation of vascular endothelial
dependent oxygenases. In 2-Oxoglutarate-Dependent Oxygen- growth factor gene transcription by hypoxia-inducible factor 1.
ases, pp. 1–58, Royal Society of Chemistry Mol. Cell. Biol. 16, 4604–4613
2. Prabhakar, N.R. and Semenza, G.L. (2012) Gaseous messen- 11. Ebert, B.L. et al. (1995) Hypoxia and mitochondrial inhibitors
gers in oxygen sensing. J. Mol. Med. (Berl.) 90, 265–272 regulate expression of glucose transporter-1 via distinct cis-
3. Marsin, A.S. et al. (2000) Phosphorylation and activation of heart acting sequences. J. Biol. Chem. 270, 29083–29089
PFK-2 by AMPK has a role in the stimulation of glycolysis during 12. Wang, G.L. and Semenza, G.L. (1995) Purification and charac-
ischaemia. Curr. Biol. 10, 1247–1255 terization of hypoxia-inducible factor 1. J. Biol. Chem. 270,
4. Semenza, G.L. et al. (1994) Transcriptional regulation of genes 1230–1237
encoding glycolytic enzymes by hypoxia-inducible factor 1. J. 13. Epstein, A.C. et al. (2001) C. elegans EGL-9 and mammalian
Biol. Chem. 269, 23757–23763 homologs define a family of dioxygenases that regulate HIF by
5. Liu, Y. et al. (1995) Hypoxia regulates vascular endothelial prolyl hydroxylation. Cell 107, 43–54
growth factor gene expression in endothelial cells. Identification 14. Bruick, R.K. and McKnight, S.L. (2001) A conserved family of
of a 5' enhancer. Circ. Res. 77, 638–643 prolyl-4-hydroxylases that modify HIF. Science 294, 1337–1340
6. Wang, G.L. et al. (1995) Hypoxia-inducible factor 1 is a basic- 15. Ivan, M. et al. (2001) HIFa targeted for VHL-mediated destruc-
helix–loop–helix–PAS heterodimer regulated by cellular O2 ten- tion by proline hydroxylation: implications for O2 sensing. Sci-
sion. Proc. Natl. Acad. Sci. U. S. A. 92, 5510–5514 ence 292, 464–468
7. Jiang, L. et al. (2016) Reductive carboxylation supports redox 16. Maxwell, P.H. et al. (1999) The tumour suppressor protein VHL
homeostasis during anchorage-independent growth. Nature targets hypoxia-inducible factors for oxygen-dependent prote-
532, 255–258 olysis. Nature 399, 271–275
8. Wenger, R.H. et al. (2005) Integration of oxygen signaling at the 17. Huang, L.E. et al. (1998) Regulation of hypoxia-inducible factor
consensus HRE. Sci. STKE 2005, re12 1a is mediated by an O2-dependent degradation domain via the
9. Semenza, G.L. and Wang, G.L. (1992) A nuclear factor induced ubiquitin–proteasome pathway. Proc. Natl. Acad. Sci. U. S. A.
by hypoxia via de novo protein synthesis binds to the human 95, 7987–7992
erythropoietin gene enhancer at a site required for transcrip- 18. Mahon, P.C. et al. (2001) FIH-1: a novel protein that interacts
tional activation. Mol. Cell. Biol. 12, 5447–5454 with HIF-1a and VHL to mediate repression of HIF-1 transcrip-
tional activity. Genes Dev. 15, 2675–2686

256 Trends in Endocrinology & Metabolism, April 2018, Vol. 29, No. 4
19. Tsukada, Y. et al. (2006) Histone demethylation by a family of 44. Yang, J. et al. (2012) Human CHCHD4 mitochondrial proteins
JmjC domain-containing proteins. Nature 439, 811–816 regulate cellular oxygen consumption rate and metabolism and
20. Klose, R.J. et al. (2006) JmjC-domain-containing proteins and provide a critical role in hypoxia signaling and tumor progres-
histone demethylation. Nat. Rev. Genet. 7, 715–727 sion. J. Clin. Invest. 122, 600–611

21. Kohli, R.M. and Zhang, Y. (2013) TET enzymes, TDG and the 45. Semenza, G.L. et al. (1996) Hypoxia response elements in the
dynamics of DNA demethylation. Nature 502, 472–479 aldolase A, enolase 1, and lactate dehydrogenase A gene
promoters contain essential binding sites for hypoxia-inducible
22. Wu, X. and Zhang, Y. (2017) TET-mediated active DNA demeth-
factor 1. J. Biol. Chem. 271, 32529–32537
ylation: mechanism, function and beyond. Nat. Rev. Genet. 18,
517–534 46. Iyer, N.V. et al. (1998) Cellular and developmental control of O2
homeostasis by hypoxia-inducible factor 1 alpha. Genes Dev.
23. Xu, Y. et al. (2011) Genome-wide regulation of 5hmC, 5mC, and
12, 149–162
gene expression by Tet1 hydroxylase in mouse embryonic stem
cells. Mol. Cell 42, 451–464 47. Kim, J.W. et al. (2006) HIF-1-mediated expression of pyruvate
dehydrogenase kinase: a metabolic switch required for cellular
24. Clifton, I.J. et al. (2006) Structural studies on 2-oxoglutarate
adaptation to hypoxia. Cell Metab. 3, 177–185
oxygenases and related double-stranded beta-helix fold pro-
teins. J. Inorg. Biochem. 100, 644–669 48. Papandreou, I. et al. (2006) HIF-1 mediates adaptation to hyp-
oxia by actively downregulating mitochondrial oxygen con-
25. McDonough, M.A. et al. (2010) Structural studies on human 2-
sumption. Cell Metab. 3, 187–197
oxoglutarate dependent oxygenases. Curr. Opin. Struct. Biol.
20, 659–672 49. Fukuda, R. et al. (2007) HIF-1 regulates cytochrome oxidase
subunits to optimize efficiency of respiration in hypoxic cells. Cell
26. Ozer, A. and Bruick, R.K. (2007) Non-heme dioxygenases:
129, 111–122
cellular sensors and regulators jelly rolled into one? Nat. Chem.
Biol. 3, 144–153 50. Hwang, H.J. et al. (2015) Hypoxia inducible factors modulate
mitochondrial oxygen consumption and transcriptional regula-
27. McNeill, L.A. et al. (2005) Hypoxia-inducible factor prolyl hydrox-
tion of nuclear-encoded electron transport chain genes. Bio-
ylase 2 has a high affinity for ferrous iron and 2-oxoglutarate.
chemistry 54, 3739–3748
Mol. Biosyst. 1, 321–324
51. Chan, S.Y. et al. (2009) MicroRNA-210 controls mitochondrial
28. Martinez, S. and Hausinger, R.P. (2015) Catalytic mechanisms
metabolism during hypoxia by repressing the iron-sulfur cluster
of Fe(II)- and 2-oxoglutarate-dependent oxygenases. J. Biol.
assembly proteins ISCU1/2. Cell Metab. 10, 273–284
Chem. 290, 20702–20711
52. Favaro, E. et al. (2010) MicroRNA-210 regulates mitochondrial
29. Hausinger, R.P. (2004) FeII/alpha-ketoglutarate-dependent
free radical response to hypoxia and Krebs cycle in cancer cells
hydroxylases and related enzymes. Crit. Rev. Biochem. Mol.
by targeting iron sulfur cluster protein ISCU. PLoS One 5,
Biol. 39, 21–68
e10345
30. Schofield, C.J. and Ratcliffe, P.J. (2005) Signalling hypoxia by
53. Chandel, N.S. et al. (1998) Mitochondrial reactive oxygen spe-
HIF hydroxylases. Biochem. Biophys. Res. Commun. 338, 617–
cies trigger hypoxia-induced transcription. Proc. Natl. Acad. Sci.
626
U. S. A. 95, 11715–11720
31. Webb, J.D. et al. (2009) Hypoxia, hypoxia-inducible factors
54. Guzy, R.D. et al. (2005) Mitochondrial complex III is required for
(HIF), HIF hydroxylases and oxygen sensing. Cell. Mol. Life
hypoxia-induced ROS production and cellular oxygen sensing.
Sci. 66, 3539–3554
Cell Metab. 1, 401–408
32. Ponnaluri, V.K. et al. (2013) A mechanistic overview of TET-
55. Mansfield, K.D. et al. (2005) Mitochondrial dysfunction resulting
mediated 5-methylcytosine oxidation. Biochem. Biophys. Res.
from loss of cytochrome c impairs cellular oxygen sensing and
Commun. 436, 115–120
hypoxic HIF-a activation. Cell Metab. 1, 393–399
33. Li, D. et al. (2015) TET family of dioxygenases: crucial roles and
56. Brunelle, J.K. et al. (2005) Oxygen sensing requires mitochon-
underlying mechanisms. Cytogenet. Genome Res. 146, 171–
drial ROS but not oxidative phosphorylation. Cell Metab. 1, 409–
180
414
34. Accari, S.L. and Fisher, P.R. (2015) Emerging roles of JmjC
57. Shadel, G.S. and Horvath, T.L. (2015) Mitochondrial ROS sig-
domain-containing proteins. Int. Rev. Cell Mol. Biol. 319, 165–
naling in organismal homeostasis. Cell 163, 560–569
220
58. Briggs, K.J. et al. (2016) Paracrine induction of HIF by glutamate
35. Hancock, R.L. et al. (2015) Epigenetic regulation by histone
in breast cancer: EglN1 senses cysteine. Cell 166, 126–139
demethylases in hypoxia. Epigenomics 7, 791–811
59. Lee, G. et al. (2016) Oxidative dimerization of PHD2 is respon-
36. Aragones, J. et al. (2009) Oxygen sensors at the crossroad of
sible for its inactivation and contributes to metabolic reprogram-
metabolism. Cell Metab. 9, 11–22
ming via HIF-1a activation. Sci. Rep. 6, 18928
37. Sharma, U. and Rando, O.J. (2017) Metabolic inputs into the
60. Masson, N. et al. (2012) The FIH hydroxylase is a cellular
epigenome. Cell Metab. 25, 544–558
peroxide sensor that modulates HIF transcriptional activity.
38. Vafai, S.B. and Mootha, V.K. (2012) Mitochondrial disorders as EMBO Rep. 13, 251–257
windows into an ancient organelle. Nature 491, 374–383
61. Hewitson, K.S. et al. (2007) Structural and mechanistic studies
39. Agani, F.H. et al. (2000) The role of mitochondria in the regulation on the inhibition of the hypoxia-inducible transcription factor
of hypoxia-inducible factor 1 expression during hypoxia. J. Biol. hydroxylases by tricarboxylic acid cycle intermediates. J. Biol.
Chem. 275, 35863–35867 Chem. 282, 3293–3301
40. Bastian, A. et al. (2017) AG311, a small molecule inhibitor of 62. Gimenez-Roqueplo, A.P. et al. (2001) The R22X mutation of the
complex I and hypoxia-induced HIF-1a stabilization. Cancer SDHD gene in hereditary paraganglioma abolishes the enzy-
Lett. 388, 149–157 matic activity of complex II in the mitochondrial respiratory chain
41. DeHaan, C. et al. (2004) Mutation in mitochondrial complex I and activates the hypoxia pathway. Am. J. Hum. Genet. 69,
ND6 subunit is associated with defective response to hypoxia in 1186–1197
human glioma cells. Mol. Cancer 3, 19 63. Isaacs, J.S. et al. (2005) HIF overexpression correlates with
42. Brown, S.T. and Nurse, C.A. (2008) Induction of HIF-2a is biallelic loss of fumarate hydratase in renal cancer: novel role
dependent on mitochondrial O2 consumption in an O2-sensitive of fumarate in regulation of HIF stability. Cancer Cell 8, 143–153
adrenomedullary chromaffin cell line. Am. J. Physiol. Cell Phys- 64. Selak, M.A. et al. (2005) Succinate links TCA cycle dysfunction
iol. 294, C1305–C1312 to oncogenesis by inhibiting HIF-a prolyl hydroxylase. Cancer
43. Hagen, T. et al. (2003) Redistribution of intracellular oxygen in Cell 7, 77–85
hypoxia by nitric oxide: effect on HIF1a. Science 302, 1975–
1978

Trends in Endocrinology & Metabolism, April 2018, Vol. 29, No. 4 257
65. Letouze, E. et al. (2013) SDH mutations establish a hyperme- 89. Nadtochiy, S.M. et al. (2016) Acidic pH is a metabolic switch for
thylator phenotype in paraganglioma. Cancer Cell 23, 739–752 2-hydroxyglutarate generation and signaling. J. Biol. Chem.
66. Xiao, M. et al. (2012) Inhibition of a-KG-dependent histone and 291, 20188–20197
DNA demethylases by fumarate and succinate that are accu- 90. Anso, E. et al. (2017) The mitochondrial respiratory chain is
mulated in mutations of FH and SDH tumor suppressors. Genes essential for haematopoietic stem cell function. Nat. Cell Biol.
Dev. 26, 1326–1338 19, 614–625
67. Sciacovelli, M. et al. (2016) Fumarate is an epigenetic modifier 91. Burr, S.P. et al. (2016) Mitochondrial protein lipoylation and the
that elicits epithelial-to-mesenchymal transition. Nature 537, 2-oxoglutarate dehydrogenase complex controls HIF1a stability
544–547 in aerobic conditions. Cell Metab. 24, 740–752
68. MacKenzie, E.D. et al. (2007) Cell-permeating alpha-ketogluta- 92. Rohle, D. et al. (2013) An inhibitor of mutant IDH1 delays growth
rate derivatives alleviate pseudohypoxia in succinate dehydro- and promotes differentiation of glioma cells. Science 340, 626–
genase-deficient cells. Mol. Cell. Biol. 27, 3282–3289 630
69. Hou, P. et al. (2014) Intermediary metabolite precursor dimethyl- 93. Kernytsky, A. et al. (2015) IDH2 mutation-induced histone and
2-ketoglutarate stabilizes hypoxia-inducible factor-1a by inhib- DNA hypermethylation is progressively reversed by small-mole-
iting prolyl-4-hydroxylase PHD2. PLoS One 9, e113865 cule inhibition. Blood 125, 296–303
70. Carey, B.W. et al. (2015) Intracellular alpha-ketoglutarate main- 94. Wang, F. et al. (2013) Targeted inhibition of mutant IDH2 in
tains the pluripotency of embryonic stem cells. Nature 518, 413– leukemia cells induces cellular differentiation. Science 340, 622–
416 626
71. Hwang, I.Y. et al. (2016) Psat1-dependent fluctuations in alpha- 95. Stein, E.M. et al. (2017) Enasidenib in mutant IDH2 relapsed or
ketoglutarate affect the timing of ESC differentiation. Cell Metab. refractory acute myeloid leukemia. Blood 130, 722–731
24, 494–501 96. Dang, L. et al. (2016) IDH mutations in cancer and progress
72. TeSlaa, T. et al. (2016) a-Ketoglutarate accelerates the initial toward development of targeted therapeutics. Ann. Oncol. 27,
differentiation of primed human pluripotent stem cells. Cell 599–608
Metab. 24, 485–493 97. FDA (2017) FDA Granted Regular Approval to Enasidenib for the
73. Mullen, A.R. et al. (2014) Oxidation of alpha-ketoglutarate is Treatment of Relapsed or Refractory AML, FDA
required for reductive carboxylation in cancer cells with mito- 98. Maxwell, P.H. and Eckardt, K.U. (2016) HIF prolyl hydroxylase
chondrial defects. Cell Rep. 7, 1679–1690 inhibitors for the treatment of renal anaemia and beyond. Nat.
74. Wise, D.R. et al. (2011) Hypoxia promotes isocitrate dehydro- Rev. Nephrol. 12, 157–168
genase-dependent carboxylation of alpha-ketoglutarate to cit- 99. Sheldon, R.A. et al. (2014) Hypoxic preconditioning protection is
rate to support cell growth and viability. Proc. Natl. Acad. Sci. U. eliminated in HIF-1a knockout mice subjected to neonatal
S. A. 108, 19611–19616 hypoxia–ischemia. Pediatr. Res. 76, 46–53
75. Metallo, C.M. et al. (2011) Reductive glutamine metabolism by 100. McCafferty, K. et al. (2014) The challenge of translating ischemic
IDH1 mediates lipogenesis under hypoxia. Nature 481, 380–384 conditioning from animal models to humans: the role of comor-
76. Mullen, A.R. et al. (2011) Reductive carboxylation supports bidities. Dis. Model. Mech. 7, 1321–1333
growth in tumour cells with defective mitochondria. Nature 101. Semenza, G.L. (2014) Oxygen sensing, hypoxia-inducible fac-
481, 385–388 tors, and disease pathophysiology. Annu. Rev. Pathol. 9, 47–71
77. Dang, L. et al. (2009) Cancer-associated IDH1 mutations pro- 102. Hill, P. et al. (2008) Inhibition of hypoxia inducible factor hydrox-
duce 2-hydroxyglutarate. Nature 462, 739–744 ylases protects against renal ischemia–reperfusion injury. J. Am.
78. Gross, S. et al. (2010) Cancer-associated metabolite 2-hydrox- Soc. Nephrol. 19, 39–46
yglutarate accumulates in acute myelogenous leukemia with 103. Jain, I.H. et al. (2016) Hypoxia as a therapy for mitochondrial
isocitrate dehydrogenase 1 and 2 mutations. J. Exp. Med. disease. Science 352, 54–61
207, 339–344
104. Gorres, K.L. and Raines, R.T. (2010) Prolyl 4-hydroxylase. Crit.
79. Achouri, Y. et al. (2004) Identification of a dehydrogenase acting Rev. Biochem. Mol. Biol. 45, 106–124
on D-2-hydroxyglutarate. Biochem. J. 381, 35–42
105. Minor, E.A. et al. (2013) Ascorbate induces ten–eleven translo-
80. Rzem, R. et al. (2006) The gene mutated in l-2-hydroxyglutaric cation (Tet) methylcytosine dioxygenase-mediated generation of
aciduria encodes l-2-hydroxyglutarate dehydrogenase. Biochi- 5-hydroxymethylcytosine. J. Biol. Chem. 288, 13669–13674
mie 88, 113–116
106. Blaschke, K. et al. (2013) Vitamin C induces Tet-dependent DNA
81. Losman, J.A. et al. (2013) (R)-2-Hydroxyglutarate is sufficient to demethylation and a blastocyst-like state in ES cells. Nature
promote leukemogenesis and its effects are reversible. Science 500, 222–226
339, 1621–1625
107. Chen, J. et al. (2013) Vitamin C modulates TET1 function during
82. Chowdhury, R. et al. (2011) The oncometabolite 2-hydroxyglu- somatic cell reprogramming. Nat. Genet. 45, 1504–1509
tarate inhibits histone lysine demethylases. EMBO Rep. 12,
108. Agathocleous, M. et al. (2017) Ascorbate regulates haemato-
463–469
poietic stem cell function and leukaemogenesis. Nature 549,
83. Xu, W. et al. (2011) Oncometabolite 2-hydroxyglutarate is a 476–481
competitive inhibitor of alpha-ketoglutarate-dependent dioxyge-
109. Cimmino, L. et al. (2017) Restoration of TET2 function blocks
nases. Cancer Cell 19, 17–30
aberrant self-renewal and leukemia progression. Cell 170,
84. Koivunen, P. et al. (2012) Transformation by the (R)-enantiomer 1079–1095.e20
of 2-hydroxyglutarate linked to EGLN activation. Nature 483,
110. Hore, T.A. et al. (2016) Retinol and ascorbate drive erasure of
484–488
epigenetic memory and enhance reprogramming to naive plu-
85. Oldham, W.M. et al. (2015) Hypoxia-mediated increases in L-2- ripotency by complementary mechanisms. Proc. Natl. Acad.
hydroxyglutarate coordinate the metabolic response to reduc- Sci. U. S. A. 113, 12202–12207
tive stress. Cell Metab. 22, 291–303
111. Zhao, B. et al. (2014) Redox-active quinones induces genome-
86. Intlekofer, A.M. et al. (2015) Hypoxia induces production of L-2- wide DNA methylation changes by an iron-mediated and Tet-
hydroxyglutarate. Cell Metab. 22, 304–311 dependent mechanism. Nucleic Acids Res. 42, 1593–1605
87. Tyrakis, P.A. et al. (2016) S-2-Hydroxyglutarate regulates CD8+ 112. Unlu, A. et al. (2016) High-dose vitamin C and cancer. J. Oncol.
T-lymphocyte fate. Nature 540, 236–241 Sci. 1, 10–12
88. Intlekofer, A.M. et al. (2017) L-2-Hydroxyglutarate production 113. PDQ Integrative, Alternative, and Complementary Therapies
arises from noncanonical enzyme function at acidic pH. Nat. Editorial Board (2002) High-Dose Vitamin C (PDQ1): Health
Chem. Biol. 13, 494–500

258 Trends in Endocrinology & Metabolism, April 2018, Vol. 29, No. 4
Professional Version. PDQ Cancer Information Summaries, US 118. Welsh, J.L. et al. (2013) Pharmacological ascorbate with gem-
National Cancer Institute citabine for the control of metastatic and node-positive pancre-
114. Chen, Q. et al. (2008) Pharmacologic doses of ascorbate act as atic cancer (PACMAN): results from a Phase I clinical trial.
a prooxidant and decrease growth of aggressive tumor xeno- Cancer Chemother. Pharmacol. 71, 765–775
grafts in mice. Proc. Natl. Acad. Sci. U. S. A. 105, 11105–11109 119. Joberty, G. et al. (2016) Interrogating the druggability of the 2-
115. Du, J. et al. (2010) Mechanisms of ascorbate-induced cytotox- oxoglutarate-dependent dioxygenase target class by chemical
icity in pancreatic cancer. Clin. Cancer Res. 16, 509–520 proteomics. ACS Chem. Biol. 11, 2002–2010

116. Verrax, J. and Calderon, P.B. (2009) Pharmacologic concen- 120. Kiriakidis, S. et al. (2017) Complement C1q is hydroxylated by
trations of ascorbate are achieved by parenteral administration collagen prolyl 4 hydroxylase and is sensitive to off-target inhi-
and exhibit antitumoral effects. Free Radic. Biol. Med. 47, 32–40 bition by prolyl hydroxylase domain inhibitors that stabilize hyp-
oxia-inducible factor. Kidney Int. 92, 900–908
117. Monti, D.A. et al. (2012) Phase I evaluation of intravenous
ascorbic acid in combination with gemcitabine and erlotinib in 121. Kinnaird, A. et al. (2016) Metabolic control of epigenetics in
patients with metastatic pancreatic cancer. PLoS One 7, cancer. Nat. Rev. Cancer 16, 694–707
e29794

Trends in Endocrinology & Metabolism, April 2018, Vol. 29, No. 4 259

Anda mungkin juga menyukai