1 s2.0 S0147651321012100 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Ecotoxicology and Environmental Safety 229 (2022) 113098

Contents lists available at ScienceDirect

Ecotoxicology and Environmental Safety


journal homepage: www.elsevier.com/locate/ecoenv

Review

Metabolic reprogramming in the arsenic carcinogenesis


Yihui Ruan a, Xin Fang a, Tingyue Guo a, Yiting Liu a, Yu Hu a, Xuening Wang a, Yuxin Hu c,
Lanyue Gao c, Yongfang Li b, Jingbo Pi b, d, Yuanyuan Xu a, b, *
a
Group of Chronic Disease and Environmental Genomics, School of Public Health, China Medical University, P.R. China
b
The Key Laboratory of Liaoning Province on Toxic and Biological Effects of Arsenic, China Medical University, P.R. China
c
Experimental Teaching Center, School of Public Health, China Medical University, P.R. China
d
Program of Environmental Toxicology, School of Public Health, China Medical University, P.R. China

A R T I C L E I N F O A B S T R A C T

Edited by Professor Bing Yan Chronic exposure to arsenic has been associated with a variety of cancers with the mechanisms undefined.
Arsenic exposure causes alterations in metabolites in bio-samples. Recent research progress on cancer biology
Keywords: suggests that metabolic reprogramming contributes to tumorigenesis. Therefore, metabolic reprogramming
Arsenic provides a new clue for the mechanisms of arsenic carcinogenesis. In the present manuscript, we review the latest
Carcinogenic mechanism
findings in reprogramming of glucose, lipids, and amino acids in response to arsenic exposure. Most studies
Glucose metabolism
focused on glucose reprogramming and found that arsenic exposure enhanced glycolysis. However, in vivo studies
Lipid metabolism
Amino acid metabolism observed “reverse Warburg effect” in some cases due to the complexity of the disease evolution and microen­
Oxidative stress vironment. Arsenic exposure has been reported to disturb lipid deposition by inhibiting lipolysis, and induce
serine-glycine one-carbon pathway. As a dominant mechanism for arsenic toxicity, oxidative stress is considered
to link with metabolism reprogramming. Few studies analyzed the causal relationship between metabolic
reprogramming and arsenic-induced cancers. Metabolic alterations may vary with exposure doses and periods.
Identifying metabolic alterations common among humans and experiment models with human-relevant exposure
characteristics may guide future investigations.

1. Introduction Metabolic reprogramming is a hallmark of malignancy. It refers to


commonly altered metabolic pathways observed in highly proliferative
Arsenic is an element that universally exists in the environment cancer cells (Faubert et al., 2020). Nobel laureate Otto Warburg first
(ATSDR, 2015). Typically, one of the most crucial routes of human noticed that cancer cells relied primarily on glycolysis as the primary
exposure to arsenic is through drinking water (IARC, 2012). Approxi­ pathway for producing adenosine triphosphate (ATP) to support rapid
mately 220 million people in more than 50 countries and regions are growth, even with sufficient oxygen, in a phenomenon called the
potentially exposed to groundwater containing arsenic levels above 10 “Warburg effect” (Warburg, 1956). Over the past two decades, in-depth
μg/L, the guideline value set by the World Health Organization (Podg­ research has not only confirmed that carcinogenic damage largely leads
orski and Berg, 2020). Long-term exposure to arsenic poses a great to the Warburg effect but also suggested that the evolution of metabolic
threat to human health. In severe cases, it leads to cancers of the skin, reprogramming has far exceeded initial expectations (Pavlova and
lung, liver, bladder, and kidney (Chen et al., 2005; Mead, 2005; Yu et al., Thompson, 2016). Currently, metabolic reprogramming refers to
2001). Arsenic, therefore, has been categorized as a Class I human changes in several metabolic pathways, such as glycolysis, gluta­
carcinogen by the International Agency for Research on Cancer (IARC, minolysis, fatty acid de novo synthesis, and fatty acid oxidation (Zhou
2012). Chromosomal aberrations, epigenetic modifications, oxidative et al., 2019). These pathways occur collectively in cancer cells to in­
stress, and DNA damage have been proposed as carcinogenic modes of crease the acquisition of essential nutrients, prioritize the allocation of
this metalloid (Bjorklund et al., 2018; Cui et al., 2021; Eckstein et al., nutrients to metabolic pathways and facilitate tumor transformation
2017; Mandal, 2017). However, the underlying mechanisms of arsenic (Pavlova and Thompson, 2016). Reprogrammed metabolic characteris­
carcinogenesis are not fully clear. tics may inform therapeutic approaches for selectively targeting cancer

* Correspondence to: School of Public Health, China Medical University, No. 77 Puhe Road, Shenbei New District, Shenyang, Liaoning 110122, P.R. China.
E-mail addresses: yyxu@cmu.edu.cn, laurel1214@hotmail.com (Y. Xu).

https://doi.org/10.1016/j.ecoenv.2021.113098
Received 17 August 2021; Received in revised form 6 December 2021; Accepted 14 December 2021
Available online 21 December 2021
0147-6513/© 2021 The Authors. Published by Elsevier Inc. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
Y. Ruan et al. Ecotoxicology and Environmental Safety 229 (2022) 113098

cells. Recently, the importance of metabolic reprogramming in the elevated in tumorigenesis (Pavlova and Thompson, 2016). PPP gener­
initiation and progression of various cancers has been intensively ates various biosynthetic precursors for nucleotides and lipids using
studied (Azoitei et al., 2016; Gu et al., 2019; Senni et al., 2019; Sun et al., glycolytic intermediates (Pinweha et al., 2016). SGOC pathway metab­
2021). In addition, the metabolic reprogramming of immune cells is olites (e.g., phenylalanine and formate) contribute to a number of
crucial to the maintenance of the tumor microenvironment (TME) and cellular biosynthetic and regulatory processes (Coco et al., 2020).
the fate of the tumor (Somarribas et al., 2021). While cancer cells opt for Methylene tetrahydrofolate dehydrogenase 2 (MTHFD2) was found to
metabolic alterations such as glycolysis to support their biosynthetic and be among the top three most frequently overexpressed metabolic en­
energetic needs for rapid proliferation and survival in hypoxic TME, zymes in cancer, suggesting that alterations in the SGOC pathway may
evidence suggests that the infiltrating immune cells (e.g., B cells, T cells, be universally selected in tumorigenesis (Nilsson et al., 2014). Mutations
myeloid-derived suppressor cells, tumor-associated macrophages, neu­ in metabolic enzymes leading to the accretion of specific intermediate
trophils, and natural killer cells) also undergo metabolic changes that metabolites also drive cancer initiation (Fig. 1). Mutations of succinate
contribute to their pro- or anti-tumor functions, as reviewed by Biswas dehydrogenase and fumarate hydratase (SHDmut and FHmut) induce
(2015). elevated levels of succinate and fumarate (Eijkelenkamp et al., 2020; Ye
A number of experimental studies have demonstrated that long-term et al., 2013), which can activate hypoxia-inducible factor (HIF) 1
arsenic exposure causes alterations in metabolites, such as glucose signaling via inhibition of prolyl hydroxylase domain (PHD) proteins,
(Garcia-Sevillano et al., 2014; Liu et al., 2014; Xu et al., 2019), lipids demethylation of DNA via inhibition of ten-eleven translocation family
(Chen et al., 2017; Rivas-Santiago et al., 2019), and amino acids (Liu (TET), and histone demethylation via inhibition of jumonji C
et al., 2014; Zhang et al., 2010). Research on the effect of arsenic on domain-containing histone lysine demethylases (JMJD) (Carlson and
glucose metabolism is the earliest and the most in-depth. In arsenic Van Beneden, 2014; Hawkins et al., 2016; Tannahill et al., 2013).
carcinogenesis, there are also the most reports on glucose metabolism. Similarly, mutations of isocitrate dehydrogenase (IDH1mut and IDH2mut)
Current studies have found that arsenic exposure causes the Warburg lead to the generation and accumulation of the oncometabolite 2-hy­
effect, which may contribute to carcinogenesis. Chronic arsenic expo­ droxyglutarate (2-HG), which results in demethylation of DNA and
sure enhanced glycolysis in human bronchial epithelial cells (BEAS-2B) histones via inhibition of JMJD and TET, respectively (Baylin and Her­
and human hepatic epithelial cells (L-02) (Luo et al., 2017; Zhao et al., man, 2000; Ye et al., 2013). Consequently, the activation of HIF1A and
2014). This metabolic pattern was required for non-anchor-dependent persistent histone and DNA methylation impair cellular differentiation,
growth properties, one of the typical characteristics of malignant cells. contributing to cancer initiation (Baylin and Herman, 2000; Figueroa
Thus, reprogramming of glucose metabolism has been considered to et al., 2010; Sajnani et al., 2017) (Fig. 1).
contribute to malignant cell transformation. However, it is more An overarching theme in cell metabolism is improving cellular
complicated in vivo, as arsenic exposure was reported to reduce glyco­ fitness to provide a selective advantage during tumorigenesis. Thus, the
lytic gene expression in the mouse liver. Moreover, arsenic exposure has metabolites produced by oncogene-regulated metabolic pathway alter­
been reported to disturb lipid deposition by inhibiting lipolysis in ations may be the basis for tumorigenesis, and some pathways might be
human renal tubular epithelial cells (HK-2) (Fang et al., 2018), and suitable therapeutic targets.
induce the serine-glycine one-carbon (SGOC) pathway in BEAS-2B cells
(Bi et al., 2020). However, the evidence is limited. Additional studies on 3. Effect of arsenic exposure on metabolism
metabolic alterations due to arsenic exposure are needed to verify these
findings. Of note, oxidative stress, a key mechanism underlying arsenic 3.1. Arsenic exposure and glucose metabolism
toxicity, is linked to metabolic reprogramming via same key factors,
such as reduced glutathione (GSH), nicotine adenine dis­ Glycolysis is a key pathway of glucose metabolism in cancer cells
phosphonucleotide (NADPH), nuclear factor erythroid 2-related factor 2 (Liao et al., 2021). Pyruvate, as the end product of glycolysis, is con­
(NRF2), and adenosine 5’-monophosphate-activated protein kinase verted into lactate and secreted extracellularly (Liao et al., 2021; Zhu
(AMPK). and Thompson, 2019). Glycolysis provides the precursors for three
This paper reviews the evidence for metabolic reprogramming in major classes of macromolecules (nucleic acids, lipids, and proteins),
response to arsenic exposure and the probable role of metabolic which are required at high levels in proliferating cells (Birts et al., 2020;
reprogramming in arsenic carcinogenesis. This information may help to Dang, 2012). Meanwhile, higher lactate levels are notably associated
formulate better preventative and therapeutic strategies for arsenic with tumor recurrence and promote the emergence of an immunosup­
poisoning. pressive microenvironment (Goetze et al., 2011; J. Li et al., 2021). The
currently available literature on alteration of glucose metabolism in
2. Metabolic reprogramming and carcinogenesis response to arsenic mainly focuses on glycolysis (Table 1). Long-term
arsenic exposure enhances glycolysis in cells, contributing to malig­
Studies on metabolism in cancer cells have deepened our under­ nant transformation. However, the effects of arsenic exposure on
standing of the functional consequences and mechanisms of metabolic glycolysis in animal models appear species specific and sometimes
changes during carcinogenesis. Most cancer cells reprogram their opposite to the in vitro observations (Fig. 2A).
nutritional metabolism due to mutations in oncogenes and tumor sup­ Arsenic-induced transformed cells presented with a metabolic switch
pressor factors, resulting in constitutive activation of signaling pathways to aerobic glycolysis (Warburg effect) (He et al., 2019). Multiple studies
involved in cell growth (Papa et al., 2019). Multiple growth signaling conducted in BEAS-2B cells have found that glycolysis and oxidative
nodes such as PI3K/Akt, p53, and Ras are aberrantly activated in cancer, phosphorylation (OXPHOS) in glucose homeostasis are reciprocally
all of which have the ability to promote the cellular acquisition of regulated during the first week of arsenite exposure (Bi et al., 2020; Zhao
glucose (Ansary et al., 2019; Bravatà et al., 2021; Hoxhaj and Manning, et al., 2013). In BEAS-2B cells treated with 1 μM arsenite, three
2020; Wang et al., 2019). The transcription factor MYC contributes to glycolysis-related genes were repressed during the first week, but 14
metabolic reprogramming by shifting the uptake and metabolism of glycolysis-related genes were highly expressed by the end of the second
glutamine in cancer cells (Venkateswaran et al., 2019). week, including enolase (ENO3), hexokinase (HK2), and phosphoglyc­
Under environmental stress conditions, the metabolic reprogram­ erate kinase (PGK2) (Zhao et al., 2013). Increased concentrations of
ming of cells not only gives them a survival advantage to adapt to the citrate and α-ketoglutarate (α-KG) but decreased concentrations of
environment, but also leads to accelerated cell proliferation and fumarate and malate in a time-dependent manner were observed in
abnormal differentiation, thus promoting tumorigenesis (Lee and Kim, BEAS-2B cells exposed to 0.25 μM arsenite for 3 days (Bi et al., 2020).
2016). Utilization of pentose phosphate pathway (PPP) is frequently With the prolongation of arsenic exposure, glycolysis gradually became

2
Y. Ruan et al. Ecotoxicology and Environmental Safety 229 (2022) 113098

Fig. 1. The role of metabolic reprogramming in


tumorigenesis. Metabolic enzyme mutations
result in the accretion of specific intermediate
metabolites that contribute to the initiation of
cancer. Blue lines represent downregulated
pathways. α-KG, α-ketoglutarate; FH, fumarate
hydratase; 2-HG, 2-hydroxyglutarate; 5hmC, 5-
hydroxymethylcytosine; IDH, isocitrate dehy­
drogenase; JMJD, jumonji C domain-containing
histone lysine demethylases; 5mC, 5-methylcy­
tosine; PHD, prolyl hydroxylase domain; TCA,
tricarboxylic acid; TET, ten-eleven trans­
location; SDH, succinate dehydrogenase.

Table 1
Effects of arsenic exposure on glucose metabolism.
Subject/model Reagent Dose Expo Metabolic changes Phenotype References
Time

in vitro
BEAS-2B NaAsO2 1 μM 4 weeks HIF1A, lactate ↑; glycolysis related genes ↑ Metabolic Zhao et al. (2013)
disruption
BEAS-2B NaAsO2 0.25 μM 3 days Citrate, isocitrate, succinate, fumarate and malate ↓; Metabolic Bi et al. (2020)
α-KG ↑ disruption
BEAS-2B NaAsO2 0.25 μM 6 months Fumarate, succinate, α-KG, isocitrate, citrate and malate Malignant Bi et al. (2020)
↓ transformation
HDF, HUC, A549, NaAsO2 0.1, 0.5, 1, 2 or 2 weeks lactate ↑ Metabolic Zhao et al. (2013)
and RWPE-1 4 μM disruption
L-02 and THLE-3 NaAsO2 2 μM 15 weeks HIF1A and MCT4↑ Malignant Luo et al. (2017)
transformation
BEAS-2B NaAsO2 1 μM 52 weeks HIF1A, 3-phosphoglycerate, isocitrate, pyruvate, G6P, Malignant Zhao et al. (2014)
and lactate ↑; glucose and α-KG ↓ transformation
in vivo
Zebrafish Na3AsO4 20 mg/L 96 hours Liver: G6P and pyruvate ↑; ursodeoxycholic acid and Liver injury Li et al. (2016)
(Aquatic chenodeoxycholic acid ↑, hypotaurine ↓
environment)
Zebrafish Na3AsO4 5 or 50 µg/L 14 days Liver: lactate, alanine, succinate, and ATP ↓; Metabolic Xu et al. (2019)
(Aquatic phosphocholine, inosine ↑ disruption
environment)
BALB/c mice NaAsO2 3 mg/kg/day 12 days Plasma: isocitrate, α-KG and citrate ↑ Metabolic Garcia-Sevillano et al.
(Gavage) disruption (2014)
C57BL/6 J mice NaAsO2 1, 2, or 4 mg/L 12 weeks Liver: glycolytic genes ↓, glycogen ↓ and Insulin resistance Gao et al. (2019)
(Drinking water) gluconeogenesis genes ↑

↑ Increase, ↓ decrease.

the main cellular energy supply pathway. Several in vitro studies of one in culture medium, including in primary human urothelial cells (HUC),
research group found that exposure to 1 μM arsenite for 2 weeks induced human prostate epithelial cells (RWPE-1), BEAS-2B cells and human
lactate accumulation in the cell accompanied by increased lactate levels dermal fibroblasts (HDF) (Zhao et al., 2013). Increased expression levels

3
Y. Ruan et al. Ecotoxicology and Environmental Safety 229 (2022) 113098

Fig. 2. Metabolic reprogramming by arsenic exposure. Arsenic exposure causes metabolic reprogramming of glucose, lipid and amino acid. Part A, B, and C
correspond to glucose, lipid and amino acid metabolism, respectively. Red, black, and blue fonts indicate upregulated, sustained, and inhibited metabolic compo­
nents, respectively. CPT, carnitine palmitoyltransferase; GLUT, glucose transporter; Gly, glycine; HIF, hypoxia-inducible factor; IDH, isocitrate dehydrogenase; LDH,
lactic dehydrogenase; LXR, liver X receptor; MCT, monocarboxylate transporter; PHD, prolyl hydroxylase domain; PHGDH, phosphoglycerate dehydrogenase; PPAR,
peroxisome proliferator-activated receptor; PSAT, phosphoserine aminotransferase; Ser, serine; SHMT, serine hydroxymethyltransferase; SREBP, sterol regulatory
element-binding protein.

of the glycolysis-related genes HK2, ENO1, and glucose transporter hydroxylate HIF1A, were reduced (Bi et al., 2020; Wong et al., 2013;
(GLUT4) were found in L-02 cells treated with 2 µM arsenite for Zhao et al., 2014). These alterations lead to the sustained elevation of
24 hours (Luo et al., 2016). At 15 weeks, a metabolic switch to glycolysis HIF1A protein levels. In L-02 cells and non-transformed human liver
occurred in arsenite-transformed malignant L-02 cells (Luo et al., 2017). epithelial cells (THLE-3) exposed to arsenite (2 μM, 15 weeks), the
The same metabolic shift was observed in arsenite-transformed malig­ expression levels of HIF1A and lactate transporter protein mono­
nant BEAS-2B cells, in which the expression levels of 45 out of 54 genes carboxylate transporter (MCT) 4 coordinated by HIF1A were upregu­
associated with OXPHOS were downregulated, while most lated (Luo et al., 2017). MCT4 sustains a higher level of arsenic-induced
glycolysis-related genes were upregulated (Bi et al., 2020). Therefore, a glycolysis by exporting lactate to the extracellular environment and
key alteration of cells to adapt to long-term arsenic exposure is the maintaining the continuous pyruvate-to-lactate conversion (Pinheiro
metabolic switch from the tricarboxylic acid (TCA) cycle to glycolysis. et al., 2012). With the excretion of lactate, an acidic microenvironment
Arsenic-induced glycolysis may promote malignant transformation. gradually develops, facilitating the generation of cytokines (e.g., IL-8
Glycolysis is controlled by master regulatory genes, most significantly and IL-6) (Tan et al., 2015). These cytokines further promote
hypoxia-responsive transcription factors of the HIF family, such as arsenic-induced carcinogenesis (Johnson et al., 2018). In summary,
HIF1A (Chen et al., 2019; Yeung et al., 2008). Accumulating evidence arsenic exposure leads to aerobic glycolysis by increasing the stability of
suggests that HIF1A can be induced by arsenite (Bi et al., 2020; Zhao HIF1A protein and sustains glycolysis by increasing HIF1A-induced
et al., 2014). A marked increase in HIF1A protein levels was observed in MCT4 levels, which contributes to malignant transformation.
BEAS-2B cells exposed to 1 μM arsenite for 2 weeks (Zhao et al., 2014). In vivo studies demonstrated that arsenic-induced changes in glucose
Moreover, the accumulation of HIF1A induced by arsenite is persistent, metabolic pathways partially resemble glycolysis. Zebrafish exposed to
which is different from the transient activation of HIF1A induced by arsenite (20 mg/L, 96 hours) exhibited elevated expression of
hypoxia (Ginouvès et al., 2008; He et al., 2011; Zhao et al., 2014). Under glycolysis-related enzymes, such as glucokinase (GCK) and pyruvate
normal conditions, HIF1A is hydroxylated at specific proline residues by kinase (PK), increased levels of glycolytic intermediates (glucose-6-
PHDs, tagging it for ubiquitination and subsequent degradation by the phosphate (G6P) and pyruvate), and increased activity of PPP in the
proteasome pathway (Kaelin and Ratcliffe, 2008). In arsenite-exposed liver (Li et al., 2016). The PPP is composed of nonoxidative and oxida­
BEAS-2B cells, the levels of pyruvate and isocitrate, which inhibit PHD tive branches, both of which promote glycolysis (Chiara et al., 2012). In
function, were elevated, whereas the levels of α-KG, a cofactor of PHD to the contrast, arsenic exposure was reported to reduce the levels of

4
Y. Ruan et al. Ecotoxicology and Environmental Safety 229 (2022) 113098

glycolytic products and gene transcription. In BALB/c mice gavaged sterol regulatory element-binding protein (SREBP), are inhibited by
with arsenite (3 mg/kg/day, 12 days), metabolome analyses demon­ arsenic. In arsenite-exposed (5 μM, 72 hours) human hepatocellular
strated that plasma concentrations of lactate and alanine were signifi­ carcinoma cells (HepG2), the mRNA levels of LXR-β and SREBP1c were
cantly decreased, while the levels of isocitrate, α-KG, glutamate, and decreased (Padovani et al., 2010). In arsenite-exposed human promye­
citrate were increased (Garcia-Sevillano et al., 2014; Garciafigueroa locytic leukemia cells (HL-60) (5 μM, 48 hours), the expression of FASN
et al., 2013). The expression of glycolysis genes (e.g., Gck and Pk) in the was decreased at the protein level (Lei and Yinsheng, 2010). The protein
livers of C57BL/6J mice, remained decreased after 12 weeks of arsenite expression levels of FASN and SREBP1c were downregulated in the
exposure (1, 2, and 4 mg/L). This may indicate a “reverse Warburg ef­ livers of arsenic-exposed C57BKS/J (3 mg/L, 16 weeks) and C57BL/6J
fect”. During cancer development, cancer cells induce surrounding (100 μg/L, 5 weeks) mice (Adebayo et al., 2015; Liu et al., 2014).
stromal fibroblasts to undergo aerobic glycolysis, producing large Therefore, much evidence suggests that arsenic exposure suppresses
amounts of energy metabolites such as pyruvate and lactate, which are lipid synthesis.
transported to cancer cells and then enter the mitochondria to produce Arsenic exposure inhibits lipid transport. Arsenite exposure was
high levels of ATP through OXPHOS and fuel the TCA cycle to resist associated with significantly downregulated expression of carnitine
apoptosis and promote cancer cell proliferation (Pavlides et al., 2010, palmitoyltransferase (CPT) 1A, a fatty acid beta-oxidation (FAO) rate-
2009). Thus, a rational explanation for the role of arsenic-induced limiting enzyme, and thus accumulation of lipids, leading to malig­
glycolysis in carcinogenesis could be made by the Warburg effect or nant transformation in different cell lines (Du et al., 2017; Fang et al.,
the “reverse Warburg effect”. 2018; Stueckle et al., 2012). CPT1A is a direct target gene of HIF2A (Du
The enhancement of glycolysis by arsenic exposure has been vali­ et al., 2017). During arsenite-induced malignant transformation (5 μM,
dated in a variety of cellular models as mentioned above. However, the 30 weeks) of HK-2 cells, HIF2A was increased while CPT1A was
expression of glycolysis genes was inhibited by short-term arsenic decreased at the mRNA level at 20 weeks, which led to cellular lipid
exposure, but induced by long-term exposure. Cell metabolism, as an deposition. Thus, it was inferred that chronic arsenite exposure induced
adaptive response to the environment, may change with exposure time. abnormal lipid metabolism through HIF2A/CPT1A and resulted in the
Thus, a time-course study of metabolic reprogramming induced by formation of malignant cells with renal clear cell carcinoma (Fang et al.,
arsenic based on human internal exposure characteristics (time and 2018). In BEAS-2B cells exposed to arsenite (2.5 μM, 6 months),
dose) is needed. The lack of a comparable time frame for the experi­ decreased CPT1A expression at the protein level was associated with
ments and complicated cell-cell interactions in the microenvironment suppression of peroxisome proliferator-activated receptor (PPAR) α,
may help to explain these differences. As mentioned above, the "reverse which is crucial to the formation of tumors (Pirat et al., 2012; Stueckle
Warburg effect" observed in animal models may be caused by lactate et al., 2012). There were similar situations in the low-dose arsenic
shuttle, transferring lactate from the stroma to cancer cells for utiliza­ exposure in vivo experiments. Under low-dose arsenic exposure (0.5 or
tion (Whitaker-Menezes et al., 2011). Therefore, future studies consid­ 2 mg/L), serum carnitine levels increased in Sprague-Dawley rats (Wang
ering exposure characteristics and more advanced in vitro study systems et al., 2015). However, when the arsenic exposure dose reached
(such as coculture and 3D culture) are critical to collect more reliable 10 mg/L, the opposite results appeared in Sprague-Dawley rats. The
data. expression level of Cpt2 in the liver was increased, and serum carnitine
levels were decreased (Wang et al., 2015).
The limited existing evidence suggests that arsenic exposure induces
3.2. Arsenic exposure and lipid metabolism
lipid accumulation in some target cells. However, the conclusions are
dramatically different among reports with varied study designs,
Lipids, as important energy storage substances, play a major role in
including the applied experimental models and treatment protocols.
the energy supply (Long et al., 2018). In addition to aerobic glycolysis,
lipid metabolism has become a key energy source for cancer initiation
and progression (Beloribi-Djefaflia et al., 2016; Zhu et al., 2016). Ar­ 3.3. Arsenic exposure and amino acid metabolism
senic’s effects in adipose tissue also manifest in increased ectopic lipid
deposition in both the liver and skeletal muscle (Renu et al., 2018). The Disorders in amino acid metabolism appear to be a general reaction
incidence of fatty liver was significantly higher in arsenic-exposed to a variety of poisons in mammals (Connor et al., 2004). However, there
zebrafish (Bambino et al., 2018). Lipid deposition may be due to inhi­ are few reports on the topic of arsenic and amino acid metabolism.
bition of lipid transport and catabolism instead of enhanced lipid syn­ Experimental investigations concerning the impacts of arsenic exposure
thesis (Table 2) (Fig. 2B). on amino acid metabolism are summarized in Table 3. Arsenic exposure
Arsenic exposure inhibits lipid synthesis. Many lipid synthesis en­ leads to enhancement of the SGOC pathway and depletion of S-adeno­
zymes, such as liver X receptors (LXRs), fatty acid synthase (FASN), and sylmethionine (SAM) (Fig. 2C).

Table 2
Effects of arsenic exposure on lipid metabolism.
Subject/model Reagent Dose Expo time Metabolic changes Phenotype References

in vitro
HK-2 NaAsO2 2 or 5 μM 30 weeks Fatty acid ↑, HIF2A ↑, CPT1A ↓ Malignant transformation Fang et al. (2018)
BEAS-2B NaAsO2 2.5 μM 6 months PPARα/δ ↓ Malignant transformation Stueckle et al. (2012)
HepG2 NaAsO2 1, 3 or 5 μM 72 hours LXR-β and SREBP1c ↓ Cholesterol efflux attenuation Padovani et al. (2010)
HL-60 NaAsO2 5 μM 48 hours FASN ↓ Growth inhibition Lei and Yinsheng (2010)
in vivo
C57BL/6 J mice NaAsO2 100 μg/L 5 weeks Liver: SREBP1c, FASN, and LDL receptor↓ Atherosclerosis Adebayo et al. (2015)
(Drinking water)
C57BKS/J mice NaAsO2 3 mg/L 16 weeks Liver: PPARγ ↑; FASN and CD36 ↓ Type 2 diabetes Liu et al. (2014)
(Drinking water) Adipose: CD36, FASN and PPARγ ↑
Sprague-Dawley rats NaAsO2 10 mg/L 3 months Liver: Cpt2 ↑ Metabolic disruption (Wang et al., 2015)
(Drinking water) Serum: carnitine ↓
Sprague-Dawley rats NaAsO2 0.5 or 2 mg/L 3 months Serum: carnitine ↑ Metabolic disruption (Wang et al., 2015)
(Drinking water)

↑ Increase, ↓ decrease.

5
Y. Ruan et al. Ecotoxicology and Environmental Safety 229 (2022) 113098

Table 3
Effects of arsenic exposure on amino acid metabolism.
Subject/ Reagent Dose Expo time Metabolic changes Phenotype References
model

in vitro
BEAS-2B NaAsO2 0, 0.25, 0.5,1 or 6 months Serine, glycine and methionine ↑ Malignant Bi et al. (2020)
2 μM transformation
in vivo
Sprague- NaAsO2 0.5, 2, or 10 mg/L 3 months Serum: methionine and SAM ↓ Metabolic Wang et al. (2015)
Dawley rats (Drinking water) disruption
BALB/c mice NaAsO2 3 mg/kg/day 12 days Liver: arginine and tryptophan ↓; cysteine, glutamate, and Metabolic Garcia-Sevillano et al.
(Drinking water) methionine ↑ disruption (2014)
Kidney: arginine ↓
Plasma: arginine ↓; cysteine, glutamate ↑
Wistar rats NaAsO2 10 or 50 mg/L 6 months Serum: L-palmitoylcarnitine ↓ Metabolic Wang et al. (2014)
(Drinking water) disruption
ICR mice NaAsO2 5, 10, or 20 mg/ 24, Serum: valine, leucine, proline, phenylalanine, tryptophan, Metabolic Chen et al. (2017)
kg 48 hours tyrosine, asparagine, glutamate, and arginine ↑ disruption
(Gavage)
Zebrafish Na3AsO4 20 mg/L 96 hours Liver: threonine, glycine, and 2-aminobutanoate ↓ Early liver damages Liu et al. (2014)
(Aquatic
environment)

↑ Increase, ↓ decrease.

Synthetic serine is a vital precursor for the composite of lipids, considerable part of cancer metabolism (Zhao et al., 2020) and SAM is
proteins and nucleic acids and is essential for cancer proliferation. the key methyl donor for arsenic biometabolism (Tellez-Plaza et al.,
Glycine supplies methyls for the one-carbon metabolism that is used in 2014). Available studies on arsenic and amino acids are relatively
synthetic lipids, nucleic acids, and proteins, as well as DNA methylation limited. In addition, no systematic profile study has been conducted.
(Labuschagne et al., 2014). Glycine can be further converted from newly
synthesized serine from 3-phosphoglycerate and imported serine cata­ 4. Link between oxidative stress and metabolic reprogramming
lyzed by serine hydroxymethyltransferases (SHMTs) (Nikiforov et al., in response to arsenic exposure
2002). The bioactivities of glycine and serine and their associated
metabolic pathways are integral to malignant transformation and rapid Oxidative stress is defined as a disproportionate increase in reactive
cell proliferation, and play an important role as drivers of cancer initi­ oxygen species (ROS) in relation to antioxidant capacity (Hayes et al.,
ation and migration (Jain et al., 2012; Locasale, 2013). The SGOC 2020). Excess ROS damage biomolecules including lipids, proteins and
pathway partially intersects with the glycolysis pathway (Fig. 2). In nucleic acids, causing malfunction of the enzymes involved in meta­
arsenite-transformed malignant BEAS-2B cells, the major metabolic in­ bolism, and even cellular injury and cell death (B. Li et al., 2021).
termediates in the glycolytic pathway were diverted to glycolytic sub­ Typically, glucose utilization, fatty acid metabolism, and one-carbon
pathways, most significantly, the serine-glycine pathway (Bi et al., metabolism are affected by oxidative stress (Assies et al., 2014; Butter­
2020). The levels of the metabolites in the SGOC pathway were signif­ field and Halliwell, 2019). Oxidative stress caused by arsenic exposure is
icantly upregulated due to increased expression of the metabolic en­ suggested as the primary mechanism behind increased cancer risks
zymes SHMT, phosphoserine aminotransferase (PSAT), (Ganyc et al., 2007). Therefore, it is reasonable to conceive that oxida­
phosphoglycerate dehydrogenase (PHGDH), and glycolate oxidase tive modification of enzymes in metabolism gives rise to metabolic
(GOX) (Bi et al., 2020). Arsenic-exposed cells seem to prefer enhancing reprograming in arsenic-induced malignancies. However, emerging ev­
the uptake and synthesis of growth-essential amino acids (e.g., glycine idence suggests that exposure to arsenic at human-relevant doses causes
and serine) for proliferation (Fig. 2). Variations in the contents of amino no obvious elevation in intracellular ROS levels (Dodson et al., 2018;
acids, including glycine and serine, in response to arsenic have been Lau et al., 2013; Wu et al., 2019). Despite these controversial reports, it
observed in in vivo studies. However, due to the differences in experi­ is well accepted that oxidative stress occurs in response to arsenic
mental animal species, arsenic concentrations, administration routes, exposure. On the other hand, certain metabolic pathways or metabolites
and exposure times, the reported results are inconsistent and even are anti-oxidative or pro-oxidative. Up to now, few studies have been
contradictory. devoted to providing direct links between oxidative stress and metabolic
SAM is an essential methyl donor for many methyltransferases reprogramming under arsenic-exposed conditions (Hu et al., 2020). To
(Goering et al., 1999). The biometabolism of arsenic shares SAM with maintain stable control over ROS production, organisms have evolved a
the DNA methylation process (Goering et al., 1999). The level of SAM in complex array of defense systems, typically transcriptionally regulated
serum was decreased in arsenite-exposed (10 mg/L, 3 months) by NRF2, composed of antioxidants such as GSH and NADPH, and
Sprague-Dawley rats (Wang et al., 2015). Therefore, arsenic exposure is facilitated by key signaling pathway nodes such as AMPK (Filomeni
suspected to cause DNA hypomethylation. Hypomethylation of the et al., 2015; Hayes et al., 2020; Panieri and Santoro, 2016). Meanwhile,
promoter regions of the oncogenes c-Myc and c-Hras has been observed these factors are key to maintaining cell metabolism.
under arsenic-exposed conditions (Takahashi et al., 2002). DNA hypo­
methylation enhances gene expression while hypermethylation sup­ 4.1. NADPH and GSH
presses it (Zong et al., 2019). Global hypomethylation, which induces
genomic instability, also contributes to cell transformation (Klutstein NADPH and GSH serve as the bridge between oxidative stress and
et al., 2016). Thus, DNA hypomethylation by depletion of SAM may metabolism (Cioffi et al., 2019). NADPH is a reducing equivalent, which
contribute to arsenic carcinogenesis under some circumstances. is utilized by hundreds of reductive enzymes such as glutathione
Activation of SGOC pathway activation and reduced SAM levels have reductase (GR) as an electron donor to maintain the reduced form
been found in many cancers. These particular pathways or amino acids (Miller et al., 2018; Zhang et al., 2019). GSH, one of the prime antiox­
were first studied in response to arsenic exposure probably because they idants, is reversibly oxidized to glutathione disulfide (GSSG) during the
had been predicted to be changed. SGOC pathway activation is a scavenging of ROS. The link between energy production and oxidative

6
Y. Ruan et al. Ecotoxicology and Environmental Safety 229 (2022) 113098

stress becomes evident because GSH is regenerated from GSSG as an 4.1.3. FAO
NADPH-dependent process. NADPH is mainly derived from PPP, FAO is an important source of NADPH. A consistent fraction of the
glycolysis, SGOC, and FAO pathways (Cioffi et al., 2019; Lotocki and acetyl-CoA produced by FAO enters the TCA cycle and generates citrate,
Kakkar, 2020; Meister and Tate, 1976; Panieri and Santoro, 2016). GSH which is therefore exported into the cytosol and funneled into metabolic
maintenance is also closely related to the above four pathways (Fig. 3). reactions catalyzed by malic enzyme (ME) 1 and IDH1, ultimately pro­
ducing large amounts of NADPH (Carracedo et al., 2013). Arsenic was
4.1.1. PPP and glycolysis reported to inhibit the key FAO enzyme CPT1A in HK2 cells, leading to
Arsenic activates the PPP and enhances glycolysis in a variety of cells malignant transformation (Fang et al., 2018). However, more evidence
(Bi et al., 2020; Luo et al., 2017, 2016; Zhao et al., 2013). Glycolytic is needed for the effect of arsenic exposure on FAO regulation and the
intermediates can be shuttled into metabolic pathways that directly or involvement of FAO in arsenic carcinogenesis.
indirectly contribute to the generation of reducing equivalents, mainly
PPP-derived NADPH or glutaminolysis-derived GSH (Panieri and San­
4.2. NRF2
toro, 2016). PPP is a major catabolic pathway of glucose through which
cancer cells produce large amounts of ribose-5-phosphate, a precursor of
NRF2 is a well-known transcription factor regulating antioxidant
nucleotide synthesis and NADPH (Chiara et al., 2012). Thus, NADPH
defense. It is held in the cytoplasm under physiological conditions by
generated by PPP is found at the crossroad between unrestricted pro­
interacting with Kelch-like ECH-associated protein (KEAP) 1 and is
liferation and scavenging of excessive ROS (Patra and Hay, 2014).
subsequently degraded via the ubiquitin–proteasome pathway (Ville­
neuve et al., 2010). Arsenic was found to inhibit the Keap1-dependent
4.1.2. SGOC pathway
E3 ubiquitin ligase by augmenting the association of Keap1 with Cul3,
The SGOC pathway integrates inputs from amino acids (mainly
resulting in NRF2 nuclear translocation and accumulation (Wang et al.,
serine and glycine) and generates multiple carbon units (tetrahy­
2008). Recently, it was indicated that NRF2 activation by a
drofolate (THF) and its derivatives) (Panieri and Santoro, 2016). This
human-relevant dose of arsenic is p62-dependent due to blockade of
pathway has traditionally been associated with cancer cells due to its
autophagic flux (Lau et al., 2013; Wu et al., 2019).
importance in the synthesis of nucleic acids, lipids and proteins in
Once activated, NRF2 transcriptionally regulates genes involved in
proliferating cells. Serine-derived one-carbon units are primarily used
antioxidation (Mitsuishi et al., 2012). With respect to GSH and NADPH,
for nucleotide synthesis (Ducker and Rabinowitz, 2017), while serine is
NRF2 promotes GSH production and regeneration by activating genes
predominantly utilized in the mitochondria of mammalian cells to
encoding glutamate-cysteine ligase catalytic subunit (GCLC),
generate NADPH and stimulate GSH synthesis (Fan et al., 2014). In
glutamate-cysteine ligase modifier subunit (GCLM) and solute carrier
arsenite-transformed malignant BEAS-2B cells, the levels of metabolites,
family 7 member 11 (SLC7A11) (Habib et al., 2015; Shanmugam et al.,
including serine, glycine, and methionine in SGOC pathway were
2016; Yamamoto et al., 2018), and regulates the expression of enzymes
significantly upregulated (Bi et al., 2020). Given the emerging roles of
that synthesize NADPH, such as IDH1, ME1, 6-phosphogluconate de­
metabolic alterations in cancer cells and the intimate connection be­
hydrogenase (PGD), and glucose6-phosphate dehydrogenase (G6PD)
tween the SGOC pathway and redox homeostasis of human tumors, key
(He et al., 2020; Heiss et al., 2013). In addition, NRF2 transcriptionally
nodes in one-carbon metabolism might represent potential preventive
regulates enzymes that catalyze rate-limiting steps or are situated at
targets of carcinogenesis.
branching points in major metabolic pathways, such as G6PD, PGD,
transaldolase (TALDO) 1, and transketolase (TKT) in PPP, MTHFD2 and

Fig. 3. The link between oxidative stress and


metabolic reprogramming in response to
arsenic exposure. GSH, NADPH, NRF2, and
AMPK mediate metabolic reprogramming to
accommodate the disturbed redox status
induced by arsenic. Red lines represent upre­
gulated pathways. Blue lines represent down­
regulated pathways. ACC, acetyl-CoA
carboxylase; AMPK, adenosine 5’-mono­
phosphate-activated protein kinase; ENO,
enolase; GCLC, glutamate-cysteine ligase cata­
lytic subunit; GCLM, glutamate-cysteine ligase
modifier subunit; G6PD, glucose6-phosphate
dehydrogenase; GSH, glutathione; GSSG,
glutathione disulfide; HK, hexokinase; IDH,
isocitrate dehydrogenase; ME, malic enzyme;
MTHFD, methylenetetrahydrofolate dehydro­
genase; NADPH, nicotine adenine dis­
phosphonucleotide; NRF2, nuclear factor
erythroid 2-related factor 2; PGD, 6-phospho­
gluconate dehydrogenase; PGK, phosphoglyc­
erate kinase; PPAT, phosphoribosyl
pyrophosphate amidotransferase; PPP, pentose
phosphate pathway; SGOC, serine-glycine one-
carbon; TALDO, transaldolase; TKT,
transketolase.

7
Y. Ruan et al. Ecotoxicology and Environmental Safety 229 (2022) 113098

phosphoribosyl pyrophosphate amidotransferase (PPAT) in nucleotide more reliable information on how metabolism alters and thereby con­
synthesis, and FASN, acetyl-CoA carboxylase (ACC) 1, stearoyl CoA tributes to arsenic carcinogenesis. Moreover, no consistent metabolic
desaturase (SCD) 1, and ATP-citrate lyase (ACL) in lipid metabolism (He profile for arsenic-induced cancers has been reported because most
et al., 2020; Heiss et al., 2013) (Fig. 3). In the malignant transformation previous studies analyzed particular metabolites or metabolic pathways.
of BEAS-2B cells, arsenic induced NRF2-dependent activation of HIF1A, The evaluation of globally altered metabolites will help to identify new
which in turn coordinated with NRF2 for the transcription of genes in candidates and comprehensively target the key metabolic changes in
the PPP, glycolysis, and SGOC pathways, enhancing glucose uptake and response to arsenic exposure.
NADPH production (Bi et al., 2020; Heiss et al., 2013; Macleod et al., Perhaps by solving the above issues, we will gain insight into the
2009; Malhotra et al., 2010) (Fig. 3). NRF2 is considered to play a dual carcinogenic mechanisms of arsenic and identify the best way to
role in cancers. How NRF2 is activated and how NRF2 activation is perform prompt prevention to protect the population exposed to arsenic.
involved in arsenic carcinogenesis are still debated. The evolution of
redox status and metabolism in chronic arsenic exposure is probably Funding
coordinated by this particular transcription factor.
This work was supported by National Natural Science Foundation of
4.3. AMPK China (82022063 and 81573187), National Key R&D Program of China
(2018YFC1311600), and Key R&D Plan Guidance Project (2018225098)
Previous examinations have shown that arsenic causes activation of from Department of Science and Technology of Liaoning Province,
the cellular energy sensor AMPK (Chanda et al., 2008; Lee et al., 2006). China.
AMPK maintains NADPH levels by decreasing fatty acid synthesis and
increasing FAO through the inhibition of ACC1/2 (Jeon et al., 2012) Declaration of Competing Interest
(Fig. 3). Fatty acid synthesis is a major NADPH-consuming process
(Walsh et al., 2018). In contrast, FAO produces NADPH (Carracedo The authors declare that they have no known competing financial
et al., 2013). FAO activation and FAS inhibition can prevent early stages interests or personal relationships that could have appeared to influence
but promote late stages of carcinogenesis (Jeon, 2016). This suggests the work reported in this paper.
that AMPK activation is beneficial for cancer prevention. Moreover,
AMPK plays a determinant role in maintaining NADPH homeostasis in References
cancer cells upon energy stress, which is critical for cancer cell survival
(Jeon et al., 2012). Thus, caution is needed when considering the role of Adebayo, A.O., Zandbergen, F., Kozul-Horvath, C.D., Gruppuso, P.A., Hamilton, J.W.,
2015. Chronic exposure to low-dose arsenic modulates lipogenic gene expression in
AMPK in cancer because it performs both anti- and pro-tumorigenic mice. J. Biochem. Mol. Toxicol. 29, 1–9. https://doi.org/10.1002/jbt.21600.
roles, depending in part on the regulation of lipid metabolism. Ansary, T.M., Nakano, D., Nishiyama, A., 2019. Diuretic effects of sodium glucose
In summary, oxidative stress or redox disturbance induced by long- cotransporter 2 inhibitors and their influence on the renin-angiotensin system. Int. J.
Mol. Sci. 20, 629. https://doi.org/10.3390/ijms20030629.
term arsenic exposure may participate in metabolic reprogramming Assies, J., Mocking, R.J.T., Lok, A., Ruhé, H.G., Pouwer, F., Schene, A.H., 2014. Effects of
directly or indirectly and may subsequently be involved in carcinogen­ oxidative stress on fatty acid- and one-carbon-metabolism in psychiatric and
esis. This reminds us of the reciprocal crosstalk between redox homeo­ cardiovascular disease comorbidity. Acta Psychiatr. Scand. 130, 163–180. https://
doi.org/10.1111/acps.12265.
stasis and metabolism, which requires further study. ATSDR, 2015. Addendum To the Toxicological Profile for Dinitrophenols. Agency for
toxic substances and disease registry.
5. Conclusion and future perspectives Azoitei, N., Becher, A., Steinestel, K., Rouhi, A., Diepold, K., Genze, F., Simmet, T.,
Seufferlein, T., 2016. PKM2 promotes tumor angiogenesis by regulating HIF-1aα
through NF-κB activation. Mol. Cancer 15, 3. https://doi.org/10.1186/s12943-015-
Metabolic reprogramming is a hallmark of cancer, providing energy 0490-2.
and metabolites to support the rapid growth and proliferation of cancer Bambino, K., Zhang, C., Austin, C., Amarasiriwardena, C., Arora, M., Chu, J., Sadler, K.C.,
2018. Inorganic arsenic causes fatty liver and interacts with ethanol to cause
cells. Chronic exposure of the cells to arsenic induces malignant trans­
alcoholic liver disease in zebrafish. Dis. Model. Mech. 11, 31575. https://doi.org/
formation. Similar to other cancer cells, arsenic-transformed cells have 10.1242/DMM.031575.
an increased need for nutrients, energy, and biosynthetic activities to Baylin, S.B., Herman, J.G., 2000. DNA hypermethylation in tumorigenesis: epigenetics
joins genetics. Trends Genet. 16, 168–174. https://doi.org/10.1016/S0168-9525
produce all macromolecular components. The mechanism of arsenic
(99)01971-X.
carcinogenesis has been a focus of decades. However, little has been Beloribi-Djefaflia, S., Vasseur, S., Guillaumond, F., 2016. Lipid metabolic reprogramming
done to determine how changes in metabolism are related to arsenic in cancer cells. Oncogenesis 5, 189. https://doi.org/10.1038/oncsis.2015.49.
carcinogenesis. Arsenic exposure interferes with metabolic transduction Bi, Z., Zhang, Q., Fu, Y., Wadgaonkar, P., Zhang, W., Almutairy, B., Xu, L., Rice, M.,
Qiu, Y., Thakur, C., Chen, F., 2020. Nrf2 and HIF1α converge to arsenic-induced
pathways at different levels, including the gene expression, intracellular metabolic reprogramming and the formation of the cancer stem-like cells.
protein, protein function, and metabolite levels (Bi et al., 2020; Fang Theranostics 10, 4134–4149. https://doi.org/10.7150/thno.42903.
et al., 2018; Luo et al., 2016; Wu et al., 2019). Birts, C.N., Banerjee, A., Darley, M., Dunlop, C.R., Nelson, S., Nijjar, S.K., Parker, R.,
West, J., Tavassoli, A., Rose-Zerilli, M.J.J., Blaydes, J.P., 2020. p53 is regulated by
Although the role of metabolic reprogramming in arsenic carcino­ aerobic glycolysis in cancer cells by the CtBP family of NADH-dependent
genesis has been studied, the full range of influences has yet to be transcriptional regulators. Sci. Signal. 13, 9529. https://doi.org/10.1126/scisignal.
elucidated and requires further investigation. An in vivo experimental aau9529.
Biswas, S.K., 2015. Metabolic reprogramming of immune cells in cancer progression.
model for arsenic carcinogenesis has not been well established to date, Immunity 43, 435–449. https://doi.org/10.1016/j.immuni.2015.09.001.
and many mechanistic studies have been conducted in vitro. Cell culture Bjorklund, G., Aaseth, J., Chirumbolo, S., Urbina, M.A., Uddin, R., 2018. Effects of
models facilitate the characterization of malignant transformation by arsenic toxicity beyond epigenetic modifications. Environ. Geochem. Health 40,
955–965. https://doi.org/10.1007/s10653-017-9967-9.
arsenic. However, they have failed to provide information on metabolic
Bravatà, V., Tinganelli, W., Cammarata, F.P., Minafra, L., Calvaruso, M., Sokol, O.,
reprogramming with the microenvironment. Therefore, it is imperative Petringa, G., Cirrone, G.A.P., Scifoni, E., Forte, G.I., Russo, G., 2021. Hypoxia
to develop more advanced in vitro study system or appropriate in vivo transcriptomic modifications induced by proton irradiation in u87 glioblastoma
multiforme cell line. J. Pers. Med. 11, 308. https://doi.org/10.3390/jpm11040308.
models for study. The wide variations in applied concentration, treat­
Butterfield, D.A., Halliwell, B., 2019. Oxidative stress, dysfunctional glucose metabolism
ment duration, and cell types/tissues have led to inconsistent results and and Alzheimer disease. Nat. Rev. Neurosci. 20, 148–160. https://doi.org/10.1038/
conclusions. Arsenic-induced tumorigenesis is a chronic process. Only a s41583-019-0132-6.
small proportion of studies have investigated the metabolic activities Carlson, P., Van Beneden, R.J., 2014. Arsenic exposure alters expression of cell cycle and
lipid metabolism genes in the liver of adult zebrafish (Danio rerio). Aquat. Toxicol.
after long-term exposure to arsenic. Studies mimicking the characteris­ 153, 66–72. https://doi.org/10.1016/j.aquatox.2013.10.006.
tics of arsenic exposure in humans (long-term and internal exposure
doses) should be fully considered in the future. Such studies will provide

8
Y. Ruan et al. Ecotoxicology and Environmental Safety 229 (2022) 113098

Carracedo, A., Cantley, L.C., Pandolfi, P.P., 2013. Cancer metabolism: fatty acid Garcia-Sevillano, M.A., Contreras-Acuna, M., Garcia-Barrera, T., Navarro, F., Gomez-
oxidation in the limelight. Nat. Rev. Cancer 13, 227–232. https://doi.org/10.1038/ Ariza, J.L., 2014. Metabolomic study in plasma, liver and kidney of mice exposed to
nrc3483. inorganic arsenic based on mass spectrometry. Anal. Bioanal. Chem. 406,
Chanda, D., Kim, S.J., Lee, I.K., Shong, M., Choi, H.S., 2008. Sodium arsenite induces 1455–1469. https://doi.org/10.1007/s00216-013-7564-z.
orphan nuclear receptor SHP gene expression via AMP-activated protein kinase to Garciafigueroa, D.Y., Klei, L.R., Ambrosio, F., Barchowsky, A., 2013. Arsenic-stimulated
inhibit gluconeogenic enzyme gene expression. Am. J. Physiol. Endocrinol. Metab. lipolysis and adipose remodeling is mediated by G-protein-coupled receptors.
295, 368–379. https://doi.org/10.1152/ajpendo.00800.2007. Toxicol. Sci. 134, 335–344. https://doi.org/10.1093/toxsci/kft108.
Chen, C., Lin, P., Tsai, M., Lee, H., 2017. Targeted lipidomics profiling of acute arsenic Ginouvès, A., Ilc, K., Macías, N., Pouysségur, J., Berra, E., 2008. PHDs overactivation
exposure in mice serum by on-line solid-phase extraction stable-isotope dilution during chronic hypoxia “desensitizes” HIFalpha and protects cells from necrosis.
liquid chromatography-tandem mass spectrometry. Arch. Toxicol. 91, 3079–3091. Proc. Natl. Acad. Sci. USA 105, 4745–4750. https://doi.org/10.1073/
https://doi.org/10.1007/s00204-017-1937-6. pnas.0705680105.
Chen, C.J., Hsu, L.I., Wang, C.H., Shih, W.L., Hsu, Y.H., Tseng, M.P., Lin, Y.C., Chou, W. Goering, P.L., Aposhian, H.V., Mass, M.J., Cebrián, M., Beck, B.D., Waalkes, M.P., 1999.
L., Chen, C.Y., Lee, C.Y., Wang, L.H., Cheng, Y.C., Chen, C.L., Chen, S.Y., Wang, Y.H., The enigma of arsenic carcinogenesis: role of metabolism. Toxicol. Sci. 49, 5–14.
Hsueh, Y.M., Chiou, H.Y., Wu, M.M., 2005. Biomarkers of exposure, effect, and https://doi.org/10.1093/toxsci/49.1.5.
susceptibility of arsenic-induced health hazards in Taiwan. Toxicol. Appl. Goetze, K., Walenta, S., Ksiazkiewicz, M., Kunz-Schughart, L.A., Mueller-Klieser, W.,
Pharmacol. 206, 198–206. https://doi.org/10.1016/j.taap.2004.10.023. 2011. Lactate enhances motility of tumor cells and inhibits monocyte migration and
Chen, F., Chen, J., Yang, L., Liu, J., Zhang, X., Zhang, Y., Tu, Q., Yin, D., Lin, D., cytokine release. Int. J. Oncol. 39, 453–463. https://doi.org/10.3892/
Wong, P.-P., Huang, D., Xing, Y., Zhao, J., Li, M., Liu, Q., Su, F., Su, S., Song, E., ijo.2011.1055.
2019. Extracellular vesicle-packaged HIF-1α-stabilizing lncRNA from tumour- Gu, Z., Liu, Y., Cai, F., Patrick, M., Zmajkovic, J., Cao, H., Zhang, Y., Tasdogan, A.,
associated macrophages regulates aerobic glycolysis of breast cancer cells. Nat. Cell Chen, M., Qi, L., Liu, X., Li, K., Lyu, J., Dickerson, K.E., Chen, W., Ni, M., Merritt, M.
Biol. 21, 498–510. https://doi.org/10.1038/s41556-019-0299-0. E., Morrison, S.J., Skoda, R.C., Deberardinis, R.J., Xu, J., 2019. Loss of EZH2
Chiara, R., Elena, G., Manuela, P., Elisabetta, A., Dario, G., 2012. The pentose phosphate reprograms BCAA metabolism to drive leukemic transformation. Cancer Discov. 9,
pathway: an antioxidant defense and a crossroad in tumor cell fate. Free Radic. Biol. 1228–1247. https://doi.org/10.1158/2159-8290.CD-19-0152.
Med. 53, 421–436. https://doi.org/10.1016/J.FREERADBIOMED.2012.05.006. Habib, E., Linher-Melville, K., Lin, H.X., Singh, G., 2015. Expression of xCT and activity
Cioffi, F., Adam, R.H.I., Broersen, K., 2019. Molecular mechanisms and genetics of of system xc- are regulated by NRF2 in human breast cancer cells in response to
oxidative stress in Alzheimer’s disease. J. Alzheimer’s Dis. 72, 981–1017. https:// oxidative stress. Redox Biol. 5, 33–42. https://doi.org/10.1016/j.
doi.org/10.3233/JAD-190863. redox.2015.03.003.
Coco, L.Del, Majellaro, M., Boccarelli, A., Cellamare, S., Altomare, C.D., Fanizzi, F.P., Hawkins, K.E., Joy, S., Delhove, J.M.K.M.K.M., Kotiadis, V.N., Fernandez, E.,
2020. Novel antiproliferative biphenyl nicotinamide: NMR metabolomic study of its Fitzpatrick, L.M., Whiteford, J.R., King, P.J., Bolanos, J.P., Duchen, M.R.,
effect on the MCF-7 cell in comparison with cisplatin and vinblastine. Molecules 25, Waddington, S.N., McKay, T.R., 2016. NRF2 orchestrates the metabolic shift during
3502. https://doi.org/10.3390/molecules25153502. induced pluripotent stem cell reprogramming. Cell Rep. 4, 1883–1891. https://doi.
Connor, S.C., Wu, W., Sweatman, B.C., Manini, J., Haselden, J.N., Crowther, D.J., org/10.1016/j.celrep.2016.02.003.
Waterfield, C.J., 2004. Effects of feeding and body weight loss on the 1H NMR-based Hayes, J.D., Dinkova-Kostova, A.T., Tew, K.D., 2020. Oxidative stress in cancer. Cancer
urine metabolic profiles of male Wistar Han rats: implications for biomarker Cell 38, 167–197. https://doi.org/10.1016/j.ccell.2020.06.001.
discovery. Biomarkers 9, 156–179. https://doi.org/10.1080/ He, F., Antonucci, L., Yamachika, S., Zhang, Z., Taniguchi, K., Umemura, A.,
13547500410001720767. Hatzivassiliou, G., Roose-Girma, M., Reina-Campos, M., Duran, A., Diaz-Meco, M.T.,
Cui, Y.H., Yang, S., Wei, J., Shea, C.R., Zhong, W., Wang, F., Shah, P., Kibriya, M.G., Moscat, J., Sun, B., Karin, M., 2020. NRF2 activates growth factor genes and
Cui, X., Ahsan, H., He, C., He, Y.Y., 2021. Autophagy of the m6A mRNA demethylase downstream AKT signaling to induce mouse and human hepatomegaly. J. Hepatol.
FTO is impaired by low-level arsenic exposure to promote tumorigenesis. Nat. 72, 1182–1195. https://doi.org/10.1016/j.jhep.2020.01.023.
Commun. 12, 2183. https://doi.org/10.1038/s41467-021-22469-6. He, J., Liu, W., Ge, X., Wang, G.-C., Desai, V., Wang, S., Mu, W., Bhardwaj, V., Seifert, E.,
Dang, C.V., 2012. Links between metabolism and cancer. Genes Dev. 26, 877–890. Liu, L.-Z., Bhushan, A., Peiper, S.C., Jiang, B.-H., 2019. Arsenic-induced metabolic
https://doi.org/10.1101/gad.189365.112. shift triggered by the loss of miR-199a-5p through Sp1-dependent DNA methylation.
Dodson, M., de la Vega, M.R., Harder, B., Castro-Portuguez, R., Rodrigues, S.D., Wong, P. Toxicol. Appl. Pharmacol. 378, 114606 https://doi.org/10.1016/j.
K., Chapman, E., Zhang, D.D., 2018. Low-level arsenic causes proteotoxic stress and taap.2019.114606.
not oxidative stress. Toxicol. Appl. Pharmacol. 341, 106–113. https://doi.org/ He, Q., Gao, Z., Yin, J., Zhang, J., Yun, Z., Ye, J., 2011. Regulation of HIF-1α activity in
10.1016/j.taap.2018.01.014. adipose tissue by obesity-associated factors: adipogenesis, insulin, and hypoxia. Am.
Du, W., Zhang, L., Brett-Morris, A., Aguila, B., Kerner, J., Hoppel, C.L., Puchowicz, M., J. Physiol. Endocrinol. Metab. 300, 877–885. https://doi.org/10.1152/
Serra, D., Herrero, L., Rini, B.I., Campbell, S., Welford, S.M., 2017. HIF drives lipid ajpendo.00626.2010.
deposition and cancer in ccRCC via repression of fatty acid metabolism. Nat. Heiss, E.H., Schachner, D., Zimmermann, K., Dirsch, V.M., 2013. Glucose availability is a
Commun. 8, 1769. https://doi.org/10.1038/s41467-017-01965-8. decisive factor for Nrf2-mediated gene expression. Redox Biol. 1, 359–365. https://
Ducker, G.S., Rabinowitz, J.D., 2017. One-carbon metabolism in health and disease. Cell doi.org/10.1016/j.redox.2013.06.001.
Metab. 25, 27–42. https://doi.org/10.1016/j.cmet.2016.08.009. Hoxhaj, G., Manning, B.D., 2020. The PI3K–AKT network at the interface of oncogenic
Eckstein, M., Eleazer, R., Rea, M., Fondufe-Mittendorf, Y., 2017. Epigenomic signalling and cancer metabolism. Nat. Rev. Cancer 20, 74–88. https://doi.org/
reprogramming in inorganic arsenic-mediated gene expression patterns during 10.1038/s41568-019-0216-7.
carcinogenesis. Rev. Environ. Health 32, 93–103. https://doi.org/10.1515/reveh- Hu, Y., Li, J., Lou, B., Wu, R., Wang, G., Lu, C., Wang, H., Pi, J., Xu, Y., 2020. The role of
2016-0025. reactive oxygen species in arsenic toxicity. Biomolecules 10, 240. https://doi.org/
Eijkelenkamp, K., Osinga, T.E., Links, T.P., van der Horst-Schrivers, A.N.A.A., 2020. 10.3390/BIOM10020240.
Clinical implications of the oncometabolite succinate in SDHx-mutation carriers. IARC, 2012. IARC Monographs: Arsenic, Metals, Fibres, and Dusts. IARC Monogr. Eval.
Clin. Genet. 97, 39–53. https://doi.org/10.1111/cge.13553. Carcinog. Risks to Humans 100C, 1–526.
Fan, J., Ye, J., Kamphorst, J.J., Shlomi, T., Thompson, C.B., Rabinowitz, J.D., 2014. Jain, M., Nilsson, R., Sharma, S., Madhusudhan, N., Kitami, T., Souza, A.L., Kafri, R.,
Quantitative flux analysis reveals folate-dependent NADPH production. Nature 510, Kirschner, M.W., Clish, C.B., Mootha, V.K., M, J., R, N., S, S., N, M., T, K., AL, S., R,
298–302. https://doi.org/10.1038/nature13236. K., MW, K., CB, C., VK, M., 2012. Metabolite profiling identifies a key role for glycine
Fang, X., Sun, R., Hu, Y., Wang, H., Guo, Y., Yang, B., Pi, J., Xu, Y., 2018. miRNA-182-5p, in rapid cancer cell proliferation. Science 336, 1040–1044. https://doi.org/10.1126/
via HIF2α, contributes to arsenic carcinogenesis: evidence from human renal science.1218595.
epithelial cells. Metallomics 10, 1607–1617. https://doi.org/10.1039/c8mt00251g. Jeon, S.M., 2016. Regulation and function of AMPK in physiology and diseases. Exp. Mol.
Faubert, B., Solmonson, A., DeBerardinis, R.J., 2020. Metabolic reprogramming and Med. 48, e245 https://doi.org/10.1038/emm.2016.81.
cancer progression. Science 368, 5473. https://doi.org/10.1126/science.aaw5473. Jeon, S.M., Chandel, N.S., Hay, N., 2012. AMPK regulates NADPH homeostasis to
Figueroa, M.E., Abdel-Wahab, O., Lu, C., Ward, P.S., Patel, J., Shih, A., Li, Y., promote tumour cell survival during energy stress. Nature 485, 661–665. https://
Bhagwat, N., Vasanthakumar, A., Fernandez, H.F., Tallman, M.S., Sun, Z., doi.org/10.1038/nature11066.
Wolniak, K., Peeters, J.K., Liu, W., Choe, S.E., Fantin, V.R., Paietta, E., Johnson, D.E., O’Keefe, R.A., Grandis, J.R., 2018. Targeting the IL-6/JAK/STAT3
Löwenberg, B., Licht, J.D., Godley, L.A., Delwel, R., Valk, P.J.M., Thompson, C.B., signalling axis in cancer. Nat. Rev. Clin. Oncol. 15, 234–248. https://doi.org/
Levine, R.L., Melnick, A., 2010. Leukemic IDH1 and IDH2 mutations result in a 10.1038/nrclinonc.2018.8.
hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic Kaelin, W.G., Ratcliffe, P.J., 2008. Oxygen sensing by metazoans: the central role of the
differentiation. Cancer Cell 18, 553–567. https://doi.org/10.1016/j. HIF hydroxylase pathway. Mol. Cell 30, 392–402. https://doi.org/10.1016/j.
ccr.2010.11.015. molcel.2008.04.009.
Filomeni, G., De Zio, D., Cecconi, F., 2015. Oxidative stress and autophagy: the clash Klutstein, M., Nejman, D., Greenfield, R., Cedar, H., 2016. DNA methylation in cancer
between damage and metabolic needs. Cell Death Differ. 22, 377–388. https://doi. and aging. Cancer Res. 76, 3446–3450. https://doi.org/10.1158/0008-5472.CAN-
org/10.1038/cdd.2014.150. 15-3278.
Ganyc, D., Talbot, S., Konate, F., Jackson, S., Schanen, B., Cullen, W., Self, W.T., 2007. Labuschagne, C.F., van den Broek, N.J.F., Mackay, G.M., Vousden, K.H., Maddocks, O.D.
Impact of trivalent arsenicals on selenoprotein synthesis. Environ. Health Perspect. K., 2014. Serine, but not glycine, supports one-carbon metabolism and proliferation
115, 346–353. https://doi.org/10.1289/ehp.9440. of cancer cells. Cell Rep. 7, 1248–1258. https://doi.org/10.1016/j.
Gao, N., Yao, X., Jiang, L., Yang, L., Qiu, T., Wang, Z., Pei, P., Yang, G., Liu, X., Sun, X., celrep.2014.04.045.
2019. Taurine improves low-level inorganic arsenic-induced insulin resistance by Lau, A., Zheng, Y., Tao, S., Wang, H., Whitman, S.A., White, E., Zhang, D.D., 2013.
activating PPARγ-mTORC2 signalling and inhibiting hepatic autophagy. J. Cell. Arsenic inhibits autophagic flux, activating the Nrf2-Keap1 pathway in a p62-
Physiol. 234, 5143–5152. https://doi.org/10.1002/jcp.27318. dependent manner. Mol. Cell. Biol. 33, 2436–2446. https://doi.org/10.1128/
mcb.01748-12.

9
Y. Ruan et al. Ecotoxicology and Environmental Safety 229 (2022) 113098

Lee, M., Hwang, J.T., Yun, H., Kim, E.J., Kim, M.J., Kim, S.S., Ha, J., 2006. Critical roles Patra, K.C., Hay, N., 2014. The pentose phosphate pathway and cancer. Trends Biochem.
of AMP-activated protein kinase in the carcinogenic metal-induced expression of Sci. 39, 347–354. https://doi.org/10.1016/j.tibs.2014.06.005.
VEGF and HIF-1 proteins in DU145 prostate carcinoma. Biochem. Pharmacol. 72, Pavlides, S., Tsirigos, A., Vera, I., Flomenberg, N., Frank, P.G., Casimiro, M.C., Wang, C.,
91–103. https://doi.org/10.1016/j.bcp.2006.03.021. Pestell, R.G., Martinez-Outschoorn, U.E., Howell, A., Sotgia, F., Lisanti, M.P., 2010.
Lee, N., Kim, D., 2016. Cancer metabolism: fueling more than just growth. Mol. Cells 39, Transcriptional evidence for the “Reverse Warburg Effect” in human breast cancer
847–854. https://doi.org/10.14348/molcells.2016.0310. tumor stroma and metastasis: Similarities with oxidative stress, inflammation,
Lei, X., Yinsheng, W., 2010. Quantitative proteomic analysis reveals the perturbation of Alzheimer’s disease, and “Neuron-Glia Metabolic Coupling.”. Aging 2, 185–199.
multiple cellular pathways in HL-60 cells induced by arsenite treatment. J. Proteome https://doi.org/10.18632/aging.100134.
Res. 9, 1129–1137. https://doi.org/10.1021/pr9011359. Pavlides, S., Whitaker-Menezes, D., Castello-Cros, R., Flomenberg, N., Witkiewicz, A.K.,
Li, B., Sun, C., Lin, X., Busch, W., 2021. The emerging role of GSNOR in oxidative stress Frank, P.G., Casimiro, M.C., Wang, C., Fortina, P., Addya, S., Pestell, R.G., Martinez-
regulation. Trends Plant Sci. 26, 156–168. https://doi.org/10.1016/j. Outschoorn, U.E., Sotgia, F., Lisanti, M.P., 2009. The reverse Warburg effect: aerobic
tplants.2020.09.004. glycolysis in cancer associated fibroblasts and the tumor stroma. Cell Cycle 8,
Li, C., Li, P., Tan, Y.M., Lam, S.H., Chan, E.C., Gong, Z., 2016. Metabolomic 3984–4001. https://doi.org/10.4161/cc.8.23.10238.
characterizations of liver injury caused by acute arsenic toxicity in zebrafish. PLoS Pavlova, N.N., Thompson, C.B., 2016. The emerging hallmarks of cancer metabolism.
One 11, e0151225. https://doi.org/10.1371/journal.pone.0151225. Cell Metab. 23, 27–47. https://doi.org/10.1016/j.cmet.2015.12.006.
Li, J., Liu, H., Liu, X., Hao, S., Zhang, Z., Xuan, H., 2021. Chinese poplar propolis inhibits Pinheiro, C., Longatto-Filho, A., Azevedo-Silva, J., Casal, M., Schmitt, F.C., Baltazar, F.,
MDA-MB-231 cell proliferation in an inflammatory microenvironment by targeting 2012. Role of monocarboxylate transporters in human cancers: state of the art.
enzymes of the glycolytic pathway. J. Immunol. Res. 2021, 6641341 https://doi. J. Bioenerg. Biomembr. 44, 127–139. https://doi.org/10.1007/s10863-012-9428-1.
org/10.1155/2021/6641341. Pinweha, P., Rattanapornsompong, K., Charoensawan, V., Jitrapakdee, S., 2016.
Liao, Y., Zhao, T., Li, L.Y., Wang, F.Q., 2021. Elevated Sad1 and UNC84 Domain MicroRNAs and oncogenic transcriptional regulatory networks controlling metabolic
Containing 2 (SUN2) level inhibits cell growth and aerobic glycolysis in oral cancer reprogramming in cancers. Comput. Struct. Biotechnol. J. 14, 223–233. https://doi.
through reducing the expressions of glucose transporter 1 (GLUT1) and lactate org/10.1016/j.csbj.2016.05.005.
dehydrogenase A (LDHA). J. Dent. Sci. 16, 460–466. https://doi.org/10.1016/j. Pirat, C., Farce, A., Lebègue, N., Renault, N., Furman, C., Millet, R., Yous, S., Speca, S.,
jds.2020.08.007. Berthelot, P., Desreumaux, P., Chavatte, P., 2012. Targeting peroxisome proliferator-
Liu, S., Guo, X., Wu, B., Yu, H., Zhang, X., Li, M., 2014. Arsenic induces diabetic effects activated receptors (PPARs): development of modulators. J. Med. Chem. 55,
through beta-cell dysfunction and increased gluconeogenesis in mice. Sci. Rep. 4, 4027–4061. https://doi.org/10.1021/jm101360s.
6894. https://doi.org/10.1038/srep06894. Podgorski, J., Berg, M., 2020. Global threat of arsenic in groundwater. Science 368,
Locasale, J.W., 2013. Serine, glycine and the one-carbon cycle: cancer metabolism in full 845–850. https://doi.org/10.1126/science.aba1510.
circle. Nat. Rev. Cancer 13, 572–583. https://doi.org/10.1038/nrc3557. Renu, K., Madhyastha, H., Madhyastha, R., Maruyama, M., Arunachlam, S., Abilash, V.
Long, J., Zhang, C., Zhu, N., Du, K., Yin, Y., Tan, X., Liao, D., Qin, L., 2018. Lipid G., 2018. Role of arsenic exposure in adipose tissue dysfunction and its possible
metabolism and carcinogenesis, cancer development. Am. J. Cancer Res. 8, 778–791. implication in diabetes pathophysiology. Toxicol. Lett. 284, 86–95. https://doi.org/
Lotocki, V., Kakkar, A., 2020. Miktoarm star polymers: branched architectures in drug 10.1016/j.toxlet.2017.11.032.
delivery. Pharmaceutics 12, 1–37. https://doi.org/10.3390/ Rivas-Santiago, C., Gonzalez-Curiel, I., Zarazua, S., Murgu, M., Ruiz Cardona, A.,
pharmaceutics12090827. Lazalde, B., Lara-Ramirez, E.E., Vazquez, E., Castaneda-Delgado, J.E., Rivas-
Luo, F., Liu, X., Ling, M., Lu, L., Shi, L., Lu, X., Li, J., Zhang, A., Liu, Q., 2016. The lncRNA Santiago, B., Lopez, J.A., Cervantes-Villagrana, A.R., Lopez-Hernandez, Y., 2019.
MALAT1, acting through HIF-1α stabilization, enhances arsenite-induced glycolysis Lipid metabolism alterations in a rat model of chronic and intergenerational
in human hepatic L-02 cells. Biochim. Biophys. Acta 1862, 1685–1695. https://doi. exposure to arsenic. Biomed. Res. Int. 2019, 4978018 https://doi.org/10.1155/
org/10.1016/j.bbadis.2016.06.004. 2019/4978018.
Luo, F., Zou, Z., Liu, X., Ling, M., Wang, Q., Wang, Q., Lu, L., Shi, L., Liu, Y., Liu, Q., Sajnani, K., Islam, F., Smith, R.A., Gopalan, V., Lam, A.K.-Y., 2017. Genetic alterations in
Zhang, A., 2017. Enhanced glycolysis, regulated by HIF-1alpha via MCT-4, promotes Krebs cycle and its impact on cancer pathogenesis. Biochimie 135, 164–172. https://
inflammation in arsenite-induced carcinogenesis. Carcinogenesis 38, 615–626. doi.org/10.1016/j.biochi.2017.02.008.
https://doi.org/10.1093/carcin/bgx034. Senni, N., Savall, M., Cabrerizo Granados, D., Alves-Guerra, M.C., Sartor, C., Lagoutte, I.,
Macleod, A.K., Mcmahon, M., Plummer, S.M., Higgins, L.G., Penning, T.M., Igarashi, K., Gougelet, A., Terris, B., Gilgenkrantz, H., Perret, C., Colnot, S., Bossard, P., 2019.
Hayes, J.D., 2009. Characterization of the cancer chemopreventive NRF2-dependent β-catenin-activated hepatocellular carcinomas are addicted to fatty acids. Gut 68,
gene battery in human keratinocytes: demonstration that the KEAP1-NRF2 pathway, 322–334. https://doi.org/10.1136/gutjnl-2017-315448.
and not the BACH1-NRF2 pathway, controls cytoprotection against electrophiles as Shanmugam, T., Selvaraj, M., Poomalai, S., 2016. Epigallocatechin gallate potentially
well as redox-cycling compounds. Carcinogenesis 30, 1571–1580. https://doi.org/ abrogates fluoride induced lung oxidative stress, inflammation via Nrf2/Keap1
10.1093/carcin/bgp176. signaling pathway in rats: an in-vivo and in-silico study. Int. Immunopharmacol. 39,
Malhotra, D., Portales-Casamar, E., Singh, A., Srivastava, S., Arenillas, D., Happel, C., 128–139. https://doi.org/10.1016/j.intimp.2016.07.022.
Shyr, C., Wakabayashi, N., Kensler, T.W., Wasserman, W.W., Biswal, S., 2010. Global Somarribas Patterson, L.F., Vardhana, S.A., 2021. Metabolic regulation of the cancer-
mapping of binding sites for Nrf2 identifies novel targets in cell survival response immunity cycle. Trends Immunol. 42, 975–993. https://doi.org/10.1016/j.
through chip-seq profiling and network analysis. Nucleic Acids Res. 38, 5718–5734. it.2021.09.002.
https://doi.org/10.1093/nar/gkq212. Stueckle, T.A., Lu, Y., Davis, M.E., Wang, L., Jiang, B.H., Holaskova, I., Schafer, R.,
Mandal, P., 2017. Molecular insight of arsenic-induced carcinogenesis and its prevention. Barnett, J.B., Rojanasakul, Y., 2012. Chronic occupational exposure to arsenic
Naunyn. Schmiede. Arch. Pharmacol. 390, 443–455. https://doi.org/10.1007/ induces carcinogenic gene signaling networks and neoplastic transformation in
s00210-017-1351-x. human lung epithelial cells. Toxicol. Appl. Pharmacol. 261, 204–216. https://doi.
Mead, M.N., 2005. Arsenic: in search of an antidote to a global poison. Environ. Health org/10.1016/j.taap.2012.04.003.
Perspect. 113, 378–386. https://doi.org/10.1289/ehp.113-a378. Sun, L., Yang, X., Huang, X., Yao, Y., Wei, X., Yang, S., Zhou, D., Zhang, W., Long, Z.,
Meister, A., Tate, S.S., 1976. Glutathione and related gamma-glutamyl compounds: Xu, X., Zhu, X., He, S., Su, X., 2021. 2-Hydroxylation of fatty acids represses
biosynthesis and utilization. Annu. Rev. Biochem. 45, 559–604. https://doi.org/ colorectal tumorigenesis and metastasis via the YAP transcriptional axis. Cancer Res.
10.1146/annurev.bi.45.070176.003015. 81, 289–302. https://doi.org/10.1158/0008-5472.CAN-20-1517.
Miller, C.G., Holmgren, A., Arnér, E.S.J., Schmidt, E.E., 2018. NADPH-dependent and Takahashi, M., Barrett, J.C., Tsutsui, T., 2002. Transformation by inorganic arsenic
-independent disulfide reductase systems. Free Radic. Biol. Med. 127, 248–261. compounds of normal Syrian hamster embryo cells into a neoplastic state in which
https://doi.org/10.1016/j.freeradbiomed.2018.03.051. they become anchorage-independent and cause tumors in newborn hamsters. Int. J.
Mitsuishi, Y., Motohashi, H., Yamamoto, M., 2012. The Keap1–Nrf2 system in cancers: Cancer 99, 629–634. https://doi.org/10.1002/ijc.10407.
stress response and anabolic metabolism. Front. Oncol. 2, 200. https://doi.org/ Tan, Z., Xie, N., Banerjee, S., Cui, H., Fu, M., Thannickal, V.J., Liu, G., 2015. The
10.3389/fonc.2012.00200. monocarboxylate transporter 4 is required for glycolytic reprogramming and
Nikiforov, M.A., Chandriani, S., O’Connell, B., Petrenko, O., Kotenko, I., Beavis, A., inflammatory response in macrophages. J. Biol. Chem. 290, 46–55. https://doi.org/
Sedivy, J.M., Cole, M.D., 2002. A functional screen for myc-responsive genes reveals 10.1074/jbc.M114.603589.
serine hydroxymethyltransferase, a major source of the one-carbon unit for cell Tannahill, G.M., Curtis, A.M., Adamik, J., Palsson-McDermott, E.M., McGettrick, A.F.,
metabolism. Mol. Cell. Biol. 22, 5793–5800. https://doi.org/10.1128/ Goel, G., Frezza, C., Bernard, N.J., Kelly, B., Foley, N.H., Zheng, L., Gardet, A.,
mcb.22.16.5793-5800.2002. Tong, Z., Jany, S.S., Corr, S.C., Haneklaus, M., Caffrey, B.E., Pierce, K., Walmsley, S.,
Nilsson, R., Jain, M., Madhusudhan, N., Sheppard, N.G., Strittmatter, L., Kampf, C., Beasley, F.C., Cummins, E., Nizet, V., Whyte, M., Taylor, C.T., Lin, H., Masters, S.L.,
Huang, J., Asplund, A., Mootha, V.K., 2014. Metabolic enzyme expression highlights Gottlieb, E., Kelly, V.P., Clish, C., Auron, P.E., Xavier, R.J., O’Neill, L.A.J., 2013.
a key role for MTHFD2 and the mitochondrial folate pathway in cancer. Nat. Succinate is an inflammatory signal that induces IL-1β through HIF-1α. Nature 496,
Commun. 5, 3128. https://doi.org/10.1038/ncomms4128. 238–242. https://doi.org/10.1038/nature11986.
Padovani, A.M.S., Molina, M.F., Mann, K.K., 2010. Inhibition of liver x receptor/retinoid Tellez-Plaza, M., Tang, W.Y., Shang, Y., Umans, J.G., Francesconi, K.A., Goessler, W.,
X receptor-mediated transcription contributes to the proatherogenic effects of Ledesma, M., Leon, M., Laclaustra, M., Pollak, J., Guallar, E., Cole, S.A., Fallin, M.D.,
arsenic in macrophages in vitro. Arterioscler. Thromb. Vasc. Biol. 30, 1228–1236. Navas-Acien, A., 2014. Association of global DNA methylation and global DNA
https://doi.org/10.1161/ATVBAHA.110.205500. hydroxymethylation with metals and other exposures in human blood DNA samples.
Panieri, E., Santoro, M.M., 2016. Ros homeostasis and metabolism: a dangerous liason in Environ. Health Perspect. 122, 946–954. https://doi.org/10.1289/ehp.1306674.
cancer cells. Cell Death Dis. 7, 2253. https://doi.org/10.1038/cddis.2016.105. Venkateswaran, N., Lafita-Navarro, M.C., Hao, Y.H., Kilgore, J.A., Perez-Castro, L.,
Papa, S., Choy, P.M., Bubici, C., 2019. The ERK and JNK pathways in the regulation of Braverman, J., Borenstein-Auerbach, N., Kim, M., Lesner, N.P., Mishra, P.,
metabolic reprogramming. Oncogene 38, 2223–2240. https://doi.org/10.1038/ Brabletz, T., Shay, J.W., DeBerardinis, R.J., Williams, N.S., Yilmaz, O.H., Conacci-
s41388-018-0582-8. Sorrell, M., 2019. MYC promotes tryptophan uptake and metabolism by the

10
Y. Ruan et al. Ecotoxicology and Environmental Safety 229 (2022) 113098

kynurenine pathway in colon cancer. Genes Dev. 33, 1236–1251. https://doi.org/ Yamamoto, M., Kensler, T.W., Motohashi, H., 2018. The KEAP1-NRF2 system: a thiol-
10.1101/gad.327056.119. based sensor-effector apparatus for maintaining redox homeostasis. Physiol. Rev. 98,
Villeneuve, N.F., Lau, A., Zhang, D.D., 2010. Regulation of the Nrf2-Keap1 antioxidant 1169–1203. https://doi.org/10.1152/physrev.00023.2017.
response by the ubiquitin proteasome system: an insight into cullin-ring ubiquitin Ye, D., Ma, S., Xiong, Y., Guan, K.L., 2013. R-2-hydroxyglutarate as the key effector of
ligases. Antioxid. Redox Signal. 13, 1699–1712. https://doi.org/10.1089/ IDH mutations promoting oncogenesis. Cancer Cell 23, 274–276. https://doi.org/
ars.2010.3211. 10.1016/j.ccr.2013.03.005.
Walsh, C.T., Tu, B.P., Tang, Y., 2018. Eight kinetically stable but thermodynamically Yeung, S.J., Pan, J., Lee, M.H., 2008. Roles of p53, MYC and HIF-1 in regulating
activated molecules that power cell metabolism. Chem. Rev. 118, 1460–1494. glycolysis - the seventh hallmark of cancer. Cell. Mol. Life Sci. 65, 3981–3999.
https://doi.org/10.1021/acs.chemrev.7b00510. https://doi.org/10.1007/s00018-008-8224-x.
Wang, C., Feng, R., Li, Y., Zhang, Y., Kang, Z., Zhang, W., Sun, D.-J., 2014. The Yu, H.S., Lee, C.H., Jee, S.H., Ho, C.K., Guo, Y.L., 2001. Environmental and occupational
metabolomic profiling of serum in rats exposed to arsenic using UPLC/Q-TOF MS. skin diseases in Taiwan. J. Dermatol. 28, 628–631. https://doi.org/10.1111/j.1346-
Toxicol. Lett. 229, 474–481. https://doi.org/10.1016/j.toxlet.2014.06.001. 8138.2001.tb00049.x.
Wang, X., Mu, X., Zhang, J., Huang, Q., Alamdar, A., Tian, M., Liu, L., Shen, H., 2015. Zhang, C., Liu, C., Li, D., Yao, N., Yuan, X., Yu, A., Lu, C., Ma, X., 2010. Intracellular
Serum metabolomics reveals that arsenic exposure disrupted lipid and amino acid redox imbalance and extracellular amino acid metabolic abnormality contribute to
metabolism in rats: a step forward in understanding chronic arsenic toxicity. arsenic-induced developmental retardation in mouse preimplantation embryos.
Metallomics 7, 544–552. https://doi.org/10.1039/c5mt00002e. J. Cell. Physiol. 222, 444–455. https://doi.org/10.1002/jcp.21966.
Wang, X.J., Sun, Z., Chen, W., Li, Y., Villeneuve, N.F., Zhang, D.D., 2008. Activation of Zhang, J., Zhang, B., Li, X., Han, X., Liu, R., Fang, J., 2019. Small molecule inhibitors of
Nrf2 by arsenite and monomethylarsonous acid is independent of Keap1-C151: mammalian thioredoxin reductase as potential anticancer agents: an update. Med.
enhanced Keap1-Cul3 interaction. Toxicol. Appl. Pharmacol. 230, 383–389. https:// Res. Rev. 39, 5–39. https://doi.org/10.1002/med.21507.
doi.org/10.1016/j.taap.2008.03.003. Zhao, E., Hou, J., Cui, H., 2020. Serine–glycine-one-carbon metabolism: vulnerabilities
Wang, Z., Li, Y., Wang, Y., Zhao, K., Chi, Y., Wang, B., 2019. Pyrroloquinoline quinine in MYCN-amplified neuroblastoma. Oncogenesis 9, 14. https://doi.org/10.1038/
protects HK-2 cells against high glucose-induced oxidative stress and apoptosis s41389-020-0200-9.
through Sirt3 and PI3K/Akt/FoxO3a signaling pathway. Biochem. Biophys. Res. Zhao, F., Malm, S.W., Hinchman, A.N., Li, H., Beeks, C.G., Klimecki, W.T., 2014.
Commun. 508, 398–404. https://doi.org/10.1016/j.bbrc.2018.11.140. Arsenite-induced pseudo-hypoxia results in loss of anchorage-dependent growth in
Warburg, O., 1956. On the origin of cancer cells. Science 123, 309–314. https://doi.org/ BEAS-2B pulmonary epithelial cells. PLoS One 9, 114549. https://doi.org/10.1371/
10.1126/science.123.3191.309. journal.pone.0114549.
Whitaker-Menezes, D., Martinez-Outschoorn, U.E., Lin, Z., Ertel, A., Flomenberg, N., Zhao, F., Severson, P., Pacheco, S., Futscher, B.W., Klimecki, W.T., 2013. Arsenic
Witkiewicz, A.K., Birbe, R.C., Howell, A., Pavlides, S., Gandara, R., Pestell, R.G., exposure induces the Warburg effect in cultured human cells. Toxicol. Appl.
Sotgia, F., Philp, N.J., Lisanti, M.P., 2011. Evidence for a stromal-epithelial “lactate Pharmacol. 271, 72–77. https://doi.org/10.1016/j.taap.2013.04.020.
shuttle” in human tumors: MCT4 is a marker of oxidative stress in cancer-associated Zhou, X., Yang, X., Sun, X., Xu, X., Li, Xi’an, Guo, Y., Wang, J., Li, Xia, Yao, L., Wang, H.,
fibroblasts. Cell Cycle 10, 1772–1783. https://doi.org/10.4161/cc.10.11.15659. Shen, L., 2019. Effect of PTEN loss on metabolic reprogramming in prostate cancer
Wong, B.W., Kuchnio, A., Bruning, U., Carmeliet, P., 2013. Emerging novel functions of cells. Oncol. Lett. 17, 2856–2866. https://doi.org/10.3892/ol.2019.9932.
the oxygen-sensing prolyl hydroxylase domain enzymes. Trends Biochem. Sci. 38, Zhu, J., Thompson, C.B., 2019. Metabolic regulation of cell growth and proliferation.
3–11. https://doi.org/10.1016/j.tibs.2012.10.004. Nat. Rev. Mol. Cell Biol. 20, 436–450. https://doi.org/10.1038/s41580-019-0123-5.
Wu, X., Sun, R., Wang, H., Yang, B., Wang, F., Xu, H., Chen, S., Zhao, R., Pi, J., Xu, Y., Zhu, Y., Aupperlee, M.D., Zhao, Y., Tan, Y.S., Kirk, E.L., Sun, X., Troester, M.A.,
2019. Enhanced p62-NRF2 feedback loop due to impaired autophagic flux Schwartz, R.C., Haslam, S.Z., 2016. Pubertal and adult windows of susceptibility to a
contributes to arsenic-induced malignant transformation of human keratinocytes. high animal fat diet in Trp53-null mammary tumorigenesis. Oncotarget 7,
Oxid. Med. Cell. Longev. 2019, 1–12. https://doi.org/10.1155/2019/1038932. 83409–83423. https://doi.org/10.18632/oncotarget.13112.
Xu, L., Lu, Z., Ji, C., Cong, M., Li, F., Shan, X., Wu, H., 2019. Toxicological effects of As Zong, D., Liu, X., Li, J., Ouyang, R., Chen, P., 2019. The role of cigarette smoke-induced
(V) in juvenile rockfish Sebastes schlegelii by a combined metabolomic and epigenetic alterations in inflammation. Epigenet. Chromatin 12, 65. https://doi.org/
proteomic approach. Environ. Pollut. 255, 113333 https://doi.org/10.1016/j. 10.1186/s13072-019-0311-8.
envpol.2019.113333.

11

You might also like