ArticlePDF AvailableLiterature Review

Regulation of glycolysis by the hypoxia‐inducible factor (HIF): implications for cellular physiology

Wiley
The Journal of Physiology
Authors:

Abstract and Figures

Under conditions of hypoxia, most eukaryotic cells can shift their primary metabolic strategy from predominantly mitochondrial respiration towards increased glycolysis to maintain ATP levels. This hypoxia‐induced reprogramming of metabolism is key to satisfying cellular energetic requirements during acute hypoxic stress. At a transcriptional level, this metabolic switch can be regulated by several pathways including the hypoxia inducible factor‐1α (HIF‐1α) which induces an increased expression of glycolytic enzymes. While this increase in glycolytic flux is beneficial for maintaining bioenergetic homeostasis during hypoxia, the pathways mediating this increase can also be exploited by cancer cells to promote tumour survival and growth, an area which has been extensively studied. It has recently become appreciated that increased glycolytic metabolism in hypoxia may also have profound effects on cellular physiology in hypoxic immune and endothelial cells. Therefore, understanding the mechanisms central to mediating this reprogramming are of importance from both physiological and pathophysiological standpoints. In this review, we highlight the role of HIF‐1α in the regulation of hypoxic glycolysis and its implications for physiological processes such as angiogenesis and immune cell effector function. image
This content is subject to copyright. Terms and conditions apply.
J Physiol 00.0 (2020) pp 1–15 1
The Journal of Physiology
TOPICAL REVIEW
Regulation of glycolysis by the hypoxia-inducible factor
(HIF): implications for cellular physiology
S. J. Kierans1,2 and C. T. Taylor1,2
1Conway Institute of Biomolecular and Biomedical Research, University College Dublin, Belfield, Dublin 4, Ireland
2School of Medicine, University College Dublin, Belfield, Dublin 4, Ireland
Edited by: Ian Forsythe & Paul Greenhaff
Cormac Taylor is Professor of Cellular Physiology at the School of Medicine and the Conway Institute,
University College Dublin. Previous appointments include a postdoctoral fellowship at Brigham and
Wome n s Ho s p i t a l a n d Har va rd Me d i c a l S choo l (Bost o n ) . P rof e s s o r Ta y lor was t h e r e c i p ient o f the Na t u r e
midcareer mentorship award (2014). The Taylor lab investigates the impact of hypoxia and hypercapnia
on cellular pathways and function in the context of inflammation and immunity. Sarah Kierans is a PhD
student working in the Taylor laboratory on the pathways regulating glycolysis in hypoxia in tumour cells
and immune cells.
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society DOI: 10.1113/JP280572
2 S. J. Kierans and C. T. Taylor J Physiol 00.0
Abstract Under conditions of hypoxia, most eukaryotic cells can shift their primary metabolic
strategy from predominantly mitochondrial respiration towards increased glycolysis to maintain
ATP levels. This hypoxia-induced reprogramming of metabolism is key to satisfying cellular
energetic requirements during acute hypoxic stress. At a transcriptional level, this metabolic
switch can be regulated by several pathways including the hypoxia inducible factor-1α(HIF-1α)
which induces an increased expression of glycolytic enzymes. While this increase in glycolytic flux
is beneficial for maintaining bioenergetic homeostasis during hypoxia, the pathways mediating
this increase can also be exploited by cancer cells to promote tumour survival and growth,
an area which has been extensively studied. It has recently become appreciated that increased
glycolytic metabolism in hypoxia may also have profound effects on cellular physiology in hypoxic
immune and endothelial cells. Therefore, understanding the mechanisms central to mediating this
reprogramming are of importance from both physiological and pathophysiological standpoints.
In this review, we highlight the role of HIF-1αin the regulation of hypoxic glycolysis and its
implications for physiological processes such as angiogenesis and immune cell effector function.
(Received 6 August 2020; accepted after revision 25 September 2020; first published online 2 October 2020)
Corresponding author C. T. Taylor: Conway Institute of Biomolecular and Biomedical Research, University College
Dublin, Belfield, Dublin D4, Ireland. Email: cormac.taylor@ucd.ie
Abstract figurelegend Schematic outlining the functional physiological consequences of enhanced glycolytic metabolism
in endothelial cells and immune cells.
Introduction
Glycolysis, which takes place in the cytoplasm, describes
the anaerobic conversion of a single molecule of glucose
into two molecules of pyruvate. The metabolism of glucose
is achieved through the oxygen-independent activities
of 10 metabolic enzymes which comprise the glycolytic
pathway to result in the generation of biochemical energy
equivalents in the form of adenosine triphosphate (ATP),
the concurrent reduction of nicotinamide adenosine
dinucleotide (NAD) to NADH and the generation of
pyruvate (Fig. 1). In the presence of sufficient levels
of molecular oxygen (O2), the pyruvate generated is
further metabolised within the mitochondria by the
tricarboxylic acid (TCA) cycle which provides electron
carriers to the electron transport chain (ETC) to generate
afurther36moleculesofATPpermoleculeofglucose
entering the glycolytic pathway. In the steady state, the
combined activities of glycolysis, the TCA cycle and
the ETC usually results in the generation of sufficient
levels of ATP to meet the bioenergetic requirements
of the cell, tissue and organism and thereby maintain
homeostasis.
The glycolytic pathway is evolutionarily conserved
across all kingdoms, including archaeal (Selig et al. 1997),
fungal (Moses, 1959), protozoal (Becker & Geiman, 1955;
Bragg & Reeves, 1962), bacterial (Kornberg, 1976) and
metazoan species (Wan et al. 2015), in addition to being
well described in plant metabolism (Plaxton, 1996). The
rate of glycolysis is reliant on the availability and sub-
sequent uptake of glucose. In mammalian cells, glucose
uptake is achieved by two means: the active transport
of glucose as evident in the mammalian small intestinal
epithelial cells via sodium-dependent glucose trans-
porters (SGLT), or alternatively, via facultative diffusion
through glucose transporters (GLUT) located within
the cell membrane of multiple cell types (Pessin &
Bell, 1992). Once taken up into the cell, glucose is
oxidised in a stepwise manner, at a rate defined by
the rate-limiting enzymes of glycolysis. These enzymes;
hexokinase, phosphofructokinase and pyruvate kinase,
are allosterically regulated and are understood to have
important roles in the regulation of glycolytic flux (Jurica
et al.1998;Kemp&Gunasekera,2002;Chanetonet al.
2012).
Glycolysis in hypoxia
Under conditions where oxygen demand by a tissue
exceeds its supply (hypoxia), cells reduce their
reliance upon O2-dependent mitochondrial oxidative
phosphorylation (OXPHOS) and preferentially use
the O2-independent glycolytic pathway to maintain
sufficient ATP production in order satisfy bio-energetic
requirements. This metabolic switch in response to hypo-
xia is initiated by the allosteric control of glycolytic
enzymes by ATP. Under hypoxic conditions, whereby
OXPHOS has become inhibited, a low level of ATP
is produced which effectively decreases the ATP/AMP
(adenosine monophosphate) ratio. This reduction in
energy charge reduces the allosteric inhibition of
ATP on glycolytic enzyme phosphofructokinase (PFK)
(Henderson, 1969). Activated PFK utilises ATP to
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
J Physiol 00.0 HIF-dependent regulation of glycolysis 3
produce fructose-1,6-bisphosphate and promote flux
through the glycolytic pathway (Fig. 1). ATP further
allosterically inhibits pyruvate kinase to reduce the rate
of glycolysis when energy charge is high. Therefore,
under conditions of hypoxia, the reduction in energy
charge reduces the allosteric inhibition of ATP on
pyruvate kinase to result in the generation of pyruvate
from phosphoenolpyruvate (Fig. 1). Pyruvate kinase is
also activated by fructose-1,6-bisphosphate (feed-forward
regulation) to promote complete flux through the
glycolytic pathway.
The allosteric regulation of glycolysis under hypo-
xic conditions is subsequently followed by the trans-
criptional upregulation of glucose transporters and
glycolytic enzymes by the hypoxia inducible factor
(HIF) transcription factor. This upregulation of glycolytic
enzymes allows for an increased flux though the pathway,
and is thus advantageous in maintaining cellular ATP
production under hypoxic conditions. When oxygen
balance (and ATP/AMP ratio) is restored, cells can usually
revert their primary metabolic strategy to OXPHOS, a
phenomenon known as the Pasteur effect (Pasteur, 1861).
Figure 1. Schematic overview of the glycolytic pathway
Glucose is taken up into the cell from the extracellular space via glucose transporters (GLUT) located within the
cellular membrane. Glucose is oxidised in a stepwise manner via the actions of glycolytic enzymes (orange text).
The glycolytic pathway relies on the investment of 2 molecules of ATP to oxidize glucose (preparatory phase, A)
and allows for the concurrent generation of 2 molecules of NADH, 2 H+,2moleculesofH
2O, 2 molecules of
pyruvate and a net gain of 2 ATP per unit of glucose (pay-off phase; B).
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
4 S. J. Kierans and C. T. Taylor J Physiol 00.0
While this hypoxia-induced increase in glycolytic flux
is beneficial in promoting cell survival during a hypoxic
insult, and is thus adaptive in nature, the mechanisms
surrounding this increased glycolytic flux can also be
manipulated by cancer cells to promote their survival
even in the presence of sufficient levels of O2(known
as aerobic glycolysis or the Warburg effect; Warburg,
1956; Fig. 2). Therefore, the mechanisms central to
mediating this metabolic switch are of key importance
in both health and disease. To date multiple mechanisms
have been described as important regulators of glycolysis
under hypoxic conditions, including those dependent
on the actions of the HIF pathway (Seagroves et al.
2001), the PI3K/Akt pathway (Xie et al. 2019) and
hypoxia-dependent post-translational modification of
glycolytic enzymes (Agbor et al.2011),amongstothers.
In this review, we will focus on the role of HIF in the
physiological regulation of glycolysis in hypoxic cells.
The HIF pathway. HIF is a heterodimeric protein
consisting of an oxygen sensitive αsubunit (HIF-1α,
HIF-2α,HIF-3α)andaconstitutivelyexpressedβsubunit
[HIF-1β/Aryl hydrocarbon receptor nuclear translocator
(ARNT)]. Of the three HIFα-homologues identified in
higher metazoans, HIF-1αis responsible for regulating
glycolysis in response to hypoxia (Hu et al.2003).
In the presence of sufficient levels of O2(normoxia),
HIF-1αhas a short half-life, estimated at about 5–10 min
(Wang et al.1995).Innormoxicconditions,theHIF-1α
protein becomes rapidly degraded as a result of the
oxygen-dependent hydroxylation at conserved proline
residues (Pro402 and Pro564 )oftheαsubunit (Ivan
et al.2001;Jaakkolaet al.2001;Yuet al.2001).This
hydroxylation is carried out by members of the prolyl
hydroxylase domain (PHD) family, of which three major
isoforms have been characterised in mammalian cells.
Hydroxylation of HIF-1αpromotes its binding to the
von Hippel-Lindau (VHL) protein, a component of the
E3 ubiquitin ligase complex, which leads to the poly-
ubiquitination and subsequent proteasomal degradation
of the αsubunit. HIF-1αis also hydroxylated by
factor inhibiting HIF (FIH) at a C-terminal trans-
criptional activation domain asparagine residue (Asn803)
to further hinder the transcriptional activation of HIF
in normoxic conditions through inhibition of binding
with coactivator cyclic AMP response element-binding
protein (CBP)/p300 (Mahon et al.2001).Therefore,
O2-dependent hydroxylation serves to repress HIF-1α
under normoxic conditions.
The hydroxylation of HIF-1αby PHD and FIH
dioxygenases is dependent on the availability of both
O2and co-factors 2-oxoglutarate, Fe (II) and ascorbate
(Knowles et al.2003;Pageet al.2008).Accordingly,the
hydroxylation of HIF-1αis suppressed under conditions
of hypoxia, resulting in the HIF-1αsubunit becoming
stabilised and accumulating within the cytoplasm when
acellbecomesdeprivedofanadequateO
2supply.
Accumulated HIF-1αsubsequently translocates to the
nucleus where it dimerises with HIF-1β/ARNT (Wang
et al.1995).BindingoftheHIFheterodimertohypoxia
responsive elements (HRE) within the promoter/enhancer
regions of target genes initiates the transcription of those
genes which promote adaptation to hypoxia (Wang et al.
1995; Semenza, 2011).
HIF-1-dependent effect on glycolytic flux. Amongst the
adaptive responses mediated by HIF-1 to promote cell
survival under hypoxic conditions is the HIF-1-dependent
reprogramming of glucose metabolism to reduce reliance
on O2-dependent energy production (Papandreou
et al.2006).Thisreprogrammingisachievedbythe
HIF-1-dependent upregulation of genes encoding glucose
transporters (e.g. GLUT1 and GLUT 3) and enzymes of
the glycolytic pathway (Semenza et al.1994;Ebertet al.
1995; Iyer et al. 1998; Seagroves et al. 2001), as well as
the HIF-1 mediated suppression of OXPHOS to further
promote an increased glycolytic flux (Kim et al.2006;
Papandreou et al. 2006). The increased glucose uptake
and its subsequent metabolism under conditions of hypo-
xia help maintain cellular ATP production concommitent
with reduced oxidative metabolism.
The upregulation of glucose transporters and glycolytic
enzymes is mediated by HIF-1 binding to the
hypoxia-responsive element (HRE) consensus sequence
(5"-(A/G)CGTG-3")withinthepromoterregionofthe
genes encoding these proteins to promote their increased
expression (Semenza et al. 1996). The genes encoding
glycolytic enzymes PFK-liver type (PFKL), aldolase
(ALDA), phosphoglycerate kinase-1 (PGK1), enolase
(ENOL) and lactate dehydrogenase-A (LDHA) contain
HIF-1 binding sites in their enhancer regions and are
therefore directly upregulated by HIF in response to hypo-
xia (Semenza et al.1994).Interestingly,theconsensus
sequence central to mediating HIF binding present in
human ALDA and human PGK1 are evolutionarily
conserved between humans and mice, suggesting that the
hypoxia-induced, HIF-1-dependent upregulation of these
genes is of central importance to mammalian glycolytic
regulation (Semenza et al.1994).
In addition to the HIF-1-dependent upregulation
of glycolytic enzymes in response to hypoxia, HIF-1α
can suppress OXPHOS under hypoxic conditions by
reducing metabolite entry into the TCA cycle via the
HIF-1α-dependent induction of pyruvate dehydrogenase
kinase 1 (PDK1) (Kim et al. 2006; Papandreou et al. 2006)
and LDHA (Yang et al. 2014a). PDK1 phosphorylates
the pyruvate dehydrogenase (PDH) complex, preventing
pyruvate from being converted to acetyl Co-A, thus
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
J Physiol 00.0 HIF-dependent regulation of glycolysis 5
Figure 2. Schematic comparing and contrasting the steady state, Warburg and Pasteur effects
In the presence of sufficient levels of oxygen, healthy cells (A) generate cellular ATP primarily by oxidative
phosphorylation by shuttling the majority of pyruvate generated (indicated by continuous arrow) into the
mitochondria. In the absence of sufficient levels of oxygen, cells reprogramme their metabolism to use glycolysis
as their primary metabolic strategy for generation of ATP (Pasteur effect). The pyruvate generated from glycolysis
is fermented into lactate and transported out of the cells via the actions of the monocarboxylate transporters
(MCTs) located within the cell membrane. Malignant cells (B)canreprogrammetheirmetabolismtopreferentially
use glycolysis as their primary metabolic strategy for energy metabolism even in the presence of sufficient levels
of oxygen and functioning mitochondria (Warburg effect). The pyruvate generated is primarily oxidised to lactate
(continuous arrow), with only a small amount of lactate being transported to the mitochondria for oxidative
phosphorylation (dashed arrow).
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
6 S. J. Kierans and C. T. Taylor J Physiol 00.0
hindering the initiation of the TCA cycle. HIF-1α
also activates transcription of LDHA, which catalyses
the conversion of pyruvate generated from glycolytic
metabolism to lactate and concurrently generates NAD+,
an cofactor essential in permitting continued glycolytic
activity (Semenza et al. 1996). Together both PDK1
and LDHA divert pyruvate away from the TCA cycle
by decreasing acetyl Co-A generation from pyruvate
and instead increasing lactate production. The lactate
produced from pyruvate can then be removed from the cell
via the actions of plasma membrane transporter, mono-
carboxylate transport 4 (MCT4) to avoid competitive
inhibition of LDHA (Ullah et al.2006).
HIF-1αalso regulates OXPHOS capacity through
directly altering mitochondrial activity. HIF-1αinduces
mitochondrial autophagy in response to hypoxia thereby
conferring a protective effect in the hypoxic cell by
reducing the levels of reactive oxygen species (ROS)
generated (Zhang et al.2008).Furthermore,hypoxia
reduces respiratory capacity by regulating cytochrome
oxidase subunit expression in a HIF-1-dependent manner
to limit ROS generation under hypoxic conditions, the reby
promoting the evasion of oxidative stress-induced cell
death (Fukuda et al.2007).
Pharmacological hydroxylase inhibition and glycolysis.
In addition to the stabilisation and subsequent activation
of HIF-1αin response to hypoxia, HIF can also
be pharmacologically activated using molecules which
inhibit the actions of the PHD enzymes, such as iron
chelators (e.g. deferoxamine, DFO) (Wang & Semenza,
1993), or molecules which compete for 2-oxoglutarate
(Mole et al.2003).ModulationofPHDenzyme
activity is beneficial in promoting anti-inflammatory
responses by selectively inducing cell death in mono-
cytes (Crifo et al.2019),neutrophils(Manresaet al.
2019), and enhancing intestinal epithelial barrier function
(Cummins et al. 2008), in pro-angiogenic responses
via the HIF-dependent upregulation of vascular end-
othelial growth factor (VEGF), offering protection in
ischaemic disease (Milkiewicz et al.2004),aswellas
helping ameliorate anaemia in patients with kidney disease
by stimulating erythropoietin (EPO) production and
regulating iron absorption and metabolism (Chen et al.
2019b,c).
In the context of glucose metabolism, hydro-
xylase inhibition protects alveolar epithelial cells from
bioenergetic failure and cell death in both in vitro
and in vivo models by enhancing glycolytic metabolism
through HIF-1α(Tojo et al. 2018). Similarly, hydro-
xylase inhibition is cytoprotective in the renal proximal
tubular epithelial cell line HK2 in an oxygen-glucose
deprivation (OGD) in vitro model of ischaemia, through
upregulation of the glycolytic pathway (Ito et al.2020).
While these observations suggest that promoting glycolysis
at the cellular level through pharmacological activation
of the HIF pathway is cytoprotective, at the level of
the organism, sustained activation of the HIF pathway
arising from loss of function mutations of VHL results
in a reduced mitochondrial respiratory capacity, reduced
exercise capacity and greater blood lactate concentration
in these patients through upregulation of HIF-dependent
glycolytic genes and OXPHOS suppression under
normoxic conditions (Formenti et al. 2010; Perrotta
et al. 2020). While such findings emphasise that HIF
plays a prominent role in regulating glucose metabolism,
they also demonstrate that promoting glycolysis, for
example, through hydroxylase inhibition, may be rather
deleterious to the organism on an integrative level. The
effect of pharmacological stabilisation of HIF on the
whole organism is therefore an area worthy of further
investigation.
HIF-2αdoes not regulate glycolytic flux. HIF-2α,a
paralogue of HIF-1α,isalsonegativelyregulatedby
prolyl and asparaginyl hydroxylation (O’Rourke et al.
1999). While HIF-1αis ubiquitously expressed in all
cell types, HIF-2αexpression is cell-type specific, with
HIF-2αmRNA having been reported in vascular end-
othelial cells, kidney and gastrointestinal tract epithelial
cells, liver parenchymal cells and type II pneumocytes
(Wiesener et al. 2003). Under hypoxic conditions, HIF-2α
dimerises with constitutively expressed HIF-1βto regulate
the transcription of numerous target genes. Despite the
similarities between HIF-1αand HIF-2αin terms of
sequence homology, stabilisation and HRE recognition,
HIF-1αand HIF-2αtarget genes can differ (Tian
et al. 1997). Processes such as erythropoiesis (Gruber
et al. 2007; Rankin et al.2007)ironabsorptionand
metabolism (Mastrogiannaki et al. 2009) and carotid body
development and function (Hodson et al. 2016; Macias
et al.2018)arelargelydependentonHIF-2αexpression,
while the expression of genes encoding glucose transporter
GLUT1 and VEGF appear to be regulated by both HIF-1α
and HIF-2α(Hu et al.2003,2007).WhileHIF-1αand
HIF-2αregulated genes can overlap, multiple studies have
reported that glycolytic genes are exclusively regulated by
HIF-1α(Hu et al. 2003, 2007; Wang et al.2005).Severe
hypoxic treatment (0.1% O2)failstoinduceglycolyticgene
expression (LDHA, PGK-1, PGM-1 and PKM) in human
renal carcinoma cells (786-O) overexpressing VHL (786-O
WT-8) despite the induction of HIF-2αresponsive genes at
the same oxygen tension (Hu et al.2003).Thesefindings
support the necessary and selective role for HIF-1αin
mediating the Pasteur effect (Seagroves et al.2001).
Take n t o ge t h er, it i s e v i d e nt t h at t h e e ff e c t s m e di a ted
by HIF-1αin promoting glycolytic flux and reducing
OXPHOS capacity play a pivotal role in promoting cellular
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
J Physiol 00.0 HIF-dependent regulation of glycolysis 7
adaptation to hypoxia. Given that this response can be
both beneficial in overcoming a hypoxic insult, while
also having the capacity to be exploited in the context of
cancer cell survival and tumour growth (Warburg effect),
acompleteunderstandingoftheregulationofglycolysis
under hypoxic conditions is of crucial importance.
While HIF plays a key role in mediating this adaptive
response, the involvement of other mechanisms and/or
signalling pathways in complementing the HIF-dependent
regulation of metabolic reprogramming as evident in
hypoxia must not be excluded. Such mechanisms,
including the involvement of epigenetics (Krieg et al.
2010), post-translation modifications (Agbor et al. 2011)
as well as the spatial reorganisation of glycolytic enzymes
under hypoxic conditions (Agbor et al.2011;Jinet al.
2017) are emerging topics of interest, and may explain,
at least in part, how loss of HIF-1βdoes not perturb the
glycolytic phenotype (Troy et al.2005).However,these
HIF-independent mechanisms are beyond the scope of
this review.
Hypoxia-induced glycolytic flux and the PI3K/Akt
pathway. The regulation of glycolytic flux by HIF
can also be influenced by activity of other pathways.
The phosphatidylinositol-4,5-bisphophate 3-kinase
(PI3K)/protein kinase B (Akt) pathway is a highly
conserved signalling pathway which plays a key role
in regulating cell cycle progression in response to
extracellular signals (Burgering & Coffer, 1995; Franke
et al.1995;Kohnet al.1995).Inadditiontoitsrole
in promoting cellular survival, the PI3K pathway has
been implicated in regulating HIF-1αexpression and
transcriptional activity under hypoxic conditions in a
cell-type specific manner (Alvarez-Tejado et al.2001;
Arsham et al.2002;Liuet al. 2019). In oesophageal
carcinoma cell lines, hypoxic exposure results in an
increased level of HIF-1αprotein stabilisation, glycolytic
enzyme expression and extracellular lactate production
(Zeng et al. 2016). This response was successfully reversed
using the PI3K inhibitor Wortmannin, suggesting that
the upregulation of glycolysis in response to hypoxia
in these cells was at least partially attributed to PI3K
upregulation. Additionally, PTEN inactivation mutations,
as evident in cancers, also appear to potentiate the effect
of hypoxia on HIF-dependent gene expression through
downstream effector mammalian target of rapamycin
complex 1 (mTORC1) activation (Zhong et al. 2000) while
exhibiting little to no effect on glycolytic gene expression
under normoxic conditions, supporting the idea that the
PI3K pathway may play a role in the regulation of hypoxic
glucose metabolism (Zundel et al.2000).
However, while these findings suggest there may be a
role for the PI3K signalling pathway in regulating glucose
metabolism in response to hypoxia through HIF-1α
stabilisation, the extent to which the PI3K/Akt signalling
pathway is involved in HIF-dependent, hypoxia-induced
glucose metabolism is an area which requires further
investigation. Under normoxic conditions, for example,
PI3K/Akt activation in response to cytokine stimulation
has been shown to regulate glucose uptake via the direct
regulation of GLUT1 activity and localisation (Wieman
et al. 2007). Additionally, PI3K has been shown to regulate
normoxic glucose metabolism through its regulation of the
expression of rate limiting enzymes of glycolysis HK2 and
PFK (DeBerardinis et al.2008;Xieet al.2019),exertinga
direct influence on glucose metabolism without activating
mitochondrial OXPHOS (Elstrom et al. 2004). Such
findings suggest that the effects of the PI3K/Akt signalling
pathway on glucose metabolism are not confined to those
mechanisms which rely on HIF activity. Understanding
the potential crosstalk and complementarity between
HIF-dependent and HIF-independent mechanisms of
hypoxic glucose metabolic regulation is an area of key
importance from a physiological standpoint.
Hypoxia-induced glycolytic flux and non-coding RNAs
(ncRNA). RNAs, including coding RNAs such as
messenger RNA (mRNA) and non-coding RNAs such as
microRNA (miRNA), are heavily regulated by hypoxia in
a HIF-dependent manner (Schodel et al.2011;Choudhry
et al. 2014). Among the non-coding RNAs regulated by
hypoxia (reviewed by Shen et al.2013),longnon-coding
RNAs (lncRNAs) have emerged as a new class of trans-
cript commonly regulated by hypoxia (Choudhry et al.
2014; Choudhry & Harris, 2018). Interestingly, a number
of lncRNAs have been implicated in the regulation of
hypoxia-enhanced glucose metabolism.
Long non-coding RNAs (lncRNAs) are a large class
of heterogenous regulatory transcripts, greater than 200
nucleotides in length, which lack evidence of protein
coding potential (Rinn & Chang, 2012). Of those lncRNAs
regulated by hypoxia, there is evidence to suggest hypoxic
regulation is mediated by the binding of HIF to an HRE
consensus sequence located within the promoter region
of these lncRNAs (Choudhry et al. 2014; Choudhry &
Harris, 2018). LncRNAs lincRNA-p21, lncHIFCAR and
lncRNA-UCA1 have all been implicated in regulating the
HIF-dependent increase in glycolysis as evident under
hypoxic conditions (Li et al. 2014; Xue et al. 2014; Yang
et al.2014b;Shihet al.2017).
Long intragenic non-coding RNA (lincRNA)-p21
(lincRNA-p21) has been implicated as a critical mediator
of the hypoxia-dependent increase in glycolysis via its
role in promoting HIF-1αstabilisation (Yang et al.
2014b). LincRNA-p21, a key repressor of p53-dependent
non-transcriptional responses (Huarte et al. 2010),
is strongly induced by hypoxia in a dose- and
time-dependent manner through the binding of HIF-1αto
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
8 S. J. Kierans and C. T. Taylor J Physiol 00.0
an HRE consensus sequence located within the promoter
region of lincRNA-p21 in mammalian cells (Yang et al.
2014b). LincRNA-p21 induction increases the expression
of glycolytic enzymes LDHA and GLUT1 to result in
an increased glucose uptake and lactate production
within these cells in response to O2deprivation. Inter-
estingly, in addition to the regulation of lincRNA-p21 by
HIF-1α, lincRNA-p21 can reciprocally regulate HIF-1α
levels under hypoxic conditions by interfering with the
HIF-1α-VHL interaction, thereby preventing HIF-1α
proteasomal degradation (Yang et al.2014b).
In addition to the regulation of metabolic
reprogramming by lincRNA-p21 in response to
hypoxia, long non-coding RNA HIF-1αco-activating
RNA (lncHIFCAR) has also been implicated in regulating
hypoxia-induced glucose uptake and lactate production
in oral squamous cell carcinoma cell lines (Shih et al.
2017). LncHIFCAR is induced by hypoxic stress and is
required for the hypoxia-induced activation of HIF-1
target genes, glycolytic enzymes LDHA and PDK1 (Shih
et al.2017).Theupregulationoftheseglycolyticenzymes
is achieved via the lncHIFCAR-dependent augmentation
of the transcriptional activity of HIF-1αby lncHIFCAR
associating with the chromatin loci of HIF-1 target
proteins (Shih et al.2017).
Finally, lncRNA urothelial associated carcinoma 1
(UCA1) has also been implicated as a hypoxia-inducible
lncRNA (Xue et al.2014),whichplaysapivotalrole
in accelerating glycolysis in bladder cancer cells via the
upregulation of the HIF-1-dependent glycolytic enzyme
hexokinase 2 (HK2) (Li et al.2014).Interestingly,while
the downstream effects of HIF-dependent upregulation
of lincRNA-p21 and lncHIFCAR appear to be mediated
primarily by the HIF pathway, lncRNA-UCA1 can
mechanistically regulate glycolytic gene expression via the
mammalian target of rapamycin (mTOR) pathway (Li
et al.2014).Therefore,thisobservationfurtherhighlights
the potentially complementary nature of HIF-dependent
and HIF-independent regulation of glucose metabolism
under hypoxic conditions. However, while these findings
outline a likely role of non-coding RNAs in regulating
HIF-1-dependent reprogramming of glucose metabolism,
as evident in hypoxic conditions, the mechanisms under-
lying the regulation of lncRNAs in both normoxic and
hypoxic conditions remain to be fully elucidated.
Functional physiological consequence of a
hypoxia-induced increase in glycolysis
The hypoxia-induced enhancement of glycolysis promotes
adaptation to O2deprivation by supporting ATP
production following suppression of OXPHOS. However,
the functional consequence of an enhanced glycolytic
flux extends beyond that of provision of bioenergetic
energy equivalents in the form of ATP. These consequences
include both the regulation of the immunity and
neovascularisation, and play important roles in the body’s
adaptation to hypoxia. The implications of increased
glycolytic flux from a physiological aspect are outlined
below and highlight our crucial need to understand the
regulation of cellular glucose metabolism under hypoxic
conditions in its entirety.
Endothelial cells. Endothelial cells (ECs) are highly
plastic cells which form the inner lining of both blood and
lymphatic vessels. Despite their close proximity to a rich
oxygen supply in the blood, glycolysis is the predominant
bioenergetic pathway for ECs, generating up to 85% of the
cellular ATP content in arterial, venous, lymphatic and
microvascular ECs (Culic et al.1997;DeBocket al. 2013).
ECs can remain quiescent for years, only becoming active
during periods of O2/nutrient deprivation, in response
to tissue damage or as a result of pathological conditions
such as inflammation or cancer (Vandekeere et al.2015).
Under such conditions, ECs can rapidly reprogramme
their metabolism to double their metabolic rate via the
further upregulation of the glycolytic pathway through
phosphofructokinase-2/fructose-2,6-bisphosphatase-3
(PFKFB3) (Dobrina & Rossi, 1983). This increase in
glycolytic metabolism promotes angiogenesis at the
vascular forefront via endothelial tip cell PFKFB3, which
when silenced, impairs vessel sprouting (De Bock et al.
2013). Endothelial tip cell extension is followed by
endothelial stalk cell proliferation to promote elongation
of the vessel. The newly formed vessel then becomes lined
with is endothelial phalanx cells, with a lower (albeit, still
substantial) glycolytic rate than the actively proliferating
ECs, allowing for vessel perfusion (De Bock et al.2013).
While it seems counterintuitive for ECs to rely
on glycolytic metabolism as their primary metabolic
strategy in both quiescent and active states, there
are multiple benefits derived from this preferential
mechanism of energy metabolism. A particular benefit
in the context of ECs of relying on anaerobic glycolysis as
ameansofenergyproductionisinthevascularisation
of tissues. In the absence of O2,glycolyticECscan
continue to promote angiogenesis in the absence of
O2in attempts to improve overall tissue oxygenation.
Inhibiting the glycolytic pathway using glucose mimetic
2-deoxy-D-glucose (2-DG) inhibits angiogenesis in both
in vitro and in vivo models by interfering with
N-linked glycosylation (Merchan et al.2010),highlighting
a pertinent role for glycolysis in neovascularisation.
Glycolytic metabolism is also beneficial for ECs, as
lactate, the end product of glycolysis, functions as a
pro-angiogenic signalling molecule in highly glycolytic
ECs via the activation of the NFκB/CXCL8 pathway and
the stabilisation of HIF-1α(Dobrina & Rossi, 1983; Vegran
et al. 2011; Sonveaux et al.2012;Eelenet al.2018).
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
J Physiol 00.0 HIF-dependent regulation of glycolysis 9
The glycolytic pathway can also beneficially provide
numerous co-factors which can be diverted to
biosynthetic growth pathways to support anabolic
cell growth and proliferation. Glycolytic intermediates
glucose-6-phosphate and 3-phosphoglycerate, for
example, are supplied to the pentose phosphate and
serine biosynthetic pathways, respectively, where they
are used in the production of ribose and amino acids
which both play important roles in supporting cell
growth (Horecker, 2002). In addition to the low level
of OXPHOS evident in quiescent and active ECs, it has
been noted that angiogenesis is not affected by complete
inhibition of mitochondrial respiration (De Bock et al.
2013). A low level of OXPHOS generates fewer reactive
oxygen species (ROS), subjecting ECs to a reduced level of
oxidative stress despite their proximity to a high-oxygen
environment, therefore protecting ECs from ROS-induced
cell death.
Take n t o ge t h er, th e s e ob s e r v at i o ns s u g ge s t th a t w h i l e
glycolysis may not be the most efficient way of producing
ATP, EC s d erive be n e t f r o m g lycoly t i c metab o l i s m in
terms of angiogenic potential, with glycolysis playing a
pivotal role in promoting EC proliferation and motility
(De Bock et al.2013).Whilebeingbenecialfrom
aphysiologicalstandpointinpromotingadaptationto
O2deprivation, the angiogenic potential of ECs can be
detrimental in pathophysiological states such as cancer
development and progression. In such situations, the
hypoxic core of tumours can result in the inappropriate
initiation of angiogenesis and adversely promote tumour
growth and survival. However, the identification of
PFKFB3 as the primary driver of glycolytic in ECs
has been of great utility as small molecule compound
3-(3-pyridinyl)-1-(4-pyridinyl)-2-propen-1-one (3PO),
an inhibitor of PFKFB3, can reduce proliferation and
migration of ECs in both in vitro and in vivo models
(De Bock et al. 2013). This finding suggests that
pharmacological intervention in the glycolytic pathway
may present new opportunities for anti-angiogenic
therapies.
Immune cells. In addition to the implications of an
enhanced glycolytic flux for endothelial cell biology,
glycolytic metabolism plays a regulatory role in immune
cell function. A variety of immune cells, including
dendritic cells (Krawczyk et al. 2010), classically activated
M1 macrophages (Michl et al. 1976), neutrophils (Sbarra
& Karnovsky, 1959), lymphocytes (T cells (Frauwirth et al.
2002; Wang et al. 2011) and B cells (Doughty et al. 2006))
adopt a high-glycolytic phenotype once activated. This
high-glycolytic phenotype therefore confers a particular
advantage to immune cells in allowing for mounting of
an immune response in tissues deprived of an adequate
oxygen supply. Such tissues include those which are
inflamed, infected or tumour burdened, where hypo-
xia is common place, as well as those tissues which
experience persistent low levels of oxygenation (so called
‘physiological hypoxia’), such as the bone marrow,
placenta or gastrointestinal tract (Taylor & Colgan,
2017).
Immune cell metabolism and the tissue micro-
environment. Unlike cancer cells, whereby an increased
rate of glycolysis and reduced reliance on OXPHOS often
arises from genetic aberrations, an enhanced glycolytic
flux in immune cells is not permanent and, instead,
immune cells obtain the capacity to rapidly alter their
metabolic profile in response to external signals (Ganeshan
& Chawla, 2014). Such external signals include hypo-
xia which is commonly encountered by immune cells in
both physiological and pathophysiological settings. The
activation of glycolysis in immune cells is regulated by
HIF-1αin activated M1 macrophages, dendritic cells,
lymphocytes and neutrophils, and co-regulated by HIF-1α
and PI3K/Akt pathway in dendritic cells and CD4+Tcells
(Ganeshan & Chawla, 2014; Colgan et al.2020).
The metabolic reprogramming of immune cells
following their activation has multiple effects on
immune responses. In dendritic cells (DCs), metabolic
reprogramming to a high-glycolytic phenotype is
necessary for DC activation and regulates DC cytokine
production and antigen presenting capability (Krawczyk
et al.2010;Evertset al.2014).Inactivatedneutrophils
and classically activated macrophages, glycolysis is crucial
to support their infiltrating capacity and microbicidal
actions (Valentine & Beck, 1951; Sbarra & Karnovsky,
1959), such that inhibition of glycolysis with 2-DG pre-
vents macrophage activation and inhibits endocytosis of
complement 3 (C3) and immunoglobulin (IgG) bound
particles (Michl et al.1976;Boxeret al.1977).Inhibition
of glycolysis hinders the development of T cells into
TH17 cells, whilst promoting the production of regulatory
T(T
reg) cells, suggesting that the glycolytic pathway is
an important mediator of T cell differentiation (Shi
et al.2011).InactivatedCD4
+Tcells,glycolysisplays
aprimaryroleinpromotingtheactivationofeffector
T cells (Michalek et al.2011),anddownregulating
glycolysis by silencing of PFKFB3 in CD4 T cells renders
these cells highly susceptible to apoptosis and, counter-
intuitively, impairs autophagy (Yang et al.2013).Finally,
sustained glycolytic metabolism following the cessation of
an immune response impairs the ability of CD8+Tcells
to form long-term immune memory such that inhibiting
glycolytic flux preserves the formation of memory CD8+T
cells (Sukumar et al.2013).Hypoxiacanthereforegreatly
directly influence immune cell phenotype and function
via metabolic reprogramming to favour a glycolytic
phenotype.
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
10 S. J. Kierans and C. T. Taylor J Physiol 00.0
Immune cell metabolism and the tumour micro-
environment. While physiological hypoxia can promote
immune cell homeostasis under hypoxic conditions by
altering immune cell metabolic status, pathological hypo-
xia, as evident in the core of tumours can regulate immune
cell effector function which in turn plays a role in tumour
development (Lee et al.2010).Numerousimmunecell
types are recruited to a tumour to reside within the
tumour microenvironment. These include those immune
cells important for mediating anti-tumour effects such
as DCs and natural killer cells, as well as those which
promote pro-tumour effects, such as tumour-associated
macrophages (TAMs) and myeloid-derived suppressor
cells (MDSCs) (Reviewed by Ou & Lv, 2020).
Hypoxia within the tumour microenvironment
promotes tumour growth by numerous mechanisms.
First, hypoxia within the tumour microenvironment can
accelerate the rate of glycolysis in MDSCs via HIF-1α,to
promote proliferation of these cells and potentiate their
immunosuppressive effects (Corzo et al. 2010). MDSCs
can exert immunosuppressive effects by suppressing
T-cell responses and modulating cytokine production
from phagocytic macrophages to promote immuno-
evasion (reviewed by Millrud et al. 2017). Hypoxia
within the tumour microenvironment further promotes
the differentiation of monocytic MDSCs (M-MDSCs)
into TAMs via HIF-1α(Corzo et al.2010).Interestingly,
TAMs within the mi c r o e nvironment of b r e a s t c ancers
upregulate aerobic glycolysis of tumour cells through
shuttling of extracellular vesicles containing lncRNA
HIF-1α-stabilising non-coding RNA (HISLA), which
promotes HIF-1αstabilisation by preventing the inter-
action between HIF-1αand the PHD enzymes (Chen et al.
2019a). These highly glycolytic tumour cells positively
regulate HISLA levels in TAMs via an increased lactate
production to further promote tumour cell glycolytic
activity and contribute to the immunoevasive tumour
phenotype (Chen et al.2019a).
Secondly, highly glycolytic tumour cells suppress T
cell activity by depriving immune cells of an adequate
glucose supply (Chang et al.2015),withlowlevelsof
glucose resulting in T cell hyporesponsiveness, even when
tumours are highly antigenic (Maciver et al.2008).Glucose
restrictions within the tumour microenvironment as a
result of increased glucose uptake of hypoxic tumour cells
can also favour the activation of an M2-like phenotype
in tumour infiltrating macrophages, thereby promoting
an anti-inflammatory response and promoting tumour
growth (Chang et al.2015).
Hypoxic tumours also have a high concentration of the
glycolytic end product lactate in their microenvironment
as a result of an increased glycolytic flux in tumour cells.
High levels of lactate in the tumour microenvironment
reduce immune cell activity. Specifically, tumour derived
lactate inhibits T cell cytokine production by disturbing
T cell metabolic homeostasis (Fischer et al. 2007; Dietl
et al. 2010). Tumour-derived lactic acid also alters the
antigen-presenting capacity and cytokine production of
human monocyte-derived DCs (Gottfried et al.2006)
and inhibits monocyte activation, migration and cyto-
kine release (Dietl et al. 2010; Goetze et al. 2011).
Lactate can also have profound effect on macrophage
polarisation, by inducing arginase 1, VEGF and other
M2-associated genes in TAMs, thereby resulting in the
development of a tumour-promoting phenotype in these
cells (Colegio et al.2014).Anincreasedintracellularlactate
concentration results in an increased level of histone
lactylation. The effect of this increased lactylation is the
increased expression of genes typical of M2-macrophage
polarisation in M1-macrophages (Zhang et al.2019).
Finally, lactate inhibits the activation and survival of
natural killer cells in both in vitro and in vivo settings
(Brand et al. 2016). To summarise, high levels of lactate
within the tumour microenvironment as a result of an
enhanced glycolytic flux can have a profound effect
on immune cell function, resulting in notable tumour
evasion.
Take n t o ge t h er, th e s e fin d i n g s h ig h li g ht t h at d i f fe r e nce s
in metabolic status, be that originating from the immune
cell itself, or indeed, the immune cell microenvironment
have a significant impact on immune cell function and can
profoundly influence an immune response. The impact of
inhibiting a high-glycolytic phenotype in immune cells
has been investigated in in vivo models of autoimmune
conditions such as experimental autoimmune neuritis
(Liu et al.2018)andsystemiclupuserythematosus(Yin
et al. 2016), and the findings suggest that inhibition
of glycolysis can have profound effects on the adaptive
and innate immune response with notable reductions
in clinical symptoms. Gaining a more thorough under-
standing of the differences in metabolic profile within a
physiological or pathological immunological niche may
lead to a method of targeted treatment for immune
pathologies or reversal of tumour cell immune evasion
in coming years.
Conclusion
In summary, studies in both in vitro and in vivo
models have demonstrated a profound capacity for hypo-
xic cells to adapt to O2deprivation by altering trans-
criptional and translational responses to promote glucose
uptake and its subsequent anaerobic catabolism. While
this response is mediated primarily by HIF stabilisation
and HIF-target gene expression, it is still possible that
alternative mechanisms are also involved in promoting
this glycolytic phenotype. The integration of these
responses and the elucidation of their precise molecular
mechanisms and crosstalk are fundamental to our
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
J Physiol 00.0 HIF-dependent regulation of glycolysis 11
understanding of human physiology, and may represent
an area of important translational potential for human
disease.
References
Agbor TA, Cheong A, Comerford KM, Scholz CC, Bruning U,
Clarke A, Cummins EP, Cagney G & Taylor CT (2011).
Small ubiquitin-related modifier (SUMO)-1 promotes
glycolysis in hypoxia. J Biol Chem 286,47184726.
Alvarez-Tejado M, Naranjo-Suarez S, Jimenez C, Carrera AC,
Landazuri MO & del Peso L (2001). Hypoxia induces the
activation of the phosphatidylinositol 3-kinase/Akt cell
survival pathway in PC12 cells: protective role in apoptosis.
J Biol Chem 276,2236822374.
Arsham AM, Plas DR, Thompson CB & Simon MC (2002).
Phosphatidylinositol 3-kinase/Akt signaling is neither
required for hypoxic stabilization of HIF-1 alpha nor
sufficient for HIF-1-dependent target gene transcription.
J Biol Chem 277,1516215170.
Becker CE & Geiman QM (1955). Utilization of glucose by
two strains of Entamoeba histolytica. Exp Parasitol 4,
493–501.
Boxer LA, Baehner RL & Davis J (1977). The effect of
2-deoxyglucose on guinea pig polymorphonuclear leukocyte
phagocytosis. JCellPhysiol91,89102.
Bragg PD & Reeves RE (1962). Pathways of glucose
dissimilation in Laredo strain of Entamoeba histolytica. Exp
Para si tol 12,393400.
Brand A, Singer K, Koehl GE, Kolitzus M, Schoenhammer G,
Thiel A, Matos C, Bruss C, Klobuch S, Peter K, Kastenberger
M, Bogdan C, Schleicher U, Mackensen A, Ullrich E,
Fichtner-Feigl S, Kesselring R, Mack M, Ritter U, Schmid M,
Blank C, Dettmer K, Oefner PJ, Hoffmann P, Walenta S,
Geissler EK, Pouyssegur J, Villunger A, Steven A, Seliger B,
Schreml S, Haferkamp S, Kohl E, Karrer S, Berneburg M,
Herr W, Mueller-Klieser W, Renner K & Kreutz M (2016).
LDHA-associated lactic acid production blunts tumor
immunosurveillance by T and NK cells. Cell Metab 24,
657–671.
Burgering BM & Coffer PJ (1995). Protein kinase B (c-Akt) in
phosphatidylinositol-3-OH kinase signal transduction.
Nature 376,599602.
Chaneton B, Hillmann P, Zheng L, Martin ACL, Maddocks
ODK, Chokkathukalam A, Coyle JE, Jankevics A, Holding
FP, Vousden KH, Frezza C, O’Reilly M & Gottlieb E (2012).
Serine is a natural ligand and allosteric activator of pyruvate
kinase M2. Nature 491,458462.
Chang CH, Qiu J, O’Sullivan D, Buck MD, Noguchi T, Curtis
JD, Chen Q, Gindin M, Gubin MM, van der Windt GJ, Tonc
E, Schreiber RD, Pearce EJ & Pearce EL (2015). Metabolic
competition in the tumor microenvironment is a driver of
cancer progression. Cell 162,12291241.
Chen F, Chen J, Yang L, Liu J, Zhang X, Zhang Y, Tu Q, Yin D,
Lin D, Wong PP, Huang D, Xing Y, Zhao J, Li M, Liu Q, Su
F, Su S & Song E (2019a). Extracellular vesicle-packaged
HIF-1alpha-stabilizing lncRNA from tumour-associated
macrophages regulates aerobic glycolysis of breast cancer
cells. Nat Cell Biol 21,498510.
Chen N, Hao C, Liu BC, Lin H, Wang C, Xing C, Liang X, Jiang
G, Liu Z, Li X, Zuo L, Luo L, Wang J, Zhao MH, Liu Z, Cai
GY, Hao L, Leong R, Wang C, Liu C, Neff T, Szczech L & Yu
KP (2019b). Roxadustat treatment for anemia in patients
undergoing long-term dialysis. N Engl J Med 381,
1011–1022.
Chen N, Hao C, Peng X, Lin H, Yin A, Hao L, Tao Y, Liang X,
Liu Z, Xing C, Chen J, Luo L, Zuo L, Liao Y, Liu BC, Leong
R, Wang C, Liu C, Neff T, Szczech L & Yu KP (2019c).
Roxadustat for anemia in patients with kidney disease not
receiving dialysis. N Engl J Med 381,10011010.
Choudhry H & Harris AL (2018). Advances in
hypoxia-inducible factor biology. Cell Metab 27,
281–298.
Choudhry H, Schodel J, Oikonomopoulos S, Camps C,
Grampp S, Harris AL, Ratcliffe PJ, Ragoussis J & Mole DR
(2014). Extensive regulation of the non-coding
transcriptome by hypoxia: role of HIF in releasing paused
RNApol2. EMBO Rep 15,7076.
Colegio OR, Chu NQ, Szabo AL, Chu T, Rhebergen AM,
Jairam V, Cyrus N, Brokowski CE, Eisenbarth SC, Phillips
GM, Cline GW, Phillips AJ & Medzhitov R (2014).
Functional polarization of tumour-associated macrophages
by tumour-derived lactic acid. Nature 513,559563.
Colgan SP, Furuta GT & Taylor CT (2020). Hypoxia and innate
immunity: keeping up with the HIFsters. Annu Rev Immunol
38,341363.
Corzo CA, Condamine T, Lu L, Cotter MJ, Youn JI, Cheng P,
Cho HI, Celis E, Quiceno DG, Padhya T, McCaffrey TV,
McCaffrey JC & Gabrilovich DI (2010). HIF-1alpha
regulates function and differentiation of myeloid-derived
suppressor cells in the tumor microenvironment. JExpMed
207,24392453.
Crifo B, Schaible B, Brown E, Halligan DN, Scholz CC,
Fitzpatrick SF, Kirwan A, Roche HM, Criscuoli M, Naldini
A, Giffney H, Crean D, Blanco A, Cavadas MA, Cummins
EP, Fabian Z & Taylor CT (2019). Hydroxylase inhibition
selectively induces cell death in monocytes. J Immunol 202,
1521–1530.
Culic O, Gruwel ML & Schrader J (1997). Energy turnover of
vascular endothelial cells. Am J Physiol Cell Physiol 273,
C205–C213.
Cummins EP, Seeballuck F, Keely SJ, Mangan NE, Callanan JJ,
Fallon PG & Taylor CT (2008). The hydroxylase inhibitor
dimethyloxalylglycine is protective in a murine model of
colitis. Gastroenterology 134,156165.e1.
De Bock K, Georgiadou M, Schoors S, Kuchnio A, Wong BW,
Cantelmo AR, Quaegebeur A, Ghesquiere B, Cauwenberghs
S, Eelen G, Phng LK, Betz I, Tembuyser B, Brepoels K, Welti
J, Geudens I, Segura I, Cruys B, Bifari F, Decimo I, Blanco R,
Wyns S, Vangindertael J, Rocha S, Collins RT, Munck S,
Daelemans D, Imamura H, Devlieger R, Rider M, Van
Veldhoven PP, Schuit F, Bartrons R, Hofkens J, Fraisl P,
Te l a ng S , D e b e r a r d i n i s RJ , S c h o o n j a n s L , V i nc k i e r S ,
Chesney J, Gerhardt H, Dewerchin M & Carmeliet P (2013).
Role of PFKFB3-driven glycolysis in vessel sprouting. Cell
154,651663.
DeBerardinis RJ, Lum JJ, Hatzivassiliou G & Thompson CB
(2008). The biology of cancer: metabolic reprogramming
fuels cell growth and proliferation. Cell Metab 7,1120.
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
12 S. J. Kierans and C. T. Taylor J Physiol 00.0
Dietl K, Renner K, Dettmer K, Timischl B, Eberhart K, Dorn C,
Hellerbrand C, Kastenberger M, Kunz-Schughart LA, Oefner
PJ, Andreesen R, Gottfried E & Kreutz MP (2010). Lactic
acid and acidification inhibit TNF secretion and glycolysis of
human monocytes. J Immunol 184,12001209.
Dobrina A & Rossi F (1983). Metabolic properties of freshly
isolated bovine endothelial cells. Biochim Biophys Acta 762,
295–301.
Doughty CA, Bleiman BF, Wagner DJ, Dufort FJ, Mataraza JM,
Roberts MF & Chiles TC (2006). Antigen receptor-mediated
changes in glucose metabolism in B lymphocytes: role of
phosphatidylinositol 3-kinase signaling in the glycolytic
control of growth. Blood 107,44584465.
Ebert BL, Firth JD & Ratcliffe PJ (1995). Hypoxia and
mitochondrial inhibitors regulate expression of glucose
transporter-1 via distinct Cis-acting sequences. J Biol Chem
270,2908329089.
Eelen G, de Zeeuw P, Treps L, Harjes U, Wong BW & Carmeliet
P (2018). Endothelial cell metabolism. Physiol Rev 98,
3–58.
Elstrom RL, Bauer DE, Buzzai M, Karnauskas R, Harris MH,
Plas DR, Zhuang H, Cinalli RM, Alavi A, Rudin CM &
Thompson CB (2004). Akt stimulates aerobic glycolysis in
cancer cells. Cancer Res 64,38923899.
Everts B, Amiel E, Huang SC, Smith AM, Chang CH, Lam WY,
Redmann V, Freitas TC, Blagih J, van der Windt GJ,
Artyomov MN, Jones RG, Pearce EL & Pearce EJ (2014).
TLR-driven early glycolytic reprogramming via the kinases
TBK1-IKKvarepsilon supports the anabolic demands of
dendritic cell activation. Nat Immunol 15,323332.
Fischer K, Hoffmann P, Voelkl S, Meidenbauer N, Ammer J,
Edinger M, Gottfried E, Schwarz S, Rothe G, Hoves S,
Renner K, Timischl B, Mackensen A, Kunz-Schughart L,
Andreesen R, Krause SW & Kreutz M (2007). Inhibitory
effect of tumor cell-derived lactic acid on human T cells.
Blood 109,38123819.
Formenti F, Constantin-Teodosiu D, Emmanuel Y, Cheeseman
J, Dorrington KL, Edwards LM, Humphreys SM, Lappin TR,
McMullin MF, McNamara CJ, Mills W, Murphy JA,
O’Connor DF, Percy MJ, Ratcliffe PJ, Smith TG, Treacy M,
Frayn KN, Greenhaff PL, Karpe F, Clarke K & Robbins PA
(2010). Regulation of human metabolism by
hypoxia-inducible factor. Proc Natl Acad Sci U S A. 107,
12722–12727.
Franke TF, Yang SI, Chan TO, Datta K, Kazlauskas A, Morrison
DK, Kaplan DR & Tsichlis PN (1995). The protein kinase
encoded by the Akt proto-oncogene is a target of the
PDGF-activated phosphatidylinositol 3-kinase. Cell 81,
727–736.
Frauwirth KA, Riley JL, Harris MH, Parry RV, Rathmell JC,
Plas DR, Elstrom RL, June CH & Thompson CB (2002). The
CD28 signaling pathway regulates glucose metabolism.
Immunity 16,769777.
Fukuda R, Zhang H, Kim JW, Shimoda L, Dang CV & Semenza
GL (2007). HIF-1 regulates cytochrome oxidase subunits to
optimize efficiency of respiration in hypoxic cells. Cell 129,
111–122.
Ganeshan K & Chawla A (2014). Metabolic regulation of
immune responses. Annu Rev Immunol 32,609634.
Goetze K, Walenta S, Ksiazkiewicz M, Kunz-Schughart LA &
Mueller-Klieser W (2011). Lactate enhances motilit y of
tumor cells and inhibits monocyte migration and cytokine
release. Int J Oncol 39,453463.
Gottfried E, Kunz-Schughart LA, Ebner S, Mueller-Klieser W,
Hoves S, Andreesen R, Mackensen A & Kreutz M (2006).
Tumor-derived lactic acid modulates dendritic cell
activation and antigen expression. Blood 107,20132021.
Gruber M, Hu CJ, Johnson RS, Brown EJ, Keith B & Simon MC
(2007). Acute postnatal ablation of Hif-2alpha results in
anemia. Proc Natl Acad Sci U S A 104,23012306.
Henderson AR (1969). Biochemistry of hypoxia: current
concepts. I. An introduction to biochemical pathways and
their control. Br J Anaesth 41,245250.
Hodson EJ, Nicholls LG, Turner PJ, Llyr R, Fielding JW,
Douglas G, Ratnayaka I, Robbins PA, Pugh CW, Buckler KJ,
Ratcliffe PJ & Bishop T (2016). Regulation of ventilatory
sensitivity and carotid body proliferation in hypoxia by the
PHD2/HIF-2 pathway. JPhysiol594,11791195.
Horecker BL (2002). The pentose phosphate pathway. JBiol
Chem 277,4796547971.
Hu CJ, Sataur A, Wang L, Chen H & Simon MC (2007). The
N-terminal transactivation domain confers target gene
specificity of hypoxia-inducible factors HIF-1alpha and
HIF-2alpha. Mol Biol Cell 18,45284542.
Hu CJ, Wang LY, Chodosh LA, Keith B & Simon MC (2003).
Differential roles of hypoxia-inducible factor 1alpha
(HIF-1alpha) and HIF-2alpha in hypoxic gene regulation.
Mol Cell Biol 23,93619374.
Huarte M, Guttman M, Feldser D, Garber M, Koziol MJ,
Kenzelmann-Broz D, Khalil AM, Zuk O, Amit I, Rabani M,
Attardi LD, Regev A, Lander ES, Jacks T & Rinn JL (2010). A
large intergenic noncoding RNA induced by p53 mediates
global gene repression in the p53 response. Cell 142,
409–419.
Ito M, Tanaka T, Ishii T, Wakashima T, Fukui K & Nangaku M
(2020). Prolyl hydroxylase inhibition protects the kidneys
from ischemia via upregulation of glycogen storage. Kidney
Int 97,687701.
Ivan M, Kondo K, Yang H, Kim W, Valiando J, Ohh M, Salic A,
Asara JM, Lane WS & Kaelin WG Jr (2001). HIFalpha
targeted for VHL-mediated destruction by proline
hydroxylation: implications for O2 sensing. Science 292,
464–468.
Iyer NV, Kotch LE, Agani F, Leung SW, Laughner E, Wenger
RH, Gassmann M, Gearhart JD, Lawler AM, Yu AY &
Semenza GL (1998). Cellular and developmental control of
O2 homeostasis by hypoxia-inducible factor 1 alpha. Genes
Dev 12,149162.
Jaakkola P, Mole DR, Tian YM, Wilson MI, Gielbert J, Gaskell
SJ, von Kriegsheim A, Hebestreit HF, Mukherji M, Schofield
CJ, Maxwell PH, Pugh CW & Ratcliffe PJ (2001). Targeting of
HIF-alpha to the von Hippel-Lindau ubiquitylation complex
by O2-regulated prolyl hydroxylation. Science 292,468472.
Jin M, Fuller GG, Han T, Yao Y, Alessi AF, Freeberg MA, Roach
NP, Moresco JJ, Karnovsky A, Baba M, Yates JR 3rd, Gitler
AD, Inoki K, Klionsky DJ & Kim JK (2017). Glycolytic
enzymes coalesce in G bodies under hypoxic stress. Cell Rep
20,895908.
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
J Physiol 00.0 HIF-dependent regulation of glycolysis 13
Jurica MS, Mesecar A, Heath PJ, Shi W, Nowak T & Stoddard
BL (1998). The allosteric regulation of pyruvate kinase by
fructose-1,6-bisphosphate. Structure 6,195210.
Kemp RG & Gunasekera D (2002). Evolution of the allosteric
ligand sites of mammalian phosphofructo-1-kinase.
Biochemistry 41,94269430.
Kim JW, Tchernyshyov I, Semenza GL & Dang CV (2006).
HIF-1-mediated expression of pyruvate dehydrogenase
kinase: a metabolic switch required for cellular adaptation to
hypoxia. Cell Metab 3,177185.
Knowles HJ, Raval RR, Harris AL & Ratcliffe PJ (2003). Effect
of ascorbate on the activity of hypoxia-inducible factor in
cancer cells. Cancer Res 63,17641768.
Kohn AD, Kovacina KS & Roth RA (1995). Insulin stimulates
the kinase activity of RAC-PK, a pleckstrin homology
domain containing ser/thr kinase. EMBO J 14,
4288–4295.
Kornberg HL (1976). Carbohydrate uptake by Escherichia coli.
JCellPhysiol89,545549.
Krawczyk CM, Holowka T, Sun J, Blagih J, Amiel E,
DeBerardinis RJ, Cross JR, Jung E, Thompson CB, Jones RG
&PearceEJ(2010).Toll-likereceptor-inducedchangesin
glycolytic metabolism regulate dendritic cell activation.
Blood 115,47424749.
Krieg AJ, Rankin EB, Chan D, Razorenova O, Fernandez S &
Giaccia AJ (2010). Regulation of the histone demethylase
JMJD1A by hypoxia-inducible factor 1 alpha enhances
hypoxic gene expression and tumor growth. Mol Cell Biol 30,
344–353.
Lee CT, Mace T & Repasky EA (2010). Hypoxia-driven
immunosuppression: a new reason to use thermal therapy in
the treatment of cancer? Int J Hyperthermia 26,
232–246.
Li Z, Li X, Wu S, Xue M & Chen W (2014). Long non-coding
RNA UCA1 promotes glycolysis by upregulating hexokinase
2 through the mTOR-STAT3/microRNA143 pathway.
Cancer Sci 105,951955.
Liu F, Huang X, Luo Z, He J, Haider F, Song C, Peng L, Chen T
&WuB(2019).Hypoxia-activatedPI3K/Aktinhibits
oxidative stress via the regulation of reactive oxygen species
in human dental pulp cells. Oxid Med Cell Longev 2019,
6595189.
Liu RT, Zhang M, Yang CL, Zhang P, Zhang N, Du T, Ge MR,
Yue LT, L i X L, Li H & Duan R S ( 2 0 1 8). Enha n c e d g lyco l y s i s
contributes to the pathogenesis of experimental
autoimmune neuritis. J Neuroinflammation 15,51.
Macias D, Cowburn AS, Torres-Torrelo H, Ortega-Saenz P,
Lopez-Barneo J & Johnson RS (2018). HIF-2alpha is
essential for carotid body development and function. Elife 7,
e34681.
Maciver NJ, Jacobs SR, Wieman HL, Wofford JA, Coloff JL &
Rathmell JC (2008). Glucose metabolism in lymphocytes is a
regulated process with significant effects on immune cell
function and survival. JLeukocBiol84,
949–957.
Mahon PC, Hirota K & Semenza GL (2001). FIH-1: a novel
protein that interacts with HIF-1alpha and VHL to mediate
repression of HIF-1 transcriptional activity. Genes Dev 15,
2675–2686.
Manresa MC, Smith L, Casals-Diaz L, Fagundes RR, Brown E,
Radhakrishnan P, Murphy SJ, Crifo B, Strowitzki MJ,
Halligan DN, van den Bogaard EH, Niehues H,
Schneider M, Taylor CT & Steinhoff M (2019).
Pharmacologic inhibition of hypoxia-inducible factor
(HIF)-hydroxylases ameliorates allergic contact dermatitis.
Allergy 74,753766.
Mastrogiannaki M, Matak P, Keith B, Simon MC, Vaulont S &
Peyssonnaux C (2009). HIF-2alpha, but not HIF-1alpha,
promotes iron absorption in mice. JClinInvest119,
1159–1166.
Merchan JR, Kovacs K, Railsback JW, Kurtoglu M, Jing Y, Pina
Y, Gao N, Murray TG, Lehrman MA & Lampidis TJ (2010).
Antiangiogenic activity of 2-deoxy-D-glucose. PLoS One 5,
e13699.
Michalek RD, Gerriets VA, Jacobs SR, Macintyre AN, MacIver
NJ, Mason EF, Sullivan SA, Nichols AG & Rathmell JC
(2011). Cutting edge: distinct glycolytic and lipid oxidative
metabolic programs are essential for effector and
regulatory CD4+T cell subsets. J Immunol 186,
3299–3303.
Michl J, Ohlbaum DJ & Silverstein SC (1976). 2-Deoxyglucose
selectively inhibits Fc and complement receptor-mediated
phagocytosis in mouse peritoneal macrophages II.
Dissociation of the inhibitory effects of 2-deoxyglucose on
phagocytosis and ATP generation. JExpMed144,
1484–1493.
Milkiewicz M, Pugh CW & Egginton S (2004). Inhibition of
endogenous HIF inactivation induces angiogenesis in
ischaemic skeletal muscles of mice. JPhysiol560,
21–26.
Millrud CR, Bergenfelz C & Leandersson K (2017). On the
origin of myeloid-derived suppressor cells. Oncotarget 8,
3649–3665.
Mole DR, Schlemminger I, McNeill LA, Hewitson KS, Pugh
CW, Ratcliffe PJ & Schofield CJ (2003). 2-Oxoglutarate
analogue inhibitors of HIF prolyl hydroxylase. Bioorg Med
Chem Lett 13,26772680.
Moses V (1959). [14C] Glucose metabolism in fungal cells. J
Gen Microbiol 20,184196.
O’Rourke JF, Tian YM, Ratcliffe PJ & Pugh CW (1999).
Oxygen-regulated and transactivating domains in
endothelial PAS protein 1: comparison with
hypoxia-inducible factor-1alpha. J Biol Chem 274,
2060–2071.
Ou X & Lv W (2020). Metabolic changes and interaction of
tumor cell, myeloid-derived suppressor cell and T cell in
hypoxic microenvironment. Future Oncol 16,
383–393.
Page EL, Chan DA, Giaccia AJ, Levine M & Richard DE (2008).
Hypoxia-inducible factor-1alpha stabilization in nonhypoxic
conditions: role of oxidation and intracellular ascorbate
depletion. Mol Biol Cell 19,8694.
Papandreou I, Cairns RA, Fontana L, Lim AL & Denko NC
(2006). HIF-1 mediates adaptation to hypoxia by actively
downregulating mitochondrial oxygen consumption. Cell
Metab 3,187197.
Pasteur L (1861). Exp ´
eriences et vues nouvelles sur la nature
des fermentations. CRAcadSci52,12601264.
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
14 S. J. Kierans and C. T. Taylor J Physiol 00.0
Perrotta S, Roberti D, Bencivenga D, Corsetto P, O’Brien KA,
Caiazza M, Stampone E, Allison L, Fleck RA, Scianguetta S,
Tartaglione I, Robbins PA, Casale M, West JA,
Franzini-Armstrong C, Griffin JL, Rizzo AM, Sinisi AA,
Murray AJ, Borriello A, Formenti F & Della Ragione F
(2020). Effects of germline VHL deficiency on growth,
metabolism, and mitochondria. N Engl J Med 382,835844.
Pessin JE & Bell GI (1992). Mammalian facilitative glucose
transporter family: structure and molecular regulation.
Annu Rev Physiol 54,911930.
Plaxton WC (1996). The organization and regulation of plant
glycolysis. Annu Rev Plant Physiol Plant Mol Biol 47,
185–214.
Rankin EB, Biju MP, Liu Q, Unger TL, Rha J, Johnson RS,
Simon MC, Keith B & Haase VH (2007). Hypoxia-inducible
factor-2 (HIF-2) regulates hepatic erythropoietin in vivo. J
Clin Invest 117,10681077.
Rinn JL & Chang HY (2012). Genome regulation by long
noncoding RNAs. Annu Rev Biochem 81,145166.
Sbarra AJ & Karnovsky ML (1959). The biochemical basis of
phagocytosis. I. Metabolic changes during the ingestion of
particles by polymorphonuclear leukocytes. J Biol Chem 234,
1355–1362.
Schodel J, Oikonomopoulos S, Ragoussis J, Pugh CW, Ratcliffe
PJ & Mole DR (2011). High-resolution genome-wide
mapping of HIF-binding sites by ChIP-seq. Blood 117,
e207–e217.
Seagroves TN, Ryan HE, Lu H, Wouters BG, Knapp M,
Thibault P, Laderoute K & Johnson RS (2001). Transcription
factor HIF-1 is a necessary mediator of the pasteur effect in
mammalian cells. Mol Cell Biol 21,34363444.
Selig M, Xavier KB, Santos H & Schonheit P (1997).
Comparative analysis of Embden-Meyerhof and
Entner-Doudoroff glycolytic pathways in hyperthermophilic
archaea and the bacterium Thermotoga. Arch Microbiol 167,
217–232.
Semenza GL (2011). Oxygen sensing, homeostasis, and disease.
N Engl J Med 365,537547.
Semenza GL, Jiang BH, Leung SW, Passantino R, Concordet JP,
Maire P & Giallongo A (1996). Hypoxia response elements
in the aldolase A, enolase 1, and lactate dehydrogenase A
gene promoters contain essential binding sites for
hypoxia-inducible factor 1. JBiolChem271,3252932537.
Semenza GL, Roth PH, Fang HM & Wang GL (1994).
Transcriptional regulation of genes encoding glycolytic
enzymes by hypoxia-inducible factor 1. J Biol Chem 269,
23757–23763.
Shen G, Li X, Jia YF, Piazza GA & Xi Y (2013).
Hypoxia-regulated microRNAs in human cancer. Acta
Pharmacol Sin 34,336341.
Shi LZ, Wang R, Huang G, Vogel P, Neale G, Green DR & Chi
H (2011). HIF1alpha-dependent glycolytic pathway
orchestrates a metabolic checkpoint for the differentiation of
TH17 and Treg cells. JExpMed208,13671376.
Shih JW, Chiang WF, Wu ATH, Wu MH, Wang LY, Yu YL,
Hung YW, Wang WC, Chu CY, Hung CL, Changou CA, Yen
Y&KungHJ(2017).LongnoncodingRNA
LncHIFCAR/MIR31HG is a HIF-1alpha co-activator driving
oral cancer progression. Nat Commun 8,15874.
Sonveaux P, Copetti T, De Saedeleer CJ, Vegran F, Verrax J,
Kennedy KM, Moon EJ, Dhup S, Danhier P, Frerart F, Gallez
B, Ribeiro A, Michiels C, Dewhirst MW & Feron O (2012).
Targeting the lactate transporter MCT1 in endothelial cells
inhibits lactate-induced HIF-1 activation and tumor
angiogenesis. PLoS One 7,e33418.
Sukumar M, Liu J, Ji Y, Subramanian M, Crompton JG, Yu Z,
Roychoudhuri R, Palmer DC, Muranski P, Karoly ED,
Mohney RP, Klebanoff CA, Lal A, Finkel T, Restifo NP &
Gattinoni L (2013). Inhibiting glycolytic metabolism
enhances CD8+Tcellmemoryandantitumorfunction.J
Clin Invest 123,44794488.
Taylor CT & Colgan SP (2017). Regulation of immunity and
inflammation by hypoxia in immunological niches. Nat Rev
Immunol 17,774785.
Tian H, McKnight SL & Russell DW (1997). Endothelial PAS
domain protein 1 (EPAS1), a transcription factor selectively
expressed in endothelial cells. Genes Dev 11,7282.
To j o K , Ta m a d a N , N a ga m in e Y , Ya z a w a T , O t a S & G o t o T
(2018). Enhancement of glycolysis by inhibition of
oxygen-sensing prolyl hydroxylases protects alveolar
epithelial cells from acute lung injury. FASEB J 32,
2258–2268.
Troy H, Chun g Y , M a y r M , L y L , Wi l l i a m s K J, St r a t ford I,
Harris A, Griffiths JR & Stubbs M (2005). Metabolic
profiling of hypoxia-inducible factor-1β-deficient and wild
type Hepa-1 cells: Effects of hypoxia measured by 1H
magnetic resonance spectroscopy. Metabolomics 1,
293–303.
Ullah MS, Davies AJ & Halestrap AP (2006). The plasma
membrane lactate transporter MCT4, but not MCT1, is
up-regulated by hypoxia through a HIF-1alpha-dependent
mechanism. J Biol Chem 281,90309037.
Valen t i n e W N & B e ck WS (195 1 ) . B i o c h e mica l s t u d i e s on
leucocytes. I. Phosphatase activity in health, leucocytosis,
and myelocytic leucemia. JLabClinMed38,3955.
Vandekeere S, Dewerchin M & Carmeliet P (2015).
Angiogenesis revisited: an overlooked role of endothelial cell
metabolism in vessel sprouting. Microcirculation 22,
509–517.
Veg r an F , B o id o t R , Mi c h i e l s C , S o nv e a u x P & Fe r o n O ( 20 1 1 ).
Lactate influx through the endothelial cell monocarboxylate
transporter MCT1 supports an NF-kappaB/IL-8 pathway
that drives tumor angiogenesis. Cancer Res 71,25502560.
Wan C, Borgeson B, Phanse S, Tu F, Drew K, Clark G, Xiong X,
Kagan O, Kwan J, Bezginov A, Chessman K, Pal S, Cromar
G, Papoulas O, Ni Z, Boutz DR, Stoilova S, Havugimana PC,
Guo X, Malty RH, Sarov M, Greenblatt J, Babu M, Derry
WB, Tillier ER, Wallingford JB, Parkinson J, Marcotte EM &
Emili A (2015). Panorama of ancient metazoan
macromolecular complexes. Nature 525,339344.
Wang GL, Jiang BH, Ru e E A & S e m e n z a G L ( 1 9 9 5 ) .
Hypoxia-inducible factor 1 is a basic-helix-loop-helix-PAS
heterodimer regulated by cellular O2tension. Proc Natl Acad
Sci U S A 92,55105514.
Wang GL & Semenza GL (1993). Desferr i oxamine induces
erythropoietin gene expression and hypoxia-inducible factor
1 DNA-binding activity: implications for models of hypoxia
signal transduction. Blood 82,36103615.
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
J Physiol 00.0 HIF-dependent regulation of glycolysis 15
Wang R, Dillon CP, Shi LZ, Milasta S, Carter R, Finke l s t e i n D ,
McCormick LL, Fitzgerald P, Chi H, Munger J & Green DR
(2011). The transcription factor Myc controls metabolic
reprogramming upon T lymphocyte activation. Immunity
35,871882.
Wang V, Davis DA, Haque M, Huang LE & Yarchoan R (2005).
Differential gene up-regulation by hypoxia-inducible
factor-1alpha and hypoxia-inducible factor-2alpha in
HEK293T cells. Cancer Res 65,32993306.
Warburg O (1956). On respiratory impairment in cancer cells.
Science 124,269270.
Wieman HL, Wofford JA & Rathmell JC (2007). Cytokine
stimulation promotes glucose uptake via
phosphatidylinositol-3 kinase/Akt regulation of
Glut1 activity and trafficking. Mol Biol Cell 18,
1437–1446.
Wiesener MS, Jurgensen JS, Rosenberger C, Scholze CK,
Horstrup JH, Warnecke C, Mandriota S, Bechmann I, Frei
UA, Pugh CW, Ratcliffe PJ, Bachmann S, Maxwell PH &
Eckardt KU (2003). Widespread hypoxia-inducible
expression of HIF-2alpha in distinct cell populations of
different organs. FASEB J 17,271273.
Xie Y, Shi X, Sheng K, Han G, Li W, Zhao Q, Jiang B, Feng J, Li
J & Gu Y (2019). PI3K/Akt signaling transduction pathway,
erythropoiesis and glycolysis in hypoxia (Review). Mol Med
Rep 19,783791.
Xue M, Li X, Li Z & Chen W (2014). Urothelial carcinoma
associated 1 is a hypoxia-inducible factor-1alpha-targeted
long noncoding RNA that enhances hypoxic bladder cancer
cell proliferation, migration, and invasion. Tumor Biol 35,
6901–6912.
Yang C, Jiang L, Zhang H, Shimoda LA, DeBerardinis RJ &
Semenza GL (2014a). Analysis of hypoxia-induced
metabolic reprogramming. Methods Enzymol 542,
425–455.
Yan g F , Zh a ng H , M e i Y & Wu M ( 2 0 1 4 b). Reciprocal
regulation of HIF-1alpha and lincRNA-p21 modulates the
Warburg effect. Mol Cell 53,88100.
Yang Z, Fujii H, Mohan SV, Goronzy JJ & Weyand CM (2013).
Phosphofructokinase deficiency impairs ATP generation,
autophagy, and redox balance in rheumatoid arthritis T cells.
JExpMed210,21192134.
Yin Y, Choi SC, Xu Z, Zeumer L, Kanda N, Croker BP & Morel
L(2016).GlucoseOxidationIsCriticalforCD4+TCell
Activation in a Mouse Model of Systemic Lupus
Erythematosus. J Immunol 196,8090.
Yu F, White SB, Zhao Q & Lee FS (2001). HIF-1alpha binding
to VHL is regulated by stimulus-sensitive proline
hydroxylation. Proc Natl Acad Sci U S A 98,
9630–9635.
Zeng L, Zhou HY, Tang NN, Zhang WF, He GJ, Hao B, Feng YD
&ZhuH(2016).Wortmannininuenceshypoxia-inducible
factor-1 alpha expression and glycolysis in esophageal
carcinoma cells. World J Gastroenterol 22,48684880.
Zhang D, Tang Z, Huang H, Zhou G, Cui C, Weng Y, Liu W,
Kim S, Lee S, Perez-Neut M, Ding J, Czyz D, Hu R, Ye Z, He
M, Zheng YG, Shuman HA, Dai L, Ren B, Roeder RG, Becker
L&ZhaoY(2019).Metabolicregulationofgeneexpression
by histone lactylation. Nature 574,575580.
Zhang H, Bosch-Marce M, Shimoda LA, Tan YS, Baek JH,
Wesley JB, Gonzalez FJ & Semenza GL (2008). Mitochondrial
autophagy is an HIF-1-dependent adaptive metabolic
response to hypoxia. J Biol Chem 283,1089210903.
Zhong H, Chiles K, Feldser D, Laughner E, Hanrahan C,
Georgescu MM, Simons JW & Semenza GL (2000).
Modulation of hypoxia-inducible factor 1alpha expression
by the epidermal growth factor/phosphatidylinositol
3-kinase/PTEN/AKT/FRAP pathway in human prostate
cancer cells: implications for tumor angiogenesis and
therapeutics. Cancer Res 60,15411545.
Zundel W, Schindler C, Haas-Kogan D, Koong A, Kaper F,
Chen E, Gottschalk AR, Ryan HE, Johnson RS, Jefferson AB,
Stokoe D & Giaccia AJ (2000). Loss of PTEN facilitates
HIF-1-mediated gene expression. Genes Dev 14,391396.
Additional information
Competing interests
The authors declare no competing interests.
Author contributions
Both authors have approved the final version of the
manuscript and agree to be accountable for all aspects
of the work. All persons designated as authors qualify for
authorship, and all those who qualify for authorship are
listed.
Funding
S.J.K. is funded by a UCD Advance PhD Course Scheme
grant awarded to C.T.T. (R19448).
Keywords
glycolysis, HIF, hypoxia, metabolism
C
!2020 The Authors. The Journal of Physiology C
!2020 The Physiological Society
... The PFK is one of the most prominent rate-limiting enzymes in the glycolytic pathway. The increase in glycolytic flux is beneficial for maintaining bioenergetic homeostasis during the hypoxia (Kierans and Taylor 2021). ...
Article
Full-text available
Purpose Vibrio parahaemolyticus is a leading seafood borne pathogen worldwide. The aim of this study was to decipher the response mechanism of V. parahaemolyticus isolates of clinical and aquatic animal origins to the hypoxic condition, which challenges the bacterial survival in the host and in the environment. Methods Growth profiles of V. parahaemolyticus isolates ( n = 5) of clinical and aquatic animal origins were examined at different stress conditions (osmolality, acid, temperature, and O 2 concentrations). Draft genomes of the V. parahaemolyticus isolates were determined using the Illumina sequencing technique. Comparative genomic analysis were performed to identify and validate the hypoxic tolerance-related genes. Results The V. parahaemolyticus isolates had an oxygen concentration-dependent growth mode, and the 10% O 2 condition strongly inhibited the bacterial growth, when incubated in TSB medium (pH 8.5, 3% NaCl) at 37 °C. Unexpectedly, in marked contrast to the normal 21% O 2 condition, the 10% O 2 treatment for 24 h significantly increased biofilm formation of V. parahaemolyticus isolates ( p < 0.05). Draft genome sequences of four V. parahaemolyticus isolates of aquatic animal origins were determined (4.914–5.3530 Mb), which carried mobile genetic elements (n = 12–29). Genome-wide gene expression changes triggered by the hypoxic condition were further examined. Comparative transcriptomic analyses unveiled multiple molecular strategies employed by the bacterium to mitigate the cell damage caused by the hypoxia. Of note, the pathogenic V. parahaemolyticus ATCC17802 down-regulated and/or shut down ten metabolic pathways to reduce cell viability and maintain cell structure under the hypoxic stress. Conclusions The results of this study fill prior gaps in the response mechanism of V. parahaemolyticus to the hypoxic condition. Different tolerance to hypoxia contributes to the persistence of pathogenic V. parahaemolyticus in the niches.
... Since HIF-1a is induced at low oxygen and promotes cell survival (41)(42)(43)(44)(45), it was prudent to determine whether HIF-1a is responsible for the attenuated response of ATM to DSBs and critically short telomeres at low oxygen. An inducible shRNA that reduced HIF-1a expression detected by western blotting and IF did not affect the induction of g-H2AX foci in response to zeocin at 3% or 20% oxygen ( Fig. 4A-C). ...
Preprint
Full-text available
Programmed telomere shortening limits tumorigenesis through the induction of replicative senescence. Here we address three long-standing questions concerning senescence. First, we show that the ATM kinase is solely responsible for the induction of replicative senescence. Senescence was delayed by ATM inhibition (ATMi) or overexpression of TRF2, the shelterin subunit dedicated to ATM repression. In contrast, there was no evidence for ATR signaling contributing to replicative senescence even when ATMi was combined with ATR inhibition. Second, we show ATMi can induce apparently normal cell divisions in a subset of senescent cells, indicating that senescence can be reversed. Third, we show that the extended replicative life span at low (physiological) oxygen is due to diminished ATM activity. At low oxygen, cells show a decreased ATM response to dysfunctional telomeres and genome-wide DSBs compared to 20% oxygen. As this effect could be reversed by NAC, the attenuated response of ATM to critically short telomeres and the resulting extended life span at low oxygen is likely due to ROS-induced formation of cysteine disulfide-bridges that crosslink ATM dimers into a form that is not activated by DSBs. These findings show how primary human cells detect shortened telomeres and reveal the molecular mechanism underlying the telomere tumor suppressor pathway.
... Limiting the ischemic time is crucial for maintaining the vitality of tissues after reperfusion. Cells in hypoxia activate pro-angiogenic mechanisms by stabilizing HIF-1α to promote angiogenesis and restore tissue blood flow [10]. However, reperfusion following ischemia is also detrimental due to the increased production of reactive oxygen species (ROS) which damage cellular components promoting apoptotic or necrotic pathways [11]. ...
Article
Full-text available
MicroRNAs play a crucial role in regulating the immune responses induced by ischemia/reperfusion injury. Through their ability to modulate gene expression, microRNAs adjust immune responses by targeting specific genes and signaling pathways. This review focuses on the impact of microRNAs on the inflammatory pathways triggered during ischemia/reperfusion injury and highlights their ability to modulate inflammation, playing a critical role in the pathophysiology of ischemia/reperfusion injury. Dysregulated expression of microRNAs contributes to the pathogenesis of ischemia/reperfusion injury, therefore targeting specific microRNAs offers an opportunity to restore immune homeostasis and improve patient outcomes. Understanding the complex network of immunoregulatory microRNAs could provide novel therapeutic interventions aimed at attenuating excessive inflammation and preserving tissue integrity.
... Taken together, the results suggest that the pathogen utilizes the miR-2013-3p/SinHIF-1α axis from 6 to 24 hpi and from 48 to 72 hpi to reduce ATP production and thereby attenuate the antimicrobial capability of the host (Figure 8). Typically, the enhancement of expression and activity of key enzymes of glycolysis are key nodes linking HIF-1α homologous gene expression and ATP production under hypoxic conditions in organisms [61,[71][72][73]. In this study, we found that changes in GDH expression and enzyme activity were consistent with the trend of SinHIF-1α expression but opposite to that of miR-2013-3p in the coelomocytes of S. intermedius after pathogen infection under normoxic conditions. ...
Article
Full-text available
To investigate the functions of hypoxia inducible factor-1α (HIF-1α) homologs, we cloned and characterized the full-length cDNA of a novel HIF-1α homolog (SinHIF-1α) from the sea urchin Strongylocentrotus intermedius. We then explored the functions of SinHIF-1α and its regulatory factor (miR-2013-3p) in the pathogen-induced metabolic response of S. intermedius. The results showed the following: 1) The full-length cDNA of SinHIF-1α was 4265 bp, with a high level of sequence conservation across the phylum Echinodermata. 2) MiR-2013-3p could bind to the 3’-UTR of SinHIF-1α and negatively regulate the expression of SinHIF-1α in S. intermedius. 3) Both SinHIF-1α silencing and miR-2013-3p overexpression suppressed the relative content of adenosine triphosphate (ATP) in coelomocytes as well as the relative expression (mRNA and protein) and total enzyme activity of glutamate dehydrogenase (GDH) in S. intermedius. 4) The miR-2013-3p/SinHIF-1α axis may regulate ATP production at 6–24 h and 48–72 h post-pathogen infection through accompanied by alterations in the relative expression and enzyme activity of GDH in S. intermedius. In summary, all observations from this study indicated that the pathogen utilizes the miR-2013-3p/SinHIF-1α axis to reduce ATP production and thereby attenuate the antimicrobial capability of the host via targeting GDH.
... HIF-1 controls the expression of more than 700 different target genes, and the major target genes are related to angiogenesis and energy metabolism. Of particular concern, HIF-1 regulates the expression of erythropoietin (EPO) and vascular endothelial growth factor (VEGF), as well as genes that regulate glucose transport or glucose metabolism, such as SLC2A1, GAPDH, PDK1, and LDHA [1,9]. ...
Article
Full-text available
Hypoxia can cause a variety of diseases, including ischemic stroke and neurodegenerative diseases. Within a certain range of partial pressure of oxygen, cells can respond to changes in oxygen. Changes in oxygen concentration beyond a threshold will cause damage or even necrosis of tissues and organs, especially for the central nervous system. Therefore, it is very important to find appropriate measures to alleviate damage. MiRNAs can participate in the regulation of hypoxic responses in various types of cells. MiRNAs are involved in regulating hypoxic responses in many types of tissues by activating the hypoxia-inducible factor (HIF) to affect angiogenesis, glycolysis and other biological processes. By analyzing differentially expressed miRNAs in hypoxia and hypoxia-related studies, as well as the HT22 neuronal cell line under hypoxic stress, we found that the expression of miR-18a was changed in these models. MiR-18a could regulate glucose metabolism in HT22 cells under hypoxic stress by directly regulating the 3’UTR of the Hif1a gene. As a small molecule, miRNAs are easy to be designed into small nucleic acid drugs, so this study can provide a theoretical basis for the research and treatment of nervous system diseases caused by hypoxia. Graphical Abstract
... The importance of this process is that it allows for the regeneration of oxidized NAD + , which can then be used in repetitive cycles of glycolysis to allow for continued ATP generation via anaerobic mechanisms (Spinicci et al., 2022). To further enhance this shift in metabolism, HIF-1 also acts to upregulate expression of all enzymes in the glycolytic pathway, as well as glucose transporters GLUT1 and GLUT3 in order to increase cellular glucose uptake and maintain sufficient levels of ATP production (Kierans and Taylor, 2021;Chang et al., 2022). In summary, the HIF-mediated shift in cellular metabolism away from aerobic respiration significantly reduces oxygen demand, while also reducing mitochondrial ROS production and protecting the cell against radical-mediated damage (Samanta and Semenza, 2017) (Figure 1). ...
Article
Full-text available
The complex and dynamic interaction between cellular energy control and gene expression modulation is shown by the intersection between mitochondrial metabolism and epigenetics in hypoxic environments. Poor oxygen delivery to tissues, or hypoxia, is a basic physiological stressor that sets off a series of reactions in cells to adapt and endure oxygen-starved environments. Often called the “powerhouse of the cell,” mitochondria are essential to cellular metabolism, especially regarding producing energy through oxidative phosphorylation. The cellular response to hypoxia entails a change in mitochondrial metabolism to improve survival, including epigenetic modifications that control gene expression without altering the underlying genome. By altering the expression of genes involved in angiogenesis, cell survival, and metabolism, these epigenetic modifications help cells adapt to hypoxia. The sophisticated interplay between mitochondrial metabolism and epigenetics in hypoxia is highlighted by several important points, which have been summarized in the current article. Deciphering the relationship between mitochondrial metabolism and epigenetics during hypoxia is essential to understanding the molecular processes that regulate cellular adaptation to reduced oxygen concentrations.
Preprint
Impaired cerebral glucose metabolism is a pathologic feature of Alzheimer Disease (AD), and recent proteomic studies highlight a disruption of glial carbohydrate metabolism with disease progression. Here, we report that inhibition of indoleamine-2,3-dioxygenase 1 (IDO1), which metabolizes tryptophan to kynurenine (KYN) in the first step of the kynurenine pathway, rescues hippocampal memory function and plasticity in preclinical models of amyloid and tau pathology by restoring astrocytic metabolic support of neurons. Activation of IDO1 in astrocytes by amyloid-beta42 and tau oligomers, two major pathological effectors in AD, increases KYN and suppresses glycolysis in an AhR-dependent manner. Conversely, pharmacological IDO1 inhibition restores glycolysis and lactate production. In amyloid-producing APPSwe-PS1∆E9 and 5XFAD mice and in tau-producing P301S mice, IDO1 inhibition restores spatial memory and improves hippocampal glucose metabolism by metabolomic and MALDI-MS analyses. IDO1 blockade also rescues hippocampal long-term potentiation (LTP) in a monocarboxylate transporter (MCT)-dependent manner, suggesting that IDO1 activity disrupts astrocytic metabolic support of neurons. Indeed, in vitro mass-labeling of human astrocytes demonstrates that IDO1 regulates astrocyte generation of lactate that is then taken up by human neurons. In co-cultures of astrocytes and neurons derived from AD subjects, deficient astrocyte lactate transfer to neurons was corrected by IDO1 inhibition, resulting in improved neuronal glucose metabolism. Thus, IDO1 activity disrupts astrocytic metabolic support of neurons across both amyloid and tau pathologies and in a model of AD iPSC-derived neurons. These findings also suggest that IDO1 inhibitors developed for adjunctive therapy in cancer could be repurposed for treatment of amyloid- and tau-mediated neurodegenerative diseases.
Article
The heart is composed of a heterogeneous mixture of cellular components perfectly intermingled and able to integrate common environmental signals to ensure proper cardiac function and performance. Metabolism defines a cell context-dependent signature that plays a critical role in survival, proliferation, or differentiation, being a recognized master piece of organ biology, modulating homeostasis, disease progression, and adaptation to tissue damage. The heart is a highly demanding organ, and adult cardiomyocytes require large amount of energy to fulfill adequate contractility. However, functioning under oxidative mitochondrial metabolism is accompanied with a concomitant elevation of harmful reactive oxygen species that indeed contributes to the progression of several cardiovascular pathologies and hampers the regenerative capacity of the mammalian heart. Cardiac metabolism is dynamic along embryonic development and substantially changes as cardiomyocytes mature and differentiate within the first days after birth. During early stages of cardiogenesis, anaerobic glycolysis is the main energetic program, while a progressive switch toward oxidative phosphorylation is a hallmark of myocardium differentiation. In response to cardiac injury, different signaling pathways participate in a metabolic rewiring to reactivate embryonic bioenergetic programs or the utilization of alternative substrates, reflecting the flexibility of heart metabolism and its central role in organ adaptation to external factors. Despite the well-established metabolic pattern of fetal, neonatal, and adult cardiomyocytes, our knowledge about the bioenergetics of other cardiac populations like endothelial cells, cardiac fibroblasts, or immune cells is limited. Considering the close intercellular communication and the influence of nonautonomous cues during heart development and after cardiac damage, it will be fundamental to better understand the metabolic programs in different cardiac cells in order to develop novel interventional opportunities based on metabolic rewiring to prevent heart failure and improve the limited regenerative capacity of the mammalian heart.
Article
Malate dehydrogenase (MDH) is an essential enzyme in the tricarboxylic acid cycle that functions in cellular respiration and redox homeostasis. Recent studies indicate that MDH facilitates metabolic plasticity in tumor cells, catalyzing the formation of an oncometabolite, contributing to altered epigenetics, and maintaining redox capacity to support the rewired energy metabolism and biosynthesis that enables cancer progression. This minireview summarizes current findings on the unique supporting roles played by MDH in human cancers and provides an update on targeting MDH in cancer chemotherapy.
Article
Full-text available
Hypoxia-inducible factor 1 (HIF-1) activates erythropoietin gene transcription in Hep3B cells subjected to hypoxia. HIF-1 activity is also induced by hypoxia in non-erythropoietin-producing cells, suggesting a more general regulatory role. We now report that RNAs encoding the glycolytic enzymes aldolase A (ALDA), phosphoglycerate kinase 1 (PGK1), and pyruvate kinase M were induced by exposure of Hep3B or HeLa cells to inducers of HIF-1 (1% O2, cobalt chloride, or desferrioxamine), whereas cycloheximide blocked induction of glycolytic RNAs and HIF-1 activity. Oligonucleotides from the ALDA, PGK1, enolase 1, lactate dehydrogenase A, and phosphofructokinase L (PFKL) genes, containing sequences similar to the HIF-1 binding site in the erythropoietin enhancer, specifically bound HIF-1 present in crude nuclear extracts or affinity-purified preparations. Sequences from the ALDA, PFKL, and PGK1 genes containing HIF-1 binding sites mediated hypoxia-inducible transcription in transient expression assays. These results support the role of HIF-1 as a mediator of adaptive responses to hypoxia that underlie cellular and systemic oxygen homeostasis.
Article
Full-text available
The Warburg effect, which originally described increased production of lactate in cancer, is associated with diverse cellular processes such as angiogenesis, hypoxia, polarization of macrophages and activation of T cells. This phenomenon is intimately linked to several diseases including neoplasia, sepsis and autoimmune diseases1,2. Lactate, which is converted from pyruvate in tumour cells, is widely known as an energy source and metabolic by-product. However, its non-metabolic functions in physiology and disease remain unknown. Here we show that lactate-derived lactylation of histone lysine residues serves as an epigenetic modification that directly stimulates gene transcription from chromatin. We identify 28 lactylation sites on core histones in human and mouse cells. Hypoxia and bacterial challenges induce the production of lactate by glycolysis, and this acts as a precursor that stimulates histone lactylation. Using M1 macrophages that have been exposed to bacteria as a model system, we show that histone lactylation has different temporal dynamics from acetylation. In the late phase of M1 macrophage polarization, increased histone lactylation induces homeostatic genes that are involved in wound healing, including Arg1. Collectively, our results suggest that an endogenous ‘lactate clock’ in bacterially challenged M1 macrophages turns on gene expression to promote homeostasis. Histone lactylation thus represents an opportunity to improve our understanding of the functions of lactate and its role in diverse pathophysiological conditions, including infection and cancer. The lactylation of lysine residues on histones in mammalian cells is stimulated by hypoxia and bacterial challenges, and increased histone lactylation induces genes involved in wound healing.
Article
Mutations in VHL, which encodes von Hippel–Lindau tumor suppressor (VHL), are associated with divergent diseases. We describe a patient with marked erythrocytosis and prominent mitochondrial alterations associated with a severe germline VHL deficiency due to homozygosity for a novel synonymous mutation (c.222C→A, p.V74V). The condition is characterized by early systemic onset and differs from Chuvash polycythemia (c.598C→T) in that it is associated with a strongly reduced growth rate, persistent hypoglycemia, and limited exercise capacity. We report changes in gene expression that reprogram carbohydrate and lipid metabolism, impair muscle mitochondrial respiratory function, and uncouple oxygen consumption from ATP production. Moreover, we identified unusual intermitochondrial connecting ducts. Our findings add unexpected information on the importance of the VHL–hypoxia-inducible factor (HIF) axis to human phenotypes.
Article
It is universally acknowledged that a large number of immune cells, as well as inflammatory factors, regulatory factors and metabolites, accumulate in the tumor microenvironment to jointly promote tumor escape, development and metastasis. Hypoxia is one of the characteristics in tumor microenvironment and is a common phenomenon in all solid tumors. In tumor hypoxia response, there is a key regulator called HIF-1a, which is a key transcriptional regulatory protein that regulates many critical genes. In this paper, the effects of hypoxia on glucose metabolism of tumor cells, myeloid-derived suppressor cells and T cells in tumor microenvironment were reviewed, and the interaction among the three was also described.
Article
Recent years have witnessed an emergence of interest in understanding metabolic changes associated with immune responses, termed immunometabolism. As oxygen is central to all aerobic metabolism, hypoxia is now recognized to contribute fundamentally to inflammatory and immune responses. Studies from a number of groups have implicated a prominent role for oxygen metabolism and hypoxia in innate immunity of healthy tissue (physiologic hypoxia) and during active inflammation (inflammatory hypoxia). This inflammatory hypoxia emanates from a combination of recruited inflammatory cells (e.g., neutrophils, eosinophils, and monocytes), high rates of oxidative metabolism, and the activation of multiple oxygen-consuming enzymes during inflammation. These localized shifts toward hypoxia have identified a prominent role for the transcription factor hypoxia-inducible factor (HIF) in the regulation of innate immunity. Such studies have provided new and enlightening insight into our basic understanding of immune mechanisms, and extensions of these findings have identified potential therapeutic targets. In this review, we summarize recent literature around the topic of innate immunity and mucosal hypoxia with a focus on transcriptional responses mediated by HIF. Expected final online publication date for the Annual Review of Immunology, Volume 38 is April 26, 2020. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
Article
Hypoxia-inducible factor (HIF) mediates protection via hypoxic preconditioning in both, in vitro and in vivo ischemia models. However, the underlying mechanism remains largely unknown. Prolyl hydroxylase domain proteins serve as the main HIF regulator via hydroxylation of HIFα leading to its degradation. At present, prolyl hydroxylase inhibitors including enarodustat are under clinical trials for the treatment of renal anemia. In an in vitro model of ischemia produced by oxygen-glucose deprivation of renal proximal tubule cells in culture, enarodustat treatment and siRNA knockdown of prolyl hydroxylase 2, but not of prolyl hydroxylase 1 or prolyl hydroxylase 3, significantly increased the cell viability and reduced the levels of reactive oxygen species. These effects were offset by the simultaneous knockdown of HIF1α. In another in vitro ischemia model induced by the blockade of oxidative phosphorylation with rotenone/antimycin A, enarodustat-enhanced glycogen storage prolonged glycolysis and delayed ATP depletion. Although autophagy is another possible mechanism of prolyl hydroxylase inhibition-induced cytoprotection, gene knockout of a key autophagy associated protein, Atg5, did not affect the protection. Enarodustat increased the expression of several enzymes involved in glycogen synthesis, including phosphoglucomutase 1, glycogen synthase 1, and 1,4-α glucan branching enzyme. Increased glycogen served as substrate for ATP and NADP production and augmented reduction of glutathione. Inhibition of glycogen synthase 1 and glutathione reductase nullified enarodustat's protective effect. Enarodustat also protected the kidneys in a rat ischemia reperfusion injury model and the protection was partially abrogated by inhibiting glycogenolysis. Thus, prolyl hydroxylase inhibition protects the kidney from ischemia via upregulation of glycogen synthesis.
Article
Erythropoietin (EPO) gene transcription is activated in kidney cells in vivo and in Hep3B cells exposed to hypoxia or cobalt chloride. Hypoxia- inducible factor 1 (HIF-1) is a nuclear factor that binds to the hypoxia-inducible enhancer of the EPO gene at a site that is required for transcriptional activation. HIF-1 DNA-binding activity is induced by hypoxia or cobalt chloride treatment of Hep3B cells. We report that treatment of Hep3B cells with desferrioxamine (DFX) induced HIF-1 activity and EPO RNA expression with kinetics similar to the induction of HIF-1 by hypoxia or cobalt chloride. Induction by each of these stimuli was inhibited by cycloheximide, indicating a requirement for de novo protein synthesis. DFX appears to induce HIF-1 by chelating iron as induction was inhibited by coadministration of ferrous ammonium sulfate. DFX administration to mice transiently increased EPO RNA levels in the kidney. As previously shown for hypoxia and cobalt treatment, DFX also induced HIF-1 activity in non-EPO-producing cells, suggesting the existence of a common hypoxia signal-transduction pathway leading to HIF-1 induction in different cell types.
Article
Background: Roxadustat is an oral hypoxia-inducible factor prolyl hydroxylase inhibitor that stimulates erythropoiesis and regulates iron metabolism. Additional data are needed regarding the effectiveness and safety of roxadustat as compared with standard therapy (epoetin alfa) for the treatment of anemia in patients undergoing dialysis. Methods: In a trial conducted in China, we randomly assigned (in a 2:1 ratio) patients who had been undergoing dialysis and erythropoiesis-stimulating agent therapy with epoetin alfa for at least 6 weeks to receive roxadustat or epoetin alfa three times per week for 26 weeks. Parenteral iron was withheld except as rescue therapy. The primary end point was the mean change in hemoglobin level from baseline to the average level during weeks 23 through 27. Noninferiority of roxadustat would be established if the lower boundary of the two-sided 95% confidence interval for the difference between the values in the roxadustat group and epoetin alfa group was greater than or equal to -1.0 g per deciliter. Patients in each group had doses adjusted to reach a hemoglobin level of 10.0 to 12.0 g per deciliter. Safety was assessed by analysis of adverse events and clinical laboratory values. Results: A total of 305 patients underwent randomization (204 in the roxadustat group and 101 in the epoetin alfa group), and 256 patients (162 and 94, respectively) completed the 26-week treatment period. The mean baseline hemoglobin level was 10.4 g per deciliter. Roxadustat led to a numerically greater mean (±SD) change in hemoglobin level from baseline to weeks 23 through 27 (0.7±1.1 g per deciliter) than epoetin alfa (0.5±1.0 g per deciliter) and was statistically noninferior (difference, 0.2±1.2 g per deciliter; 95% confidence interval [CI], -0.02 to 0.5). As compared with epoetin alfa, roxadustat increased the transferrin level (difference, 0.43 g per liter; 95% CI, 0.32 to 0.53), maintained the serum iron level (difference, 25 μg per deciliter; 95% CI, 17 to 33), and attenuated decreases in the transferrin saturation (difference, 4.2 percentage points; 95% CI, 1.5 to 6.9). At week 27, the decrease in total cholesterol was greater with roxadustat than with epoetin alfa (difference, -22 mg per deciliter; 95% CI, -29 to -16), as was the decrease in low-density lipoprotein cholesterol (difference, -18 mg per deciliter; 95% CI, -23 to -13). Roxadustat was associated with a mean reduction in hepcidin of 30.2 ng per milliliter (95% CI, -64.8 to -13.6), as compared with 2.3 ng per milliliter (95% CI, -51.6 to 6.2) in the epoetin alfa group. Hyperkalemia and upper respiratory infection occurred at a higher frequency in the roxadustat group, and hypertension occurred at a higher frequency in the epoetin alfa group. Conclusions: Oral roxadustat was noninferior to parenteral epoetin alfa as therapy for anemia in Chinese patients undergoing dialysis. (Funded by FibroGen and FibroGen [China] Medical Technology Development; ClinicalTrials.gov number, NCT02652806.).