ArticlePDF AvailableLiterature Review

Therapeutic targeting of the mitochondrial one-carbon pathway: perspectives, pitfalls, and potential

Authors:

Abstract and Figures

Most of the drugs currently prescribed for cancer treatment are riddled with substantial side effects. In order to develop more effective and specific strategies to treat cancer, it is of importance to understand the biology of drug targets, particularly the newly emerging ones. A comprehensive evaluation of these targets will benefit drug development with increased likelihood for success in clinical trials. The folate-mediated one-carbon (1C) metabolism pathway has drawn renewed attention as it is often hyperactivated in cancer and inhibition of this pathway displays promise in developing anticancer treatment with fewer side effects. Here, we systematically review individual enzymes in the 1C pathway and their compartmentalization to mitochondria and cytosol. Based on these insight, we conclude that (1) except the known 1C targets (DHFR, GART, and TYMS), MTHFD2 emerges as good drug target, especially for treating hematopoietic cancers such as CLL, AML, and T-cell lymphoma; (2) SHMT2 and MTHFD1L are potential drug targets; and (3) MTHFD2L and ALDH1L2 should not be considered as drug targets. We highlight MTHFD2 as an excellent therapeutic target and SHMT2 as a complementary target based on structural/biochemical considerations and up-to-date inhibitor development, which underscores the perspectives of their therapeutic potential.
Content may be subject to copyright.
Oncogene
https://doi.org/10.1038/s41388-021-01695-8
REVIEW ARTICLE
Therapeutic targeting of the mitochondrial one-carbon pathway:
perspectives, pitfalls, and potential
Li Na Zhao 1Mikael Björklund2,3,4 Matias J. Caldez5Jie Zheng6Philipp Kaldis 1
Received: 23 October 2020 / Revised: 27 January 2021 / Accepted: 2 February 2021
© The Author(s), under exclusive licence to Springer Nature Limited 2021
Abstract
Most of the drugs currently prescribed for cancer treatment are riddled with substantial side effects. In order to develop more
effective and specic strategies to treat cancer, it is of importance to understand the biology of drug targets, particularly the
newly emerging ones. A comprehensive evaluation of these targets will benet drug development with increased likelihood
for success in clinical trials. The folate-mediated one-carbon (1C) metabolism pathway has drawn renewed attention as it is
often hyperactivated in cancer and inhibition of this pathway displays promise in developing anticancer treatment with fewer
side effects. Here, we systematically review individual enzymes in the 1C pathway and their compartmentalization to
mitochondria and cytosol. Based on these insight, we conclude that (1) except the known 1C targets (DHFR, GART, and
TYMS), MTHFD2 emerges as good drug target, especially for treating hematopoietic cancers such as CLL, AML, and T-cell
lymphoma; (2) SHMT2 and MTHFD1L are potential drug targets; and (3) MTHFD2L and ALDH1L2 should not be
considered as drug targets. We highlight MTHFD2 as an excellent therapeutic target and SHMT2 as a complementary target
based on structural/biochemical considerations and up-to-date inhibitor development, which underscores the perspectives
of their therapeutic potential.
Introduction
Historically, drugs to treat cancer patients relied on the fact
that cancer cells proliferate faster than most healthy cells in
our body. This strategy is still valid and is being pursued
around the world by academic and pharmaceutical labora-
tories. In the last decade, our interest in metabolism has
been revived by the nding that cancer cells remodel their
metabolism and that they can choose from more metabolic
pathways than healthy cells [1,2]. This has led to the
hypothesis that cancer cells can exploit many metabolic
pathways to fulll their metabolic needs, which is generally
not the case for healthy cells [3,4]. The latest reviews on
metabolic exibility exhibited by cancer cells in their pro-
gression, notably growth, invasion, and drug evasion, dis-
cussed this in details [3,4], which highlights diverse factors
responsible for remodeling of cancer metabolism. Although
this creates complications, this may provide an Achilles
heel for cancer cells that could be exploited therapeutically.
Here we are considering the 1C pathway as a major avenue
to target cancer therapeutically. Nevertheless, targeting
metabolic pathways is not necessarily straightforward since
these pathways are also being used by healthy cells.
Therefore, we have to focus on metabolic pathways that are
either redundant or not usually used in healthy cells but
essential in cancer cells. Many different metabolic path-
ways, such as serine and glycine metabolism [5,6], gluta-
mine metabolism [37], and altered glycolysis [810], are
*Li Na Zhao
lina.zhao@med.lu.se
*Philipp Kaldis
philipp.kaldis@med.lu.se
1Department of Clinical Sciences, Lund University,
Malmö, Sweden
2Zhejiang UniversityUniversity of Edinburgh (ZJU-UoE)
Institute, Haining, Zhejiang, PR China
32nd Afliated Hospital, Zhejiang University School of Medicine,
Hangzhou, Zhejiang, PR China
4Deanery of Biomedical Sciences, College of Medicine and
Veterinary Medicine, University of Edinburgh, Edinburgh, UK
5Laboratory of Host Defense, The World Premier International
Research Center Initiative (WPI) Immunology Frontier Research
Center (IFReC), Osaka University, Osaka, Japan
6School of Information Science and Technology, Shanghai Tech
University, Shanghai, PR China
1234567890();,:
1234567890();,:
being pursued therapeutically. In this review, we are
focusing on the one-carbon (1C) pathway, which starts with
folate and provides 1C units (methyl groups) for DNA
methylation as well as for a variety of anabolic pathways
including DNA, RNA, and amino acid synthesis (Fig. 1).
One feature of this pathway is that it is compartmentalized
to both mitochondria and cytosol.
Due to its essential roles in nucleic acid biosynthesis and
in proliferation, folate has become the basis for a class of
drugs (known as antifolates) that antagonise its action.
Antifolate, as an antimetabolite, has been used in the clinic
since 1948 [11]. Antifolate has been further used for appli-
cations like antibiotics and chemotherapeutics [1214].
However, these antifolate drugs are generally toxic to normal
cells due to: (a) the targets being present in both normal and
cancer cells; and (b) low specicity as they target multiple
enzymes and block both DNA and RNA synthesis. The side
effects include hair loss, bone marrow toxicity, and cardiac
anomalies, which can be life-threatening [1517]. Given the
lack of safe therapies for many cancer patients, and the
importance of folate in biosynthesis, we aim to review the
folate-mediated 1C pathway in depth to provide new per-
spectives for cancer treatment with reduced side effects.
Folate, also known as folic acid, is a generic term for
water-soluble vitamin B. It can be reduced to dihydrofolate
(DHF) and then to tetrahydrofolate (THF) by dihydrofolate
reductase (DHRF), which sequentially transfer hydrogens
from NADPH to folic acid in the positions 58 (Fig. 2). The
chemical structure of folate consists of a 2-amino-4-
hydroxy-pteridine ring, para-aminobenzoic acid (pABA),
and polyglutamate. The 1C units are covalently linked to
two nitrogen atoms from the pteridine ring and the pABA
moiety at the 5 and 10 positions (Fig. 2). The pteridine ring
can be reduced or oxidized, and different oxidation states of
THF enable it to carry activated 1C units to be readily
interconverted within the cell [18].
The folate-mediated 1C metabolism is ubiquitous in
eukaryotic cells and organs. Unlike bacteria, yeast, and plants,
animals are not able to synthesize folate and have to take it up
from food. Insufcient folate will lead to complications,
notably neural tube defects [19]andfolate-deciency anemia
[20]. Folate is an essential cofactor in the mammalian 1C
metabolic network (Fig. 1), which comprises two inter-
dependent biosynthetic pathways that are compartmentalized
to the cytoplasm and mitochondria. The 1C network supports
multiple vital physiological processes including (i) biosynth-
esis (purine and thymidine) and homocysteine methylation in
the cytoplasm, (ii) protein synthesis (formylmethionyl-trans-
fer RNA; tRNA) and amino acid homeostasis (serine, glycine,
Fig. 1 Compartmentalization of
the mammalian one-carbon
metabolism to mitochondria and
cytosol. THF: tetrahydrofolate;
DHFR: dihydrofolate reductase;
SAM: S-adenosylmethionine;
TYMS: thymidylate synthetase;
MFT: mitochondrial folate
transporter.
Fig. 2 Chemical structure of folate. Folate consists of three distinct
folate moieties: pterin (red; 2-amino-4-hydroxy-pteridine heterocyclic
ring), pABA (orange; p-aminobenzoyl group), and ployglutamate
(black).
L. N. Zhao et al.
and methionine), as well as (iii) catabolism (choline, purines,
and histidine) in mitochondria [21,22]. The fundamental
nature of 1C metabolism compartmentalization is an evolu-
tionary strategy to enable cellular function properly by opti-
mizing spatial organization of metabolic feats [23] and each
compartment harbors enzyme complexes with specic activ-
ities to support cellular processes. Understanding the specic
activities of these enzymes in each compartment, highlights
new opportunities for selective intervention of certain
enzymes activities.
1C units (essentially methyl groups) use THF as a carrier.
THF accepts the 1C units derived from histidine, serine, and
formate in the cytosol. Namely, the 5-formimino-THF gen-
erated by histidine degradation is converted to 5,10-methy-
lene-THF by the formimidoyltransferase cyclodeaminase.
Serines usually come from de novo serine synthesis [24]or
direct through dietary intake. SHMT1/2 transfers the Cβfrom
serine to THF. Furthermore, serine, but not glycine, supplies
the 1C pool for cancer-cell proliferation [25].
The typical chemical reactions involving 1C units are
CH2-THF dehydrogenase (D), methenyl-THF (CH+-THF)
cyclohydrolase (C), and CHO-THF synthetase (S). Usually,
the trifunctional methylenetetrahydrofolate dehydrogenase,
cyclohydrolase and formyltetrahydrofolate synthetase 1
(MTHFD1) will carry out the D,C, and Sactivities in the
cytoplasm. The methylenetetrahydrofolate dehydrogenase
2-like protein (MTHFD2L) will carry out the Dand C
activities in mitochondria in normal cells and is expressed at
low levels in embryonic development [26]. During
embryonic development, MTHFD2 carries out the Dand C
activities in mitochondria. It is expressed in both embryonic
and tumor cells but at low levels in normal cells.
MTHFD1L will carry out the Sactivities in mitochondria.
Facilitated by rapid transportation of serine and exchange of
glycine, mitochondria become the primary source of for-
mate as 1C units for cytoplasmic metabolism [18,27].
In normal cells, the cytosol operates at near-neutral pH
(7.07.4) in most tissues and the pH is tightly regulated by
ion transport proteins [28,29], while in the mitochondrial
matrix a more alkaline pH in the range from 7.7 to 7.9 has
been observed. Cancer cells are characterized by altered
metabolism and inammation, leading to an increased
intracellular pH (7.37.6) and a decreased extracellular pH
(6.87.0 versus ~7.4 for normal cells) [30,31]. The altered
pH promotes a favorable microenvironment for cancer-cell
growth, proliferation, and invasion [32]. Since the cytosolic
and mitochondrial 1C pathways are parallel and the key
enzymes share high sequence identity (an adaptation of the
isoforms to the respective cellular environments) with
nearly the same catalytic properties, this signicantly
complicates drug development in terms of specicity, but
also provides opportunities as there are ne differences
between the enzymes in the cytosol and the mitochondria.
The mitochondrial 1C pathway is consistently upregu-
lated in cancer cells [33]. The hyperactivated 1C metabo-
lism has been shown to play a signicant role in cancer-cell
proliferation and malignancy [34]. Specically, MTHFD2
was found to be the most commonly upregulated both at
protein and mRNA levels in various types of cancers
[8,35]; MTHFD1L is overexpressed in colorectal cancer
[36], bladder cancer [37], and hepatocellular carcinoma
[38]; SHMT2 overexpression is observed in various cancers
and the knockdown of SHMT2 yields promising results in
suppressing cancer-cell proliferation and invasion [39]. The
hyperactivity of the 1C pathway partially reects the ana-
bolic metabolism facilitated by 1C units generated from
mitochondria and transported to the cytosol, which sustain
tumorigenesis [33]. The 1C units deprivation resulting from
serine restriction or de novo serine synthesis inhibition can
be an effective strategy against tumorigenesis [5].
Key enzymes in the 1C pathway
In order to review the potential of the 1C pathway in cancer
therapy, we need to introduce the enzymes that make up the
pathway. The focus is on the enzymatic activities of these
proteins and their structural arrangement to highlight the
potential success of chemical inhibitors.
There are ve key enzymes [SHMT2 (EC:2.1.2.1),
MTHFD2 (EC:1.5.1.15), MTHFD2L (EC:1.5.1.15), MTHF
D1L (EC:6.3.4.3), and ALDH1L2 (EC:1.5.1.6)] involved in
the mitochondrial 1C pathway (see Fig. 1), and with the
exception of MTHFD2L, all are upregulated in cancer
[34,38,40,41]. To be considered as potential targets for
selective therapeutic intervention, it is important to under-
stand the fundamental characteristics of these genes and the
encoded proteins. It is also important to compare the detailed
catalytic properties of these enzymes with their closely
related cytosolic counterparts. For example, cytosolic
SHMT1 is an isoenzyme of the mitochondrial SHMT2, and
the cytosolic ALDH1L1 is related to the mitochondrial
ALDH1L2. The detailed chemical reactions are depicted in
Fig. 3. Another issue that needs consideration is the use of
cofactors, which can determine the function, activity, and
specicity of the enzymes. MTHFD1 uses NADP+as
cofactor, while MTHFD2 and MTHFD2L can use either
NAD+or NADP+but with low efciency when using
NADP+. The origin of the cofactor preference of NAD+or
NADP+is suggested to promote a more thermo-
dynamically favorable pathway to balance the pools of 10-
formyl-THF during development[42]. Sequence and
structure comparison of MTHFD1 and MTHFD2 indicates
that Ser197 is a key factor in binding 2-phosphate of
NADP+while Arg233 is important to bind the phosphate.
The Ser to Arg233 switch from MTHFD1 to MTHFD2
Therapeutic targeting of the mitochondrial one-carbon pathway: perspectives, pitfalls, and potential
could enable the cofactor preference from NADP+to
NAD+. The mitochondrial 1C metabolism, both for produ-
cing and then transferring 1C units to the cytosol, and for
generating glycine and NADPH, is usually studied as a
whole since all chemical processes are folate-mediated.
Folates, in general, share one common structure comprising
three moieties (see Fig. 2): the pteridine ring, which can be
reduced or oxidized; the pABA linker together with the
pteridine ring coordinates the 1C units; and the poly-
glutamate tail, which enhances substrate binding and enables
its retention within mitochondria. Below we follow the ow
of folate through the 1C pathway to review SHMT2,
MTHFD2, MTHFD2L, MTHFD1L, and ALDH1L2.
SHMT2: serine hydroxymethyltransferase 2
The Shmt2 gene contains 11 exons and 10 introns [43].
Alternative splicing results in three conrmed transcript var-
iants. They are the mitochondrial Shmt2, a cytosolic Shmt2
(Shmt2α), which lacks exon 1 that encodes the mitochondrial
targeting peptide, and little is known about the third isoform
which is missing residues 199208 (199PKTGLIDYNQ208).
SHMT2 catalyzes the rst pivotal reaction in the mito-
chondrial 1C metabolism pathway, by reversibly interconvert-
ingserinetoglycineandbytransferringthehydroxymethyl
group (Cβof serine) to THF thereby generating 5,10-methy-
lenetetrahydrofolate (5,10-CH2-THF) [44]. It functions as a
Fig. 3 Reaction scheme of key
enzymes in the 1C pathway.
aDHFR catalyzes folic acid
reduction to DHF; bDHF is then
futher processed to THF by
DHFR; cSHMT converts serine
into glycine; dTYMS converts
dUMP to dTMP; and eMTHFD2
carries out the
methylenetetrahydrofolate
dehydrogease and cyclohydrolase
activities.
L. N. Zhao et al.
tetramer in a pyridoxal-5-phosphate (PLP)-dependent manner,
and PLP, the active form of vitamin B6, is responsible for the
oligomeric state of human SHMT2 and SHMT1 in mito-
chondria and cytosol, respectively [45]. However, SHMT1 can
exist in solution as a tetramer while SHMT2 undergoes a
dimer-to-tetramer transition in a PLP-dependent manner [45].
PLP binds covalently to Lys-280, the succinylation at Lys-280
prevents PLP binding, and desuccinylation by SIRT5 restores
the activity and promotes cell proliferation [46].
SHMT2 generates the major source of activated 1C units
for the de novo purine and pyrimidine synthesis in mito-
chondria. It is important for the translation of the mito-
chondrial respiratory complex by providing methyl donors
to produce the taurinomethyluridine base at the wobble
position of selected mitochondrial tRNAs to maintain the
tRNA pool [47,48]. The loss of SHMT2 leads to impaired
expression of respiratory chain enzymes [47].
The cytosolic isoforms (both SHMT1 and SHMT2α)can
translocate to the nucleus in a small ubiquitin-like modier
(SUMO)-dependent manner. They assemble the metabolic
complex at sites of DNA replication to sustain de novo thy-
midylate synthesis and to promote cell proliferation during S
and G2/M phases [49,50]. Particularly, Ubc13-mediated
ubiquitination is required for SHMT1 nuclear export and its
increased stability within the nucleus, whereas Ubc9-
mediated modication with SUMO2/3 will result in its
nuclear degradation [51]. Both ubiquitin and SUMO mod-
ication compete for the same lysine residue and thus
determine the localization of SHMT1 in the nucleus [51].
SHMT2αis functionally redundant with SHMT1 in the
nuclear de novo thymidylate synthesis [49].
SHMT1 levels are cell-cycle-regulated since it undergoes
SUMO modication (Ub-SHMT1), which leads to its
degradation and to nuclear translocation in Sphase [51].
SHMT2 is expressed at a constant level throughout the cell
cycle to maintain intracellular glycine levels [51]. However,
elevated expression of SHMT2 is observed during tumor cell
proliferation and is reected by increased glycine accumu-
lation, which is also associated with poor prognosis and high
mortality in various cancers [41,52,53]. The disruption of
the mouse Shmt2 gene confers fetal liver-specic down-
regulation of the 1C pathway resulting in mitochondrial
respiration defects and growth retardation [54]. Since
SHMT2 generates the mitochondrial 10-formyl THF, which
is a cofactor for mitochondrial methionyl-tRNA for-
myltransferase to generate formylmethionyl-tRNAs (fMet-
tRNA), it is important to maintain the fMet-tRNA pool for
the translation of mitochondrially encoded proteins [54,55].
Because SHMT2 is important for embryonic development
and expression of mitochondrial-encoded genes [54,55],
targeting SHMT2 in normal cells is not desirable. However,
overexpression of SHMT2 in cancer cells promotes cell
proliferation [33]. For example, overexpression of SHMT2 is
associated with poor prognosis in intrahepatic cholangio-
carcinoma, and poor survival in patients with kidney cancer
[56,57]. Furthermore, downregulation of SHMT2 suppresses
tumorigenesis in human hepatocellular carcinoma [58].
Therapeutic targeting would require reduction of the SHMT2
activity by a weak inhibitor. There is no potent or specic
SHMT2 inhibitor available and the published inhibitors are
general inhibitors toward serine hydroxymethyltransferases,
which have not been tested in clinical trials so far [59]. These
general inhibitors include: (i) triazine antifolate (NSC127755),
which irreversibly inhibits SHMT2 and is toxic [60]; (ii)
Leucovorin (5formylTHF) inhibits both SHMT1 and
SHMT2 at low micromolar concentration and is easily con-
verted into other folic-acid derivatives, which hampers its
clinical use [61]; and (iii) Pyrazolopyran derivatives, which
preferentially inhibit SHMT1 [62].
MTHFD2: NAD+-dependent methylenetetrahydrofolate
dehydrogenase cyclohydrolase
The Mthfd2 gene contains 8 exons and 7 introns. Alter-
native splicing results in two isomers and four potential
isoforms. The identied two isomers are the mitochondrial
Mthfd2 and the cytosolic Mthfd2, which lacks exon 1
encoding the targeting peptide for its efcient translocation
into mitochondria.
MTHFD2 has emerged as a novel therapeutic target since it
is only expressed during embryonic development [40]andis
almost undetectable in adults. Importantly, MTHFD2 is
overexpressed in cancer cells, notably in acute myeloid leu-
kemia (AML) [63], pancreatic cancer [34], adenocarcinomas
of the colorectum and lung [64], colorectal cancer [65,66],
glioma [67], and renal cell carcinoma [68]. Overexpression is
associated with poor overall survival of patients suffering
from renal cell carcinoma [40,69]. The overexpression of
MTHFD2 supplies cancer cells with building blocks for pyr-
imidine and purine biosynthesis during rapid proliferation and
is broadly required for the growth of cancer cells [40]. Tar-
geting MTHFD2 will unlikely cause major side effects due to
its low expression in most adult tissue [35,66,6872]. Inhi-
bition of MTHFD2 should lead to NADPH and redox
homeostasis imbalances, and thus should suppress tumori-
genic proliferation and growth, as well as enhance cancer cell
death under hypoxia [66]. In addition, MTHFD2 is not folate
sensitive [73], if the drug targeting MTHFD2 is a folate
antagonist, then the potential side effects can be further miti-
gated by taking a folic-acid supplement [22]. The knockdown
of MTHFD2 causes a decrease in the expression of the cell-
cycle genes such as Ccna2,Mcm7,andSkp2, indicating that
this may interfere with cell-cycle progression [74].
The structure of MTHFD2 has been co-crystallized in
complex with NAD+, inorganic phosphate (Pi), and the inhi-
bitor LY345899, which provides a reliable template (5TC4.
Therapeutic targeting of the mitochondrial one-carbon pathway: perspectives, pitfalls, and potential
pdb) for starting structure-based drug discovery [75]. Subse-
quently, MTHFD2 in complex with the inhibitor DS44960156
(6JIB.pdb), Compound 1 (6JID.pdb), and Compound 18
(6KG2.pdb) as well as cofactors (NAD+and Pi) were solved at
2.25 Å, and these structures bear similarity with the rst
identied structure. In all four structures, the absolutely
required Mg2+is missing, and the last three structures are
deposited in the Protein Data Bank as dimers. Since MTHFD2
functions as dimer, we have reconstituted the active site of the
MTHFD2 in a homodimer complex with inorganic phosphate,
NAD+, and substrate using homology modeling, and per-
formed a preliminary QM/MM study (Fig. 4; unpublished
data). Our results indicate that without Mg2+, the inorganic
phosphate is drifting away, suggesting that Mg2+plays an
important role in stabilizing the complex. Monomeric human
bifunctional MTHFD2 consists of D and C domains respon-
sible for the Dand Cactivities (see Fig. 1). The D/C domain of
MTHFD2 is composed of two α/βstrands that assemble to
form a cleft. The walls of the cleft are lined with highly con-
served residues. NAD+and Piare bound along one wall, while
the substrate is bound in a deep and rather hydrophobic cleft at
the interface between the two domains. MTHFD2 functions as
a homodimer, and homodimerization occurs by antiparallel
interaction of two NAD+-binding domains. However, the
complete MTHFD2·NAD+·Mg2+·Pi·THF complex is hard to
crystalize, especially in the active state at which the reaction
takes place. This may be because the complex is unstable or
otherwise short-lived. Therefore, we have reconstructed the
MTHFD2·NAD+·Pi·THF complex using homology modeling,
and our simulation indicates that Mg2+plays an important role
in stabilizing the inorganic phosphate. More importantly, pos-
sibly two phosphates may be involved in the binding of Mg2+
(see Fig. 4) or two magnesium ions, which are further coor-
dinated by water molecules. Currently there is no experimental
conrmation for our predictions of the active site. Unlike
magesium ions which are smaller with a low electron density,
two phosphate ions may be easily detected by current experi-
mental technology. Hence, we propose the presence of two
magnesium ions that coordinate the phosphate. Our previous
work in the simulation of the system (see Fig. 5)withonlyone
phosphate group indicates that the Piis drifting away from the
NAD+, which further supports the necessity of having two
magnesium ions. Another important observation from our
simulation indicates that the exible loop from residue 281 and
285 is important in governing the ligand recruitment and
release. Currently we are progressing to validate our predictions
by quantum mechanics studies as the accuracy of the structures
is key to harness the therapeutic potential of MTHFD2.
MTHFD2 inhibitors have gained attention as an attrac-
tive option as cancer drugs [35,52,64]. The latest progress
includes: (i) carolacton: a natural product, which inhibits
both MTHFD1 and MTHFD2 at the low nM range, and
Fig. 4 Our computational model of
MTHFD2·NAD+·Pi·Mg2+·Pi·THF
system. Note that the Mg2+and
inorganic phosphate are shown as
spheres, THF and NAD+are shown
as sticks and colored white and
yellow for C atoms, respectively.
Fig. 5 Our computational
model of MTHFD2·NAD+
·P ·THF system without Mg+.
Note that the inorganic
phosphate is shown as sphere,
THF and NAD+are shown as
sticks and colored white and
blue for C atoms, respectively.
L. N. Zhao et al.
decreases the growth of cancer cell lines at mM con-
centrations. However, carolacton lacks specicity toward
MTHFD2 [76]; (2) LY345899: a folate analogue, is capable
of suppressing tumor growth in mouse xenograft models of
colorectal cancer but it inhibits both MTHFD2 (IC50:
663 nM) and MTHFD1 (IC50: 96 nM) at the same time, and
how much inhibition of the tumor growth is due to the
contribution of MTHFD2 inhibition is unclear. Therefore
LY345899 lacks specicity [66]; (3) DS44960156: a tri-
cyclic coumarin scaffold [77], shows selectivity for
MTHFD2 over MTHFD1 but with low potency (IC50:
1.6 μM); and (4) DS18561882, optimized based on
DS44960156, with >250-fold improved potency and
90-fold improved selectivity for MTHFD2 (6.3 nM) over
MTHFD1 (0.57 μΜ)[78]. DS18561882 seems a decent
drug candidate, however, considering that the preclinical
evaluation procedure and the experimental data revealed
limitations and deciencies as well as a need for further
clinical studies. Overall, more efforts are still needed for the
development of potent and specic MTHFD2 inhibitors.
The NAD+-dependent activities of MTHFD2 follows an
ordered kinetic mechanism, in which NAD+is recruited
rst followed by THF, and during the product release, 2,4-
CH+-THF is released before NADH [79]. The kcat value for
MTHFD2 in the presence of NAD+for CH2-H4PteGlu1and
CH2-H4PteGlu5is 12.4 s1and 15.4 s1, respectively, while
in the presence of NADP+,thekcat value is 1.5 s1and 6.4 s1
[80]. Based on this data, MTHFD2 displays a preference to
use NAD+as cofactor.
MTHFD2 has been described mainly as a mitochondrial
folate metabolic enzyme. However, recent studies suggest
that MTHFD2 has non-enzymatic (or uncharacterized)
functions in the nucleus where MTHFD2 may localize at
sites of DNA replication [81]. In addition, unlike other
metabolic enzymes, MTHFD2 possesses a remarkable short
half-life, which resembles enzymes in RNA metabolism,
cell cycle, or signaling proteins [82]. Furthermore, in nor-
mal tissue, MTHFD2 was reported to be transcribed but not
translated [83]. Understanding these non-enzymatic func-
tions as well as why MTHFD2 is turned over rapidly, is
important for optimal targeting of MTHFD2 in cancer cells.
MTHFD2L: NADP+-dependent
methylenetetrahydrofolate dehydrogenase
The Mthfd2l gene contains nine exons and eight introns.
Alternative splicing results in ve isomers due to skipping/
retaining of exons 2 and/or 8, which are observed in human
brain and placenta [84]. The ve isomers are the canonical
isoform 4 (full-length; 347 amino acids; skipped exon 2 and
exon 8); isoform 2 (1277 missing [encoding 70 amino
acids]), which retains exon 2 leading to the translation
start from the rst AUG codon in exon 1 and an early
termination, isoform 1 (158 missing; 289 amino acids;
retains exon 2) with translation from the AUG codon in
exon 3; isoforms 3 (158 missing, 282347 missing; 223
amino acids) and 5 (269310 missing; 305 amino acids)
both of which are skipping exon 2 and/or exon 8. While the
residues 120 are encoding for the targeting peptide for its
efcient translocation into mitochondria, only isoforms 4
and 5 include the targeting peptide and are able to be
imported into the mitochondria. However, isoform 5 is
missing the residues from 269 to 310, which may sig-
nicantly lower its afnity for binding the substrate. Since
retention of exon 2 is rare, the isomers 1 and 2 are expressed
at low levels and currently detailed studies of the isoforms 3
and 5 are missing. However, MTHFD2L is a bifunctional
enzyme with CH2-THF dehydrogenase and 5,10-methenyl-
THF cyclohydrolase activities and is expressed in both
embryonic tissues and adult tissues. It is believed to have
little importance in cancer development and no association
with proliferation has been reported [85].
MTHFD2L can use either NAD+or NADP+as cofac-
tors, and for mono- and poly-glutamylated substrates with
similar catalytic efciencies [26]. The kcat value for
MTHFD2L in the presence of NAD+for CH2-H4PteGlu1
and CH2-H4PteGlu5is 2.7 s1and 8.8 s1, respectively [15].
Using the 5,10-CH2-H4PteGlu1substrate under saturating
conditions, the dehydrogenase activity of MTHFD2L in the
presence of NAD+was 3.4-fold higher than the presence of
NADP+[80]. This data may indicate that MTHFD2L pre-
fers NAD+as a substrate.
MTHFD1L: methylenetetrahydrofolate
dehydrogenase (NADP+-dependent) 1-like protein
The current genomic annotation of the Mthfd1l gene shows
that it contains 31 exons and 30 distinct GT-AG introns, which
is different from the early characterization with 28 exons plus
one alternative exon [86]. Alternative splicing results in ve
mRNAs variants and one unspliced form. Five spliced
mRNAs encode two complete isoforms (with one missing
residue 276978) while the unspliced form is apparently
noncoding. MTHFD1L consists of the transit peptide (131)
and two major domains: an inactive methylene-THF dehy-
drogenase and cyclohydrolase N-terminal domain (32347)
and an active larger 10-CHO-THF synthetase C-terminal
domain (348977). Its 10-CHO-THF synthetase activity
complements MTHFD2L in normal (in embryonic tissue) and
transformed cells in the last step of the 1C pathway. It is
expressed at all stages of embryogenesis and in adults, parti-
cularly highly expressed in the neural tube, developing brain,
limb, and tail bud, as well as in the craniofacial structures. The
deletion of Mthfd1l gene is strongly associated with birth
defects characterized with an aberrant neural tube and cra-
niorachischisis [87]. Mthfd1l is transcriptionally activated by
Therapeutic targeting of the mitochondrial one-carbon pathway: perspectives, pitfalls, and potential
NRF2, overexpressed in multiple cancer types, and plays an
essential role in cancer-cell growth [38]. However, the func-
tion of MTHFD1L in normal tissues needs to be characterized
before evaluating the therapeutic potential of MTHFD1L in
cancer cells.
ALDH1L2: aldehyde dehydrogenase 1 family
member L2
The Aldh1l2 gene contains 23 exons and 22 introns [88].
Alternative splicing results in three isoformes including the
full-length isomer 1, isomer 2 is missing residues from 204 to
923,andisoform3ismissingresiduesfrom517to923.Since
the region 438923 is important for aldehyde dehydrogenase
activities, the uncharacterized isomer 2 and isoform 3 may be
catalytically inactive. The loss of ALDH1L2 impairs mito-
chondrial functions, leads to distorted mitochondrial mor-
phology, and is associated with accumulation of acylcarnitine
derivatives and Krebs cycle intermediates [89]. ALDH1L2 is
an integral part of mitochondria and plays an important role in
the energy balance of the cell. Because of these essential
functions, it is unlikely to be a useful therapeutic target.
Other therapeutic targets in the 1C pathway
DHFR (EC:1.5.1.3): dihydrofolate reductase
The Dhfr gene contains six exons and ve introns with strictly
conserved intron and exon boundaries. Alternative splicing
results in two isoforms of which one is the canonical one and
the other missing residues 152. DHFR is a key enzyme in
folate metabolism as it generates 5,10-methylene-THF for
the subsequent 1C metabolism [90]. Most importantly, it is
involved in the de novo mitochondrial thymidylate biosynth-
esis pathway and catalyzes de novo glycine and purine
synthesisaswellasDNAprecursorsynthesis[91].
Prior to the 1C mitochondrial pathway, tetrahydrofolic
acid is reduced from dihydrofolic acid (DHF) by DHFR
using NADPH as electron donor. DHFR is overexpressed,
notably in acute lymphoblastic leukemia and is associated
with poor survival [92]. DHFR as a drug target has been
widely studied as antibacterial, antifungal, antimalarial, and
anticancer agent, and it has been intensively studied for its
antifolate properties in cancer chemotherapy [93]. Current
DHFR inhibitors include: (a) pyrimethamine [94]: a folic-
acid antagonist, used as antimalarial drug but now mainly
used with leucovorin to treat the cystoisosporiasis, which is
potent for protozoal DHFRs but weak for mammalian
DHFRs [95]; (b) Proguanil [96]: used to treat and prevent
malaria; (c) trimethoprim [97]: an antibacterial drug com-
monly used in association with sulfonamides (altogether
known as trimethoprim-sulfamethoxazole or co-trimoxazole)
to reinforce its action and prevent development of drug
resistance. Proguanil is a potent inhibitor for bacterial
DHFRs and a weak inhibitor for mammalian DHFRs; and (d)
methotrexate (MTX): an analog of folic acid, used in the
clinical setting as a chemotherapy agent and immune system
suppressant. MTX inhibits DHFR with a high afnity by
forming an inactive ternary complex with DHFR and
NADPH through the binding of one or more glutamate
residues to form polyglutamate derivatives, thus reducing the
amount of THF required for the synthesis of pyrimidine,
purines, and several other amino acids [98,99]. MTX has
been used for leukemia and choriocarcinoma treatment.
However, MTX lacks specicity and has substantial and
undesirable cytotoxicity [100]. DHFR, as an established drug
target for many Food and Drug Administration-approved
drugs [101], would benet from further drug development in
order to reduce side effects.
Dhfr was altered in 4.4% of 91 cases from TCGA Pan-
Can 2018 dataset with deep deletion in ovarian and prostate
cancer and signicant mutations in uterine cancer. Forty-
one examined patient samples exhibit 40 Dhfr mutations
across 14 projects from GDC. We note that S119P/Y,
R77K, and E78G are within DHFR active center, which
may affect DHFR normal function and the drug efcacy.
N13S may affect the Met-20 function loop (residues 924)
of DHFR, for the other two function loop: the FG loop
(residues 116132) and GH loop (residues 142149), a
distant mutant K122N may affect DHFR function. One
extra factor will compound the difculty in predicting,
which mutation will further affect DHFR functions as a
network of dynamically coupled residues in DHFR collec-
tively promoting DHFR activities was suggested [102].
From the literature, we notice that D152V was reported to
cause megaloblastic anemia and cerebral folate deciency
resulting in severe neurologic disease [103].
From DepMap, we noticed that there are 553 (out of 789)
cancer-cell lines whose proliferation will be affected by the
loss of DHFR function. However, DHFR shows a selec-
tivity of 0.186 with lung cancer, leukemia, skin cancer,
breast cancer, ovarian cancer, and pancreatic cancer will
benet most from DHFR inhibition.
TYMS (EC:2.1.1.45): thymidylate synthetase (TS)
The Tyms gene contains 12 exons and is separated by 11
distinct introns. Alternative splicing results in three isoforms,
one of which is full length (313 amino acids), another one is
missing residues 152185 and expressed in normal and can-
cerous cells, while the third one is missing residues 69151
and is expressed only in cancerous tissues. In the presence of
serine and NADPH, mitochondrial SHMT2 and TYMS,
together with human mitochondrial DHFR, convert deoxyur-
idylate (dUMP) to deoxythymidylate (dTMP) and this
L. N. Zhao et al.
contributes to the de novo mitochondrial thymidylate bio-
synthesis pathway [91]. TS forms an essential step in DNA
synthesis and DNA repair [104]. Its expression level has been
used as a biomarker in predicting the outcome of treatment in
prostate cancer as its overexpression is associated with a poor
survival rate and aggressive tumor features [105].
Over the previous half a century, several inhibitors have
been developed to target TS including: (a) Raltitrexed
(ZD1694; tomudex) [106]: one of the strongest quinazoline
(folate-based) antimetabolite in use with an IC50 of 9 nM; (b)
Pemetrexed [12]: a folate antimetabolite, which inhibits TS,
DHFR, and glycinamide ribonucleotide transformylase; (c)
Nolatrexed (AG337): with an IC50 between 0.39 and 6.6 µM,
water-soluble, and lipophilic nonclassical folate analog with
non-glutamate-containing moieties. The lipophilicity leads to
a short retention time, which reduces the active time in
the cell. Nolatrexed is currently in a Phase 3 clinical
trial (NCT00012324) [107]; and (d) Plevitrexed (ZD9331;
AstraZeneca): a non-polyglutamated quinazoline antifolate
drug with specicity toward TS, currently is in a Phase 2
clinical trial (NCT00014690); (e) 1843U89 (GW 1843):
benzoquinazoline folate analogue specically targeting TS
[108]. Its liposomal formulation GS7904L is currently eval-
uated in the clinic [109]. It is noteworthy that the anti-
metabolite 5-uorouracil (5-FU), a pyrimidine antagonist,
will be anabolized to 5-uoro-2-deoxyuridylate (FdUMP),
and as a nucleotide analogue of dUMP will covalently bind
TS and form a ternary complex with THF to inactive this
enzyme, thereby inhibiting DNA synthesis [110].
Tyms was altered in more than 6% of 182 cases from
TCGA PanCan 2018 dataset with noticeable amplication in
esophagus, bladder, uterine, head and neck, lung, sarcoma,
pancreatic, and melanoma cancers. Sixty-one examined
patient samples exhibit involving 62 Tyms mutations across
22 projects from GDC. We note that D49N, D218E, M311L,
V79E, F80L, R78H, and S229R are not distant from active
center, which are likely to affect biochemical activities of
TYMS in terms of kcat or KMvalue and also affect the drug
efcacy if the inhibitor is targeting the active site. In addition,
A191V and P188H are located at the active loop 181197,
which may deactivate TYMS since this loopsipping is
important for substrate binding and closing of the active site
[111]. V3L and M190K were observed to stabilize the
inactive loop [112,113]. The other variants of TYMS are
important to evaluate for cancer chemotherapy since some
variants are important in determining 5-FU treatment and
cytotoxicity [114].
From DepMap, we noticed that there are 328 (out of 789)
cancer-cell lines whose proliferation will be affected by the
loss of TYMS function. However, TYMS shows a relatively
high selectivity of 0.766. Lung cancer, leukemia, skin cancer,
breast cancer, and pancreatic cancer are among the top cell
lines that are sensitive toward TYMS loss-of-function.
GARS-AIRS-GART (hmGART): glycinamide
ribonucleotide synthetase (GARS), aminoimidazole
ribonucleotide synthetase (AIRS), and glycinamide
ribonucleotide formyltransferase (GART)
The Gars-Airs-Gart gene (also abbreviated as Gart) encodes
two proteins: full-length trifunctional GARS-AIRS-GART,
composed of three activities: GARS (PurD, EC:6.3.4.13),
AIRS (PurM, EC:6.3.3.1), and GARTfase (PurN,
EC:2.1.2.2.). GARS-AIRS-GART is involved in the second,
third, and fth steps of de novo purine synthesis, and the
proximity of the three activities in one enzyme complex
and the exibility of these domains elevate the pathway
output, while the fourth step is catalyzed by phosphor-
ibosylformylglycineamide amidotransferase (FGARAT,
PurL). The other protein resulting from alternative splicing
is the monofunctional GARS protein [115]. The activity of
GART requires 10-formyltetrahydrofolate and has been the
subject of much research interest for antifolate drug devel-
opment [116]. Current inhibitors targeting GART include:
(a) Lometrexol ([6R]-dideazatetrahydrofolate, DDATHF): a
folate analog antimetabolite [117], currently in clinical trial
NCT00024310 (phase 1) and NCT00033722 (phase 2); (b)
Pelitrexol (AG 2037): developed by Pzer and has com-
pleted two phase 2 trials by 2006 in the US [118]; (c)
LY254155 and LY222306: differ from lometrexol in the
replacement of the phenylene moiety by a thiophene and a
furan, respectively, which may have clinical activity; (d)
Pemetrexed disodium hemipentahydrate (LY231514): can
inhibit GART with an IC50 of 65 nM, and inhibit TS and
DHFR at 1.3 and 7.2 nM, respectively, is being tested in
extensive clinical trials (several Phase 3 clinical trials both
active and completed) [119].
One hundred and seventy-nine examined patient samples
exhibit displaying 195 Gart mutations across 26 projects
from GDC. G157R will affect manganese and phosphate
binding thus compromising the catalytic activities. E190K/D
will directly affect the ATP binding afnity, and L193F/I will
slightly affect the KMof ATP. Detailed structural studies
have been done for the region 467794 [120]. For 10-
formyltetrahydrofolate binding regions (896899 and
947951), we note that R871C, E948G, D951V, and L899I
will affect the 10-formyltetrahydrofolate binding; for
5-phosphoribosylglycinamide binding region (818820 and
977980), K977N will signicantly affect biochemical
activities of GART in terms of kcat or KMvalue and also
affect the drug efcacy if the inhibitor is targeting the active
site. D951V mutation disables its ability to raise pKa for the
active site His residue and the frameshift in G924Vfs*23
needs to be studied carefully as it is in the proximity of the
active center. This data indicate that there are mutations in
GART that will affect drug binding and this needs to be
considered in drug development.
Therapeutic targeting of the mitochondrial one-carbon pathway: perspectives, pitfalls, and potential
From DepMap, we noticed that there are 184 (out of 789)
cancer-cell lines whose proliferation will be affected by the
loss of GART function. However, GART shows a relatively
high selectivity of 0.459. Lung cancer, leukemia, breast
cancer, colon/colorectal cancer, myeloma, lymphoma, and
gastric cancer are among the top cancer types that should be
sensitive toward GART loss-of-function.
We have described studies of the mutations of each gene
among different cancer cases, since even a single mutation
may have profound implications for cancer development and
treatment. Therefore, it is important to note that usually a set
of various mutations across multiple genes collectively
contributes to cancer pathology. Taken together, this sum-
mary of the genes in the 1C pathway provides a basis for us
to study in-depth the roles of these enzymes in proliferation,
differentiation, and metastasis. In addition, a deeper under-
standing of the 1C pathway is important for the rational
design of drug combinations in the treatment of cancer.
Discussion
Potential pitfalls in targeting folate-mediated
metabolic enzymes
In order to develop a safe and specicinhibitor,thereare
several key enzymes we need to consider besides the well-
established MTHFD2. First and foremost there is
MTHFD1, as a close analogue of MTHFD2. MTHFD1 is
expressed in normal tissues and most abundantly in the
liver [76]. Inhibition of MTHFD1 would most likely cause
unwanted side effects since (a) MTHFD1 deciency
results in a phenotype that includes severe combined
immunodeciency and megaloblastic anemia [121]; and
(b) MTHFD1 inhibition disrupts the nuclear metabolite
composition and alters gene expression, which impairs
fast-growing cells [122]. Meanwhile, MTHFD2L, sharing
6065% identity and the highly conserved substrate and
cofactor binding pocket with MTHFD2, is expressed in all
tissues examined, with the highest expression levels in
brain and lung in both humans and rodents [55,57].
Unfortunately, there are no studies whether inhibitors of
MTHFD2 or MTHFD1 affect the activity of MTHFD2L.
Therefore, a selective inhibitor for MTHFD2 without
activity against MTHFD1 and MTHFD2L would be
desirable as a lead for drug discovery to warrant successful
clinical investigations. As discussed previously, the
cofactor (NAD+) pocket and the substrate (CH2-THF)
pocket of MTHFD2 are not specic enough as NAD(P)
+-dependent enzymes are ubiquitous in metabolism and
other cellular processes [51]. This substrate (folate) pocket
is also present in enzymes of the THF-mediated 1C
metabolic pathway, which has been the basis for the
antifolate drug development [123]. Therefore inhibitors
that target either pocket will likely result in undesired side
effects. Hence, we are proposing to exploit a novel inte-
grated pocket, different from the current cofactor and
substrate clefts, which is formed at the interface between
twoMTHFD2monomerswhentheydimerize(manuscript
in preparation), which could serve as an allosteric site for
the structure-based drug discovery.
Studies have shown that MTHFD2 knockdown only
inhibits tumor growth but is not sufcient to kill cancer cells
[64]. This could be due to incomplete silencing of MTHFD2
or due to compensation by other enzymes. Therefore, it is
important to analyse CRISPR knockout cells as well as
consider additional enzymes of the 1C pathway as combi-
natorial treatments to enhance efciency. This has been done
in the DepMap project and the proliferation of only four
cancer-cell lines is affected by Mthfd2 KO (see above).
SHMT2 is such a potential candidate for combinatorial
treatment. As discussed earlier, SHMT2 is overexpressed in
multiple cancer lines similar as MTHFD2 and the knock-
down of SHMT2 was found to suppress tumorigenesis in
human hepatocellular carcinoma [58]. As a potential ther-
apeutic target in combination with MTHFD2, weak inhibi-
tion of SHMT2 could be desirable to reduce its activity and
to enhance MTHFD2 inhibition. Overall, this would provide
a new perspective in the anticancer strategy targeting the 1C
pathway.
Another avenue for intervention is at the transcriptional
level. MYC binds directly at Mthfd2,Mthfd1l, and Shmt2
promoters [124], and c-myc also targets other genes such as
Phgdh,Psph,Slc19a1,Dhfr,Tyms,Gart,Shmt1,Mthfd1,
Fgps, and Gcsh that are involved in serine, folate, and
glycine metabolism [125,126]. Since myc activation cor-
relates or is fundamentally responsible for the over-
expression of Mthfd2,Mthfd1l, and Shmt2 genes [124],
targeting myc activation may be the last resort but is riddled
with difculties. This is mostly due to the fact that it is
difcult to target transcription factors although for myc,
there has been some progress with ONCO-myc [127].
Better than MTX?
MTX is an old schoolbut mainstay drug that has been used
in the clinic since 1947 to treat cancer [128]. There are many
reasons why this drug is used, one main reason is that it was
less toxic than the then-current anticancer regimens but MTX
causes a plethora of side effects including haematopoietic- and
hepatic-toxicity, dizziness, drowsiness, etc [129131]. It is
clear that MTX exerts its efcacy on DHFR and TS from the
1C pathway; however, off-target(s) of MTX are not well
known especially since it acts as anti-inammatory agent
[132]. Furthermore, it is known that many cancers become
quickly resistant to treatment with MTX, which is a major
L. N. Zhao et al.
issue [133136]. In summary, although MTX works, it is not a
perfect drug and improvements would be desirable. In com-
parison with MTX, we would expect that inhibitors of
MTHFD2 and/or SHMT2 have the following advantages:
since MTHFD2 is only expressed in cancer cells and not adult
cells, MTHFD2 inhibitors should be potent enough to elim-
inate cancer cells while sparing healthy cells, thereby reducing
the side effects.
Compartmentalization of the 1C pathway
The rationale for compartmentalization of the 1C pathway
has been described previously [22] but there are other
considerations of the compartmentalization. As an evolu-
tionary strategy, 1C metabolism spatially orchestrates into
distinct compartments to enable cellular function properly
(or more accurately) and to decrease interference with each
other [23]. Each compartment connes/concentrates spe-
cic enzyme complexes as well as their substrates and
products with specic microenvironments to optimize
subcellular biochemical processes. This compartmentali-
zation has all the advantages it needs for its function.
However, there are advantages and challenges in drug
development when targeting mitochondrial enzymes like
MTHFD2 and SHMT2. One big challenge to deliver drugs
to the mitochondria given that they have to cross two
membranes (outer and inner mitochondrial membrane) and
the intermembrane space. In addition, the chemical condi-
tions are different in mitochondria compared to the cytosol.
However, diverse delivery mechanisms have been proven
to be successful in delivering prodrugs or organelle-specic
chemical agents to its place of action [137]. As an effective
strategy to target mitochondria, the compartmentalization
could prolong drug retention in mitochondria, which may
enhance the drug efciency. At the same time, since there
is crosstalk between the cytoplasmic and mitochondrial 1C
pathway, one has to wonder whether targeting a cyto-
plasmic 1C enzyme could simultaneously inhibit the
mitochondrial 1C pathway. One possibility is the mito-
chondrial folate transporter (MFT encoded by the Slc25a32
gene), which imports THR from the cytosol to mitochon-
drial. CRISPR KO of Slc25A32 only affected proliferation
of four cancer-cell lines (DepMap), which could indicate
that inhibiting MFT in combination with either MTHFD2
or SHMT2 could be successful. In summary, compart-
mentalization can be viewed both as a drawback or
opportunity. Since the mitochondrial and cytoplasmic 1C
pathways can compensate for each other, this causes
additional problems for drug development. Nevertheless, if
the nodes where the two pathways interact are being con-
sidered, one could envision taking advantage of this by
affecting both the cytoplasmic and mitochondrial pathway
at the same time.
Open questions
Overall, there are uncharted territories in the mitochondrial
1C metabolism, which should be considered in future drug
development:
(1) For MTHFD2, one aspect that needs clarifying is
whether the transcriptional regulation that leads to
high expression of the mRNA and protein in the
proliferation stage of tumors are the same as in
embryonic development. Enzymes that are expressed
only in proliferating cells but not in adult tissues can
be considered as a potential anti-proliferation target.
(2) A previous study suggested that proliferation and
invasion are mutually exclusive [84]. This should be
considered when designing drugs targeting proliferat-
ing cells, since cell-cycle-arrested cells could become
more prone to invasion leading to metastasis [138].
Thus it is of paramount importance when inhibiting
cell division to investigate whether cancer cells arrest
and invade or whether they undergo cell death. The
latter would be obviously more desirable.
(3) A substantial pitfall in therapeutic inhibition of the
mitochondrial 1C pathway is that the loss of the
mitochondrial activity generating 1C units could be
compensated by cytosolic metabolism [139]. Both
cytosolic and mitochondrial pathways could sustain
tumorigenesis while the mitochondrial pathway is
essential in nutrient-poor conditions [139]. Thus the
interplay of the mitochondrial and cytosolic 1C
pathway needs to be investigated in more detail (as
discussed above).
(4) To what extent overexpressionmatters for metabolic
enzymes in general and particularly for MTHFD2, which
has a catalytic efciency of 12.4 s1?Doesover-
expressionreally elevate catalytic output of MTHFD2?
In addition, the half-life of proteins are generally in the
range of 0.535 h in dividing mammalian cells (~24 h
cell cycle) and ~43 h in nondividing cells [140].
Nevertheless, the time for protein synthesis takes
between 20 s to several minutes (average 2 min) [141],
which would indicate that protein synthesis may be more
important than the half-life of a protein. Therefore, a
quantitative understanding of protein levels, turn-over,
and catalytic activity is desirable.
(5) In general, a tumor biomarker should be very specic,
for example the FLT3 internal tandem duplication
(FLT3/ITD) has been reported to be a biomarker for the
response to MTHFD2 inhibition in AML [124].
However, for other tumor types, a simple and rapid
tests to predict whether a tumor will respond clinically
to inhibition of MTHFD2 would be very useful for
drug development.
Therapeutic targeting of the mitochondrial one-carbon pathway: perspectives, pitfalls, and potential
Concluding remarks
In this review, we are not only making an attempt to
investigate comprehensively the key enzymes in the 1C
mitochondrial pathway but also consider their counterparts
in the cytosol, based on conrmatory evidence and their
therapeutic perspectives. By integrating experimental data,
and literature information regarding the key enzymes in
mitochondrial 1C pathway in tumors and microorganisms,
we come to the following conclusions: (1) MTHFD2 is an
excellent therapeutic target; (2) SHMT2 is a good drug
target; (3) MTHFD1L is a potential drug target; and (4)
MTHFD2L and ALDH1L2 should not be considered for
drug development. Meanwhile, based on structural/bio-
chemical and up-to-date inhibitor development, we have
given special attention to MTHFD2 and SHMT2 as a
complementary target, which will deepen our understanding
of their therapeutic potential. In addition, we have provided
a brief review of other important enzymes in the 1C path-
way, which are well-established drug targets, some of which
are widely used in the clinic. Even for the established drugs
more efforts are needed to decrease the side effects. Atten-
tion shall also be given to less developed inhibitors, which
could be potential game changers in the future. Rather than
restricting our attention to gene overexpression,we have
focused on the functional importance in normal cells to
identify which should be therapeutically targeted.
Overall, we have discussed some important pitfalls and
possible ways to move forward in targeting the 1C meta-
bolism for cancer therapy. Since the 1C pathway is a
dynamic, interdependent network, the chemical steps cata-
lyzed by the 1C enzymes are coupled to the ow of the
folic-acid derivatives in this entire network. It is important
to systematically study the 1C pathway to capture the
dynamics of these biochemical processes, which will pro-
vide a better understanding of this pathway in terms of
overexpression as well as hyperactivity in cancer cells. This
may eventually open new therapeutic perspectives for
treating various cancers safely and efciently.
Acknowledgements LNZ was supported by an A*STAR International
Fellowship (AIF) from Singapore. This work is supported by the
IngaBritt och Arne Lundbergs Forskningsstiftelse LU2020-0013; PK
is supported by the Faculty of Medicine, Lund University; the Swedish
Foundation for Strategic Research Dnr IRC15-0067, and Swedish
Research Council, Strategic Research Area EXODIAB, Dnr
20091039. LNZ would like to thank Prof. Chew Lock Yue (NTU)
and Prof. Ulf Ryde (LU) for comments on the manuscript.
Compliance with ethical standards
Conict of interest The authors declare no competing interests.
Publishers note Springer Nature remains neutral with regard to
jurisdictional claims in published maps and institutional afliations.
References
1. Pavlova NN, Thompson CB. The emerging hallmarks of cancer
metabolism. Cell Metab. 2016;23:2747.
2. Vander Heiden MG. Targeting cancer metabolism: a therapeutic
window opens. Nat Rev Drug Discov. 2011;10:67184.
3. McGuirk S, Audet-Delage Y, St-Pierre J. Metabolic tness and
plasticity in cancer progression. Trends Cancer. 2020;6:4961.
4. Kreuzaler P, Panina Y, Segal J, Yuneva M. Adapt and conquer:
metabolic exibility in cancer growth, invasion and evasion. Mol
Metab. 2020;33:83101.
5. Amelio I, Cutruzzolá F, Antonov A, Agostini M, Melino G.
Serine and glycine metabolism in cancer. Trends Biochem Sci.
2014;39:1918.
6. Locasale JW. Serine, glycine and one-carbon units: cancer
metabolism in full circle. Nat Rev Cancer. 2013;13:57283.
7. Choi Y-K, Park K-G. Targeting glutamine metabolism for cancer
treatment. Biomol Ther. 2018;26:1928.
8. Fadaka A, Ajiboye B, Ojo O, Adewale O, Olayide I, Emuo-
whochere R. Biology of glucose metabolization in cancer cells. J
Oncological Sci. 2017;3:4551.
9. Shiratori R, Furuichi K, Yamaguchi M, Miyazaki N, Aoki H,
Chibana H, et al. Glycolytic suppression dramatically changes
the intracellular metabolic prole of multiple cancer cell lines
in a mitochondrial metabolism-dependent manner. Sci Rep.
2019;9:18699.
10. Lin J, Xia L, Liang J, Han Y, Wang H, Oyang L, et al. The roles
of glucose metabolic reprogramming in chemo- and radio-
resistance. J Exp Clin Cancer Res. 2019;38:218.
11. Farber S, Diamond LK, Mercer RD, Sylvester RF, Wolff JA.
Temporary remissions in acute leukemia in children produced by
folic acid antagonist, 4-aminopteroyl-glutamic acid (ami-
nopterin). N Engl J Med. 1948;238:78793.
12. Chattopadhyay S, Moran RG, Goldman ID. Pemetrexed: bio-
chemical and cellular pharmacology, mechanisms, and clinical
applications. Mol Cancer Ther. 2007;6:40417.
13. Estrada A, Wright DL, Anderson AC. Antibacterial antifolates:
from development through resistance to the next generation.
Cold Spring Harb Perspect Med. 2016;6. https://doi.org/10.1101/
cshperspect.a028324.
14. Purcell WT, Ettinger DS. Novel antifolate drugs. Curr Oncol
Rep. 2003;5:11425.
15. Carver JR, Szalda D, Ky B. Asymptomatic cardiac toxicity in
long-term cancer survivors: dening the population and recom-
mendations for surveillance. Semin Oncol. 2013;40:22938.
16. Yeh Edward TH, Tong Ann T, Lenihan Daniel J, Wamique
YusufS, Joseph Swafford, Christopher Champion, et al. Car-
diovascular complications of cancer therapy. Circulation.
2004;109:312231.
17. Wang Y, Probin V, Zhou D. Cancer therapy-induced residual
bone marrow injury-mechanisms of induction and implication
for therapy. Curr Cancer Ther Rev. 2006;2:2719.
18. Appling DR. Compartmentation of folate-mediated one-carbon
metabolism in eukaryotes. FASEB J.1991;5(12):264551.
19. Toriello HV. Folic acid and neural tube defects. Genet Med.
2005;7:2834.
20. Kujovich JL. Evaluation of anemia. Obstet Gynecol Clin North
Am. 2016;43:24764.
21. Fox JT, Stover PJ. Folate-mediated one-carbon metabolism.
Vitam Horm. 2008;79:144.
22. Ducker GS, Rabinowitz JD. One-carbon metabolism in health
and disease. Cell Metab. 2017;25:2742.
23. Agapakis CM, Boyle PM, Silver PA. Natural strategies for the
spatial optimization of metabolism in synthetic biology. Nat
Chem Biol. 2012;8:52735.
L. N. Zhao et al.
24. Kit S. The biosynthesis of free glycine and serine by tumors.
Cancer Res. 1955;15:7158.
25. Labuschagne CF, van den Broek NJF, Mackay GM, Vousden
KH, Maddocks ODK. Serine, but not glycine, supports one-
carbon metabolism and proliferation of cancer cells. Cell Rep.
2014;7:124858.
26. Shin M, Bryant JD, Momb J, Appling DR. Mitochondrial
MTHFD2L is a dual redox cofactor-specic methylenetetrahy-
drofolate dehydrogenase/methenyltetrahydrofolate cyclohy-
drolase expressed in both adult and embryonic tissues. J Biol
Chem. 2014;289:1550717.
27. Cybulski RL, Fisher RR. Intramitochondrial localization and
proposed metabolic signicance of serine transhydrox-
ymethylase. Biochemistry. 1976;15:31837.
28. Balut C, vandeVen M, Despa S, Lambrichts I, Ameloot M,
Steels P, et al. Measurement of cytosolic and mitochondrial pH
in living cells during reversible metabolic inhibition. Kidney Int.
2008;73:22632.
29. Boron WF. Regulation of intracellular pH. Adv Physiol Educ.
2004;28:16079.
30. Shirmanova MV, Druzhkova IN, Lukina MM, Matlashov ME,
Belousov VV, Snopova LB, et al. Intracellular pH imaging in
cancer cells in vitro and tumors in vivo using the new genetically
encoded sensor SypHer2. Biochim Biophys Acta (BBA) - Gen
Subj. 2015;1850:190511.
31. White KA, Grillo-Hill BK, Barber DL. Cancer cell behaviors
mediated by dysregulated pH dynamics at a glance. J Cell Sci.
2017;130:6639.
32. Hu Y, Li Y. Effect of low pH treatment on cell cycle and cell
growth. FASEB J. 2018;32:804.49.
33. Vazquez A, Tedeschi PM, Bertino JR. Overexpression of the
mitochondrial folate and glycineserine pathway: a new determinant
of methotrexate selectivity in tumors. Cancer Res. 2013;73:47882.
34. Noguchi K, Konno M, Koseki J, Nishida N, Kawamoto K,
Yamada D, et al. The mitochondrial one-carbon metabolic
pathway is associated with patient survival in pancreatic cancer.
Oncol Lett. 2018;16:182734.
35. Nilsson R, Jain M, Madhusudhan N, Sheppard NG, Strittmatter
L, Kampf C, et al. Metabolic enzyme expression highlights a key
role for MTHFD2 and the mitochondrial folate pathway in
cancer. Nat Commun. 2014;5:3128.
36. Agarwal S, Behring M, Hale K, Al Diffalha S, Wang K, Manne
U, et al. MTHFD1L, a folate cycle enzyme, is involved in pro-
gression of colorectal cancer. Transl Oncol. 2019;12:14617.
37. Eich M-L, Rodriguez Pena MDC, Chandrashekar DS, Chaux A,
Agarwal S, Gordetsky JB, et al. Expression and role of methy-
lenetetrahydrofolate dehydrogenase 1 like (MTHFD1L) in
bladder cancer. Transl Oncol. 2019;12:141624.
38. Lee D, Xu IM-J, Chiu DK-C, Lai RK-H, Tse AP-W, Lan Li L, et al.
Folate cycle enzyme MTHFD1L confers metabolic advantages in
hepatocellular carcinoma. J Clin Investig. 2017;127:185672.
39. Wu M, Wanggou S, Li X, Liu Q, Xie Y. Overexpression of
mitochondrial serine hydroxyl-methyltransferase 2 is associated
with poor prognosis and promotes cell proliferation and invasion
in gliomas. Onco Targets Ther. 2017;10:37818.
40. Tedeschi PM, Vazquez A, Kerrigan JE, Bertino JR. Mitochondrial
methylenetetrahydrofolate dehydrogenase (MTHFD2) over-
expression is associated with tumor cell proliferation and is a
novel target for drug development. Mol Cancer Res. 2015;13:
13616.
41. Liu Y, Yin C, Deng M-M, Wang Q, He X-Q, Li M-T, et al. High
expression of SHMT2 is correlated with tumor progression and
predicts poor prognosis in gastrointestinal tumors. Eur Rev Med
Pharm Sci. 2019;23:937992.
42. Di Pietro E, Sirois J, Tremblay ML, MacKenzie RE. Mitochon-
drial NAD-dependent methylenetetrahydrofolate dehydrogenase-
methenyltetrahydrofolate cyclohydrolase is essential for embryo-
nic development. Mol Cell Biol. 2002;22:415866.
43. Stover PJ, Chen LH, Suh JR, Stover DM, Keyomarsi K, Shane
B. Molecular cloning, characterization, and regulation of the
human mitochondrial serine hydroxymethyltransferase gene. J
Biol Chem. 1997;272:18428.
44. Florio R, di Salvo ML, Vivoli M, Contestabile R. Serine
hydroxymethyltransferase: a model enzyme for mechanistic,
structural, and evolutionary studies. Biochim Biophys Acta.
2011;1814:148996.
45. Giardina G, Brunotti P, Fiascarelli A, Cicalini A, Costa MGS,
Buckle AM, et al. How pyridoxal 5-phosphate differentially
regulates human cytosolic and mitochondrial serine hydro-
xymethyltransferase oligomeric state. FEBS J. 2015;282:122541.
46. Yang X, Wang Z, Li X, Liu B, Liu M, Liu L, et al. SHMT2
Desuccinylation by SIRT5 drives cancer cell proliferation.
Cancer Res. 2018;78:37286.
47. Morscher RJ, Ducker GS, Li SH-J, Mayer JA, Gitai Z, Sperl W,
et al. Mitochondrial translation requires folate-dependent tRNA
methylation. Nature. 2018;554:12832.
48. Lucas S, Chen G, Aras S, Wang J. Serine catabolism is essential
to maintain mitochondrial respiration in mammalian cells. Life
Sci Alliance. 2018;1:e201800036.
49. Anderson DD, Stover PJ. SHMT1 and SHMT2 are functionally
redundant in nuclear de novo thymidylate biosynthesis. PLoS
ONE. 2009;4:e5839.
50. Guiducci G, Paone A, Tramonti A, Giardina G, Rinaldo S,
Bouzidi A, et al. The moonlighting RNA-binding activity of
cytosolic serine hydroxymethyltransferase contributes to control
compartmentalization of serine metabolism. Nucleic Acids Res.
2019;47:424054.
51. Anderson DD, Eom JY, Stover PJ. Competition between sumoy-
lation and ubiquitination of serine hydroxymethyltransferase 1
determines its nuclear localization and its accumulation in the
nucleus. J Biol Chem. 2012;287:47909.
52. Jain M, Nilsson R, Sharma S, Madhusudhan N, Kitami T, Souza
AL, et al. Metabolite proling identies a key role for glycine in
rapid cancer cell proliferation. Science. 2012;336:10404.
53. Lee GY, Haverty PM, Li L, Kljavin NM, Bourgon R, Lee J, et al.
Comparative oncogenomics identies PSMB4 and SHMT2 as
potential cancer driver genes. Cancer Res. 2014;74:311426.
54. Tani H, Ohnishi S, Shitara H, Mito T, Yamaguchi M, Yone-
kawa H, et al. Mice decient in the Shmt2 gene have mito-
chondrial respiration defects and are embryonic lethal. Sci Rep.
2018;8:425.
55. Minton DR, Nam M, McLaughlin DJ, Shin J, Bayraktar EC,
Alvarez SW, et al. Serine catabolism by SHMT2 is required for
proper mitochondrial translation initiation and maintenance of
formylmethionyl-tRNAs. Mol Cell. 2018;69:61021.e5.
56. Ning S, Ma S, Saleh AQ, Guo L, Zhao Z, Chen Y. SHMT2
overexpression predicts poor prognosis in intrahepatic cho-
langiocarcinoma. Gastroenterol Res Pract. 2018;2018:4369253.
57. Wang H, Chong T, Li B-Y, Chen X-S, Zhen W-B. Evaluating
the clinical signicance of SHMT2 and its co-expressed gene in
human kidney cancer. Biol Res. 2020;53:46.
58. Woo CC, Chen WC, Teo XQ, Radda GK, Teck Hock Lee P.
Downregulating serine hydroxymethyltransferase 2 (SHMT2)
suppresses tumorigenesis in human hepatocellular carcinoma.
Oncotarget. 2016;7:5300517.
59. NonakaH,NakanishiY,KunoS,OtaT,MochidomeK,Saito
Y, et al. Design strategy for serine hydroxymethyltransferase
probes based on retro-aldol-type reaction. Nat Commun.
2019;10:876.
60. Snell K, Riches D. Effects of a triazine antifolate (NSC 127755)
on serine hydroxymethyltransferase in myeloma cells in culture.
Cancer Lett. 1989;44:21720.
Therapeutic targeting of the mitochondrial one-carbon pathway: perspectives, pitfalls, and potential
61. Stover P, Schirch V. 5-Formyltetrahydrofolate polyglutamates are
slow tight binding inhibitors of serine hydroxymethyltransferase. J
Biol Chem. 1991;266:154350.
62. Marani M, Paone A, Fiascarelli A, Macone A, Gargano M, Rinaldo
S, et al. A pyrazolopyran derivative preferentially inhibits the activity
of human cytosolic serine hydroxymethyltransferase and induces cell
death in lung cancer cells. Oncotarget. 2016;7:457083.
63. Gu Y, Si J, Xiao X, Tian Y, Yang S. miR-92a inhibits proliferation
and induces apoptosis by regulating methylenetetrahydrofolate
dehydrogenase 2 (MTHFD2) expression in acute myeloid leuke-
mia. Oncol Res. 2017;25:106979.
64. Koseki J, Konno M, Asai A, Colvin H, Kawamoto K, Nishida N,
et al. Enzymes of the one-carbon folate metabolism as anticancer
targets predicted by survival rate analysis. Sci Rep. 2018;8:303.
65. Yan Y, Zhang D, Lei T, Zhao C, Han J, Cui J, et al. MicroRNA-
33a-5p suppresses colorectal cancer cell growth by inhibiting
MTHFD2. Clin Exp Pharm Physiol. 2019;46:92836.
66. Ju H-Q, Lu Y-X, Chen D-L, Zuo Z-X, Liu Z-X, Wu Q-N, et al.
Modulation of redox homeostasis by inhibition of MTHFD2 in
colorectal cancer: mechanisms and therapeutic implications. J
Natl Cancer Inst. 2018. https://doi.org/10.1093/jnci/djy160.
67. Xu T, Zhang K, Shi J, Huang B, Wang X, Qian K, et al.
MicroRNA-940 inhibits glioma progression by blocking mito-
chondrial folate metabolism through targeting of MTHFD2. Am
J Cancer Res. 2019;9:25069.
68. Nishimura T, Nakata A, Chen X, Nishi K, Meguro-Horike M,
Sasaki S, et al. Cancer stem-like properties and getinib resistance
are dependent on purine synthetic metabolism mediated by the
mitochondrial enzyme MTHFD2. Oncogene. 2019;38:246481.
69. Lin H, Huang B, Wang H, Liu X, Hong Y, Qiu S, et al. MTHFD2
overexpression predicts poor prognosis in renal cell carcinoma and
is associated with cell proliferation and vimentin-modulated
migration and invasion. Cell Physiol Biochem. 2018;51:9911000.
70. Asai A, Koseki J, Konno M, Nishimura T, Gotoh N, Satoh T,
et al. Drug discovery of anticancer drugs targeting methylene-
tetrahydrofolate dehydrogenase 2. Heliyon. 2018;4:e01021.
71. Mejia NR, MacKenzie RE. NAD-dependent methylenetetrahy-
drofolate dehydrogenase is expressed by immortal cells. J Biol
Chem. 1985;260:1461620.
72. Mejia NR, MacKenzie RE. NAD-dependent methylenetetrahy-
drofolate dehydrogenase-methenyltetrahydrofolate cyclohydrolase
in transformed cells is a mitochondrial enzyme. Biochem Biophys
Res Commun. 1988;155:16.
73. Patel H, Pietro ED, MacKenzie RE. Mammalian broblasts
lacking mitochondrial NAD+-dependent methylenetetrahy-
drofolate dehydrogenase-cyclohydrolase are glycine auxotrophs.
J Biol Chem. 2003;278:1943641.
74. Yu C, Yang L, Cai M, Zhou F, Xiao S, Li Y, et al. Down-
regulation of MTHFD2 inhibits NSCLC progression by sup-
pressing cycle-related genes. J Cell Mol Med. 2020;24:156877.
75. Gustafsson R, Jemth A-S, Gustafsson NMS, Färnegårdh K,
Loseva O, Wiita E, et al. Crystal structure of the emerging cancer
target MTHFD2 in complex with a substrate-based inhibitor.
Cancer Res. 2017;77:93748.
76. Fu C, Sikandar A, Donner J, Zaburannyi N, Herrmann J, Reck
M, et al. The natural product carolacton inhibits folate-
dependent C1 metabolism by targeting FolD/MTHFD. Nat
Commun. 2017;8:1529.
77. Kawai J, Ota M, Ohki H, Toki T, Suzuki M, Shimada T, et al.
Structure-based design and synthesis of an isozyme-selective
MTHFD2 inhibitor with a tricyclic coumarin scaffold. ACS Med
Chem Lett. 2019;10:8938.
78. Kawai J, Toki T, Ota M, Inoue H, Takata Y, Asahi T, et al.
Discovery of a potent, selective, and orally available MTHFD2
inhibitor (DS18561882) with in vivo antitumor activity. J Med
Chem. 2019. https://doi.org/10.1021/acs.jmedchem.9b01113.
79. Rios-Orlandi EM, MacKenzie RE. The activities of the
NAD-dependent methylenetetrahydrofolate dehydrogenase-
methenyltetrahydrofolate cyclohydrolase from ascites tumor cells
are kinetically independent. J Biol Chem. 1988;263:46627.
80. Shin M, Momb J, Appling DR. Human mitochondrial MTHFD2
is a dual redox cofactor-specic methylenetetrahydrofolate
dehydrogenase/methenyltetrahydrofolate cyclohydrolase. Cancer
Metab. 2017;5:11.
81. Gustafsson Sheppard N, Jarl L, Mahadessian D, Strittmatter L,
Schmidt A, Madhusudan N, et al. The folate-coupled enzyme
MTHFD2 is a nuclear protein and promotes cell proliferation. Sci
Rep. 2015;5:15029.
82. Koufaris C, Nilsson R. Protein interaction and functional data
indicate MTHFD2 involvement in RNA processing and trans-
lation. Cancer Metab. 2018;6:12.
83. Kg P, Re M. NAD+-dependent methylenetetrahydrofolate
dehydrogenase-cyclohydrolase: detection of the mRNA in nor-
mal murine tissues and transcriptional regulation of the gene in
cell lines. Biochim Biophys Acta. 1993;1171:2817.
84. Bolusani S, Young BA, Cole NA, Tibbetts AS, Momb J, Bryant
JD, et al. Mammalian MTHFD2L encodes a mitochondrial
methylenetetrahydrofolate dehydrogenase isozyme expressed in
adult tissues. J Biol Chem. 2011;286:516674.
85. Nilsson R, Nicolaidou V, Koufaris C. Mitochondrial MTHFD
isozymes display distinct expression, regulation, and association
with cancer. Gene. 2019;716:144032.
86. Prasannan P, Pike S, Peng K, Shane B, Appling DR. Human
mitochondrial C1-tetrahydrofolate synthase: gene structure, tissue
distribution of the mRNA, and immunolocalization in Chinese
hamster ovary calls. J Biol Chem. 2003;278:4317887.
87. Momb J, Lewandowski JP, Bryant JD, Fitch R, Surman DR,
Vokes SA, et al. Deletion of Mthfd1l causes embryonic lethality
and neural tube and craniofacial defects in mice. PNAS.
2013;110:54954.
88. Krupenko NI, Dubard ME, Strickland KC, Moxley KM, Oleinik
NV, Krupenko SA. ALDH1L2 is the mitochondrial homolog of
10-formyltetrahydrofolate dehydrogenase. J Biol Chem. 2010;
285:2305663.
89. Sarret C, Ashkavand Z, Paules E, Dorboz I, Pediaditakis P, Sumner
S, et al. Deleterious mutations in ALDH1L2 suggest a novel cause
for neuro-ichthyotic syndrome. NPJ Genom Med. 2019;4:17.
90. Schober AF, Mathis AD, Ingle C, Park JO, Chen L, Rabinowitz
JD, et al. A two-enzyme adaptive unit within bacterial folate
metabolism. Cell Rep. 2019;27:335970.e7.
91. Anderson DD, Quintero CM, Stover PJ. Identication of a de
novo thymidylate biosynthesis pathway in mammalian mito-
chondria. PNAS. 2011;108:151638.
92. Organista-Nava J, mez-Gómez Y, Illades-Aguiar B, Rivera-
Ramírez AB, Saavedra-Herrera MV, Leyva-Vázquez MA. Over-
expression of dihydrofolate reductase is a factor of poor survival
in acute lymphoblastic leukemia. Oncol Lett. 2018;15:840511.
93. Raimondi MV, Randazzo O, La Franca M, Barone G, Vignoni E,
Rossi D, et al. DHFR inhibitors: reading the past for discovering
novel anticancer agents. Molecules. 2019;24:1140.
94. Raimondi MV, Randazzo O, La Franca M, Barone G, Vignoni E,
Rossi D. et al. DHFR Inhibitors: Reading the Past for Dis-
covering Novel Anticancer Agents. Molecules. 2019;24:1140
95. Schweitzer BI, Dicker AP, Bertino JR. Dihydrofolate reductase
as a therapeutic target. The FASEB J. 1990;4:244152.
96. Patel SN, Kain KC. Atovaquone/proguanil for the prophylaxis
and treatment of malaria. Expert Rev Anti Infect Ther.
2005;3:84961.
97. Wróbel A, Arciszewska K, Maliszewski D, Drozdowska D.
Trimethoprim and other nonclassical antifolates an excellent
template for searching modications of dihydrofolate reductase
enzyme inhibitors. J Antibiotics. 2020;73:527.
L. N. Zhao et al.
98. Srinivasan B, Skolnick J. Insights into the slow-onset tight-
binding inhibition of Escherichia coli dihydrofolate reductase:
detailed mechanistic characterization of pyrrolo [3,2-f] quina-
zoline-1,3-diamine and its derivatives as novel tight-binding
inhibitors. FEBS J. 2015;282:192238.
99. Rajagopalan PTR, Zhang Z, McCourt L, Dwyer M, Benkovic SJ,
Hammes GG. Interaction of dihydrofolate reductase with meth-
otrexate: ensemble and single-molecule kinetics. Proc Natl Acad
Sci USA. 2002;99:134816.
100. Kremer JM. Toward a better understanding of methotrexate.
Arthritis Rheum. 2004;50:137082.
101. Yuthavong Y, Tarnchompoo B, Vilaivan T, Chitnumsub P,
Kamchonwongpaisan S, Charman SA, et al. Malarial dihy-
drofolate reductase as a paradigm for drug development against a
resistance-compromised target. PNAS. 2012;109:168238.
102. Agarwal PK, Billeter SR, Rajagopalan PTR, Benkovic SJ,
Hammes-Schiffer S. Network of coupled promoting motions in
enzyme catalysis. PNAS. 2002;99:27949.
103. Cario H, Smith DEC, Blom H, Blau N, Bode H, Holzmann K,
et al. Dihydrofolate reductase deciency due to a homozygous
DHFR mutation causes megaloblastic anemia and cerebral folate
deciency leading to severe neurologic disease. Am J Hum Genet.
2011;88:22631.
104. Kotoula V, Krikelis D, Karavasilis V, Koletsa T, Eleftheraki AG,
Televantou D, et al. Expression of DNA repair and replication
genes in non-small cell lung cancer (NSCLC): a role for thy-
midylate synthetase (TYMS). BMC Cancer. 2012;12:342.
105. Burdelski C, Strauss C, Tsourlakis MC, Kluth M, Hube-Magg
C, Melling N, et al. Overexpression of thymidylate synthase
(TYMS) is associated with aggressive tumor features and early
PSA recurrence in prostate cancer. Oncotarget. 2015;6:
837787.
106. Calvert AH, Alison DL, Harland SJ, Robinson BA, Jackman AL,
Jones TR, et al. A phase I evaluation of the quinazoline antifolate
thymidylate synthase inhibitor, N10-propargyl-5,8-dideazafolic
acid, CB3717. J Clin Oncol. 1986;4:124552.
107. Webber S, Bartlett CA, Boritzki TJ, Hillard JA, Howland EF,
Johnston AL, et al. AG337, a novel lipophilic thymidylate syn-
thase inhibitor: in vitro and in vivo preclinical studies. Cancer
Chemother Pharm. 1996;37:50917.
108. Duch DS, Banks S, Dev IK, Dickerson SH, Ferone R, Heath LS,
et al. Biochemical and cellular pharmacology of 1843U89, a
novel benzoquinazoline inhibitor of thymidylate synthase. Can-
cer Res. 1993;53:8108.
109. Chu E, Callender MA, Farrell MP, Schmitz JC. Thymidylate
synthase inhibitors as anticancer agents: from bench to bedside.
Cancer Chemother Pharm. 2003;52:809.
110. Peters GJ, Backus HHJ, Freemantle S, van Triest B, Codacci-
Pisanelli G, van der Wilt CL, et al. Induction of thymidylate
synthase as a 5-uorouracil resistance mechanism. Biochim
Biophys Acta. 2002;1587:194205.
111. Chen D, Jansson A, Sim D, Larsson A, Nordlund P. Structural
analyses of human thymidylate synthase reveal a site that may
control conformational switching between active and inactive
states. J Biol Chem. 2017;292:1344958.
112. Lovelace LL, Johnson SR, Gibson LM, Bell BJ, Berger SH,
Lebioda L. Variants of human thymidylate synthase with loop
181-197 stabilized in the inactive conformation. Protein Sci.
2009;18:162836.
113. Huang X, Gibson LM, Bell BJ, Lovelace LL, Peña MMO,
Berger FG, et al. Replacement of Val3 in human thymidylate
synthase affects its kinetic properties and intracellular stability.
Biochemistry. 2010;49:247582.
114. Peters EJ, Kraja AT, Lin SJ, Yen-Revollo JL, Marsh S, Pro-
vince MA, et al. Association of thymidylate synthase variants
with 5-uorouracil cytotoxicity. Pharmacogenet Genom.
2009;19:399401.
115. Brodsky G, Barnes T, Bleskan J, Becker L, Cox M, Patterson D.
The human GARS-AIRS-GART gene encodes two proteins
which are differentially expressed during human brain develop-
ment and temporally overexpressed in cerebellum of individuals
with Down syndrome. Hum Mol Genet. 1997;6:204350.
116. Deis SM, Doshi A, Hou Z, Matherly LH, Gangjee A, Dann CE.
Structural and enzymatic analysis of tumor-targeted antifolates
that inhibit glycinamide ribonucleotide formyltransferase. Bio-
chemistry. 2016;55:457482.
117. Bronder JL, Moran RG. Antifolates targeting purine synthesis
allow entry of tumor cells into S phase regardless of p53 func-
tion. Cancer Res. 2002;62:523641.
118. Robert F, Garrett C, Dinwoodie WR, Sullivan DM, Bishop M,
Amantea M, et al. Results of 2 phase I studies of intravenous (iv)
pelitrexol (AG2037), a glycinamide ribonucleotide for-
myltransferase (GARFT) inhibitor, in patients (pts) with solid
tumors. J Clin Onocol. 2004;22:3075.
119. Shih C, Chen VJ, Gossett LS, Gates SB, MacKellar WC, Habeck
LL. et al. LY231514, a pyrrolo[2,3-d]pyrimidine-based antifolate
that inhibits multiple folate-requiring enzymes. Cancer Res.
1997;57:111623.
120. Welin M, Grossmann JG, Flodin S, Nyman T, Stenmark P,
Trésaugues L, et al. Structural studies of tri-functional human
GART. Nucleic Acids Res. 2010;38:730819.
121. Burda P, Kuster A, Hjalmarson O, Suormala T, Bürer C, Lutz S,
et al. Characterization and review of MTHFD1 deciency: four
new patients, cellular delineation and response to folic and
folinic acid treatment. J Inherit Metab Dis. 2015;38:86372.
122. Sdelci S, Rendeiro AF, Rathert P, You W, Lin J-MG, Ringler A,
et al. MTHFD1 interaction with BRD4 links folate metabolism to
transcriptional regulation. Nat Genet. 2019;51:9908.
123. New Antifolates in Clinical Development. Cancer Network.
https://www.cancernetwork.com/view/new-antifolates-clinical-
development.
124. Pikman Y, Puissant A, Alexe G, Furman A, Chen LM, Frumm
SM, et al. Targeting MTHFD2 in acute myeloid leukemia. J Exp
Med. 2016;213:1285306.
125. Zeller KI, Jegga AG, Aronow BJ, ODonnell KA, Dang CV. An
integrated database of genes responsive to the Myc oncogenic
transcription factor: identication of direct genomic targets.
Genome Biol. 2003;4:R69.
126. Vazquez A, Markert EK, Oltvai ZN. Serine biosynthesis with
one carbon catabolism and the glycine cleavage system
represents a novel pathway for ATP generation. PLoS ONE.
2011;6:e25881.
127. Dang CV, Reddy EP, Shokat KM, Soucek L. Drugging the
undruggablecancer targets. Nat Rev Cancer. 2017;17:5028.
128. Sneader W. Drug discovery: a history. UK: John Wiley & Sons;
2005. https://books.google.se/books?id=jglFsz5EJR8C&printsec=
frontcover&source=gbs_ge_summary_r&cad=0#v=onepa
ge&q&f=false.
129. Wang W, Zhou H, Liu L. Side effects of methotrexate therapy
for rheumatoid arthritis: a systematic review. Eur J Med Chem.
2018;158:50216.
130. Zachariae H. Methotrexate side-effects. Br J Dermatol.
1990;122:12733.
131. Albrecht K, Müller-Ladner U. Side effects and management of
side effects of methotrexate in rheumatoid arthritis. Clin Exp
Rheumatol. 2010;28:S95101.
132. Chan ESL, Cronstein BN. Methotrexatehow does it really
work? Nat Rev Rheumatol. 2010;6:1758.
133. Bertino JR, Göker E, Gorlick R, Li WW, Banerjee D. Resistance
mechanisms to methotrexate in tumors. Stem Cells. 1996;14:59.
Therapeutic targeting of the mitochondrial one-carbon pathway: perspectives, pitfalls, and potential
134. Nagura E. [Methotrexate and its drug resistance]. Gan Kagaku
Ryoho. 1988;15:28827.
135. Wang J, Li G. Mechanisms of methotrexate resistance in
osteosarcoma cell lines and strategies for overcoming this
resistance. Oncol Lett. 2015;9:9404.
136. Wojtuszkiewicz A, Peters GJ, van Woerden NL, Dubbelman B,
Escherich G, Schmiegelow K, et al. Methotrexate resistance in
relation to treatment outcome in childhood acute lymphoblastic
leukemia. J Hematol Oncol. 2015;8:61.
137. Frantz M-C, Wipf P. Mitochondria as a target in treatment.
Environ Mol Mutagen. 2010;51:46275.
138. Medwig-Kinney TN, Smith JJ, Palmisano NJ, Tank S, Zhang W,
Matus DQ. A developmental gene regulatory network for C.
elegans anchor cell invasion. Development. 2020;147. https://
doi.org/10.1242/dev.185850.
139. Ducker GS, Chen L, Morscher RJ, Ghergurovich JM, Esposito
M, Teng X, et al. Reversal of cytosolic one-carbon ux com-
pensates for loss of the mitochondrial folate pathway. Cell
Metab. 2016;23:114053.
140. Toyama BH, Hetzer MW. Protein homeostasis: live long, wont
prosper. Nat Rev Mol Cell Biol. 2013;14:5561.
141. Alberts B, Johnson A, Lewis J, Raff M, Roberts K, Walter P.
From RNA to protein. In: Molecular biology of the cell. 4th ed.
New York: Garland Science; 2002. https://www.ncbi.nlm.nih.
gov/books/NBK26829/.
L. N. Zhao et al.
... Because MTHFD2 is upregulated and drives cancer progression in a wide variety of tumors, and its expression is low in proliferative and differentiated adult tissues 12,15 , this enzyme has become very attractive for the development of novel, metabolically-targeted cancer therapies. In fact, many authors have highlighted the potential of MTHFD2 as a promising therapeutic target in cancer 8,[48][49][50][51] . ...
... The study of the structure of MTHFD2 has been of great importance for rational drug design 8,9,49 . MTHFD2 forms a dimer, which is achieved by key stabilizing interactions in the interaction of two NAD + -binding domains 8,52 , and its crystal structure was reported in 2017 by Gustafsson and colleagues 53 . ...
... The study of the structure of MTHFD2 has been of great importance for rational drug design 8,9,49 . MTHFD2 forms a dimer, which is achieved by key stabilizing interactions in the interaction of two NAD + -binding domains 8,52 , and its crystal structure was reported in 2017 by Gustafsson and colleagues 53 . However, the high degree of structural similarity between the isoenzymes MTHFD2, MTHFD2L and MTHFD1 makes it highly unlikely that competitive inhibitors will achieve enzyme specificity 49 . ...
Article
Full-text available
Methylenetetrahydrofolate dehydrogenase 2 (MTHFD2) is a mitochondrial enzyme of the folate-mediated one-carbon metabolism pathway. MTHFD2 has become a highly attractive therapeutic target due to its consistent upregulation in cancer tissues and its major contribution to tumor progression, although it also performs vital functions in proliferating healthy cells. Here, we review the diversity of canonical and non-canonical functions of this key metabolic enzyme under physiological conditions and in carcinogenesis. We provide an overview of its therapeutic potential and describe its regulatory mechanisms. In addition, we discuss the recently described non-canonical functions of MTHFD2 and the mechanistic basis of its oncogenic function. Finally, we speculate on novel therapeutic approaches that take into account subcellular compartmentalization and outline new research directions that would contribute to a better understanding of the fundamental roles of this metabolic enzyme in health and disease.
... In the process of 1C metabolism, TYMS converts deoxyuridylate to deoxythymidine-5′monophosphate in a 5,10-methylene-THF-dependent manner. 63 Many drugs have been developed by targeting TYMS, including 5-FU, pemetrexed and raltitrexed, which have already been applied in the clinic for years. Additionally, TYMS should be a potential target for immunotherapy. ...
... 196 By influencing the synthesis of GSH and NADPH, 1C metabolism is crucial for controlling the redox steady state. 63 Because of its participation in these biological functions, 1C metabolism regulates various downstream pathways that contribute to the progression of cancer. Thus, thoroughly understanding the function of 1C metabolism contributes to providing precise targets and specific pathways that are essential for cancer therapy. ...
Article
Full-text available
Background One‐carbon (1C) metabolism is a metabolic network that plays essential roles in biological reactions. In 1C metabolism, a series of nutrients are used to fuel metabolic pathways, including nucleotide metabolism, amino acid metabolism, cellular redox defence and epigenetic maintenance. At present, 1C metabolism is considered the hallmark of cancer. The 1C units obtained from the metabolic pathways increase the proliferation rate of cancer cells. In addition, anticancer drugs, such as methotrexate, which target 1C metabolism, have long been used in the clinic. In terms of immunotherapy, 1C metabolism has been used to explore biomarkers connected with immunotherapy response and immune‐related adverse events in patients. Methods We collected numerous literatures to explain the roles of one‐carbon metabolism in cancer immunotherapy. Results In this review, we focus on the important pathways in 1C metabolism and the function of 1C metabolism enzymes in cancer immunotherapy. Then, we summarise the inhibitors acting on 1C metabolism and their potential application on cancer immunotherapy. Finally, we provide a viewpoint and conclusion regarding the opportunities and challenges of targeting 1C metabolism for cancer immunotherapy in clinical practicability in the future. Conclusion Targeting one‐carbon metabolism is useful for cancer immunotherapy.
... Targeting MTHFD2, which is primarily expressed in tumor cells and notably absent in most normal tissues, has emerged as a promising strategy for anticancer therapies [23,27,32]. The use of drugs designed to selectively target MTHFD2 has the potential to effectively eliminate cancer cells dependent on this enzyme while minimizing adverse effects on healthy cells with low levels of the MTHFD2 protein [33,34]. ...
Article
Full-text available
This study aimed to explore the potential of Urtica dioica (U. dioica) ethanolic leaf extract for cancer treatment by identifying its components, evaluating its effects on cancer cell lines, and analyzing its molecular docking. The objective of this study was to investigate the anticancer properties of U. dioica ethanolic leaf extract and assess its potential as a therapeutic strategy for cancer treatment. This study utilized high-performance liquid chromatography (HPLC) to analyze the chemical composition of U. dioica ethanolic leaf extract. The anticancer effects of the extract were evaluated by assessing cell viability, determining IC50 values, and conducting ADMET analysis after oral administration. U. dioica ethanolic leaf extract was found to contain methyl hexadecanoate as its primary component, along with flavonoids and polyphenols. It effectively reduced cell viability in various tested cancer cell lines, with IC50 values varying for each cell line. The duration of treatment significantly influenced cell viability, with the most significant reduction observed after 48 h. Molecular docking studies suggested that catechin, kaempferol, and quercetin-3-O-rutinoside may have potential as inhibitors of the MTHFD2 enzyme. This study revealed the potential of U. dioica and its compounds in cancer treatment. Ethanolic leaf extract has been shown to have anticancer effects on various cancer cell lines, with catechin and kaempferol showing promise as inhibitors of the MTHFD2 enzyme. Further research is warranted to explore the therapeutic implications of U. dioica in cancer treatment.
... The folate metabolism pathway contributes to cell proliferation and survival by producing one-carbon formyl groups for various cellular processes, including de novo purine and thymidine synthesis [11,33,34]. Additionally, the folate cycle produces NADPH for use in maintaining redox homeostasis which is necessary for cellular detoxification [12]. ...
Article
Full-text available
Objectives Cancer cells undergo metabolic reprogramming to adapt to high oxidative stress, but little is known about how metabolic remodeling enables gastric cancer cells to survive stress associated with aberrant reactive oxygen species (ROS) production. Here, we aimed to identify the key metabolic enzymes that protect gastric cancer (GC) cells from oxidative stress. Methods ROS level was detected by DCFH-DA probes. Multiple cell biological studies were performed to identify the underlying mechanisms. Furthermore, cell-based xenograft and patient-derived xenograft (PDX) model were performed to evaluate the role of MTHFD2 in vivo. Results We found that overexpression of MTHFD2, but not MTHFD1, is associated with reduced overall and disease-free survival in gastric cancer. In addition, MTHFD2 knockdown reduces the cellular NADPH/NADP+ ratio, colony formation and mitochondrial function, increases cellular ROS and cleaved PARP levels and induces in cell death under hypoxia, a hallmark of solid cancers and a common inducer of oxidative stress. Moreover, genetic or pharmacological inhibition of MTHFD2 reduces tumor burden in both tumor cell lines and patient-derived xenograft-based models. Discussion our study highlights the crucial role of MTHFD2 in redox regulation and tumor progression, demonstrating the therapeutic potential of targeting MTHFD2.
... This enzyme is highly expressed in more than 85% of cancer types compared with normal tissues [13], induced an increased flux of serine catabolism and one-carbon cycle. Highly SHMT2 expression promotes malignant behavior of cancer cells through multiple biological processes, including cell proliferation, chemoresistance, angiogenesis, migration and invasion [14][15][16]. Clinical research has shown that SHMT2 is highly expressed in malignant thyroid tissues compare to normal tissues and is associated with poor clinical outcomes [17], but the functions and underlying mechanisms of SHMT2 in PTC remain unclear. Thus, elucidating the functions and molecular mechanism of SHMT2 in PTC would benefit its diagnosis and therapy. ...
Article
Full-text available
Cancer cells alter their metabolism and epigenetics to support cancer progression. However, very few modulators connecting metabolism and epigenetics have been uncovered. Here, we reveal that serine hydroxymethyltransferase-2 (SHMT2) generates S-adenosylmethionine (SAM) to epigenetically repress phosphatase and tensin homolog (PTEN), leading to papillary thyroid cancer (PTC) metastasis depending on activation of AKT signaling. SHMT2 is elevated in PTC, and is associated with poor prognosis. Overexpressed SHMT2 promotes PTC metastasis both in vitro and in vivo. Proteomic enrichment analysis shows that AKT signaling is activated, and is positively associated with SHMT2 in PTC specimens. Blocking AKT activation eliminates the effects of SHMT2 on promoting PTC metastasis. Furthermore, SHMT2 expression is negatively associated with PTEN, a negative AKT regulator, in PTC specimens. Mechanistically, SHMT2 catalyzes serine metabolism and produces activated one-carbon units that can generate SAM for the methylation of CpG islands in PTEN promoter for PTEN suppression and following AKT activation. Importantly, interference with PTEN expression affects SHMT2 function by promoting AKT signaling activation and PTC metastasis. Collectively, our research demonstrates that SHMT2 connects metabolic reprogramming and epigenetics, contributing to the poor progression of PTC.
Article
Full-text available
One-carbon metabolism is a universal metabolic process that mediates the transfer of one-carbon units for purine and thymidine synthesis. One-carbon metabolism has been found to be dysregulated in various cancer types due to its role in production of purine and pyrimidine nucleotides, epigenetic program, and redox homeostasis. One-carbon metabolism is composed a network of one-carbon metabolic enzymes. Disturbing the expression and enzymatic activity of these one-carbon metabolic enzymes could lead to fluctuations of metabolites in the tumor microenvironment. Serine hydroxymethyltransferases (SHMTs) and methylenetetrahydrofolate dehydrogenases (MTHFDs) are gradually recognized as important one-carbon metabolic enzymes for regulating tumor initiation and development, representing potential therapeutic targets for anti-tumor strategies. In the review, we primarily focused on the role of SHMTs and MTHFDs in cancer. Several inhibitors targeting MTHFDs and SHMTs have exert its potential to decrease tumor burden and inhibit tumor proliferation, highlighting the potential of targeting one-carbon metabolic enzymes for anti-cancer strategies.
Preprint
Full-text available
Background: Neuroblastoma is the common extracranial solid tumor in children, with a poor prognosis for high-risk patients. MYCN amplification is the most important molecular biomarker in the high-risk category. For MYCN’s “undruggable” properties--lack of enzymatic pocket for conventional small molecules to bind and inaccessibility for antibodies due to the predominant nucleus localization of MYCN, current therapeutic strategies have not significantly improved. It is known that MYCN can affect a variety of tumor metabolism and is closely related to tumor differentiation. The objective of this study is to investigate the impact of purine metabolism on the maintenance of stemness in MYCN high-risk neuroblastoma and identify potential small molecule inhibitors that can enhance the differentiation of neuroblastoma cells. Methods: Metabolic mass spectrometry data were used to analyze the differences in metabolites between neuroblastoma cell lines with high and low MYCN, and key metabolic enzymes were analyzed in conjunction with public RNA databases. Different neuroblastoma cell lines were treated with the small molecule inhibitor for cell differentiation, proliferation, colony formation, and cell migration analysis, to find inhibitors that are promotive for cell differentiation and validate them in vivo and in vitro. Results: In our study, we found significant changes in nucleotide metabolism in NB cell lines with high MYCN compared to those with low MYCN. The expression of purine metabolic enzymes was positively correlated with MYCN expression, prognosis, and differentiation status in neuroblastoma. The small molecular lometrexol, a phosphoribosylglycinamide formyltransferase (GART) inhibitor, which blocks the de novo pathway of purine metabolism, can promote a pro-differentiation effect on neuroblastoma cells through in vivo and in vitro experiments and inhibit tumorigenesis. Discussion: Our study suggested that lometrexol, the small molecule inhibitor of nucleotide metabolism, may contribute to improving treatment strategies for pediatric neuroblastoma, enhancing patient prognosis, and improving overall quality of life.
Article
Full-text available
Background: Kidney cancer is one of the most common cancers in the world. It is necessary to clarify its underlying mechanism and find its prognostic biomarkers. Current studies showed that SHMT2 may be participated in several kinds of cancer. Methods: Our studies investigated the expression of SHMT2 in kidney cancer by Oncomine, Human Protein Atlas database and ULCAN database. Meanwhile, we found its co-expression gene by cBioPortal online tool and validated their relationship in A498 and ACHN cells by cell transfection, western blot and qRT-PCR. Besides these, we also explored their prognostic values via the Kaplan-Meier plotter database in different types of kidney cancer patients. Results: SHMT2 was found to be increased in 7 kidney cancer datasets, compared to normal renal tissues. For the cancer stages, ages and races, there existed significant difference in the expression of SHMT2 among different groups by mining of the UALCAN database. High SHMT2 expression is associated with poor overall survival in patients with kidney cancer. Among all co-expressed genes, NDUFA4L2 and SHMT2 had a high co-expression efficient. SHMT2 overexpression led to the increased expression of NDUFA4L2 at both mRNA and protein levels. Like SHMT2, overexpressed NDUFA4L2 also was associated with worse overall survival in patients with kidney cancer. Conclusion: Based on above results, overexpressed SHMT2 and its co-expressed gene NDUFA4L2 were all correlated with the prognosis in kidney cancer. The present study might be benefit for better understanding the clinical significance of SHMT2 and provided a potential therapeutic target for kidney cancer in future.
Article
Full-text available
The aggressiveness and recurrence of glioma are major obstacles for the treatment of this type of tumor. Further understanding of the molecular mechanisms of glioma is necessary to improve the efficacy of therapy. MicroRNAs have been widely studied in many human cancers. Here, we found that miR-940 was one of the primary downregulated miRNAs in clinical samples and glioma cell lines through bioinformatics analysis and qRT-PCR. Upregulating miR-940 expression significantly inhibited the proliferation and invasion and promoted apoptosis of U87 and U118 cells. In addition, experiments in vivo showed that upregulation of miR-940 expression inhibited xenograft growth. Methylenetetrahydrofolate dehydrogenase (MTHFD2), a dual-functional metabolic enzyme, is involved in the one-carbon metabolism of folate in mitochondria. We found MTHFD2 to be overexpressed in glioma tissues and our clinical samples by qRT-PCR and Western blot assays. Through TargetScan prediction and luciferase assays, we found that miR-940 directly targets MTHFD2. Upregulation of miR-940 expression inhibited the expression of MTHFD2 and led to intracellular one-carbon metabolism dysfunction. Furthermore, the antitumor effects of miR-940 could be attenuated by overexpression of MTHFD2. Together, the results of our study suggest that miR-940 may be a new therapeutic target for the treatment of glioma through targeting of MTHFD2.
Article
Full-text available
Most cancer cells rely on glycolysis to generate ATP, even when oxygen is available. However, merely inhibiting the glycolysis is insufficient for the eradication of cancer cells. One main reason for this is that cancer cells have the potential to adapt their metabolism to their environmental conditions. In this study, we investigated how cancer cells modify their intracellular metabolism when glycolysis is suppressed, using PANC-1 pancreatic cancer cells and two other solid tumor cell lines, A549 and HeLa. Our study revealed that glycolytically suppressed cells upregulated mitochondrial function and relied on oxidative phosphorylation (OXPHOS) to obtain the ATP necessary for their survival. Dynamic changes in intracellular metabolic profiles were also observed, reflected by the reduced levels of TCA cycle intermediates and elevated levels of most amino acids. Glutamine and glutamate were important for this metabolic reprogramming, as these were largely consumed by influx into the TCA cycle when the glycolytic pathway was suppressed. During the reprogramming process, activated autophagy was involved in modulating mitochondrial function. We conclude that upon glycolytic suppression in multiple types of tumor cells, intracellular energy metabolism is reprogrammed toward mitochondrial OXPHOS in an autophagy-dependent manner to ensure cellular survival.
Article
Full-text available
Cellular invasion is a key part of development, immunity, and disease. Using the in vivo model of C. elegans anchor cell invasion, we characterize the gene regulatory network that promotes cell invasion. The anchor cell is initially specified in a stochastic cell fate decision mediated by Notch signaling. Previous research has identified four conserved transcription factors, fos-1a (Fos), egl-43 (EVI1/MEL), hlh-2 (E/Daughterless) and nhr-67 (NR2E1/TLX), that mediate anchor cell specification and/or invasive behavior. Connections between these transcription factors and the underlying cell biology that they regulate are poorly understood. Here, using genome editing and RNA interference, we examine transcription factor interactions before and after anchor cell specification. Initially, these transcription factors function independently of one another to regulate LIN-12 (Notch) activity. Following anchor cell specification, egl-43, hlh-2, and nhr-67, function largely parallel to fos-1 in a type I coherent feed-forward loop with positive feedback to promote invasion. Together, these results demonstrate that the same transcription factors can function in cell fate specification and differentiated cell behavior, and that a gene regulatory network can be rapidly assembled to reinforce a post-mitotic, pro-invasive state.
Article
Full-text available
Methylenetetrahydrofolate dehydrogenase 2 (MTHFD2) is a bifunctional enzyme located in the mitochondria. It has been reported to be overexpressed in several malignancies. However, the relationship between the expression of MTHFD2 and non‐small cell lung cancer (NSCLC) remains largely unknown. In this study, we found that MTHFD2 was significantly overexpressed in NSCLC tissues and cell lines. Knockdown of MTHFD2 resulted in reduced cell growth and tumorigenicity in vitro and in vivo. Besides, the mRNA and protein expression level of cell cycle genes, such as CCNA2, MCM7 and SKP2, was decreased in MTHFD2 knockdown H1299 cells. Our results indicate that the inhibitory effect of MTHFD2 knockdown on NSCLC may be mediated via suppressing cell cycle‐related genes. These findings delineate the role of MTHFD2 in the development of NSCLC and may have potential applications in the treatment of NSCLC.
Article
Full-text available
Identification of new molecular targets is needed for the treatment of colorectal cancer (CRC). Methylenetetrahydrofolate dehydrogenase 1 like (MTHFD1L), an enzyme in the folate cycle, is involved in formate generation and therefore in one-carbon metabolism. Here, we examined the expression and the role of MTHFD1L in CRC progression. Bioinformatics analysis of several public databases showed overexpression of MTHFD1L in CRC tissues as compared to normal tissues. Quantitative real-time PCR and Western blotting revealed that expressions of MTHFD1L RNA and protein were higher in CRC tissues compared to their corresponding normal tissues of CRC patients. Immunohistochemical staining demonstrated higher cytoplasmic MTHFD1L reactivity in tumor tissues compared to paired normal tissues. Further, to determine the functional relevance of MTHFD1L, it was knocked down by an siRNA in CRC cells. Silencing of MTHFD1L inhibited CRC cell proliferation, colony formation, invasion, and migration. Thus, to our knowledge for the first time in the literature, we show that MTHFD1L is involved in CRC progression and that blocking of MTHFD1L decreases the growth of colon cancer cells, thus providing an avenue to target this enzyme with small molecule inhibitors.
Article
Cancer growth affects the pH of microenvironment that becomes acidified and supports local invasion. Studies have reported that pH is a factor in cancer growth, division, and spread. Although cancer cells produce low extracellular pH (6.5–6.9), they are able to maintain their intracellular pH at favorable ranges (7.2–7.4). Our preliminary results showed that growth of cancer cells changes after low pH treatment. The objective of our research is to study how low pH affects cell morphology and behavior. In my preliminary experiments, HeLa cells were incubated with low pH solution for 10 min and harvested using trypsin into medium with normal pH. After 24 hours in culture, cell number was assessed using Trypan Blue Exclusion Assay. As a result, low pH treated group doubled 1.4 times; whereas control group doubled 2.3 times, suggesting a division latency in low pH treated group on Day 1. Cell morphology on Day 1 was recorded as images and was categorized into round, spindle, triangle, and irregular groups, among which spindle cells were likely to be at G1 phase. Percentage of the spindle cells gradually increased during 16–24 hours in experimental group; however, in control group, percentage of spindle cells decreased, suggesting that low pH had induced cell cycle arrest at G1 phase. In conclusion, low pH treatment affects cell growth potentially through G1 cell cycle arrest. Future work is to confirm cell cycle arrest at G1 phase after low pH treatment and to study the mechanism of low pH effect on cancer cell growth. Support or Funding Information Funding is provided partially by NIH (#NS081629) to YVL. Thanks to the Osteopathic Heritage Foundations, Graduate Assistantship Program at Ohio University Heritage College of Osteopathic Medicine for supporting graduate student YH. This abstract is from the Experimental Biology 2018 Meeting. There is no full text article associated with this abstract published in The FASEB Journal .
Article
Cancer cells have enhanced metabolic needs due to their rapid proliferation. Moreover, throughout their progression from tumor precursors to metastases, cancer cells face challenging physiological conditions, including hypoxia, low nutrient availability, and exposure to therapeutic drugs. The ability of cancer cells to tailor their metabolic activities to support their energy demand and biosynthetic needs throughout disease progression is key for their survival. Here, we review the metabolic adaptations of cancer cells, from primary tumors to therapy resistant cancers, and the mechanisms underpinning their metabolic plasticity. We also discuss the metabolic coupling that can develop between tumors and the tumor microenvironment. Finally, we consider potential metabolic interventions that could be used in combination with standard therapeutic approaches to improve clinical outcome.
Article
Objective: Gastrointestinal tumors are malignant tumors with high morbidity. Mitochondrial serine hydroxymethyltransferase 2 (SHMT2) is a key enzyme in the synthesis of serine and glycine, which has prognostic and therapeutic value for many malignant tumors. However, the role of SHMT2 in gastric cancer (GC), esophageal cancer (ESCC), and colorectal cancer (CC) has not been clarified. Patients and methods: The expression of SHMT2 was detected in GC, ESCC, and CC by immunohistochemistry and reverse real time transcription-polymerase chain reaction. The relationships between SHMT2 expression and clinicopathologic characteristics, recurrence-free survival (RFS), and disease-specific survival (DSS) were analyzed by the survival analysis and correlation analysis. Results: The positive expression rate of SHMT2 in GC, ESCC, and CC was 74.1%, 69.2%, and 71.7%, respectively. Patients with high expression of SHMT2 had a worse prognosis. In GC, high SHMT2 expression had positive correlation with lymph node metastasis (p=0.005) and histological grade (p=0.002). In ESCC, high SHMT2 expression had positive correlation with pT classification (p=0.033) and pM classification (p=0.029). In CC, high SHMT2 expression had positive correlation with tumor size (p=0.004), lymph node metastasis (p=0.035), TNM stage (p=0.007), and histological grade (p=0.020). Notably, SHMT2 expression was an independent prognostic factor for RFS and DSS in GC, ESCC, and CC (p<0.05). Conclusions: SHMT2 is upregulated in GC, ESCC, and CC. The high expression of SHMT2 is correlated with gastrointestinal tumors progression, and poor prognosis, which is a potential new target for the diagnosis and treatment of gastrointestinal tumors.