ISSN 0006-2979, Biochemistry (Moscow), 2024, Vol. 89, No. 5, pp. 817-838 © Pleiades Publishing, Ltd., 2024.
817
REVIEW
Metabolic Adaptations and Functional Activity
of Macrophages in Homeostasis and Inflammation
Taisiya R. Yurakova
1
, Ekaterina A. Gorshkova
1
, Maxim A. Nosenko
2
,
and Marina S. Drutskaya
1,3,a
*
1
Engelhardt Institute of Molecular Biology, Russian Academy of Sciences, 119991 Moscow, Russia
2
Trinity Biomedical Sciences Institute, Trinity College Dublin, D02F306 Dublin, Ireland
3
Division of Immunobiology and Biomedicine, Center of Genetics and Life Sciences,
Sirius University of Science and Technology, 354340 Federal Territory Sirius, Russia
a
e-mail: marinadru@gmail.com
Received November 27, 2023
Revised February 6, 2024
Accepted February 8, 2024
AbstractIn recent years, the role of cellular metabolism in immunity has come into the focus of many studies.
These processes form a basis for the maintenance of tissue integrity and homeostasis, as well as represent an
integral part of the immune response, in particular, inflammation. Metabolic adaptations not only ensure energy
supply for immune response, but also affect the functions of immune cells by controlling transcriptional and
post-transcriptional programs. Studying the immune cell metabolism facilitates the search for new treatment
approaches, especially for metabolic disorders. Macrophages, innate immune cells, are characterized by a high
functional plasticity and play a key role in homeostasis and inflammation. Depending on the phenotype and
origin, they can either perform various regulatory functions or promote inflammation state, thus exacerbating
the pathological condition. Furthermore, their adaptations to the tissue-specific microenvironment influence the
intensity and type of immune response. The review examines the effect of metabolic reprogramming in macro-
phages on the functional activity of these cells and their polarization. The role of immunometabolic adaptations
of myeloid cells in tissue homeostasis and in various pathological processes in the context of inflammatory and
metabolic diseases is specifically discussed. Finally, modulation of the macrophage metabolism-related mecha-
nisms reviewed as a potential therapeutic approach.
DOI: 10.1134/S0006297924050043
Keywords: proinflammatory cytokines, macrophage polarization, immunometabolism
Abbreviations: 2-DG,2-deoxyglucose; 2-HG,2-hydroxyglutarate; ARG1,arginase 1; ACLY,ATP citrate lyase; AMPK,AMP-ac-
tivated protein kinase; CPT,carnitine palmitoyltransferase; DMF,dimethyl fumarate; ETC,electron transport chain; FA,fat-
ty acid; HIF-1α, hypoxia-inducible factor 1-alpha; IDH, isocitrate dehydrogenase; iNOS, inducible nitric oxide synthase;
LDHA,lactate dehydrogenase; LPS,lipopolysaccharide; mTOR,mammalian target of rapamycin; OXPHOS,oxidative phos-
phorylation; PGE
2
, prostaglandin E2; PKM2, pyruvate kinase M2; PPP, pentose phosphate pathway; ROS, reactive oxygen
species; SDH,succinate dehydrogenase.
* To whom correspondence should be addressed.
INTRODUCTION
The majority of immunological studies in cell and
animal models have been focused on the functions
of immune cells, in particular, antigen recognition,
as well as elimination and production of active com-
pounds and cytokines. Because of this, immune re-
sponse has become increasingly viewed as a systemic
body response leading to metabolic rewiring and func-
tional changes in tissues and organs. The association
between metabolism and immune system was first dis-
covered in the last century by physiologist Otto War-
burg, who noticed that upregulation of glycolysis was
characteristic not only of tumor cells, but also of the
activated leukocytes [1, 2]. Now, it is commonly recog-
nized that metabolic adaptations are an integral part
YURAKOVA et al.818
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
of the immune response [3-8]. The discovery of regula-
tory pathways involved in the metabolic control of im-
mune processes during infection, cancer, and tissue re-
generation at the cellular, tissue, and organ levels have
led to the emergence of a new interdisciplinary field of
research, immunometabolism [3, 8]. Immunometabolic
disorders are responsible for many diseases typical
for the modern human population, including obesity,
diabetes, sepsis, autoimmune and autoinflammatory
diseases [3]. Hence, understanding the mechanisms
by which physiological microenvironment of immune
cells in tissues and organs regulates their functions is
very important. Among immune cells, macrophages
are of a special interest because they can be found in
almost all animal tissues and represent an integral
component in the maintenance of tissue homeostasis
[9]. The review describes the relationship between the
functional and metabolic features of macrophages in
homeostasis and immunometabolic disorders, as well
as discusses possible therapeutic approaches based on
the modulation of macrophage metabolism.
FUNCTIONAL AND PHENOTYPIC
DIVERSITY OF MACROPHAGES
Macrophages had been initially considered as a
part of the mononuclear phagocytic system responsi-
ble for the removal of pathogens and apoptotic cells,
but the current understanding of the functions and
Fig. 1. Macrophage differentiation and polarization. a)Tissue-resident macrophages differentiate from three types of hemato-
poietic stem cell precursors found in the yolk sac, embryonic liver, and bone marrow. b)Origin of tissue-resident macrophages
in different organs. Tissue-resident macrophages are formed in closed niches during embryogenesis; in adults, peripheral
blood monocytes do not migrate into these niches. In open niches, monocytes are recruited at different rates (slow or fast)
and replace tissue-resident macrophage pool formed during embryogenesis. c)Phenotype of tissue-resident macrophages is a
result of their differentiation and polarization. Macrophages polarize into pro-inflammatory(M1) or anti- inflammatory(M2)
phenotypes under the influence of external stimuli. Created with the BioRender.com based on previously published
data[20, 24-26].
METABOLIC CHANGES IN MACROPHAGE POLARIZATION 819
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
the origin of macrophages have been significantly ex-
panded. Macrophages were shown to play an import-
ant role in homeostasis, tissue formation during em-
bryogenesis, and wound healing, as well as in various
tissue-specific functions [10]. Thus, tissue-specific mac-
rophages determine normal development and func-
tioning of the brain, bones, ovaries, and adipose tissue.
Macrophages are involved in the control of systemic
metabolism and temperature adaptation [11, 12]. Fi-
nally, macrophages play an important role in the de-
velopment of metabolic diseases, such as atherosclero-
sis [13], osteoporosis, obesity, and type  2 diabetes [14,
15]. Pathogenic contribution of macrophages to fibro-
sis [16] and tumor development has been recognized
as well [17].
Tissue-resident macrophages can vary dramati-
cally in both phenotype and functions. A population of
macrophages in an organ or tissue is characterized by
the functional heterogeneity that depends on the cell
location, tissue condition, cell origin, and many other
factors. Tissue-resident macrophages can differentiate
from the precursors of hematopoietic stem cells found
in the yolk sac, embryonic liver, and bone marrow [18,
19]. The proportion of macrophages originating from
different migration waves varies depending on the or-
gan [20] (Fig.1, a and b).
The phenotype of a macrophage largely depends
on the surrounding signals [21]. The two main mac-
rophage phenotypes are pro-inflammatory (M1) and
anti-inflammatory (M2). M1 macrophages induce and
maintain inflammation via release of pro-inflammato-
ry cytokines, activation of endothelial cells, and attrac-
tion of other immune cells to the site of inflammation.
M2 macrophages counteract inflammation by phago-
cytosis of apoptotic cells, triggering of collagen depo-
sition, coordination of tissue repair processes, and re-
lease of anti-inflammatory cytokines [22]. It should be
noted that the M1 and M2 phenotypes could be clear-
ly distinguished mainly in in  vitro experiments using
certain combinations of stimuli. Under physiological
conditions, the phenotype of macrophages is greatly
influenced by various signaling molecules and tissue
microenvironment, so pure M1 or M2 phenotypes are
rarely observed in vivo [23]. Depending on the con-
text, macrophages acquire an intermediate or unique
phenotype and are typically represented by a mixed
heterogeneous population of M1 and M2 cells [23].
Therefore, macrophage polarization in  vivo should be
viewed as a complex tissue-specific process developing
in time (Fig.1) [24-26].
There are two stages of “decision-making” at the
molecular level in polarization [27]. The first one is al-
teration in the functional state of M1/M2 macrophages
that involves activation of signaling pathways associat-
ed with either innate immunity, such as signaling via
pattern recognition receptors (PRR), or stimuli from
T cells [28, 29]. Signals leading to macrophage polar-
ization in  vivo are often difficult to identify [21]. Bac-
terial endotoxin, lipopolysaccharide (LPS), or a com-
bination of LPS and interferon gamma (IFNγ, main
Th1-associated cytokine) are generally used to polarize
M1 macrophages in  vitro. M2 macrophages are polar-
ized using Th2-associated cytokine interleukin-4 (IL-4)
or a combination of IL-4 and IL-13 [22]. IRF, STAT,
and NF-κB transcription factors play a key role in the
control of M1/M2 polarization, M1 regulators (NF-κB-
p50-p65, STAT1, IRF5, SOCS3, and HIF-1α) being antag-
onists of M2 regulators (NF-κB-p50-p50, STAT6, IRF4,
SOCS1, and HIF-2). The second stage of “decision-mak-
ing” involves interaction of transcription factors [30]
and is closely related to the activity of the two main
regulators of cellular metabolism– AMP-activated pro-
tein kinase (AMPK) and mammalian target of rapamy-
cin (mTOR). The most important outcome of the second
stage is altered expression of metabolism-associated
genes, which affects a variety of metabolic processes,
such as glycolysis, Krebs cycle, oxidative phosphoryla-
tion (OXPHOS), pentose phosphate pathway (PPP), for-
mation of NADPH and reactive oxygen species (ROS),
biosynthesis of nucleotides and amino acids, as well as
nitrogen metabolism.
METABOLIC ADAPTATIONS
OF MACROPHAGES DURING POLARIZATION
To meet the demands for ATP and NADPH, M1
macrophages use aerobic glycolysis and PPP, leading to
the breaks in the Krebs cycle at two points and reduc-
tion in the cellular respiration, in particular, OXPHOS,
as well changes in the metabolism of fatty acids (FAs)
(upregulation of beta oxidation and suppression of
synthesis of FAs). M2 macrophages do not depend on
glycolysis and mainly use OXPHOS as a source of en-
ergy, so that there are no breaks in the Krebs cycle.
During M2 polarization, glucose is utilized mainly by
the PPP and used for the synthesis of uridine diphos-
phate N-acetylglucosamine (UDP-GlcNAc) [31]. The
changes in the metabolism of FAs in M2 macrophages
are opposite to those observed in M1 cells, i.e., the
synthesis of FAs is decreased and their beta oxida-
tion is upregulated [22, 30]. Based on the changes in
the characteristic properties of M1 or M2 phenotypes
(Table 1) acquired as a result of pharmaceutical or ge-
netic suppression of metabolic enzymes, it is possible
to assess the contribution of particular metabolic path-
ways to the macrophage polarization. Below, we dis-
cuss in detail the metabolic changes that occur during
macrophage polarization.
Glucose metabolism. Glucose is important in both
M1 and M2 polarization [32, 33], but its distribution be-
tween metabolic pathways is different for these two
YURAKOVA et al.820
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
Table 1. Properties of M1 and M2 macrophages
Polarization property M1 (LPS + IFNγ) M2 (IL-4 + IL-13)
Stimuli
LPS + IFNγ;
bacterial products; LPS and other ligands
of Toll-like receptors (TLRs);
cytokines of Th1 lymphocytes (IFN-γ, TNF)
IL-4 + IL-13;
basophiles;
mast cells; Th2 lymphocytes
Cell-derived
mediators
TNF, IL-1β, IL-6, IL-12, IL-23;
ROS, PGE
2
IL-10, TGFβ, IGF-1, VEGF-A,
EGF, PDGF
Cell surface markers CD80, CD86, CIITA, MHCII
mannose receptor (CD206),
CD36, IL1Rα, CD163 CD36,
RELMα (FIZZ1in humans), MMP
Signaling pathways
NF-κB (p50-p65), STAT1, IRF5, HIF-1α, SOCS3, AP1;
inflammasome activation
NF-κB (p50-p50), STAT6, IRF4,
HIF2, SOCS1, GATA3, PPARγ,
(in mice also YM1)
Functions
elimination of bacteria; tumor resistance;
Th1 response; pathogen elimination and antigen
presentation to T lymphocytes
anti-inflammatory response;
tissue remodeling;
wound healing; angiogenesis
types of macrophage activation. M1 macrophages sig-
nificantly increase glucose entry via upregulation of
the GLUT1 transporter [34], while glucose utilization
by M2 macrophages remains unchanged [35]. Glycol-
ysis has been considered as the main metabolic path-
way conjugated with ATP production in immunome-
tabolism. However, recent data indicate an important
role of PPP and glycosylation in the metabolism of acti-
vated immune cells. The studies of glucose metabolism
typically use glucose analogue 2-deoxyglucose (2-DG).
However, the effect of this compound in immune cells
is nonspecific, as it blocks all glucose catabolism path-
ways, not only glycolysis [36]. As a result, 2-DG dis-
rupts polarization of both M1 and M2 macrophages,
which has led to erroneous conclusion on a significant
contribution of glycolysis to both types of macrophage
polarization.
A more detailed study has demonstrated a criti-
cal importance of glycolysis for M1 macrophages [31].
Activation of M1 macrophages is associated with the
increase in the expression of several variants of the
key glycolysis enzyme– phosphofructokinase (PFK1-M
and PFK2) [35, 37]. The reaction catalyzed by this en-
zyme is irreversible and represents a commitment step
for glycolysis. Upregulation of the PFK-M expression
is associated with the suppression of production of mi-
croRNA-21 upon macrophage stimulation by IFNγ [37].
PFK2 suppression leads to a decrease in the expression
of inducible NO synthase (iNOS) and mitogen-inducible
form of cyclooxygenase (COX2) by M1 macrophages[35].
Only M1 (but not M2) macrophages demonstrate in-
creased expression of the phosphofructokinase regu-
lator (PFKFB3); inhibition of this protein prevents up-
regulation of glycolysis and activation of M1 macro-
phages [38]. Inhibition of glyceraldehyde 3-phosphate
dehydrogenase (GAPDH), one of the key glycolytic en-
zymes, suppresses IL-1β production similar to 2-DG
[32, 39]. Finally, suppression of lactate dehydrogenase
(LDHA) and pyruvate dehydrogenase  1 (PDK1) genes
also caused a decrease in the IL-1β production by the
macrophages [40]. The critical importance of glycolysis
for M1 macrophages remains poorly understood. Pre-
sumably, it might be due to the OXPHOS suppression
during M1 polarization, which requires a compen-
satory upregulation of glycolysis to ensure ATP pro-
duction. OXPHOS remains active in M2 macrophages,
making glycolysis nonessential for maintaining ener-
gy balance [36].
The most important transcription factor regu-
lating glycolysis is hypoxia-inducible factor 1-alpha
(HIF-1α) [41]. HIF-1α controls expression of genes
coding for glycolytic enzymes, glucose transporter
GLUT1, inflammatory mediators, LDHA, and PDK.
LDHA catalyzes pyruvate conversion to lactate; PDK
inactivates pyruvate dehydrogenase (PDH) and reduc-
es pyruvate entry into the Krebs cycle. Pyruvate con-
version to lactate is important for M1 macrophages
as it provides restoration of the NAD
+
pool in order
to maintain glycolytic reactions. HIF-1α expression is
regulated by signals from the innate immune recep-
tors and proinflammatory cytokines via NF-κB [42-44]
and growth factors. For example, granulocyte-macro-
phage colony-stimulating factor (GM-CSF) activates
HIF-1α through the PI3K/AKT/mTOR pathway [45-49].
Although HIF-1α is inactivated in the presence of
oxygen, it can be stabilized by succinate produced
METABOLIC CHANGES IN MACROPHAGE POLARIZATION 821
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
in the mitochondria due to breaks in the Krebs cycle
[22, 32].
Expression of different isoforms of pyruvate kinase
M2 (PKM2) accounts for different regulation of glycol-
ysis in M1 and M2 macrophages [35, 50, 51]. PKM2 has
two conformations: the tetramer is enzymatically ac-
tive and participates in glycolysis, while the dimer,
which lacks the enzymatic activity, localizes to the
nucleus and functions as a transcription factor that
regulates expression of HIF-1α-dependent genes, in-
cluding Il1b. Stabilization of the PKM2 tetramer with
the small TEPP-46 molecule reduces activation of M1
macrophages and makes them phenotypically similar
to M2 cells [50].
Pentose phosphate pathway takes place in the
cytoplasm. The oxidative phase of PPP plays an im-
portant role in the metabolism of M1 macrophages
[32, 52]. During this phase, glucose-6-phosphate is con-
verted to ribulose-5-phosphate, which is accompanied
by the NADPH formation from H and NADP
+
. NADPH
is involved in the synthesis of glutathione, an antiox-
idant that protects cells from oxidative stress. On the
other hand, NADPH is a substrate for many ROS-pro-
ducing enzymes, for example, NADPH oxidases (NOXs)
are essential for the pathogen destruction by M1 mac-
rophages [22, 53]. NADPH is also required for the syn-
thesis of FAs and prostaglandins.
PPP is less important in M2 macrophages and is
involved in cell metabolism to a much lesser extent.
On the one hand, UDP-GlcNAc synthesized from ri-
bose-5-phosphate produced during the PPP oxidation
stage, is used in N-glycosylation of the mannose re-
ceptor abundantly present on the surface of M2 mac-
rophages [54]. UDP-GlcNAc can be synthesized from
glucose-1-phosphate independently of PPP during M2
polarization [31]. The importance of N-glycosylation for
the M2 phenotype development has been confirmed in
the experiments using the inhibitor of N-glycosylation
tunicamycin [31]. On the other hand, it was recently
demonstrated that activation of PPP suppresses effero-
cytosis (phagocytosis of tolerogenic apoptotic cell) that
is characteristic for M2 macrophages [55].
Oxidative phosphorylation. OXPHOS is the main
source of ATP in the cells under normoxic conditions,
including inactive macrophages (M0). Polarization in-
creases energy requirements of the cells regardless of
polarization type, but the activity of OXPHOS in M1
and M2 macrophages differs. M1 macrophages are
characterized by OXPHOS suppression, while M2 po-
larization, on the contrary, increases the activity of the
electron transport chain (ETC) [56]. Nitric oxide (NO)
suppresses OXPHOS in M1 macrophages by blocking
the ETC and prevents their repolarization into M2
macrophages [57]. Besides, recognition of live bacteria
by the macrophages results in the reduction of active
complexes  I and III of the ETC due to the ROS produc-
tion [58]. Activation of STAT6 during M2 polarization
leads to the upregulated expression of genes involved
in the FA oxidation and OXPHOS [59]. OXPHOS sup-
pression, in turn, interferes with the expression of M2
markers, such as arginase 1 (ARG1) and mannose re-
ceptor.
The functioning of the ETC is inevitably accompa-
nied by the generation of ROS. Thus, ROS generation
in the mitochondria is a result of electron leakage
through the complexes I, II, and III in the case of dis-
ruption of the ETC or mitochondrial membrane po-
tential. ROS production in the mitochondria depends
on the metabolic state of these organelles [60,  61].
Itshould be noted that the decrease in the respiration
during M1 polarization is associated with the increase
in the production of mitochondrial ROS, which stim-
ulates secretion of proinflammatory cytokines [62].
Thus, activation of macrophages with LPS leads to the
ROS production, apparently, via reverse electron flow
from the succinate dehydrogenase (SDH) complex in
the ETC. Moreover, activation of the toll-like receptors
(TLRs) 1, 2, and 4 stimulates mitochondrial ROS gen-
eration due to the TRAF6 translocation into the mito-
chondria [63]. Macrophages produce ROS in response
to a pathogen or activation by proinflammatory cyto-
kines; ROS generation is associated with the activation
of NOXs and is independent of cyanides that block the
complex IV of the respiratory chain. ROS production
by cells of the innate immune system, including mac-
rophages, is necessary to fight pathogens, however, a
prolonged inflammatory process increases ROS pro-
duction, resulting in oxidative stress, which leads to
the dysfunction of the vascular endothelium and other
disorders [64].
The Krebs cycle, also known as the tricarboxyl-
ic acid cycle, is a central metabolic pathway that pro-
vides cells with ATP. During macrophage activation,
many processes, such as proliferation, synthesis of
cytokines and other proinflammatory mediators, mi-
gration, and phagocytosis, require ATP. Moreover, M2
macrophages need ATP to maintain high levels of gly-
cosylation of lectin and mannose receptors. Intermedi-
ate metabolites of the Krebs cycle are also important
for the macrophage activation and polarization. Their
accumulation in the cell can result from the breaks in
the Krebs cycle due to the inhibition of isocitrate de-
hydrogenase (IDH) or SDH. The impact of Krebs cycle
intermediates (citrate, itaconate, succinate, fumarate,
and alpha-ketoglutarate) on the macrophage polariza-
tion is described in detail in [65, 66].
Citrate is synthesized in the Krebs cycle via con-
densation of oxaloacetate and acetyl-CoA formed from
pyruvate or as a product of FA catabolism. M1 macro-
phages accumulate citrate in the cytosol due to the IDH
suppression at the transcriptional level and overex-
pression of the mitochondrial citrate carrier (CIC) [67].
YURAKOVA et al.822
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
Despite its ability to suppress glycolysis by affecting
phosphofructokinases directly and pyruvate kinase in-
directly, as well as to stimulate gluconeogenesis [68],
citrate seems to be essential in the regulation of FA
synthesis in M1 macrophages. Citrate is a substrate of
ATP-citrate lyase (ACLY), an important source of ace-
tyl-CoA utilized in the FA synthesis cycle. Inhibition of
ACLY leads to a decrease in the production of prosta-
glandin E2 (PGE
2
), NO, and ROS, that play an important
role in the inflammation and oxidative stress [69].
Itaconate is one of the most abundant metabolites
in M1 polarization [31]. It is produced from cis-aconi-
tate by aconitate decarboxylase 1 (ACOD1), whose ex-
pression is upregulated in M1 macrophages. On one
hand, itaconate is a competitive SDH inhibitor that
interrupts the Krebs cycle and suppresses OXPHOS in
proinflammatory macrophages [70,  71]. On the other
hand, itaconate binds covalently to cysteine residues
in proteins [72], thus suppressing their biological func-
tions. For example, itaconate alkylates cysteines in
fructose-1,6-bisphosphate aldolase (ALDOA), GAPDH,
and LDHA, i.e., the key glycolytic enzymes, thus block-
ing their activity and leading to the reduction of glu-
cose consumption and lactate production by the cells
[72]. By alkylating KEAP1, itaconate activates NRF2
transcription factor which stimulates expression of
genes involved in cell protection from the oxidative
stress [73]. Finally, itaconate inhibits the production of
NO and proinflammatory cytokines (TNF, IL-6, IL-1β,
IL-12, and IL-18) by M1 macrophages [70, 74] partially
due to the suppression of IκBζ, a regulator of transcrip-
tional activity of NF-κB. Itaconate promotes Activating
transcription factor 3 (ATF3) production in cultured
macrophages, which, in turn, leads to IκBζ suppression
at the post-transcriptional level [75]. In the case of
IL-1β, itaconate not only suppresses expression of the
Il1b gene, but also interferes with the cytokine pro-
duction at the post-translational level by reducing
caspase-1 activity [76]. Moreover, some itaconate deriv-
atives alkylate NLRP3 directly, preventing the assem-
bly of the inflammasome [77]. Itaconate presumably
participates in the feedback loop involved in acquisi-
tion of the M1 phenotype by slowing down and limit-
ing the activation of M1-associated genetic programs,
which may play an important role in the control of
immune response. M2 macrophages do not produce
itaconate, but can uptake it from the medium [78].
Itwas recently shown that exogenous itaconate inhib-
its differentiation of M2 macrophages by suppressing
the IL-4 signaling pathway mediated by JAK1/STAT6
[79]. There are also data that macrophage-derived
itaconate affects tumor cells [80] and T cells [81]. Al-
though itaconate is a specific metabolite of M1 po-
larization, it exerts a wide range of effects on sur-
rounding cells, contributing to the formation of the
M1-favorable microenvironment.
Succinate is generated in the Krebs cycle from
succinyl-CoA by succinyl-CoA-ligase and is converted
to fumarate by SDH. SDH is a part of the Krebs cycle
and a component of the complex II of the ETC. Itacon-
ate accumulating during M1 polarization reduces SDH
activity, leading to the accumulation of succinate in
the mitochondria [70]. Succinate can be transport-
ed by the dicarboxylate transporter (DIC, SLC25A10)
from the mitochondria to the cell cytoplasm, where it
has multiple functions. First, succinate inhibits prolyl
hydroxylase (PHD) and thereby blocks HIF-1α degra-
dation in the presence of oxygen. This results in the
HIF-1α accumulation and activation of aerobic gly-
colysis, also known as the Warburg effect, as well as
production of proinflammatory cytokines (e.g., IL-1β)
[32]. Second, it causes post-translational changes in
proteins via succinylation of lysine residues. For ex-
ample, succinylation of pyruvate kinase in M1 mac-
rophages leads to the HIF-1α-mediated increase in
the IL-1β production [82]. Finally, succinate serves as
a cell–cell mediator of inflammation; it is secreted by
inflammatory macrophages and interacts with G pro-
tein-coupled SUCNR1/GPR91 receptors, thus promot-
ing IL-1β production [83, 84]. It should be noted that
expression of succinate receptors on the macrophage
surface increases in response to inflammatory signals
(LPS, TNF, and IFNγ) [67].
Fumarate is a product of SDH activity. However,
fumarate levels are significantly elevated in M1 mac-
rophages despite the inhibition of SDH, which is due
to the increased expression of argininosuccinate syn-
thase (ASS1), an enzyme of the urea cycle, in proin-
flammatory macrophages. ASS1 catalyzes the synthesis
of argininosuccinate, which is then decomposed into
arginine and fumarate. Fumarate, a by-product of the
urea cycle, has a significant effect on the total fuma-
rate level which increases with inflammation. More-
over, according to some data, fumarate and arginino-
succinate are among the most produced metabolites in
activated macrophages compared to inactive cells [85].
Fumarate inhibits IL-10 production, thereby increas-
ing TNF biosynthesis. The effect of fumarate on IL-10
production may be associated with suppression of the
ERK signaling pathway and PI3K signaling [85].
Alpha-ketoglutarate (α-KG) is formed by IDH via
isocitrate conversion in the Krebs cycle. It is involved
in many immune processes [86], e.g., epigenetic repro-
gramming. First of all, α-KG is the main cofactor in
several families of histone demethylases, such as JMJD
(Jumonji C-domain-containing histone demethylases)
and TET (ten-eleven translation) enzymes [87, 88].
α-KG promotes transcription of IL-4-dependent genes
due to its action on JMJD3. On the other hand, exoso-
mal α-KG is also involved in M2 polarization by acti-
vating TET-mediated DNA demethylation, which leads
to the suppression of the STAT3/NF-κB pathways [89].
METABOLIC CHANGES IN MACROPHAGE POLARIZATION 823
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
Finally, α-KG stimulates FA oxidation, which is a com-
mon feature of M2 macrophages. Interestingly, a high
α-KG/succinate ratio contributes to M2 polarization,
while the low one, on the contrary, determines the
proinflammatory M1 phenotype. Since α-KG is a co-
factor of JMJD3, while succinate inhibits this enzyme,
their combined action allows to regulate H3K27 de-
methylation [90].
2-Hydroxyglutarate (2-HG), which exists in both
L- and D-isoforms, is also involved in the epigenetic
regulation. 2-HG can be synthesized by the nonspecific
activity of several enzymes, such as malate dehydro-
genase (MDH), LDHA, and mutant IDH1/2mut isoform
found in tumor cells [91, 92]. 2-HG, succinate, and fu-
marate inhibit α-KG at the epigenetic level [93,  94],
therefore, the ratio between these compounds, espe-
cially between 2-HG and α-KG, is important in several
immune processes. 2-HG accumulates in tissues under
hypoxic conditions or at low pH [95], as well as in M1
macrophages in response to LPS activation [96]. In  vitro
experiments have shown that L-2-HG inactivates HIF
prolyl hydroxylase, stabilizes HIF-1α, and thus pro-
motes IL-1β production and activation of glycolysis
[96]. In contrast, D-2-HG, also formed in M1 cells, con-
tributes to the suppression of inflammatory processes
at the late stage of LPS-induced response in  vitro and
is a regulator of local and systemic inflammatory re-
actions in  vivo [97].
Amino acid metabolism is important not only
in cell homeostasis and protein synthesis, but also in
many immune processes, including macrophage po-
larization [98]. The deficit of amino acids in the me-
dium impairs migration, proliferation, maturation,
and effector functions of immune cells. The effects of
arginine, glutamine, glycine, and serine on the macro-
phage functions have been studied in more detail.
Activated macrophages require arginine as a sub-
strate for two competing enzymes – ARG1 and iNOS.
Typically, ARG1 expression in M2 macrophages is in-
creased. ARG1 converts arginine into urea and orni-
thine; ornithine initiates the synthesis of polyamines
involved in tissue repair. Overall, ARG1 contributes to
the anti-inflammatory phenotype of macrophages and
thereby suppression of T cell proliferation and cyto-
kine production [94]. Ornithine is also essential for
the immune functions of macrophages in the context
of Mycobacterium tuberculosis infection [99, 100]. The
expression of iNOS in macrophages is upregulated by
proinflammatory stimuli (LPS, TNF, IFNγ). iNOS con-
verts arginine to nitric oxide and citrulline; NO spon-
taneously reacts with oxygen and ROS, resulting in
the generation nitrogen and oxygen species with the
antimicrobial and regulatory activities. ASS1 converts
citrulline into argininosuccinate, which is further de-
graded to arginine and thus maintains NO production.
In addition to the ROS generation, NO is involved in
the remodeling of mitochondrial ETC during M1 polar-
ization. Thus, the treatment of macrophages with LPS/
IFNγ leads to the induction of NO synthesis along with
the decrease in the activity of complexesI and II; the
short-term action of NO on the macrophages results in
the reversible inhibition of complex IV, due to the NO
competition with oxygen for the enzyme catalytic site
[101, 102]. Although disturbances in the functioning of
complex I contribute to the increase in the ROS pro-
duction in the mitochondria and expression of proin-
flammatory factors such as IL-1β and TNF [103], recent
studies have shown that this process is not directly re-
lated to the NO action [104, 105]. Apparently, at later
stages of activation, NO has the regulatory functions
due to the ability to inhibit mitochondrial complexes
and to reduce their number [106]. Moreover, the effect
NO on the ETC leads to changes in the mitochondrial
morphology and is one of the factors preventing re-
polarization of M1 macrophages to the respiration-de-
pendent M2 phenotype [57].
Another important compound for the macro-
phages is glutamine which is required for the syn-
thesis of amino acids and nucleotides, production
of NADPH and energy, and many other biosynthet-
ic processes [107, 108]. Depending on the metabolic
pathway, glutamine stimulates either M1 or M2 mac-
rophage polarization. On one hand, glutamine can en-
ter the Krebs cycle through α-KG, thereby stimulating
the synthesis of succinate in M1 macrophages [32],
which is also significant for the HIF-1α stabilization
and glycolysis maintenance [109]. At the same time,
upregulation of succinate biosynthesis is accompanied
by an increase in the expression of the SLC3A2 glu-
tamine transporter gene and activation of glutamine
uptake [110, 111]. Interestingly, some of succinate in
LPS-activated macrophages is produced by the gam-
ma-aminobutyric acid (GABA) shunt. In this pathway,
which bypasses the Krebs cycle, glutamine is used for
the sequential synthesis of glutamate, GABA, succinic
semialdehyde, and eventually succinate. Inhibition of
GABA transaminase, the key enzyme of this pathway,
significantly reduces the amount of succinate produced
from glutamine and, as a result, prevents HIF-1α sta-
bilization and IL-1β secretion in response to LPS [32].
On the other hand, the deficiency of glutamine in
the medium or inhibition of glutaminase by its se-
lective blocker bis-2-(5-phenylacetamido-1,3,4-thiadi-
azole-2-diyl) ethyl sulfide (BPTES) during macrophage
activation by LPS prevents the development of endo-
toxin tolerance [90], similar being demonstrated in a
mouse model of toxic shock [112]. Endotoxin tolerance
is an important mechanism of homeostasis mainte-
nance, as acquisition of the tolerance toward repeated
LPS stimulation by the macrophages helps to protect
the body against possible excessive immune system ac-
tivation [113]. Therefore, glutamine is involved in both
YURAKOVA et al.824
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
inflammatory response of M1 macrophages to LPS and
their subsequent negative regulation through the tol-
erance formation.
Glutamine metabolism can also contribute to the
M2 polarization, mainly by stimulating accumulation
of α-KG [90]. Like glucose, glutamine is necessary for
the synthesis of UDP-GlcNAc that is used by M2 mac-
rophages for glycosylation of mannose receptor and
RELMα [31]. M2 macrophages not only import gluta-
mine from the environment, but also synthesize it
from glutamate and ammonia using glutamine syn-
thetase (GS). GS is fundamental for the M2 phenotype.
It is almost absent in M1 macrophages and is highly
expressed in M2 macrophages, especially in response
to IL-10 [114]. Therefore, both the lack of glutamine in
the medium and the inhibition of glutaminolysis by
BPTES lead to disturbances in the alternative polariza-
tion of macrophages [31, 90].
Serine and glycine play an important role in the
regulation of redox balance, as both of them are in-
volved in the synthesis of glutathione. The mainte-
nance of the redox balance is important for the M1
polarization of macrophages, since ROS produced by
these cells cause oxidative stress [63, 115, 116]. This
triggers the activity of the NRF2 transcription factor
and results in the reduction of the LPS-induced NF-κB
activity, in particular, due to the regulation of gluta-
thione metabolism [117,  118]. Serine conversion to
glycine may be necessary for glutathione production
[119]. At the same time, endogenous synthesis of ser-
ine in LPS-activated macrophages is known to lead
to the production of S-adenosylmethionine, which is
involved in the epigenetic regulation of IL-1β expres-
sion[119].
Fatty acid synthesis and oxidation. FAs are es-
sential for the synthesis of cell membranes and vari-
ous bioactive compounds, e.g., eicosanoids. Moreover,
FAs are another important “energy currency” of the
body; their oxidation results in the synthesis of ATP.
Macrophage polarization is highly dependent on
FA metabolism. Thus, FA anabolism is typical for M1
macrophages: activation of TLRs, e.g., by IFNγ, caus-
es accumulation of triacylglycerides (in a form of fat
droplets), diacylglycerides, and cholesterol esters [120,
121], which correlates with the phenotype of foam
cells, i.e., macrophages associated with atherosclero-
sis, granulomas, and other inflammatory pathologies
[122]. Accumulation of fats is preceded by the activa-
tion of lipogenesis due to the activity of the transcrip-
tion factor SREBP-1 [123] and ACLY, which produces
acetyl-CoA from citrate for subsequent utilization by
fatty acid synthase (FAS) [124]. Macrophages from
mice deficient for these transcription factors, enzymes,
and some carrier proteins involved in FA anabolism,
such as uncoupling protein  2 (UCP2) [125] and ace-
tyl-CoA carboxylase (ACC) [126]), demonstrated a de-
creased ability for proinflammatory response, mainly
due to the inflammasome disruption. It was recently
shown that fat droplets in M1 macrophages serve as a
source of PGE
2
production. PGE
2
is responsible for the
activation of phagocytosis and production of IL-1β and
IL-6 in macrophages. The blockade of DGAT1 (carrier
involved in FA deposition) causes suppression of the
in  vitro and in  vivo inflammatory responses of macro-
phages [120].
As mentioned above, the Krebs cycle remains
undisturbed in alternatively activated macrophages,
which maintains higher OXPHOS levels; the main
source of acetyl-CoA in M2 macrophages is FA. Beta ox-
idation occurs in the mitochondria, and the entry of
FAs into these organelles is provided by carnitine pal-
mitoyltransferases 1 (CPT1, located in the outer mito-
chondrial membrane) and 2 (CPT2, located in the inner
mitochondria membrane). Recent experiments with
pharmacological and genetic blockade of CPT1/2 have
shown that the inhibition of beta oxidation does not
block the acquisition of the M2 phenotype by mouse
[127] and human [128] macrophages, contrary to the
previous views [129]. Both synthesis and beta oxida-
tion of FAs are important for the inflammasome func-
tioning. Thus, NLRP3 activation can be inhibited by the
activity of NOX4, which, in turn, regulates CPT1A [130].
The balance between the FA synthesis and oxidation
is the most important regulator of inflammasome ac-
tivation, and, therefore, of the inflammatory response
of M1 macrophages [131]. Although, the role of FA me-
tabolism in the M2 phenotype is currently undergoing
revision, modulation of FA metabolism is still consid-
ered as a promising therapy for reprogramming tu-
mor-associated M2-like macrophages [132, 133].
REGULATION OF METABOLIC
REPROGRAMMING OF MACROPHAGES
DURING POLARIZATION
Reprogramming of metabolic pathways during
macrophage activation largely depends on the key met-
abolic master regulators mTOR and AMPK [134-136].
mTOR forms two types of complexes: mTORC1 and
mTORC2 [137]. mTORC1 integrates information coming
from the extracellular (growth factors, cellular stress)
and intracellular stimuli (concentration of amino acids
in the lysosomes, sugar levels, and lipid content) and
transmits it further, activating HIF-1α, PPARγ, SREBP-1,
and MYC, which promotes cell division, activates bio-
synthesis of nucleic acids, proteins, and lipids, and
suppresses catabolic processes and autophagy [136,
137]. The balance of the mTORC1/mTORC2-mediated
signaling is critically important for the macrophage
differentiation and polarization and is controlled
by signaling cascades mediated by innate immunity
METABOLIC CHANGES IN MACROPHAGE POLARIZATION 825
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
Table 2. Correlation between metabolic changes and macrophage polarization
Metabolic pathway and factors M1 (LPS + IFNγ) M2 (IL-4 + IL-13)
Glycolysis
highly activated;
involved in maintaining the pro-inflamma-
tory properties of macrophages
activated;
important for macrophage
response to IL-4
GLUT1
PFK
6-Phosphofructo-2-kinase/
fructose-2,6-bisphosphatase 3 (PFKFB3)
↑–
PKM2 dimer tetramer
PPP
activated; essential for ROS generation
by NOX, NO production, nucleotide
and protein synthesis
participates in the synthesis
of UDP-GlcNAC
Oxidizing phase of PPP:
phosphogluconate dehydrogenase (PGD)
↑–
Non-oxidizing phase of PPP:
sedoheptulose kinase carbohydrate
kinase-like protein (CARKL)
↓↓
Krebs cycle and OXPHOS
weakened by NO and itaconate;
reverse electron transport, ROS production,
HIF-1α stabilization; IL-1β expression
OXPHOS activation;
intact Krebs cycle
Mitochondrial morphology
predominance
of fragmented mitochondria
predominance
of mitochondrial networks
Citrate = first cycle break
IDH
CIC
ACLY
Itaconate
ACOD1
↑–
Succinate = second cycle break
SDH
↓–
Amino acid metabolism
Arginine
iNOS
ARG1
Tryptophan
Indoleamine 2,3-dioxygenase (IDO)
Glutamine
GS
PHD
Lipid metabolism fatty acids synthesis ↑ fatty acids oxidation ↑
SREBP
Liver X receptor (LXR)
Note. “↑”, green, increased compared to inactivated macrophages (M0); “–”, grey, no changes compared to M0; “↓”, red, decreased
compared to M0.
YURAKOVA et al.826
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
receptors, cytokines, and growth factors [138]. mTORC1
is essential for M1 polarization, while mTORC2 appears
to be important for both types of macrophage activa-
tion [137]. The involvement of mTORC1 is largely asso-
ciated with its ability to regulate anabolism via coor-
dination of production of many factors mediating the
functions of activated macrophages. Thus, SREBP-1 ac-
tivation by mTORC1 is critically important for cytokine
production, synthesis of lipids and lipid mediators,
and phagocytosis. SREBP-1 also controls NADH produc-
tion in the PPP necessary for the ROS generation [136].
The PI3K/AKT/mTOR signaling pathway is important
for the stimulation of glycolysis and cell proliferation
[136, 139]. Finally, mTOR is associated with glutamine
metabolism. The mTORC2/IRF4 signaling pathway is in-
volved in metabolic reprogramming occurring during
alternative macrophage activation [140]. mTORC1 also
promotes M2 polarization, in particular, by stimulat-
ing the synthesis of UDP-GlcNAc and N-glycosylation
of lectins [31]. Moreover, the mTORC1-dependent syn-
thesis of lipid mediators (which is associated with the
expression of COX2 in M1 macrophages and of COX1
in M2 macrophages) is involved in the production of
both proinflammatory and anti-inflammatory media-
tors [141].
AMPK, along with mTOR, participates in the inte-
gration of various signals and regulates cell metabo-
lism through the enzyme-mediated post-translational
modification of target proteins [142]. These master
regulators complement each other. While mTOR senses
the level of nutrients and triggers anabolic processes,
AMPK detects the deficit in energy and activates cat-
abolic processes aimed at ATP synthesis [143]. Thus,
in the case of energy deficit (e.g., during starvation),
AMPK responds to the intracellular AMP content and
phosphorylation state of RAPTOR and TSC2 proteins
and suppresses signaling pathways associated with
mTORC1 [144]. Therefore, AMPK activation in macro-
phages typically results in the suppression of inflam-
matory response and M1 polarization [145-149]. The
absence of AMPK leads to the hyperpolarization of
macrophages towards the M1 phenotype in response
to the LPS stimulation and to the upregulation of ex-
pression of the key enzymes involved in the metabo-
lism of glucose, arginine, prostaglandins, and choles-
terol and synthesis of itaconate. This results in the
increased production of PGE
2
, NO, IL-6, and IL-12 and
reduction in the biosynthesis of IL-10 [150]. In addi-
tion, AMPK is necessary for the acquisition of the M2
phenotype by the macrophages, for example, during
muscle regeneration [146].
Summarizing all the above, metabolic changes
play an essential role in the polarization of macro-
phages and their immunological functions (Table 2).
These processes are not only interconnected, but also
dependent on each other. On one hand, regulation
of immunometabolic enzymes is mediated by nutri-
ents and metabolites (amino acids, citrate, succinate,
itaconate, FAs, etc.). This method of control appeared
at the early stages of evolution and has been function-
ing successfully so far. On the other hand, the regula-
tion of cell functions, including immune response, in
accordance with the body’s needs and environmen-
tal conditions is impossible without more complex
mechanisms that had evolved later. Such mechanisms
mediated, for example, by mTOR and AMPK, include
integration of intracellular signals with subsequent
post-translational modification of proteins and epigen-
etic modification of genes [142]. Immunometabolic ad-
aptation of macrophages to incoming signals underlies
the maintenance of homeostasis not only in individual
tissues, but in an entire organism as well. As a result,
disruption of such regulation can lead to the develop-
ment of various diseases and maladaptive states.
MODULATION OF MACROPHAGE METABOLISM
AS A THERAPEUTIC STRATEGY
Disruption of macrophage immunometabolic
functions as a pathology factor. Recently, the under-
standing of the role of macrophages in various patho-
logical processes has significantly expanded [23, 151]
(Fig.2). On one hand, M2 macrophages are involved in
the metabolic control, while M1 macrophages, on the
contrary, are associated with metabolic pathologies,
such as diabetes, obesity, metabolic syndrome, and in-
sulin resistance. On the other hand, M1 macrophages
mediate the antitumor effects, while M2 macrophages
create an immunosuppressive environment and pro-
mote tumor development. M2 macrophages are respon-
sible for maintaining tissue homeostasis and wound
healing; however, excessive or prolonged activation of
both M1 and M2 macrophages results in tissue damage
and impaired function in such diseases as arthritis, ath-
erosclerosis, glomerulonephritis, and atopic dermatitis.
Macrophages, which provide body protection against
bacteria, viruses, and macroparasites, are at the same
time able to cause a cytokine storm and an unbalanced
response to harmless environmental factors, thus par-
ticipating in the pathogenesis of allergies and asthma.
Since macrophages play a key role in many diseases,
they have become an important subject in the explora-
tion of new therapeutic strategies.
As described above, activated and polarized mac-
rophages undergo metabolic adaptations. Therefore,
modification of macrophage metabolism entails chang-
es in the direction and strength of their polariza-
tion. At the same time, due to the unique features of
metabolism in the activated macrophages, such effects
will be specific and should not affect other body cells,
which can open up the prospects for using this ap-
METABOLIC CHANGES IN MACROPHAGE POLARIZATION 827
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
Fig. 2. Macrophage polarization in normal and pathological conditions. Pink,polarization associated with pathology development;
green,physiological functions of macrophages. Created with BioRender software based on previously published data [23, 151].
proach in the therapy of inflammatory diseases with
substantial role of macrophage. Below, we discuss how
modulation of certain metabolic pathways in macro-
phages could be used as a therapeutic approach.
mTOR and AMPK modulation. As mentioned
above, mTOR is involved in the polarization of both
M1 and M2 macrophages [134]. Treatment of cul-
tured macrophages with rapamycin, an inhibitor of
mTORC1, contributed to the acquisition of the anti-in-
flammatory phenotype, while activation of mTORC1,
on the contrary, promoted formation of the inflam-
matory phenotype [138]. In the model of the CLP (ce-
cal ligation and puncture)-induced sepsis, rapamycin
protected the mice from death by acting on M1 mac-
rophages [152]. Apparently, the effect of rapamy-
cin is due to its ability to cause autophagy, leading
to a decrease in the IL-1β and IL-18 production by
M1 macrophages. This regulation takes place at the
post-transcriptional level and is associated with a de-
crease in the mitochondrial ROS and pro-IL-1β [153].
Modulation of mTOR can also be used in the therapy
of ulcerative colitis. Thus, dioscin (steroid saponin)
simultaneously inhibits signals through the mTORC1/
HIF-1α pathway and activates mTORC2/PPAR-γ signal-
ing, which results in the suppression of aerobic gly-
colysis, increase in the FA oxidation, and shift in mac-
rophage polarization from M1 to M2, resulting in the
mouse protection against the DSS (dextran sodium
sulfate)-induced ulcerative colitis [154]. On the other
hand, in many studies, the use of rapamycin has led
to the shift toward the M1 phenotype. A combination
of rapamycin and hydroxychloroquine caused the
elimination of anti-inflammatory M2 macrophages
in the glioblastoma model and increased the effec-
tiveness of checkpoint inhibitors [155]. Moreover, the
blockade of mTOR contributed to the M1 polarization
of macrophages and ensured plaque stabilization in
the arteries during progressive atherosclerosis [156].
The same effect was achieved by another mTOR block-
er, arsenic trioxide [157].
Modulation of AMPK provides another approach
to the regulation of macrophage metabolism. Met-
formin is used in the clinic for targeting AMPK. It in-
hibits the ETC complex I, which controls ATP and ROS
YURAKOVA et al.828
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
production. Blocking ATP production leads to the in-
crease in the AMP content and ADP/ATP ratio, thereby
activating AMPK [158]. Metformin is known to inhib-
it the LPS-induced IL-1β production and to promote
IL-10 production [115]. Some studies have shown that
metformin regulates the functions of macrophages
in atherosclerosis by suppressing monocyte differen-
tiation and reducing inflammation, oxidative stress,
and apoptosis [159]. It is important to note that met-
formin can regulate macrophage functions through
the AMPK- independent pathways (NF-κB, ABCG5/8,
SIRT1, FOXO1/FABP4, and HMGB1). In the future, a
combination of metformin with drugs affecting the
macrophage functions (e.g., SGLT2 inhibitors, statins,
IL-1β inhibitors) can enhance and expand its thera-
peutic potential [159].
Glycolysis blockade. Most conclusions on the role
of glycolysis in the polarization of macrophages have
been made based on the studies using 2-DG, an inhibi-
tor of glycolysis that also induces various other effects
[54]. New blockers based on 2-DG may have a higher
specificity [160]. Currently, such compounds are inves-
tigated for the anti-cancer activity, while their effect
on the macrophage metabolism remains unclear. Re-
cent studies have shown that cancer and immune cells
can accumulate a dimeric form of the PKM2 capable of
regulating the expression of HIF-1α-dependent genes
[50, 161]. The small-molecule inhibitor TEPP-46, which
inhibits this process and restores the PKM2 enzymatic
activity, has shown promising results in experimen-
tal models of cancer and infectious diseases [50,  162].
TEPP-46 suppresses activation of M1 macrophages and
promotes mitochondrial biogenesis [163]. Another ap-
proach is the direct HIF-1α inhibition. PX-478 (selective
HIF-1α blocker) suppressed tumor cell growth [164]
and was found to be efficient in the mouse model of
atherosclerosis [165]. However, the detailed mecha-
nism of this inhibitor on macrophage polarization still
has to be elucidated.
Application of Krebs cycle metabolites in ther-
apy. Inflammatory processes and malignant transfor-
mation are accompanied with the reprogramming of
the Krebs cycle, associated genes, and metabolites [70,
166, 167]. Dimethylmalonate and 4-octilitaconate (4-OI)
are small molecules that target SDH and were found
to reduce inflammation in several experimental mod-
els [73,  168]. As mentioned above, itaconate and its
derivatives activate NRF2, leading to the stimulation of
the antioxidant response and upregulation of expres-
sion of cytokine-limiting ATF3 [73, 75]. In particular,
4-OI targets the NRF2/KEAP1 pathway and prevents
cytokine storm during acute infection in mice [73].
Moreover, itaconate has recently been shown to inhibit
NLRP3 inflammasome [77]. Such significant immuno-
modulatory potential of itaconate and its derivatives
can be used in therapy. It was shown recently that
itaconate affects M2 polarization of macrophages by in-
hibiting the JAK1/STAT6 signaling pathway. A decrease
in the severity of steroid-resistant asthma in mice ad-
ministered with 4-OI indicates the therapeutic poten-
tial of this metabolite in Th2-dependent diseases [79].
As a compound affecting the NRF2 pathway and
glycolysis, dimethyl fumarate (DMF, a derivative of fu-
marate) is currently used as an immunomodulatory
drug in the treatment of multiple sclerosis and psoria-
sis [169-171]. DMF also inhibits NF-κB, ERK, and other
signaling pathways. It activates NRF2, resulting in cell
protection against oxidative stress, and promotes the
anti-inflammatory phenotype in macrophages [172].
DMF also affects glycolysis by reducing the GAPDH
activity, suggesting that DMF can also be used to mod-
ulate immunometabolism during infectious diseas-
es[170].
Modulation of amino acid and FA metabolism.
The difference between arginine consumption by M1
and M2 macrophages is a basis for a potential selec-
tive therapy. Thus, the inhibition of iNOS with ami-
noguanidine prevented NO production by the macro-
phages and reduced the symptoms of disease in the
mouse model of multiple sclerosis [173]. Other iNOS
inhibitors have shown efficacy in the mouse models
of ischemic kidney disease and lung inflammation
[174,  175]. At the same time, inhibition of ARG1 by
CB-1158 shifted macrophage polarization towards the
M1 phenotype in the mouse model of carcinogenesis.
This resulted in the inhibition of the immunosuppres-
sive microenvironment and increased the efficacy of
antitumor therapy [176,  177]. Currently, the effica-
cy of ARG1 inhibition in cancer therapy is tested in a
number of clinical trials (NCT02903914, NCT03910530,
NCT03314935, NCT03837509, NCT03361228). Inhibi-
tion of enzymes involved in glutamine metabolism by
BPTES or CB-839 has shown promising results in the
models of multiple sclerosis [178] and rheumatoid ar-
thritis [179]. Finally, amino acid deficiency can also be
used for targeting the macrophages. For example, ser-
ine restriction attenuates excessive macrophage acti-
vation during endotoxemia [180], so it could be useful
in the treatment of sepsis. Halofuginone, which mim-
ics amino acid deficiency by activating GCN2 (general
control nonderepressible 2 kinase), reduces intestinal
inflammation [181] and induces autophagy [182].
The role of FA metabolism in the formation of
pathogenetic macrophage phenotype is only partly un-
derstood. In recent years, special attention has been
focused on the drugs that suppress expression or block
the activity of carnitine palmitoyltransferases (e.g.,
etomoxir) [183]. The effect of these compounds on the
development of colitis-associated cancer [184,  185],
bronchial asthma [186] and obstructive pulmonary
disease [187] is currently under study. Some of these
studies have emphasized the correlation between the
METABOLIC CHANGES IN MACROPHAGE POLARIZATION 829
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
FA synthesis/oxidation and activity of NLRP3 inflam-
masome, since the inflammasome is significantly in-
volved in the chronic inflammation and carcinogene-
sis [184].
CONCLUSION
Taken together, the results of studies in recent de-
cades clearly suggest a close relationship between the
metabolism and functional characteristics of immune
cells, including macrophages, upon their activation.
The metabolic needs of activated cells are inextricably
linked to their functional features, so that there is a
significant difference between the cellular metabolism
of pro- and anti-inflammatory immune cells. At the
same time, many aspects of macrophage metabolism
remain poorly understood. The use of drugs aimed at
metabolism reprogramming is hindered by the limit-
ed selectivity of available inhibitors. Currently known
inhibitors tend to affect central processes, e.g., glucose
catabolism (in the case of 2-DG) or mTOR signaling.
The development of more specific drugs targeting cer-
tain metabolic pathways will allow to explore their
effect on the macrophage polarization in more detail.
Italso opens new prospects for immunometabolic ther-
apy through modulation of macrophage involvement
in disease pathogenesis.
Acknowledgments. The authors express their
gratitude to D.  Anisov for discussions and valuable re-
marks.
Contributions. M.S.D. supervised the study and
developed the concept of the review; T.R.Y., E.A.G.,
and M.A.N. carried out the search of literature sources
and discussed the data; T.R.Y. and M.A.N. prepared the
manuscript; M.S.D. and E.A.G. edited the manuscript;
T.R.Y. prepared the images.
Funding. This work was supported by the Russian
Science Foundation (project no.19-75-30032).
Ethics declaration. This work does not describe
any studies involving humans or animals as objects
performed by any of the authors. The authors of this
work declare that they have no conflicts of interest
infinancial or any other sphere.
REFERENCES
1. Warburg, O. (1956) On the origin of cancer cells, Sci-
ence, 123, 309-314, doi:10.1126/science.123.3191.309.
2. Warburg,O., Wind,F., and Negelein,E. (1927) Theme-
tabolism of tumors in the body, J. Gen. Physiol., 8,
519-530, doi:10.1085/jgp.8.6.519.
3. Lercher,A., Baazim,H., and Bergthaler,A. (2020) Sys-
temic immunometabolism: challenges and opportu-
nities, Immunity, 53, 496-509, doi: 10.1016/j.immuni.
2020.08.012.
4. Febbraio, M. A., Hiscock, N., Sacchetti, M., Fisch-
er, C. P., and Pedersen, B. K. (2004) Interleukin-6 is a
novel factor mediating glucose homeostasis during
skeletal muscle contraction, Diabetes, 53, 1643-1648,
doi:10.2337/diabetes.53.7.1643.
5. Hotamisligil, G.S., Shargill, N.S., and Spiegelman, B.M.
(1993) Adipose expression of tumor necrosis factor-
alpha: direct role in obesity-linked insulin resistance,
Science, 259, 87-91, doi:10.1126/science.7678183.
6. Wieman, H.L., Wofford, J.A., and Rathmell, J.C. (2007)
Cytokine stimulation promotes glucose uptake via
phosphatidylinositol-3 kinase/Akt regulation of Glut1
activity and trafficking, Mol. Biol. Cell., 18, 1437-1446,
doi:10.1091/mbc.e06-07-0593.
7. Pearce, E. L., Walsh, M. C., Cejas, P. J., Harms, G. M.,
Shen,H., Wang, L.S., Jones, R.G., and Choi,Y. (2009)
Enhancing CD8 T-cell memory by modulating fatty
acid metabolism, Nature, 460, 103-107, doi: 10.1038/
nature08097.
8. Wang,A., Luan, H.H., and Medzhitov,R. (2019) Anevo-
lutionary perspective on immunometabolism, Science,
363, eaar3932, doi:10.1126/science.aar3932.
9. Gordon, S., Crocker, P.R., Morris,L., Lee, S.H., Perry,
V.H., and Hume, D.A. (1986) Localization and function
of tissue macrophages, Ciba Found Symp., 118, 54-67,
doi:10.1002/9780470720998.ch5.
10. Okabe, Y., and Medzhitov, R. (2016) Tissue biology
perspective on macrophages, Nat. Immunol., 17, 9-17,
doi:10.1038/ni.3320.
11. Wynn, T.A., Chawla,A., and Pollard, J.W. (2013) Mac-
rophage biology in development, homeostasis and dis-
ease, Nature, 496, 445-455, doi:10.1038/nature12034.
12. Gordon,S., Pluddemann, A., and Martinez Estrada,F.
(2014) Macrophage heterogeneity in tissues: phe-
notypic diversity and functions, Immunol. Rev., 262,
36-55, doi:10.1111/imr.12223.
13. Moore, K.J., and Tabas,I. (2011) Macrophages in the
pathogenesis of atherosclerosis, Cell, 145, 341-355,
doi:10.1016/j.cell.2011.04.005.
14. Chawla,A., Nguyen, K. D., and Goh, Y. P. (2011) Mac-
rophage-mediated inflammation in metabolic disease,
Nat. Rev. Immunol., 11, 738-749, doi:10.1038/nri3071.
15. Olefsky, J. M., and Glass, C. K. (2010) Macrophages,
inflammation, and insulin resistance, Annu. Rev.
Physiol., 72, 219-246, doi: 10.1146/annurev-physi-
ol-021909-135846.
16. Wynn, T.A., and Barron,L. (2010) Macrophages: mas-
ter regulators of inflammation and fibrosis, Semin.
Liver Dis., 30, 245-257, doi:10.1055/s-0030-1255354.
17. Noy, R., and Pollard, J. W. (2014) Tumor-associated
macrophages: from mechanisms to therapy, Immunity,
41, 49-61, doi:10.1016/j.immuni.2014.06.010.
18. Ginhoux, F., and Jung, S. (2014) Monocytes and mac-
rophages: developmental pathways and tissue homeo-
YURAKOVA et al.830
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
stasis, Nat. Rev. Immunol., 14, 392-404, doi: 10.1038/
nri3671.
19. Epelman, S., Lavine, K. J., and Randolph, G. J. (2014)
Origin and functions of tissue macrophages, Immuni-
ty, 41, 21-35, doi:10.1016/j.immuni.2014.06.013.
20. Varol,C., Mildner,A., and Jung,S. (2015) Macrophages:
development and tissue specialization, Annu. Rev. Im-
munol., 33, 643-675, doi: 10.1146/annurev-immunol-
032414-112220.
21. Hussell,T., and Bell, T.J. (2014) Alveolar macrophages:
plasticity in a tissue-specific context, Nat. Rev. Immu-
nol., 14, 81-93, doi:10.1038/nri3600.
22. Viola, A., Munari, F., Sanchez-Rodriguez, R., Scola-
ro,T., and Castegna,A. (2019) Themetabolic signature
of macrophage responses, Front. Immunol., 10, 1462,
doi:10.3389/fimmu.2019.01462.
23. Sica, A., and Mantovani, A. (2012) Macrophage plas-
ticity and polarization: invivo veritas, J. Clin. Invest.,
122, 787-795, doi:10.1172/JCI59643.
24. Ginhoux, F., and Guilliams, M. (2016) Tissue-resident
macrophage ontogeny and homeostasis, Immunity,
44, 439-449, doi:10.1016/j.immuni.2016.02.024.
25. Hoeffel, G., and Ginhoux, F. (2015) Ontogeny of tis-
sue-resident macrophages, Front. Immunol., 6, 486,
doi:10.3389/fimmu.2015.00486.
26. Wculek, S. K., Dunphy, G., Heras-Murillo,I., Mastran-
gelo, A., and Sancho, D. (2022) Metabolism of tissue
macrophages in homeostasis and pathology, Cell.
Mol. Immunol., 19, 384-408, doi: 10.1038/s41423-021-
00791-9.
27. Wang,N., Liang,H., and Zen,K. (2014) Molecular mech-
anisms that influence the macrophage m1-m2 polar-
ization balance, Front. Immunol., 5, 614, doi:10.3389/
fimmu.2014.00614.
28. Mills, C. D., Kincaid, K., Alt, J. M., Heilman, M. J.,
and Hill, A. M. (2000) M-1/M-2 macrophages and
the Th1/Th2 paradigm, J. Immunol., 164, 6166-6173,
doi:10.4049/jimmunol.164.12.6166.
29. Biswas, S. K., and Mantovani, A. (2010) Macrophage
plasticity and interaction with lymphocyte subsets:
cancer as a paradigm, Nat. Immunol., 11, 889-896,
doi:10.1038/ni.1937.
30. Geiss, C., Salas, E., Guevara-Coto, J., Regnier-Vigou-
roux,A., and Mora-Rodriguez, R.A. (2022) Multistabil-
ity in macrophage activation pathways and metabolic
implications, Cells, 11, 404, doi:10.3390/cells11030404.
31. Jha, A.K., Huang, S.C., Sergushichev,A., Lampropou-
lou, V., Ivanova, Y., Loginicheva, E., Chmielewski, K.,
Stewart, K.M., Ashall,J., Everts,B., Pearce, E.J., Drig-
gers, E.M., and Artyomov, M.N. (2015) Network inte-
gration of parallel metabolic and transcriptional data
reveals metabolic modules that regulate macrophage
polarization, Immunity, 42, 419-430, doi: 10.1016/
j.immuni.2015.02.005.
32. Tannahill, G. M., Curtis, A. M., Adamik, J., Palsson-
McDermott, E.M., McGettrick, A.F., Goel,G., Frezza,C.,
Bernard, N. J., Kelly, B., Foley, N. H., Zheng, L., Gar-
det,A., Tong,Z., Jany, S.S., Corr, S.C., Haneklaus,M.,
Caffrey, B. E., Pierce, K., Walmsley, S., Beasley, F. C.,
etal. (2013) Succinate is an inflammatory signal that
induces IL-1beta through HIF-1alpha, Nature, 496,
238-242, doi:10.1038/nature11986.
33. Lin,J., Kong,Q., Hao,W., and Hu,W. (2020) High glu-
cose contributes to the polarization of peritoneal mac-
rophages to the M2phenotype invivo and invitro, Mol.
Med. Rep., 22, 127-134, doi:10.3892/mmr.2020.11130.
34. Fukuzumi, M., Shinomiya, H., Shimizu, Y., Ohishi, K.,
and Utsumi, S. (1996) Endotoxin-induced enhance-
ment of glucose influx into murine peritoneal mac-
rophages via GLUT1, Infect. Immun., 64, 108-112,
doi:10.1128/iai.64.1.108-112.1996.
35. Rodriguez-Prados, J. C., Traves, P. G., Cuenca, J.,
Rico, D., Aragones, J., Martin-Sanz, P., Cascante, M.,
and Bosca, L. (2010) Substrate fate in activated mac-
rophages: a comparison between innate, classic,
and alternative activation, J. Immunol., 185, 605-614,
doi:10.4049/jimmunol.0901698.
36. Wang,F., Zhang,S., Vuckovic,I., Jeon,R., Lerman, A.,
Folmes, C. D., Dzeja, P. P., and Herrmann, J. (2018)
Glycolytic stimulation is not a requirement for M2
macrophage differentiation, Cell Metab., 28, 463-475,
doi:10.1016/j.cmet.2018.08.012.
37. Hackett, E. E., Charles-Messance, H., O’Leary, S. M.,
Gleeson, L. E., Munoz-Wolf, N., Case, S., Wedder-
burn,A., Johnston, D.G.W., Williams, M.A., Smyth,A.,
Ouimet,M., Moore, K.J., Lavelle, E.C., Corr, S.C., Gor-
don, S.V., Keane,J., and Sheedy, F.J. (2020) Mycobac-
terium tuberculosis limits host glycolysis and IL-1beta
by restriction of PFK-M via microRNA-21, Cell Rep.,
30, 124-136, doi:10.1016/j.celrep.2019.12.015.
38. Chen,R., Wang,J., Dai,X., Wu,S., Huang,Q., Jiang,L.,
and Kong,X. (2022) Augmented PFKFB3-mediated gly-
colysis by interferon-gamma promotes inflammato-
ry M1 polarization through the JAK2/STAT1 pathway
in local vascular inflammation in Takayasu arteri-
tis, Arthritis Res. Ther., 24, 266, doi: 10.1186/s13075-
022-02960-1.
39. Galvan-Pena, S., Carroll, R. G., Newman, C., Hinchy,
E. C., Palsson-McDermott, E., Robinson, E. K., Covar-
rubias,S., Nadin,A., James, A.M., Haneklaus,M., Car-
penter,S., Kelly, V.P., Murphy, M.P., Modis, L.K., and
O’Neill, L.A. (2019) Malonylation of GAPDH is an in-
flammatory signal in macrophages, Nat. Commun., 10,
338, doi:10.1038/s41467-018-08187-6.
40. Xie, M., Yu, Y., Kang, R., Zhu, S., Yang, L., Zeng, L.,
Sun,X., Yang,M., Billiar, T.R., Wang,H., Cao,L., Jiang,J.,
and Tang, D. (2016) PKM2-dependent glycolysis pro-
motes NLRP3 and AIM2 inflammasome activation, Nat.
Commun., 7, 13280, doi:10.1038/ncomms13280.
41. Wang,T., Liu,H., Lian, G., Zhang, S.Y., Wang,X., and
Jiang,C. (2017) HIF1alpha-induced glycolysis metab-
olism is essential to the activation of inflammatory
METABOLIC CHANGES IN MACROPHAGE POLARIZATION 831
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
macrophages, Mediators Inflamm., 2017, 9029327,
doi:10.1155/2017/9029327.
42. van Uden,P., Kenneth, N.S., and Rocha,S. (2008) Reg-
ulation of hypoxia-inducible factor-1α by NF-κB, Bio-
chem.J., 412, 477-484, doi:10.1042/BJ20080476.
43. Blouin, C.C., Page, E.L., Soucy, G.M., and Richard, D.E.
(2004) Hypoxic gene activation by lipopolysaccharide
in macrophages: implication of hypoxia-inducible
factor 1alpha, Blood, 103, 1124-1130, doi: 10.1182/
blood-2003-07-2427.
44. Rius, J., Guma, M., Schachtrup, C., Akassoglou, K.,
Zinkernagel, A. S., Nizet, V., Johnson, R. S., Haddad,
G. G., and Karin, M. (2008) NF-kappaB links innate
immunity to the hypoxic response through transcrip-
tional regulation of HIF-1alpha, Nature, 453, 807-811,
doi:10.1038/nature06905.
45. Joshi, S., Singh, A. R., Zulcic, M., and Durden, D. L.
(2014) Amacrophage-dominant PI3K isoform controls
hypoxia-induced HIF1alpha and HIF2alpha stability
and tumor growth, angiogenesis, and metastasis, Mol.
Cancer Res., 12, 1520-1531, doi: 10.1158/1541-7786.
MCR-13-0682.
46. Arranz, A., Doxaki, C., Vergadi, E., Martinez de la
Torre,Y., Vaporidi,K., Lagoudaki, E.D., Ieronymaki,E.,
Androulidaki,A., Venihaki,M., Margioris, A.N., Statho-
poulos, E. N., Tsichlis, P. N., and Tsatsanis, C. (2012)
Akt1 and Akt2 protein kinases differentially contrib-
ute to macrophage polarization, Proc. Natl. Acad. Sci.
USA, 109, 9517-9522, doi:10.1073/pnas.1119038109.
47. Cheng, S. C., Quintin, J., Cramer, R. A., Shepardson,
K. M., Saeed, S., Kumar, V., Giamarellos-Bourboulis,
E.J., Martens, J.H., Rao, N.A., Aghajanirefah,A., Man-
jeri, G. R., Li,Y., Ifrim, D. C., Arts, R.J., van der Veer,
B.M., Deen, P.M., Logie,C., O’Neill, L.A., Willems,P.,
van de Veerdonk, F.L., etal. (2014) mTOR- and HIF-1al-
pha-mediated aerobic glycolysis as metabolic basis for
trained immunity, Science, 345, 1250684, doi:10.1126/
science.1250684.
48. Kelley, T.W., Graham, M.M., Doseff, A.I., Pomerantz,
R. W., Lau, S. M., Ostrowski, M. C., Franke, T. F., and
Marsh, C. B. (1999) Macrophage colony-stimulating
factor promotes cell survival through Akt/protein ki-
nase B, J.Biol. Chem., 274, 26393-26398, doi: 10.1074/
jbc.274.37.26393.
49. Vergadi, E., Ieronymaki, E., Lyroni, K., Vaporidi, K.,
and Tsatsanis,C. (2017) Akt signaling pathway in mac-
rophage activation and M1/M2 polarization, J.Immu-
nol., 198, 1006-1014, doi:10.4049/jimmunol.1601515.
50. Palsson-McDermott, E.M., Curtis, A.M., Goel,G., Lauter-
bach, M.A., Sheedy, F.J., Gleeson, L.E., vandenBosch,
M.W., Quinn, S.R., Domingo-Fernandez,R., Johnston,
D.G., Jiang, J.K., Israelsen, W.J., Keane,J., Thomas,C.,
Clish,C., Vander Heiden,M., Xavier, R.J., and O’Neill,
L.A. (2015) Pyruvate kinase M2 regulates Hif-1alpha
activity and IL-1beta induction and is a critical de-
terminant of the warburg effect in LPS-activated
macrophages, Cell Metab., 21, 65-80, doi: 10.1016/
j.cmet.2014.12.005.
51. Mills, E. L., and O’Neill, L. A. (2016) Reprogramming
mitochondrial metabolism in macrophages as an an-
ti-inflammatory signal, Eur. J. Immunol., 46, 13-21,
doi:10.1002/eji.201445427.
52. Baardman, J., Verberk, S. G. S., Prange, K.H.M., van
Weeghel, M., van der Velden, S., Ryan, D. G., Wust,
R.C.I., Neele, A.E., Speijer,D., Denis, S.W., Witte, M.E.,
Houtkooper, R.H., O’Neill, L.A., Knatko, E. V., Dinko-
va-Kostova, A.T., Lutgens,E., de Winther, M.P.J., and
Van den Bossche, J. (2018) A defective pentose phos-
phate pathway reduces inflammatory macrophage
responses during hypercholesterolemia, Cell Rep.,
25, 2044-2052, doi:10.1016/j.celrep.2018.10.092.
53. Yi,L., Liu,Q., Orandle, M.S., Sadiq-Ali,S., Koontz, S.M.,
Choi, U., Torres-Velez, F. J., and Jackson, S. H. (2012)
p47(phox) directs murine macrophage cell fate de-
cisions, Am. J. Pathol.,
180, 1049-1058, doi: 10.1016/
j.ajpath.2011.11.019.
54. Haschemi, A., Kosma, P., Gille, L., Evans, C. R., Bu-
rant, C.F., Starkl,P., Knapp,B., Haas,R., Schmid, J.A.,
Jandl, C., Amir, S., Lubec, G., Park, J., Esterbauer, H.,
Bilban, M., Brizuela, L., Pospisilik, J. A., Otterbein,
L.E., and Wagner,O. (2012) Thesedoheptulose kinase
CARKL directs macrophage polarization through con-
trol of glucose metabolism, Cell Metab., 15, 813-826,
doi:10.1016/j.cmet.2012.04.023.
55. He,D., Mao,Q., Jia,J., Wang,Z., Liu,Y., Liu,T., Luo,B.,
and Zhang,Z. (2021) Pentose phosphate pathway regu-
lates tolerogenic apoptotic cell clearance and immune
tolerance, Front. Immunol., 12, 797091, doi: 10.3389/
fimmu.2021.797091.
56. Huang, S.C., Everts,B., Ivanova,Y., O’Sullivan,D., Na-
scimento,M., Smith, A.M., Beatty,W., Love-Gregory,L.,
Lam, W. Y., O’Neill, C. M., Yan, C., Du, H., Abumrad,
N.A., Urban, J.F.,Jr., Artyomov, M.N., Pearce, E.L., and
Pearce, E.J. (2014) Cell-intrinsic lysosomal lipolysis is
essential for alternative activation of macrophages,
Nat. Immunol., 15, 846-855, doi:10.1038/ni.2956.
57. Van den Bossche,J., Baardman,J., Otto, N.A., van der
Velden,S., Neele, A.E., van den Berg, S.M., Luque-Mar-
tin,R., Chen, H.J., Boshuizen, M.C., Ahmed,M., Hoek-
sema, M.A., de Vos, A.F., and de Winther, M.P. (2016)
Mitochondrial dysfunction prevents repolarization
of inflammatory macrophages, Cell Rep., 17, 684-696,
doi:10.1016/j.celrep.2016.09.008.
58. Garaude, J., Acin-Perez, R., Martinez-Cano, S., En-
amorado, M., Ugolini, M., Nistal-Villan, E., Hervas-
Stubbs, S., Pelegrin, P., Sander, L. E., Enriquez, J. A.,
and Sancho,D. (2016) Mitochondrial respiratory-chain
adaptations in macrophages contribute to antibac-
terial host defense, Nat. Immunol., 17, 1037-1045,
doi:10.1038/ni.3509.
59. Vats, D., Mukundan, L., Odegaard, J. I., Zhang, L.,
Smith, K.L., Morel, C.R., Wagner, R.A., Greaves, D.R.,
YURAKOVA et al.832
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
Murray, P. J., and Chawla,A. (2006) Oxidative metab-
olism and PGC-1beta attenuate macrophage-mediat-
ed inflammation, Cell Metab., 4, 13-24, doi: 10.1016/
j.cmet.2006.05.011.
60. Turrens, J. F. (2003) Mitochondrial formation of
reactive oxygen species, J. Physiol., 552, 335-344,
doi:10.1113/jphysiol.2003.049478.
61. Suski, J. M., Lebiedzinska, M., Bonora, M., Pinton, P.,
Duszynski, J., and Wieckowski, M. R. (2012) Rela-
tion between mitochondrial membrane potential
and ROS formation, Methods Mol. Biol., 810, 183-205,
doi:10.1007/978-1-61779-382-0_12.
62. Bulua, A. C., Simon, A., Maddipati, R., Pelletier, M.,
Park, H., Kim, K. Y., Sack, M. N., Kastner, D. L., and
Siegel, R. M. (2011) Mitochondrial reactive oxygen
species promote production of proinflammatory cy-
tokines and are elevated in TNFR1-associated peri-
odic syndrome (TRAPS), J. Exp. Med., 208, 519-533,
doi:10.1084/jem.20102049.
63. West, A. P., Brodsky, I. E., Rahner, C., Woo, D.K., Erd-
jument-Bromage,H., Tempst,P., Walsh, M.C., Choi,Y.,
Shadel, G.S., and Ghosh,S. (2011) TLR signalling aug-
ments macrophage bactericidal activity through mi-
tochondrial ROS, Nature, 472, 476-480, doi: 10.1038/
nature09973.
64. Agita, A., and Alsagaff, M. T. (2017) Inflammation,
immunity, and hypertension, Acta Med. Indones., 49,
158-165.
65. He, W., Miao, F. J., Lin, D. C., Schwandner, R. T.,
Wang, Z., Gao, J., Chen, J. L., Tian, H., and Ling, L.
(2004) Citric acid cycle intermediates as ligands for
orphan G-protein-coupled receptors, Nature, 429,
188-193, doi:10.1038/nature02488.
66. Hooftman, A., and O’Neill, L. A. J. (2019) The immu-
nomodulatory potential of the metabolite itacon-
ate, Trends Immunol., 40, 687-698, doi: 10.1016/j.it.
2019.05.007.
67. Infantino, V., Iacobazzi, V., Menga, A., Avantaggiati,
M. L., and Palmieri, F. (2014) A key role of the mito-
chondrial citrate carrier (SLC25A1) in TNFα- and
IFNγ-triggered inflammation, Biochim. Biophys. Acta,
1839, 1217-1225, doi:10.1016/j.bbagrm.2014.07.013.
68. Icard, P., Coquerel, A., Wu, Z., Gligorov, J., Fuks, D.,
Fournel, L., Lincet, H., and Simula, L. (2021) Under-
standing the central role of citrate in the metabolism
of cancer cells and tumors: AnUpdate, Int.J. Mol. Sci.,
22, 6587, doi:10.3390/ijms22126587.
69. Infantino, V., Pierri, C. L., and Iacobazzi, V. (2019)
Metabolic routes in inflammation: the citrate
pathway and its potential as therapeutic tar-
get, Curr. Med. Chem., 26, 7104-7116, doi: 10.2174/
0929867325666180510124558.
70. Lampropoulou, V., Sergushichev, A., Bambousko-
va,M., Nair,S., Vincent, E.E., Loginicheva,E., Cervant-
es-Barragan, L., Ma, X., Huang, S. C., Griss, T., Wein-
heimer, C. J., Khader, S., Randolph, G. J., Pearce, E. J.,
Jones, R.G., Diwan,A., Diamond, M.S., and Artyomov,
M. N. (2016) Itaconate links inhibition of succinate
dehydrogenase with macrophage metabolic remod-
eling and regulation of inflammation, Cell Metab., 24,
158-166, doi:10.1016/j.cmet.2016.06.004.
71. Nemeth, B., Doczi, J., Csete, D., Kacso, G., Ravasz, D.,
Adams, D., Kiss, G., Nagy, A. M., Horvath, G., Tret-
ter, L., Mocsai, A., Csepanyi-Komi, R., Iordanov, I.,
Adam-Vizi, V., and Chinopoulos, C. (2016) Abolition
of mitochondrial substrate-level phosphorylation by
itaconic acid produced by LPS-induced Irg1 expres-
sion in cells of murine macrophage lineage, FASEBJ.,
30, 286-300, doi:10.1096/fj.15-279398.
72. Qin,W., Qin,K., Zhang,Y., Jia,W., Chen,Y., Cheng,B.,
Peng, L., Chen, N., Liu, Y., Zhou, W., Wang, Y. L.,
Chen, X., and Wang, C. (2019) S-glycosylation-based
cysteine profiling reveals regulation of glycolysis by
itaconate, Nat. Chem. Biol., 15, 983-991, doi: 10.1038/
s41589-019-0323-5.
73. Mills, E. L., Ryan, D. G., Prag, H. A., Dikovskaya, D.,
Menon, D., Zaslona, Z., Jedrychowski, M. P., Costa,
A.S.H., Higgins,M., Hams,E., Szpyt,J., Runtsch, M.C.,
King, M.S., McGouran, J.F., Fischer,R., Kessler, B.M.,
McGettrick, A.F., Hughes, M.M., Carroll, R.G., Booty,
L. M., et al. (2018) Itaconate is an anti-inflammatory
metabolite that activates Nrf2 via alkylation of KEAP1,
Nature, 556, 113-117, doi:10.1038/nature25986.
74. Zhu,X., Guo,Y., Liu,Z., Yang,J., Tang,H., and Wang,Y.
(2021) Itaconic acid exerts anti-inflammatory and
antibacterial effects via promoting pentose phos-
phate pathway to produce ROS, Sci. Rep., 11, 18173,
doi:10.1038/s41598-021-97352-x.
75. Bambouskova, M., Gorvel, L., Lampropoulou, V., Ser-
gushichev, A., Loginicheva, E., Johnson, K., Koren-
feld, D., Mathyer, M. E., Kim, H., Huang, L. H., Dun-
can, D., Bregman, H., Keskin, A., Santeford, A., Apte,
R. S., Sehgal, R., Johnson, B., Amarasinghe, G. K.,
Soares, M. P., Satoh, T., et al. (2018) Electrophilic
properties of itaconate and derivatives regulate the
IkappaBzeta-ATF3 inflammatory axis, Nature, 556,
501-504, doi:10.1038/s41586-018-0052-z.
76. Bambouskova, M., Potuckova, L., Paulenda, T.,
Kerndl, M., Mogilenko, D. A., Lizotte, K., Swain, A.,
Hayes, S., Sheldon, R. D., Kim, H., Kapadnis, U., Ellis,
A.E., Isaguirre,C., Burdess,S., Laha,A., Amarasinghe,
G.K., Chubukov,V., Roddy, T.P., Diamond, M.S., Jones,
R. G., et al. (2021) Itaconate confers tolerance to late
NLRP3 inflammasome activation, Cell Rep., 34, 108756,
doi:10.1016/j.celrep.2021.108756.
77. Hooftman, A., Angiari, S., Hester, S., Corcoran, S. E.,
Runtsch, M.C., Ling,C., Ruzek, M.C., Slivka, P.F., Mc-
Gettrick, A. F., Banahan, K., Hughes, M. M., Irvine,
A.D., Fischer,R., and O’Neill, L.A.J. (2020) Theimmu-
nomodulatory metabolite itaconate modifies NLRP3
and inhibits inflammasome activation, Cell Metab., 32,
468-478, doi:10.1016/j.cmet.2020.07.016.
METABOLIC CHANGES IN MACROPHAGE POLARIZATION 833
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
78. Puchalska, P., Huang, X., Martin, S. E., Han, X., Patti,
G. J., and Crawford, P. A. (2018) Isotope tracing un-
targeted metabolomics reveals macrophage polar-
ization-state-specific metabolic coordination across
intracellular compartments, iScience, 9, 298-313,
doi:10.1016/j.isci.2018.10.029.
79. Runtsch, M.C., Angiari,S., Hooftman,A., Wadhwa,R.,
Zhang,Y., Zheng,Y., Spina, J.S., Ruzek, M.C., Argiriadi,
M.A., McGettrick, A.F., Mendez, R.S., Zotta,A., Peace,
C.G., Walsh,A., Chirillo,R., Hams,E., Fallon, P.G., Jaya-
maran,R., Dua, K., Brown, A. C., et al. (2022) Itacon-
ate and itaconate derivatives target JAK1 to suppress
alternative activation of macrophages, Cell Metab.,
34, 487-501, doi:10.1016/j.cmet.2022.02.002.
80. Weiss, J.M., Davies, L.C., Karwan,M., Ileva,L., Ozaki,
M. K., Cheng, R. Y., Ridnour, L. A., Annunziata, C. M.,
Wink, D. A., and McVicar, D. W. (2018) Itaconic acid
mediates crosstalk between macrophage metabolism
and peritoneal tumors, J.Clin. Invest., 128, 3794-3805,
doi:10.1172/JCI99169.
81. (2023) Myeloid-derived itaconate suppresses antitu-
mor immunity, Cancer Discov., 13, 255, doi: 10.1158/
2159-8290.CD-RW2022-213.
82. Wang, F., Wang, K., Xu, W., Zhao, S., Ye, D., Wang, Y.,
Xu,Y., Zhou,L., Chu,Y., Zhang,C., Qin,X., Yang,P., and
Yu,H. (2017) SIRT5 desuccinylates and activates pyru-
vate kinaseM2 to block macrophage IL-1beta produc-
tion and to prevent DSS-induced colitis in mice, Cell
Rep., 19, 2331-2344, doi:10.1016/j.celrep.2017.05.065.
83. Kelly, B., and O’Neill, L. A. (2015) Metabolic repro-
gramming in macrophages and dendritic cells in in-
nate immunity, Cell Res., 25, 771-784, doi: 10.1038/
cr.2015.68.
84. Littlewood-Evans, A., Sarret, S., Apfel, V., Loesle, P.,
Dawson,J., Zhang,J., Muller,A., Tigani,B., Kneuer,R.,
Patel,S., Valeaux,S., Gommermann,N., Rubic-Schnei-
der, T., Junt, T., and Carballido, J. M. (2016) GPR91
senses extracellular succinate released from inflam-
matory macrophages and exacerbates rheumatoid
arthritis, J. Exp. Med., 213, 1655-1662, doi: 10.1084/
jem.20160061.
85. Hooftman, A., Peace, C. G., Ryan, D. G., Day, E. A.,
Yang, M., McGettrick, A. F., Yin, M., Montano, E. N.,
Huo, L., Toller-Kawahisa, J. E., Zecchini, V., Ryan,
T.A.J., Bolado-Carrancio,A., Casey, A.M., Prag, H.A.,
Costa, A.S.H., De Los Santos,G., Ishimori,M., Wallace,
D.J., Venuturupalli,S., etal. (2023) Macrophage fuma-
rate hydratase restrains mtRNA-mediated interferon
production, Nature, 615, 490-498, doi:10.1038/s41586-
023-05720-6.
86. Zaslona, Z., and O’Neill, L. A. J. (2020) Cytokine-like
roles for metabolites in immunity, Mol. Cell, 78,
814-823, doi:10.1016/j.molcel.2020.04.002.
87. Tsukada, Y., Fang, J., Erdjument-Bromage, H., War-
ren, M. E., Borchers, C. H., Tempst, P., and Zhang, Y.
(2006) Histone demethylation by a family of JmjC
domain-containing proteins, Nature, 439, 811-816,
doi:10.1038/nature04433.
88. Tahiliani,M., Koh, K.P., Shen,Y., Pastor, W.A., Banduk-
wala,H., Brudno,Y., Agarwal,S., Iyer, L.M., Liu, D.R.,
Aravind,L., and Rao,A. (2009) Conversion of 5-meth-
ylcytosine to 5-hydroxymethylcytosine in mammali-
an DNA by MLL partner TET1, Science, 324, 930-935,
doi:10.1126/science.1170116.
89. Liu, Z., Gan, L., Zhang, T., Ren, Q., and Sun, C. (2018)
Melatonin alleviates adipose inflammation through
elevating alpha-ketoglutarate and diverting adipose-
derived exosomes to macrophages in mice, J. Pineal
Res., 64, e12455, doi:10.1111/jpi.12455.
90. Liu, P.S., Wang,H., Li,X., Chao,T., Teav,T., Christen,S.,
Di Conza,G., Cheng, W.C., Chou, C. H., Vavakova,M.,
Muret, C., Debackere, K., Mazzone, M., Huang, H. D.,
Fendt, S. M., Ivanisevic, J., and Ho, P. C. (2017) al-
pha-ketoglutarate orchestrates macrophage activation
through metabolic and epigenetic reprogramming,
Nat. Immunol., 18, 985-994, doi:10.1038/ni.3796.
91. Dang,L., White, D.W., Gross,S., Bennett, B.D., Bittinger,
M.A., Driggers, E.M., Fantin, V. R., Jang, H.G., Jin, S.,
Keenan, M. C., Marks, K. M., Prins, R. M., Ward, P. S.,
Yen, K.E., Liau, L.M., Rabinowitz, J.D., Cantley, L.C.,
Thompson, C.B., Vander Heiden, M.G., and Su, S.M.
(2009) Cancer-associated IDH1 mutations produce
2-hydroxyglutarate, Nature, 462, 739-744, doi:10.1038/
nature08617.
92. Dang,L., White, D.W., Gross,S., Bennett, B.D., Bittinger,
M.A., Driggers, E.M., Fantin, V. R., Jang, H.G., Jin, S.,
Keenan, M.C., Marks, K.M., Prins, R.M., Ward, P.S., Yen,
K.E., Liau, L.M., Rabinowitz, J.D., Cantley, L.C., Thomp-
son, C. B., Vander Heiden, M. G., and Su, S. M. (2010)
Cancer-associated IDH1 mutations produce 2-hydroxy-
glutarate, Nature, 465, 966, doi:10.1038/nature09132.
93. Carey, B.W., Finley, L.W., Cross, J.R., Allis, C.D., and
Thompson, C. B. (2015) Intracellular alpha-ketogluta-
rate maintains the pluripotency of embryonic stem
cells, Nature, 518, 413-416, doi:10.1038/nature13981.
94. Sciacovelli, M., Goncalves, E., Johnson, T. I., Zecchini,
V. R., da Costa, A. S., Gaude, E., Drubbel, A. V., Theo-
bald, S. J., Abbo, S. R., Tran, M. G., Rajeeve, V., Car-
daci, S., Foster, S., Yun, H., Cutillas, P., Warren, A.,
Gnanapragasam,V., Gottlieb,E., Franze,K., Huntly,B.,
et al. (2016) Fumarate is an epigenetic modifier that
elicits epithelial-to-mesenchymal transition, Nature,
537, 544-547, doi:10.1038/nature19353.
95. Du, X., and Hu, H. (2021) The roles of 2-hydroxyglu-
tarate, Front. Cell Dev. Biol., 9, 651317, doi: 10.3389/
fcell.2021.651317.
96. Williams, N.C., Ryan, D.G., Costa, A.S.H., Mills, E.L.,
Jedrychowski, M. P., Cloonan, S. M., Frezza, C., and
O’Neill, L.A. (2022) Signaling metabolite L-2-hydroxy-
glutarate activates the transcription factor HIF-1alpha
in lipopolysaccharide-activated macrophages, J. Biol.
Chem., 298, 101501, doi:10.1016/j.jbc.2021.101501.
YURAKOVA et al.834
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
97. de Goede, K. E., Harber, K. J., Gorki, F. S., Verberk,
S. G. S., Groh, L. A., Keuning, E. D., Struys, E.A., van
Weeghel, M., Haschemi, A., de Winther, M. P. J., van
Dierendonck, X., and Van den Bossche, J. (2022)
d-2-Hydroxyglutarate is an anti-inflammatory immu-
nometabolite that accumulates in macrophages after
TLR4 activation, Biochim. Biophys. Acta Mol. Basis.
Dis., 1868, 166427, doi:10.1016/j.bbadis.2022.166427.
98. Kieler, M., Hofmann, M., and Schabbauer, G. (2021)
More than just protein building blocks: how amino ac-
ids and related metabolic pathways fuel macrophage
polarization, FEBS J., 288, 3694-3714, doi: 10.1111/
febs.15715.
99. Sivangala Thandi,R., Radhakrishnan, R.K., Tripathi,D.,
Paidipally,P., Azad, A.K., Schlesinger, L.S., Samten,B.,
Mulik, S., and Vankayalapati, R. (2020) Ornithine-A
urea cycle metabolite enhances autophagy and con-
trols Mycobacterium tuberculosis infection, Nat. Com-
mun., 11, 3535, doi:10.1038/s41467-020-17310-5.
100. Gutierrez, M.G., Master, S.S., Singh, S.B., Taylor, G.A.,
Colombo, M. I., and Deretic, V. (2004) Autophagy is a
defense mechanism inhibiting BCG and Mycobacte-
rium tuberculosis survival in infected macrophages,
Cell, 119, 753-766, doi:10.1016/j.cell.2004.11.038.
101. Brown, G. C., and Cooper, C. E. (1994) Nanomolar
concentrations of nitric oxide reversibly inhibit
synaptosomal respiration by competing with oxy-
gen at cytochrome oxidase, FEBS Lett., 356, 295-298,
doi:10.1016/0014-5793(94)01290-3.
102. Cleeter, M.W., Cooper, J.M., Darley-Usmar, V.M., Mon-
cada,S., and Schapira, A.H. (1994) Reversible inhibi-
tion of cytochrome c oxidase, the terminal enzyme of
the mitochondrial respiratory chain, by nitric oxide.
Implications for neurodegenerative diseases, FEBS
Lett., 345, 50-54, doi:10.1016/0014-5793(94)00424-2.
103. Jin,Z., Wei,W., Yang,M., Du,Y., and Wan,Y. (2014) Mi-
tochondrial complex I activity suppresses inflamma-
tion and enhances bone resorption by shifting macro-
phage-osteoclast polarization, Cell Metab., 20, 483-498,
doi:10.1016/j.cmet.2014.07.011.
104. McNeill, E., Crabtree, M. J., Sahgal, N., Patel, J., Ch-
uaiphichai, S., Iqbal, A. J., Hale, A. B., Greaves, D. R.,
and Channon, K. M. (2015) Regulation of iNOS func-
tion and cellular redox state by macrophage Gch1 re-
veals specific requirements for tetrahydrobiopterin
in NRF2 activation, Free Radic. Biol. Med., 79, 206-216,
doi:10.1016/j.freeradbiomed.2014.10.575.
105. Seim, G.L., Britt, E.C., John, S.V., Yeo, F.J., Johnson,
A.R., Eisenstein, R.S., Pagliarini, D.J., and Fan,J. (2019)
Two-stage metabolic remodelling in macrophages in
response to lipopolysaccharide and interferon-gam-
ma stimulation, Nat. Metab., 1, 731-742, doi: 10.1038/
s42255-019-0083-2.
106. Palmieri, E. M., Gonzalez-Cotto, M., Baseler, W. A.,
Davies, L.C., Ghesquiere,B., Maio,N., Rice, C.M., Rou-
ault, T. A., Cassel, T., Higashi, R. M., Lane, A. N., Fan,
T. W., Wink, D. A., and McVicar, D. W. (2020) Nitric
oxide orchestrates metabolic rewiring in M1 macro-
phages by targeting aconitase 2 and pyruvate dehy-
drogenase, Nat. Commun., 11, 698, doi:10.1038/s41467-
020-14433-7.
107. Byles, V., Covarrubias, A. J., Ben-Sahra, I., Lamming,
D. W., Sabatini, D. M., Manning, B. D., and Horng, T.
(2013) TheTSC-mTOR pathway regulates macrophage
polarization, Nat. Commun., 4, 2834, doi: 10.1038/
ncomms3834.
108. Newsholme, P., Gordon, S., and Newsholme, E. A.
(1987) Rates of utilization and fates of glucose, glu-
tamine, pyruvate, fatty acids and ketone bodies
by mouse macrophages, Biochem. J., 242, 631-636,
doi:10.1042/bj2420631.
109. Meiser, J., Kramer, L., Sapcariu, S. C., Battello, N.,
Ghelfi, J., D’Herouel, A. F., Skupin, A., and Hiller, K.
(2016) Pro-inflammatory macrophages sustain pyru-
vate oxidation through pyruvate dehydrogenase for
the synthesis of itaconate and to enable cytokine ex-
pression, J. Biol. Chem., 291, 3932-3946, doi: 10.1074/
jbc.M115.676817.
110. Funk, J. L., Feingold, K. R., Moser, A. H., and Grun-
feld, C. (1993) Lipopolysaccharide stimulation of
RAW 264.7 macrophages induces lipid accumulation
and foam cell formation, Atherosclerosis, 98, 67-82,
doi:10.1016/0021-9150(93)90224-i.
111. Hu,L., Yu,Y., Huang,H., Fan,H., Hu,L., Yin,C., Li,K.,
Fulton, D. J., and Chen, F. (2016) Epigenetic rRegula-
tion of interleukin 6 by histone acetylation in mac-
rophages and its role in paraquat-induced pulmo-
nary fibrosis, Front. Immunol., 7, 696, doi: 10.3389/
fimmu.2016.00696.
112. Zhang, Q., Zhao, K., Shen, Q., Han, Y., Gu, Y., Li, X.,
Zhao, D., Liu, Y., Wang, C., Zhang, X., Su, X., Liu, J.,
Ge,W., Levine, R.L., Li, N., and Cao,X. (2015) Tet2 is
required to resolve inflammation by recruiting Hdac2
to specifically repress IL-6, Nature, 525, 389-393,
doi:10.1038/nature15252.
113. Biswas, S. K., and Lopez-Collazo, E. (2009) Endo-
toxin tolerance: new mechanisms, molecules and
clinical significance, Trends Immunol., 30, 475-487,
doi:10.1016/j.it.2009.07.009.
114. Palmieri, E.M., Menga,A., Martin-Perez,R., Quinto,A.,
Riera-Domingo,C., De Tullio,G., Hooper, D.C., Lamers,
W.H., Ghesquiere,B., McVicar, D.W., Guarini,A., Maz-
zone,M., and Castegna,A. (2017) Pharmacologic or ge-
netic targeting of glutamine synthetase skews macro-
phages toward an M1-like phenotype and inhibits tu-
mor metastasis, Cell Rep., 20, 1654-1666, doi:10.1016/
j.celrep.2017.07.054.
115. Kelly,B., Tannahill, G.M., Murphy, M.P., and O’Neill,
L. A. (2015) Metformin inhibits the production of re-
active oxygen species from NADH:Ubiquinone ox-
idoreductase to limit induction of interleukin-1beta
(IL-1beta) and boosts interleukin-10 (IL-10) in lipopoly-
METABOLIC CHANGES IN MACROPHAGE POLARIZATION 835
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
saccharide (LPS)-activated macrophages, J.Biol. Chem.,
290, 20348-20359, doi:10.1074/jbc.M115.662114.
116. Zhou,R., Yazdi, A.S., Menu,P., and Tschopp,J. (2011)
Arole for mitochondria in NLRP3 inflammasome acti-
vation, Nature, 469, 221-225, doi:10.1038/nature09663.
117. Thimmulappa, R. K., Lee, H., Rangasamy, T., Reddy,
S. P., Yamamoto, M., Kensler, T. W., and Biswal, S.
(2006) Nrf2 is a critical regulator of the innate im-
mune response and survival during experimental sep-
sis, J.Clin. Invest., 116, 984-995, doi:10.1172/JCI25790.
118. Kobayashi, E. H., Suzuki, T., Funayama, R., Nagashi-
ma, T., Hayashi, M., Sekine, H., Tanaka, N., Morigu-
chi, T., Motohashi, H., Nakayama, K., and Yamamo-
to,M. (2016) Nrf2 suppresses macrophage inflamma-
tory response by blocking proinflammatory cytokine
transcription, Nat. Commun., 7, 11624, doi: 10.1038/
ncomms11624.
119. Yu, W., Wang, Z., Zhang, K., Chi, Z., Xu, T., Jiang, D.,
Chen, S., Li, W., Yang, X., Zhang, X., Wu, Y., and
Wang, D. (2019) One-carbon metabolism supports
S-adenosylmethionine and histone nethylation to
drive inflammatory macrophages, Mol. Cell, 75, 1147-
1160, doi:10.1016/j.molcel.2019.06.039.
120. Castoldi,A., Monteiro, L.B., van Teijlingen Bakker,N.,
Sanin, D. E., Rana, N., Corrado, M., Cameron, A. M.,
Hassler, F., Matsushita, M., Caputa, G., Klein Geltink,
R. I., Buscher, J., Edwards-Hicks, J., Pearce, E. L., and
Pearce, E.J. (2020) Triacylglycerol synthesis enhances
macrophage inflammatory function, Nat. Commun.,
11, 4107, doi:10.1038/s41467-020-17881-3.
121. Morgan, P. K., Huynh, K., Pernes, G., Miotto, P. M.,
Mellett, N.A., Giles,C., Meikle, P.J., Murphy, A.J., and
Lancaster, G.I. (2021) Macrophage polarization state
affects lipid composition and the channeling of exog-
enous fatty acids into endogenous lipid pools, J.Biol.
Chem., 297, 101341, doi:10.1016/j.jbc.2021.101341.
122. Guerrini,V., and Gennaro, M.L. (2019) Foam cells: one
size doesn’t fit all, Trends Immunol., 40, 1163-1179,
doi:10.1016/j.it.2019.10.002.
123. Im, S. S., Yousef, L., Blaschitz, C., Liu, J. Z., Edwards,
R. A., Young, S. G., Raffatellu, M., and Osborne, T. F.
(2011) Linking lipid metabolism to the innate immune
response in macrophages through sterol regulatory
element binding protein-1a, Cell Metab., 13, 540-549,
doi:10.1016/j.cmet.2011.04.001.
124. Batista-Gonzalez, A., Vidal, R., Criollo, A., and Carre-
no, L. J. (2019) New insights on the role of lipid me-
tabolism in the metabolic reprogramming of mac-
rophages, Front. Immunol., 10, 2993, doi: 10.3389/
fimmu.2019.02993.
125. van Dierendonck,X., Sancerni,T., Alves-Guerra, M.C.,
and Stienstra, R. (2020) The role of uncoupling pro-
tein 2 in macrophages and its impact on obesity-in-
duced adipose tissue inflammation and insulin resis-
tance, J. Biol. Chem., 295, 17535-17548, doi: 10.1074/
jbc.RA120.014868.
126. Yeudall, S., Upchurch, C. M., Seegren, P. V., Pavelec,
C. M., Greulich, J., Lemke, M. C., Harris, T. E., Desai,
B.N., Hoehn, K.L., and Leitinger,N. (2022) Macrophage
acetyl-CoA carboxylase regulates acute inflammation
through control of glucose and lipid metabolism, Sci.
Adv., 8, eabq1984, doi:10.1126/sciadv.abq1984.
127. Nomura, M., Liu, J., Rovira, I., Gonzalez-Hurtado, E.,
Lee,J., Wolfgang, M.J., and Finkel,T. (2016) Fatty acid
oxidation in macrophage polarization, Nat. Immunol.,
17, 216-217, doi:10.1038/ni.3366.
128. Divakaruni, A.S., Hsieh, W.Y., Minarrieta,L., Duong,
T.N., Kim, K.K.O., Desousa, B.R., Andreyev, A.Y., Bow-
man, C.E., Caradonna,K., Dranka, B.P., Ferrick, D.A.,
Liesa, M., Stiles, L., Rogers, G. W., Braas, D., Ciaraldi,
T.P., Wolfgang, M.J., Sparwasser,T., Berod,L., Bensing-
er, S.J., etal. (2018) Etomoxir inhibits macrophage po-
larization by disrupting CoA homeostasis, Cell Metab.,
28, 490-503, doi:10.1016/j.cmet.2018.06.001.
129. Van den Bossche, J., O’Neill, L. A., and Menon, D.
(2017) Macrophage immunometabolism: where are
we (going)? Trends Immunol., 38, 395-406, doi:10.1016/
j.it.2017.03.001.
130. Moon, J.S., Nakahira,K., Chung, K.P., DeNicola, G.M.,
Koo, M.J., Pabon, M.A., Rooney, K.T., Yoon, J.H., Ry-
ter, S. W., Stout-Delgado, H., and Choi, A. M. (2016)
NOX4-dependent fatty acid oxidation promotes NLRP3
inflammasome activation in macrophages, Nat. Med.,
22, 1002-1012, doi:10.1038/nm.4153.
131. Olona, A., Leishman, S., and Anand, P. K. (2022)
The NLRP3 inflammasome: regulation by metabolic
signals, Trends Immunol., 43, 978-989, doi: 10.1016/
j.it.2022.10.003.
132. Kumar, S., Mittal, S., Gupta, P., Singh, M., Chaluval-
ly-Raghavan, P., and Pradeep, S. (2022) Metabolic re-
programming in tumor-associated macrophages in the
ovarian tumor microenvironment, Cancers (Basel),
14, 5224, doi:10.3390/cancers14215224.
133. Wu,L., Zhang,X., Zheng,L., Zhao,H., Yan,G., Zhang,Q.,
Zhou, Y., Lei, J., Zhang, J., Wang, J., Xin, R., Jiang, L.,
Peng, J., Chen, Q., Lam, S. M., Shui, G., Miao, H., and
Li, Y. (2020) RIPK3 orchestrates fatty acid metabo-
lism in tumor-associated macrophages and hepa-
tocarcinogenesis, Cancer Immunol. Res., 8, 710-721,
doi:10.1158/2326-6066.CIR-19-0261.
134. Covarrubias, A.J., Aksoylar, H.I., and Horng,T. (2015)
Control of macrophage metabolism and activation by
mTOR and Akt signaling, Semin. Immunol., 27, 286-296,
doi:10.1016/j.smim.2015.08.001.
135. Kang, S., and Kumanogoh, A. (2020) The spectrum
of macrophage activation by immunometabolism,
Int. Immunol., 32, 467-473, doi: 10.1093/intimm/
dxaa017.
136. Dibble, C. C., and Manning, B. D. (2013) Signal in-
tegration by mTORC1 coordinates nutrient input
with biosynthetic output, Nat. Cell Biol., 15, 555-564,
doi:10.1038/ncb2763.
YURAKOVA et al.836
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
137. Weichhart, T., Hengstschlager, M., and Linke, M.
(2015) Regulation of innate immune cell function by
mTOR, Nat. Rev. Immunol., 15, 599-614, doi: 10.1038/
nri3901.
138. Weichhart,T., Costantino,G., Poglitsch,M., Rosner,M.,
Zeyda, M., Stuhlmeier, K. M., Kolbe,T., Stulnig, T. M.,
Horl, W. H., Hengstschlager, M., Muller, M., and Sae-
mann, M.D. (2008) TheTSC-mTOR signaling pathway
regulates the innate inflammatory response, Immuni-
ty, 29, 565-577, doi:10.1016/j.immuni.2008.08.012.
139. Robey, R. B., and Hay, N. (2009) Is Akt the “Warburg
kinase”?-Akt-energy metabolism interactions and on-
cogenesis, Semin. Cancer Biol., 19, 25-31, doi:10.1016/
j.semcancer.2008.11.010.
140. Huang, S. C., Smith, A. M., Everts, B., Colonna, M.,
Pearce, E. L., Schilling, J. D., and Pearce, E. J. (2016)
Metabolic reprogramming mediated by the mTORC2-
IRF4 signaling axis is essential for macrophage alter-
native activation, Immunity, 45, 817-830, doi:10.1016/
j.immuni.2016.09.016.
141. Biswas, S. K., and Mantovani, A. (2012) Orchestra-
tion of metabolism by macrophages, Cell Metab., 15,
432-437, doi:10.1016/j.cmet.2011.11.013.
142. Beltrao, P., Bork, P., Krogan, N. J., and van Noort, V.
(2013) Evolution and functional cross-talk of protein
post-translational modifications, Mol. Syst. Biol., 9,
714, doi:10.1002/msb.201304521.
143. Stein, S. C., Woods, A., Jones, N. A., Davison, M. D.,
and Carling,D. (2000) Theregulation of AMP-activat-
ed protein kinase by phosphorylation, Biochem. J.,
345, 437-443, doi:10.1042/bj3450437.
144. O’Neill, L.A., and Hardie, D. G. (2013) Metabolism of
inflammation limited by AMPK and pseudo-starva-
tion, Nature, 493, 346-355, doi:10.1038/nature11862.
145. Nath,N., Khan,M., Rattan,R., Mangalam,A., Makkar,
R.S., de Meester,C., Bertrand,L., Singh,I., Chen,Y., Vi-
ollet,B., and Giri,S. (2009) Loss of AMPK exacerbates
experimental autoimmune encephalomyelitis disease
severity, Biochem. Biophys. Res. Commun., 386, 16-20,
doi:10.1016/j.bbrc.2009.05.106.
146. Mounier, R., Theret, M., Arnold, L., Cuvellier, S., Bul-
tot,L., Goransson,O., Sanz,N., Ferry,A., Sakamoto,K.,
Foretz, M., Viollet, B., and Chazaud, B. (2013) AMP-
Kalpha1 regulates macrophage skewing at the time
of resolution of inflammation during skeletal muscle
regeneration, Cell Metab., 18, 251-264, doi: 10.1016/
j.cmet.2013.06.017.
147. Sag, D., Carling, D., Stout, R. D., and Suttles, J. (2008)
Adenosine 5′-monophosphate-activated protein kinase
promotes macrophage polarization to an anti-inflam-
matory functional phenotype, J.Immunol., 181, 8633-
8641, doi:10.4049/jimmunol.181.12.8633.
148. Vasamsetti, S.B., Karnewar,S., Kanugula, A.K., Thati-
palli, A. R., Kumar, J. M., and Kotamraju, S. (2015)
Metformin inhibits monocyte-to-macrophage differ-
entiation via AMPK-mediated inhibition of STAT3 ac-
tivation: potential role in atherosclerosis, Diabetes,
64, 2028-2041, doi:10.2337/db14-1225.
149. Seneviratne, A., Cave, L., Hyde, G., Moestrup, S. K.,
Carling,D., Mason, J.C., Haskard, D.O., and Boyle, J.J.
(2021) Metformin directly suppresses atherosclerosis
in normoglycaemic mice via haematopoietic adenos-
ine monophosphate-activated protein kinase, Cardio-
vasc. Res.,
117, 1295-1308, doi:10.1093/cvr/cvaa171.
150. Phair, I.R., Nisr, R.B., Howden, A.J.M., Sovakova,M.,
Alqurashi, N., Foretz, M., Lamont, D., Viollet, B., and
Rena, G. (2023) AMPK integrates metabolite and
kinase-based immunometabolic control in macro-
phages, Mol. Metab., 68, 101661, doi:10.1016/j.molmet.
2022.101661.
151. Kadomoto,S., Izumi,K., and Mizokami,A. (2021) Mac-
rophage polarity and disease dontrol, Int.J. Mol. Sci.,
23, 144, doi:10.3390/ijms23010144.
152. Wang, Z., Li, Y., Yang, X., Zhang, L., Shen, H., Xu, W.,
and Yuan, C. (2019) Protective effects of rapamycin
induced autophagy on CLP septic mice, Comp. Im-
munol. Microbiol. Infect. Dis., 64, 47-52, doi: 10.1016/
j.cimid.2019.01.009.
153. Ko, J. H., Yoon, S. O., Lee, H. J., and Oh, J. Y. (2017)
Rapamycin regulates macrophage activation by in-
hibiting NLRP3 inflammasome-p38 MAPK-NFkappaB
pathways in autophagy- and p62-dependent man-
ners, Oncotarget, 8, 40817-40831, doi: 10.18632/
oncotarget.17256.
154. Wu, M.M., Wang, Q.M., Huang, B.Y., Mai, C.T., Wang,
C.L., Wang, T.T., and Zhang, X.J. (2021) Dioscin ame-
liorates murine ulcerative colitis by regulating mac-
rophage polarization, Pharmacol. Res., 172, 105796,
doi:10.1016/j.phrs.2021.105796.
155. Hsu, S. P. C., Chen, Y. C., Chiang, H. C., Huang, Y. C.,
Huang, C.C., Wang, H. E., Wang, Y. S., and Chi, K.H.
(2020) Rapamycin and hydroxychloroquine combi-
nation alters macrophage polarization and sensitiz-
es glioblastoma to immune checkpoint inhibitors,
J. Neurooncol., 146, 417-426, doi: 10.1007/s11060-
019-03360-3.
156. Sun, R. Z., Fan, Y., Liang, X., Gong, T. T., Wang, Q.,
Liu,H., Shan, Z.Y., and Lei,L. (2018) Rapamycin and
FTY720 alleviate atherosclerosis by cross talk of mac-
rophage polarization and autophagy, Biomed. Res. Int.,
2018, 1010248, doi:10.1155/2018/1010248.
157. Fang, S., Wan, X., Zou, X., Sun, S., Hao, X., Liang, C.,
Zhang, Z., Zhang, F., Sun, B., Li, H., and Yu, B. (2021)
Arsenic trioxide induces macrophage autophagy and
atheroprotection by regulating ROS-dependent TFEB
nuclear translocation and AKT/mTOR pathway, Cell
Death Dis., 12, 88, doi:10.1038/s41419-020-03357-1.
158. Ursini,F., Russo,E., Pellino,G., D’Angelo,S., Chiaraval-
loti,A., De Sarro,G., Manfredini,R., and De Giorgio,R.
(2018) Metformin and autoimmunity: a “new deal”
of an old drug, Front. Immunol., 9, 1236, doi:10.3389/
fimmu.2018.01236.
METABOLIC CHANGES IN MACROPHAGE POLARIZATION 837
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
159. Feng, X., Chen, W., Ni, X., Little, P. J., Xu, S., Tang, L.,
and Weng,J. (2021) Metformin, macrophage dysfunc-
tion and atherosclerosis, Front. Immunol., 12, 682853,
doi:10.3389/fimmu.2021.682853.
160. Pajak, B., Siwiak, E., Soltyka, M., Priebe, A., Zielins-
ki, R., Fokt, I., Ziemniak, M., Jaskiewicz, A., Borows-
ki, R., Domoradzki, T., and Priebe, W. (2019) 2-De-
oxy-d-glucose and its analogs: from diagnostic to ther-
apeutic agents, Int. J. Mol. Sci., 21, 234, doi:10.3390/
ijms21010234.
161. Luo, W., Hu, H., Chang, R., Zhong, J., Knabel, M.,
O’Meally, R., Cole, R. N., Pandey, A., and Semenza,
G.L. (2011) Pyruvate kinase M2 is a PHD3-stimulated
coactivator for hypoxia-inducible factor 1, Cell, 145,
732-744, doi:10.1016/j.cell.2011.03.054.
162. Anastasiou,D., Yu,Y., Israelsen, W.J., Jiang, J.K., Box-
er, M.B., Hong, B.S., Tempel,W., Dimov,S., Shen,M.,
Jha,A., Yang,H., Mattaini, K.R., Metallo, C.M., Fiske,
B.P., Courtney, K.D., Malstrom,S., Khan, T.M., Kung,C.,
Skoumbourdis, A.P., Veith,H., etal. (2012) Pyruvate ki-
nase M2 activators promote tetramer formation and
suppress tumorigenesis, Nat. Chem. Biol., 8, 839-847,
doi:10.1038/nchembio.1060.
163. Yi,Z., Wu,Y., Zhang,W., Wang,T., Gong,J., Cheng,Y.,
and Miao, C. (2020) Activator-mediated pyruvate ki-
naseM2 activation contributes to endotoxin tolerance
by promoting mitochondrial biogenesis, Front. Immu-
nol., 11, 595316, doi:10.3389/fimmu.2020.595316.
164. Welsh, S., Williams, R., Kirkpatrick, L., Paine-Murri-
eta, G., and Powis, G. (2004) Antitumor activity and
pharmacodynamic properties of PX-478, an inhibitor
of hypoxia-inducible factor-1alpha, Mol. Cancer Ther.,
3, 233-244, doi:10.1158/1535-7163.233.3.3.
165. Villa-Roel,N., Ryu, K., Gu, L., Fernandez Esmerats, J.,
Kang, D.W., Kumar, S., and Jo, H. (2022) Hypoxia in-
ducible factor 1alpha inhibitor PX-478 reduces ath-
erosclerosis in mice, Atherosclerosis, 344, 20-30,
doi:10.1016/j.atherosclerosis.2022.01.002.
166. Mills, E. L., Kelly, B., Logan, A., Costa, A. S. H., Var-
ma,M., Bryant, C.E., Tourlomousis,P., Dabritz, J.H.M.,
Gottlieb, E., Latorre, I., Corr, S. C., McManus, G.,
Ryan,D., Jacobs, H.T., Szibor,M., Xavier, R.J., Braun,T.,
Frezza,C., Murphy, M.P., and O’Neill, L.A. (2016) Suc-
cinate dehydrogenase supports metabolic repurposing
of mitochondria to drive inflammatory macrophages,
Cell, 167, 457-470, doi:10.1016/j.cell.2016.08.064.
167. Yen, K. E., Bittinger, M. A., Su, S.M., and Fantin, V. R.
(2010) Cancer-associated IDH mutations: biomarker
and therapeutic opportunities, Oncogene, 29, 6409-
6417, doi:10.1038/onc.2010.444.
168. Chouchani, E.T., Pell, V.R., Gaude,E., Aksentijevic,D.,
Sundier, S.Y., Robb, E. L., Logan,A., Nadtochiy, S.M.,
Ord, E. N. J., Smith, A. C., Eyassu, F., Shirley, R., Hu,
C.H., Dare, A.J., James, A.M., Rogatti,S., Hartley, R.C.,
Eaton, S., Costa, A. S. H., Brookes, P. S., et al. (2014)
Ischaemic accumulation of succinate controls reper-
fusion injury through mitochondrial ROS, Nature,
515, 431-435, doi:10.1038/nature13909.
169. Gross, C. C., Schulte-Mecklenbeck, A., Klinsing, S.,
Posevitz-Fejfar, A., Wiendl, H., and Klotz, L. (2016)
Dimethyl fumarate treatment alters circulating
T helper cell subsets in multiple sclerosis, Neurol.
Neuroimmunol. Neuroinflamm., 3, e183, doi: 10.1212/
NXI.0000000000000183.
170. Kornberg, M. D., Bhargava, P., Kim, P. M., Putluri, V.,
Snowman, A.M., Putluri,N., Calabresi, P.A., and Sny-
der, S. H. (2018) Dimethyl fumarate targets GAPDH
and aerobic glycolysis to modulate immunity, Science,
360, 449-453, doi:10.1126/science.aan4665.
171. Mrowietz, U., Szepietowski, J. C., Loewe, R., van
de Kerkhof, P., Lamarca, R., Ocker, W. G., Tebbs,
V. M., and Pau-Charles, I. (2017) Efficacy and safe-
ty of LAS41008 (dimethyl fumarate) in adults with
moderate-to- severe chronic plaque psoriasis: a ran-
domized, double-
blind, Fumaderm((R)) - and place-
bo-controlled trial (BRIDGE), Br.J. Dermatol., 176, 615-
623, doi:10.1111/bjd.14947.
172. Han,R., Xiao,J., Zhai,H., and Hao,J. (2016) Dimeth-
yl fumarate attenuates experimental autoimmune
neuritis through the nuclear factor erythroid-
derived 2-related factor 2/hemoxygenase-1 path-
way by altering the balance of M1/M2 macrophages,
J. Neuroinflammation, 13, 97, doi: 10.1186/s12974-
016-0559-x.
173. Cross, A.H., Misko, T.P., Lin, R.F., Hickey, W.F., Trot-
ter, J. L., and Tilton, R.G. (1994) Aminoguanidine, an
inhibitor of inducible nitric oxide synthase, amelio-
rates experimental autoimmune encephalomyelitis
in SJL mice, J.Clin. Invest., 93, 2684-2690, doi:10.1172/
JCI117282.
174. Chatterjee, P.K., Patel, N.S., Sivarajah,A., Kvale, E.O.,
Dugo, L., Cuzzocrea, S., Brown, P. A., Stewart, K. N.,
Mota-Filipe,H., Britti,D., Yaqoob, M.M., and Thiemer-
mann, C. (2003) GW274150, a potent and highly se-
lective inhibitor of iNOS, reduces experimental renal
ischemia/reperfusion injury, Kidney Int., 63, 853-865,
doi:10.1046/j.1523-1755.2003.00802.x.
175. Dugo,L., Marzocco,S., Mazzon,E., Di Paola,R., Geno-
vese,T., Caputi, A.P., and Cuzzocrea,S. (2004) Effects
of GW274150, a novel and selective inhibitor of iNOS
activity, in acute lung inflammation, Br.J. Pharmacol.,
141, 979-987, doi:10.1038/sj.bjp.0705683.
176. Miret, J.J., Kirschmeier,P., Koyama,S., Zhu,M., Li, Y.Y.,
Naito,Y., Wu,M., Malladi, V.S., Huang,W., Walker,W.,
Palakurthi, S., Dranoff, G., Hammerman, P. S., Pecot,
C.V., Wong, K.K., and Akbay, E.A. (2019) Suppression
of myeloid cell arginase activity leads to therapeu-
tic response in a NSCLC mouse model by activating
anti-tumor Immunity, J. Immunother. Cancer, 7, 32,
doi:10.1186/s40425-019-0504-5.
177. Steggerda, S. M., Bennett, M. K., Chen, J., Ember-
ley,E., Huang,T., Janes, J.R., Li,W., MacKinnon, A.L.,
YURAKOVA et al.838
BIOCHEMISTRY (Moscow) Vol. 89 No. 5 2024
Makkouk, A., Marguier, G., Murray, P. J., Neou, S.,
Pan,A., Parlati,F., Rodriguez, M. L.M., Van de Velde,
L.A., Wang, T., Works, M., Zhang,J., Zhang,W., etal.
(2017) Inhibition of arginase by CB-1158 blocks my-
eloid cell-mediated immune suppression in the tu-
mor microenvironment, J.Immunother. Cancer, 5, 101,
doi:10.1186/s40425-017-0308-4.
178. Kono, M., Yoshida, N., Maeda, K., and Tsokos, G. C.
(2018) Transcriptional factor ICER promotes glu-
taminolysis and the generation of Th17 cells, Proc.
Natl. Acad. Sci. USA, 115, 2478-2483, doi: 10.1073/
pnas.1714717115.
179. Takahashi, S., Saegusa, J., Sendo, S., Okano, T.,
Akashi, K., Irino, Y., and Morinobu, A. (2017) Gluta-
minase1 plays a key role in the cell growth of fibro-
blast-like synoviocytes in rheumatoid arthritis, Arthri-
tis Res. Ther., 19, 76, doi:10.1186/s13075-017-1283-3.
180. Rodriguez, A.E., Ducker, G.S., Billingham, L.K., Marti-
nez, C.A., Mainolfi,N., Suri,V., Friedman,A., Manfre-
di, M.G., Weinberg, S.E., Rabinowitz, J.D., and Chan-
del, N. S. (2019) Serine metabolism supports macro-
phage IL-1beta production, Cell Metab., 29, 1003-1011,
doi:10.1016/j.cmet.2019.01.014.
181. Ravindran, R., Loebbermann, J., Nakaya, H. I.,
Khan,N., Ma,H., Gama,L., Machiah, D.K., Lawson,B.,
Hakimpour,P., Wang, Y.C., Li,S., Sharma,P., Kaufman,
R.J., Martinez,J., and Pulendran,B. (2016) Theamino
acid sensor GCN2 controls gut inflammation by inhib-
iting inflammasome activation, Nature, 531, 523-527,
doi:10.1038/nature17186.
182. Chen, G.Q., Gong, R.H., Yang, D.J., Zhang,G., Lu, A.P.,
Yan, S. C., Lin, S. H., and Bian, Z. X. (2017) Halofugi-
none dually regulates autophagic flux through nutri-
ent-sensing pathways in colorectal cancer, Cell Death
Dis., 8, e2789, doi:10.1038/cddis.2017.203.
183. Wang,M., Wang,K., Liao,X., Hu,H., Chen,L., Meng,L.,
Gao,W., and Li,Q. (2021) Carnitine palmitoyltransfer-
ase system: a new target for anti-inflammatory and
anticancer therapy? Front. Pharmacol., 12, 760581,
doi:10.3389/fphar.2021.760581.
184. Qiao, S., Lv, C., Tao, Y., Miao, Y., Zhu, Y., Zhang, W.,
Sun,D., Yun,X., Xia,Y., Wei,Z., and Dai,Y. (2020) Arc-
tigenin disrupts NLRP3 inflammasome assembly in
colonic macrophages via downregulating fatty acid
oxidation to prevent colitis-associated cancer, Cancer
Lett., 491, 162-179, doi:10.1016/j.canlet.2020.08.033.
185. Peng, S., Chen, D., Cai, J., Yuan, Z., Huang, B., Li, Y.,
Wang,H., Luo,Q., Kuang,Y., Liang,W., Liu,Z., Wang,Q.,
Cui,Y., Wang,H., and Liu,X. (2021) Enhancing cancer-
associated fibroblast fatty acid catabolism within a
metabolically challenging tumor microenvironment
drives colon cancer peritoneal metastasis, Mol. Oncol.,
15, 1391-1411, doi:10.1002/1878-0261.12917.
186. Esteves, P., Blanc, L., Celle, A., Dupin, I., Maurat, E.,
Amoedo, N., Cardouat, G., Ousova, O., Gales, L., Bell-
vert, F., Begueret, H., Thumerel, M., Dupuy, J. W.,
Desbenoit, N., Marthan, R., Girodet, P. O., Rossign-
ol, R., Berger, P., and Trian, T. (2021) Crucial role of
fatty acid oxidation in asthmatic bronchial smooth
muscle remodelling, Eur. Respir. J., 58, 2004252,
doi:10.1183/13993003.04252-2020.
187. Jiang, Z., Knudsen, N. H., Wang, G., Qiu, W., Naing,
Z. Z. C., Bai, Y., Ai, X., Lee, C. H., and Zhou, X. (2017)
Genetic control of fatty acid beta-oxidation in chronic
obstructive pulmonary disease, Am.J. Respir. Cell Mol.
Biol., 56, 738-748, doi:10.1165/rcmb.2016-0282OC.
Publishers Note. Pleiades Publishing remains
neutral with regard to jurisdictional claims in pub-
lished maps and institutional affiliations.