Next Article in Journal
High VEGFA Expression Is Associated with Improved Progression-Free Survival after Bevacizumab Treatment in Recurrent Glioblastoma
Next Article in Special Issue
High G9a Expression in DLBCL and Its Inhibition by Niclosamide to Induce Autophagy as a Therapeutic Approach
Previous Article in Journal
Transcriptome Meta-Analysis of Triple-Negative Breast Cancer Response to Neoadjuvant Chemotherapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regulation of Autophagy via Carbohydrate and Lipid Metabolism in Cancer

1
Department of Human Anatomy and Cell Science, College of Medicine, University of Manitoba, Winnipeg, MB R3E 0V9, Canada
2
Department of Physiology and Pathophysiology, Rady Faculty of Health Sciences, Institute of Cardiovascular Sciences, Albrechtsen Research Centre, St. Boniface Hospital, Winnipeg, MB R2H 2A6, Canada
3
Section of Thoracic Surgery, Department of Surgery, Health Sciences Centre, Rady Faculty of Health Sciences, University of Manitoba, Winnipeg, MB R3T 6C5, Canada
4
CancerCare Manitoba Research Institute, Winnipeg, MB R3E 0V9, Canada
5
Department of Radiology, Section of Radiation Oncology, Rady Faculty of Health Sciences, University of Manitoba, Winnipeg, MB R3T 2N2, Canada
6
College of Pharmacy, QU Health, Qatar University, Doha 2713, Qatar
7
Department of Pharmacology, Faculty of Medicine and Dentistry, University of Alberta, Edmonton, AB T6G 2H7, Canada
8
Department of Biochemistry & Medical Genetics, University of Manitoba, Winnipeg, MB R3E 0J9, Canada
9
Research Institute of Oncology and Hematology, Winnipeg, MB R3E 0V9, Canada
10
Faculty of Medicine in Zabrze, Academia of Silesia, 41-800 Zabrze, Poland
11
Biology of Breathing Theme, Children Hospital Research Institute of Manitoba, University of Manitoba, Winnipeg, MB R3E 3P5, Canada
*
Author to whom correspondence should be addressed.
Cancers 2023, 15(8), 2195; https://doi.org/10.3390/cancers15082195
Submission received: 4 January 2023 / Revised: 21 March 2023 / Accepted: 28 March 2023 / Published: 7 April 2023
(This article belongs to the Special Issue Autophagy–EMT Interrelations: At the Core of Tumor Transformation)

Abstract

:

Simple Summary

The metabolism of carbohydrates and lipids is essential to meet the diverse needs of cells. The autophagy cellular pathway is crucial for normal cellular function through its ability to target and degrade damaged organelles and misfolded proteins. Autophagy also provides cells with necessary energy as an end product. Cancer cells have a high rate of proliferation and, thus, require more energy compared to normal cells. Not surprisingly, cancer cells are highly dependent on these metabolic pathways. This review provides an in-depth review of recent research findings on how carbohydrate and lipid metabolism regulate autophagy and how this impacts cancer cells.

Abstract

Metabolic changes are an important component of tumor cell progression. Tumor cells adapt to environmental stresses via changes to carbohydrate and lipid metabolism. Autophagy, a physiological process in mammalian cells that digests damaged organelles and misfolded proteins via lysosomal degradation, is closely associated with metabolism in mammalian cells, acting as a meter of cellular ATP levels. In this review, we discuss the changes in glycolytic and lipid biosynthetic pathways in mammalian cells and their impact on carcinogenesis via the autophagy pathway. In addition, we discuss the impact of these metabolic pathways on autophagy in lung cancer.

1. Introduction

Autophagy is a highly regulated process and generally classified into three major types: Macroautophagy, Microautophagy and Chaperone-Mediated Autophagy (CMA) (Figure 1A) [1,2,3]. Macroautophagy (thereafter autophagy) involves the formation of double membrane structures called autophagosomes, which engulf cargoes, such as misfolded proteins, cytoplasmic proteins, or damaged organelles within the cell.
The autophagosomes then fuse with lysosomes, which contain hydrolases, resulting in the degradation and recycling of the cargo’s contents [4,5,6]. Microautophagy is the direct fusion of cargo with lysosomes. During microautophagy, the lysosomal or endosomal membranes are invaginated followed by the fusion of the cargo with these organelles and finally their degradation and recycling [7]. CMA is a very selective type of autophagy. CMA is distinct from macroautophagy and microautophagy in that it does not require membranes for the engulfment of the cargo. CMA selectively targets proteins that contain a specific KFERQ pentapeptide residue (Lys-Phe-Glu-Arg-Gln). This residue is detected by the cytosolic heat shock cognate 70 KDa protein (hsc70). The target protein and hsc70 complex are transported into the lysosome by the lysosomal-associated membrane protein 2A (LAMP2A) receptor in the lysosomal membrane [8,9]. Autophagy can also be selective in detecting cargoes using specific mediators. Therefore, this has resulted in new terminologies, such as pexophagy (for peroxisomes), mitophagy (for mitochondria), and aggrephagy (for aggregates) [10,11] (Figure 1B). The mammalian target of rapamycin (mTOR) complex I (mTORC I) is the most well-known autophagy repressor [12,13]. Autophagy inducers such as starvation, inhibit mTORC I and facilitate the activation of ULK1 and autophagy-related 13 (ATG13) and, thus, inducing autophagy [14]. In addition, AMP kinase (AMPK) is activated upon low ATP/AMP ratios and is one of the main inhibitors of mTORC I that, consequently, induces autophagy [15,16]. It also activates autophagy through phosphorylating ULK1.
Autophagy protects cells from stress caused by DNA damage, reactive oxygen species (ROS), adenosine triphosphate (ATP) deficiency, and nutrient deprivation [17,18,19,20,21]. In normal cells, autophagy occurs at a basal level, whereas it significantly increases in cancer cells to support their high metabolic energy demands [22,23,24,25].
Autophagy is tightly regulated via different mechanisms. These include the processing of autophagy-related gene (ATG) proteins, the ULK1 kinase complex (ULK1-mAtg13–FIP200–Atg101), the class III phosphatidylinositol 3-kinase (PI3K) complex, VPS34 complex, and the ATG5–ATG12–ATG16 complex interaction with light chain 3 (LC3)-II [19,26,27] (Figure 1A). Autophagy is usually part of the response to chemotherapy agents in tumor cells and it can increase or decrease chemotherapy-induced apoptosis [28,29]. This response is dependent on the type of tumour cells and chemotherapy agents used [29,30,31,32,33]. On the other hand, autophagy-mediated cell death could also be involved in response to chemotherapy agents. Autophagy-dependent cell death, also known as autophagic cell death, is defined as a tightly regulated process, which does not require other cell death pathways. Autophagic-dependent cell death is mechanistically dependent on the autophagic machinery [34,35] and is characterized by the existence of a large number of cellular vacuoles (vacuolization), lysosomal degradation, no nuclear condensation and no caspase activation [12,13].
Autophagy plays a double-edged sword role in cancer-suppression and cancer-promotion [9]. In the cancer initiation stages, autophagy removes damaged organelles or misfolded proteins to suppress factors involved in cancer initiation (such as ROS and the unfolded protein response). Conversely, in late stages of cancer, autophagy may provide energy for cancer cell metabolism, facilitate tumor cell metastasis [36,37,38,39], or induce chemotherapy resistance [9,19,40].

General Metabolism

Metabolism comprises all of the biochemical reactions that living cells use to generate the energy required for cell development, survival, and proliferation. [41,42,43,44]. Glycolysis, the Krebs cycle (glucose metabolism), and lipid metabolism are responsible for providing the required energy and building blocks for living cells. Glycolysis occurs in both aerobic and anaerobic conditions. The steps of the glycolysis pathway are summarized in Figure 2A. The Krebs cycle, also known as the citric acid cycle or the tricarboxylic acid (TCA) cycle, is summarized in Figure 2B. The interconnection of glycolysis and the TCA cycle is visualized in (Figure 2A,B). Fatty acids (FAs) are the most common form of stored energy, and triacylglycerols (TAGs) are the most common non-toxic form of FAs [45]. Generally, there are four sources from which to acquire FAs and TAGs: de novo lipogenesis (DNL), cytoplasmic TAG stores, FAs derived from TAGs of lipoprotein remnants taken up by the liver, and plasma non-esterified fatty acids (NEFA) released by adipose tissue [46]. The liver and adipose tissue are the two main sources in the body that produce FAs. FAs synthesized and exported from the liver provide an energy source and lipoprotein structural components for building cell membranes. FAs synthesized in adipose tissue through DNL contribute to long-term energy storage [47]. DNL is a metabolic pathway under tight regulation that involves a series of enzymatic reactions, occurring primarily in the liver and adipose tissue, to produce FAs from surplus carbohydrates to store energy. This pathway is summarized in Figure 2C [48,49,50].
In the presence of oxygen, pyruvate is transported into the mitochondria where it is converted into acetyl-CoA. Furthermore, many cells also utilize FA β-oxidation, which involves breaking down a long-chain acyl-CoA molecule into acetyl-CoA molecules (summarized in Figure 2D) [52,53,54,55,56]. The other lipid metabolic pathway is triglyceride biosynthesis. They are synthesized by the esterification of FAs into glycerol (summarized in Figure 2E) [47,57,58,59].
Compared to normal cells and apart from autophagy, cancer cells exhibit increased glucose uptake and lactate production even in the presence of oxygen [60]. This process of aerobic glycolysis was coined the “Warburg effect”, named after Otto Warburg, who first observed this [60]. Originally, Warburg hypothesized that this was due to a mitochondrial dysfunction in tumour cells but it has since been found that tumour cells and normal cells have similar rates of OXPHOS, suggesting that tumour cells have intact mitochondria [61]. However, if that is the case, why would tumour cells use this inefficient means of energy production?
The speed of cytosolic ATP production is about 100 times faster than in mitochondria. Therefore, aerobic glycolysis supplies more ATP per unit time than OXPHOS [62]. Cells with this higher rate of ATP production gain a selective advantage when accessing shared and limited resources such as the limited glucose in the tumour microenvironment [63]. When the demand for ATP is increased, aerobic glycolysis can be increased rapidly, while OXPHOS remains constant [62]. This supports malignant cell growth and proliferation. The Warburg effect also supports uncontrolled proliferation by feeding into biosynthesis. The increased glucose consumption serves as a carbon source for anabolic processes, such as nucleotide, lipid, and protein synthesis. The excess glucose also allows for a greater production of reducing equivalents (NADPH) through the pentose phosphate pathway (PPP), which is needed for reductive biosynthesis [61]. It has been proposed that this inefficient method of ATP production is a trade-off for maintaining the high flux through anabolic pathways.
The Warburg effect is of prominent importance in cancer cells for numerous reasons. It results in a higher rate of cytosolic ATP production compared with that in mitochondria. Therefore, aerobic glycolysis produces more ATP than oxidative phosphorylation (OXPHOS) per unit time [62]. Cells with this higher rate of ATP production gain a selective advantage when accessing shared and limited resources such as the limited glucose in the cancer microenvironment [64]. When the demand for ATP is increased, aerobic glycolysis is rapidly increased, while OXPHOS remains constant [62,64]. The Warburg effect also supports unregulated proliferation by feeding into biosynthesis. The increased glucose consumption serves as a carbon source for anabolic processes (e.g., lipid, nucleotide, and protein synthesis). The excess glucose also allows for a greater production of reducing equivalents of NADPH by the PPP, which is needed for reductive biosynthesis [64]. It has been proposed that this inefficient method of ATP production is a trade-off for maintaining the high flux through anabolic pathways in cancer cells.
Lactate may be the only oncometabolite involved in all the hallmarks of cancer progression, and it has been suggested that lactate production may be one of the main purposes of the Warburg effect [62,65]. Lactate excretion from cancer cells contributes to the production of an acidic cancer microenvironment, conferring a growth and metastatic advantage to cancer cells [66]. The low extracellular pH (pHe) promotes mutagenesis, chromosomal abnormalities, and genome instability [67], allowing cells to acquire mutations that favor malignant progression. The low pH stimulates cancer cells to secrete proteins that will degrade the extracellular matrix (ECM) and promote angiogenesis, creating an environment that favors intravasation, invasion, and metastasis [65,66]. The acidic microenvironment also represses lymphocytes, helping cancer cells with immune evasion [68]. This allows cancer cells to survive, metastasize, and colonize distant sites. Finally, lactate and the acidic microenvironment induce epithelial-to-mesenchymal transition (EMT) in cancer cells through the activation of autophagy and transforming growth factor β (TGFβ) [38,39,69] expression; thus, increasing motility and invasiveness and, subsequently, metastasis [70,71]. Several investigations have shown that the suppression of glycolysis increases autophagy, which becomes a propelling force of OXPHOS for ATP production to support cancer cell survival [72,73,74]. Recent research has shown that glycolytic enzymes and the metabolic intermediates of glycolysis are involved in autophagy regulation [72,73,74].
The reprogramming of lipid metabolism in cancer directly contributes to malignant transformation and progression [60]. The expression levels of some of the crucial enzymes in lipid metabolism are altered in cancer. For example, fatty acid synthase (FASN), which is a multi-enzyme protein complex that catalyzes FA synthesis, is upregulated in cancer [64]. Sterol regulatory element-binding proteins (SREBPs) are a family of transcription factors regulating the expression of genes involved in the uptake and synthesis of FAs [71]. SREBPs are also upregulated in various cancers and promote cancer growth. It has been shown that the overexpression of lipid metabolism-related enzymes (e.g., FASN and SREBPs) is correlated with advanced stages of different type of cancers, poor patient survival, and cancer cell proliferation, invasion, and metastasis [75]. Moreover, the fatty acid oxidation (FAO) of lipids is increased in cancer cells to provide them with ATP to promote growth and survival under metabolic stress. In addition, FAO produces NADPH, which protects cancer cells from the oxidative stress of ROS [75]. Interestingly, different studies have recently shown the intricate link between autophagy and lipid homeostasis [76,77]. Autophagy plays a crucial role in the intracellular degradation of lipids [78,79,80]. This helps produce the precursors required for the biosynthesis of other lipids, repair damaged structures, and produce the energy necessary for the anabolic processes within the cell that require lipid building blocks [60]. Lipids play a crucial role in autophagic activity in multiple tissues. Therefore, they have important implications for diseases such as cancer [81,82]. The interconnection between lipid metabolism and autophagy is further highlighted by shared regulators, such as AMPK, mTORC1, TFEB, and PPARs [83,84]. Similar to glycolysis, lipid metabolism can exert its effects on autophagy during cancer through both lipid metabolites and the enzymes involved in lipid metabolism.
Despite the use of aerobic glycolysis for ATP production, cancer cells rely on functional mitochondria. The promotion of aerobic glycolysis is activated by pro-oncogenic mutations, such as K-RAS, c-MYC, and PIK3CA, not mitochondrial inactivation. Elimination of mitochondrial DNA (mtDNA) has been shown to reduce tumour growth rates and limit tumorigenesis [85,86]. Mutations of mtDNA has been found in several cancers, including renal adenocarcinoma, prostate cancer, bladder cancer, colorectal cancer, breast cancer, and head and neck tumours [87]. These mutations allow the altered cellular metabolism to promote malignant transformation. Mitochondria are vital for tumour progression and growth through their biosynthetic and bioenergetic effects [86]. Mitochondria are important for oxidation–reduction (redox) reactions, control the production of reactive oxygen species (ROS), sequester and release Ca2+, thereby controlling cytosolic Ca2+ levels, provide building blocks for biosynthetic pathways, and are involved in anterograde and retrograde signaling with the nucleus [87].
Given the importance of mitochondria in cancer, there must be a way to regulate their production. To prevent the accumulation of defective mitochondria, mitophagy (the selective autophagy of mitochondria) selectively degrades excessive or damaged mitochondria [85]. The knockdown of autophagy in mouse models of cancer leads to the build-up of defective mitochondria, impairing mitochondrial respiration and cell growth, and increasing cell death [87]. Similarly, autophagy has been found to be crucial for mitochondrial metabolism and tumour growth in KRAS-driven lung cancer and BRAF-driven cancers [87]. Having mitochondrial quality control is important for malignant growth by preserving the supply of functioning mitochondria [85]. Defective mitochondria provide a direct signal to mitophagy machinery for degradation, whereas mitochondrial synthesis is triggered by reduced cellular energy output [86].
In the current review article, we will discuss and explain how the different steps of glycolysis and lipid metabolism are connected to autophagy, and their impact on tumorigenesis and metastasis. In the last section, we will specifically focus our discussion on the impact of autophagy and these metabolic pathways in lung cancer.

2. Glycolysis and Autophagy Regulation in Cancer

In this section, we will discuss the key enzymes involved in regulating glycolysis and their effect on autophagy.

2.1. Hexokinase 2 (HK2)

HK2 catalyzes the initial step of glycolysis, which is the conversion of glucose to glucose 6-phosphate [88,89]. In mammals, there are four hexokinase isoforms: HK1, HK2, HK3, and HK4 [90]. Among the subtypes, HK2 is considered the most important one, which is overexpressed in various tumors including breast, lung, gastric and laryngeal squamous cancer, hepatocellular carcinoma (HCC), and renal cell carcinoma [91,92,93,94,95,96,97]. In addition to its critical role in glycolysis and autophagy, HK2 also facilitates apoptosis resistance and the induction of chemo-resistance in pancreatic, epithelial ovarian and breast cancers by increasing its expression or dimerization and interaction with voltage-dependent anion channels [98,99,100,101,102]. HK2 is overexpressed in ErbB2-driven breast cancer for tumor initiation and maintenance [103]. Further studies have shown that targeting HK2 via miR-143 inhibited cell proliferation and the Warburg effect in colon [104], breast [105], prostate [106], and renal cell cancers [96]. Overall, overexpression of HK2 results in an enhanced autophagic response. In contrast, suppression of autophagy is seen with an inhibition of HK2 (Figure 3A) [107].

2.2. Phosphofructo-Kinase/Fructose Biphosphatases (PFKFBs)

PFKFB controls the intracellular steady-state concentration of fructose 2,6-bisphosphate (F2,6P2) [108]. There are four isozymes of PFKFB (PFKFB1–4), which display different properties, including tissue specific expression and the response to kinases and phosphatases [109,110]. PFKFB3 and PFKFB4 are found overexpressed in breast [107,111,112,113,114,115,116,117], bladder [118,119], pancreatic [120,121], and gastric cancers [120,121]. Mechanistically, low levels of ATP leads to the phosphorylation of the C-terminal domain at Ser461 of PFKFB3 by the cellular energy sensor AMPK [122]. In addition, this condition induces MK2 (MAPK (mitogen-activated protein kinase)-activated protein kinase 2) phosphorylation at Thr334 via p38α MAPK, which, in turn, activates PFKFB3 (phosphorylation at Ser461). Moreover, MK2 also activates SRE (serum-response element) in the promoter region of the PFKFB3 gene via the phosphorylation of SRF (serum response factor) at Ser103, resulting in the increased transcription and activity of PFKFB3 to amplified glycolytic and autophagic rates (Figure 3B) [122]. Klarer et al. demonstrated that the inhibition of PFKFB3 using either siRNA transfection or derivatives of the PFKFB3 inhibitor, 3-(3-pyridinyl)-1-(4-pyridinyl)-2-propen-1-one (3PO), initiates survival autophagy in HCT-116 colon adenocarcinoma cells [123]. Some studies showed that PFKFB4 depletion promoted oxidative stress, ultimately triggering survival autophagy as a stress-adaptive survival mechanism in lung cancer [124,125].

2.3. Glyceraldehyde-3-Phosphate Dehydrogenase (GAPDH)

GAPDH catalyzes the conversion of glyceraldehyde-3-phosphate to 1.3-bisphosphoglycerate in glycolysis, which is a reversible reaction [126,127]. As well as playing the vital role of maintaining aerobic glycolysis in the cytosol, GAPDH also participates in several non-glycolytic functions including autophagy [128]. In the cytosol, under glucose shortage, GAPDH is phosphorylated at Ser122 by AMPK. This leads to GAPDH redistributing into the nucleus, where it interacts directly with sirtuin1 (SIRT1) to stimulate SIRT1 activation and autophagy, increasing SIRT1 deacetylase activity and disassociating its inhibitor DBC1 (deleted in breast cancer-1) (Figure 3C) [129]. In general, various studies have shown that the GAPDH overexpression is associated with breast, ovarian, lung, and prostate cancers’ aggressiveness and metastases [126,130,131,132,133,134]. Therefore, GAPDH depletion is recommended as a novel strategy to control tumor cells. The GAPDH inhibitor (3-bromo-isoxazoline) induces autophagy and inhibits the proliferation of pancreatic ductal-adenocarcinoma cells (PDAC) and pancreatic cancer stem cells (CSCs) [135]. Furthermore, the ROS/Akt/mTOR axis is activated by the inhibition of mitochondrial uncoupling protein 2 (UCP2) in pancreatic adenocarcinoma cells, which, in turn, leads to tumor cell growth arrest, induction of caspase-mediated apoptosis, and the translocation of GAPDH into the nucleus to promote BECLIN1-mediated autophagy [136].

2.4. Phosphoglycerate Kinase (PGK)

PGK coordinates ATP production in glycolysis by catalyzing the conversion of 1,3-bisphosphoglycerate to 3-phosphoglycerate. PGK has two isoforms, PGK1 and PGK2 [137,138]. PGK1 has a significant role in sustaining cellular homeostasis and autophagy, making it an attractive molecular target for cancer therapy. The condition of glutamine deprivation and hypoxia inhibit mTOR-mediated arrest-defective protein 1 (ARD1) Ser228 phosphorylation, leading to ARD1 binding to PGK1 and Lys388 acetylation with subsequent PGK1-mediated BECLIN1 Ser30 phosphorylation. This activates ATG14L-associated class III phosphatidylinositol 3-kinase VPS34 to initiate autophagosome formation in glioblastoma (GBM) patients. ARD1-dependent PGK1 acetylation and PGK1-mediated BECLIN1 Ser30 phosphorylation are involved in brain tumorigenesis [139] (Figure 3D). In another investigation, it has been shown that PGK1 induced the phosphorylation of PRAS40 (the proline-rich AKT substrate of 40 kDa, encoded by the gene AKT1S1) at Thr246 in Ewing’s sarcoma cell line A673, and liver cancer cell lines HepG2 and SNU449. PRAS40 is involved in PGK1-induced cancer cell growth. It was found that PGK1 inhibits autophagy-mediated cell death, which is associated with PRAS40. In normoxic conditions, PGK1 blocks autophagic cell death via PRAS40 phosphorylation, stimulating cellular proliferation. In hypoxia, PGK1 interacts with BECLIN1 and induces its phosphorylation with subsequent autophagy induction. Therefore, PGK1 is, potentially, a novel autophagy modulator for tumorigenesis that acts between repressing autophagy-mediated cell death via PRAS40 and inducing autophagy through BECLIN1 based on oxygen levels in the environment [140].

2.5. Pyruvate Kinase (PK)

PK catalyzes the conversion of phosphoenolpyruvate (PEP) to pyruvate, which is the final step of glycolysis. PK has four isoforms: PKL (liver), PKR (erythrocytes), PKM1 (muscle and brain), and PKM2 (brain and liver) [141,142,143]. PKM2 can be a tetramer or dimer and the tetrameric form has a higher affinity for PEP, while the dimer has a lower affinity. Under physiological conditions, the tetramer form dominates, whereas cancer cells primarily prefer the dimeric form, which is necessary for the Warburg effect (aerobic glycolysis). In other words, the lower activity of the dimeric form of PKM2 induces an increase in glycolysis, and can promote tumor survival and growth [141,142,144,145,146,147,148,149]. Therefore, identifying new compounds that act as PKM2 activators could be a potential therapeutic strategy for cancer treatment.
Many investigations provide evidence that PKM2 is involved in autophagy. Overexpression of PKM2 leads to the activation of class I PI3K that could lead to activation of the Akt-mTOR signaling pathway, which, in turn, inhibits autophagy. PKM2 phosphorylates PRAS40 and releases this protein from mTORC1 with its subsequent activation in pancreatic cancer (Figure 3E) [150]. The knockdown of PKM2 in cancer cells activates AMPK signaling to maintain energy homeostasis via autophagy activation in prostate cancer cells [144]. The downregulation of PKM2 acts as a tumor suppressor in acute myeloid leukemia (AML) by increasing BECLIN1-mediated autophagy [150]. PKM2 is also overexpressed in HCC cells and is induced through the JAK/STAT3 pathway, which might be involved in the response of HCC tumor cells to chemotherapy-induced apoptosis [151].

2.6. Lactate Dehydrogenase (LDH)

LDH catalyzes the reduction of pyruvate to lactate with the associated interconversion of NADH and NAD+. It has two isomers, including LDHA and LDHB, that support the conversion of pyruvate to lactate in cells and back, respectively [152,153,154]. LDHA is enhanced in many types of cancers (lung, liver, testicular, and breast) to establish metabolic reprogramming [155,156,157,158,159]. In tamoxifen-resistant breast cancer cells, ATP production decreased in parallel with increased glycolytic pathway activity [160]. In this scenario, LDHA is associated with tamoxifen resistance via BECLIN1-induced pro-survival autophagy [160].
LDHB is a crucial contributor to lysosomal activity and basal autophagy in cancer. This enzyme converts lactate and NAD+ to pyruvate, NADH and H+ to create an acidic environment. This is essential for vesicle maturation and protease activation during autophagy; however, the regulation of this has not yet been defined [161]. Liang Shi et al. demonstrated that the protein sirtuin5 (SIRT5) caused hyperactivity of LDHB by the deacetylation at Lys329 that triggered autophagy and accelerated the growth of colorectal cancer cells (CRC) [150].

2.7. Lactate

One of the most important consequences of the Warburg effect in cancer cells is the increase in lactate production, which, in turn, leads to the acidification of the tumor microenvironment, creating an extracellular pH ranging from 6.0 to 6.5 (Figure 4) [162,163].
Emerging studies have shown that lactate-induced acidosis leads to processes such as metastasis [164,165], angiogenesis [166,167,168], invasion [165,169], and immunosuppression [162,170]. Lactate itself cannot freely diffuse across the plasma membrane and monocarboxylate transporters 1–4 (MCT1-4) control the exchange of lactate across membranes [171,172]. Among the MCTs, MCT1 and MCT4 are extensively expressed in cancer cell lines, playing an important role in metabolic reprogramming and tumor aggression [173,174,175,176,177]. Regulating the expression of MCT1 via Osimertinib (OSI) activates liver kinase B1 (LKB1)/AMPK signaling, leading to autophagy induction in CRC cells [178]. On the other hand, blocking lactic acid export by inhibiting MCT4 provides an efficient anticancer strategy. MCT4 depletion by different doses of 7-aminocarboxycoumarin derivatives (7acc1) enhanced the cytotoxicity of NK cells in breast carcinoma by interfering with lactate flux, reversing the acidification of the tumor microenvironment, and inducing autophagy [179].

2.8. Hypoxia-Inducible Factor-1 Alpha (HIF-1α) and Glycolysis

Adaptation to hypoxic conditions is achieved primarily through excessive accumulation of HIF-1α in tumor cells. HIF-1α increases the expression of the glycolysis pathway and compensates for the ATP demand of eukaryotic cells under hypoxic conditions [180]. Beyond its role in glycolysis, HIF-1α participates in autophagy by increasing the transcription of BNIP3 and BNIP3L genes, which leads to the dissociation of the Bcl-XL and Bcl-2 complex and the activation of BECLIN1-induced autophagy [181,182] (Figure 4). Notably, glucose deprivation could induce HIF1α-independent autophagy through the induction of AMPK and the inhibition of mTOR (Figure 3F) [183,184]. The mTOR signaling pathway is inactivated under hypoxic conditions, leading to activation of ULK1, which is required for the initiation of autophagy (Figure 3F). Therefore, hypoxia-induced autophagy may be considered as a target to reverse the radio resistance in cancer cells. Yueming Shen et al. showed that autophagy was inhibited by the knockdown of lincRNA-p21. This occurred due to the downregulation of HIF-1α protein levels and activation of the Akt/mTOR/P70S6K signaling pathways. This is a suggested mechanism to enhance the radiosensitivity of hypoxic tumor cells in human HCC and glioma cells [185]. A recent survey indicated that the expression of HIF-1α is commonly increased in a variety of human tumors, including lung, breast, and osteosarcoma [186].
We have summarized the impact of the glycolysis-related enzymes involved in the regulation of autophagy in different cancers (excluding lung cancer, which is discussed separately below) in Table 1.

3. Lipid Metabolism and Regulation of Autophagy in Cancer

Historically, lipids have been challenging to study. However, the advancements in technology and methodologies have made progress in the research on the role of lipids in autophagy and cancer progression possible. Despite recent discoveries about this relationship, our understanding of how lipid metabolism affects autophagy alterations during cancer is still rather limited. In this section, we aim to provide the most current findings and understanding of how changes in lipid metabolism influence autophagy during cancer. Lipids play a crucial role in autophagic activity in multiple tissues; therefore, they have important implications for diseases such as cancer [78,190]. The connection between lipid metabolism and autophagy is further strengthened by the common regulators that they both share, including AMPK, mTORC1, Transcription Factor EB (TFEB), and PPARs [191,192,193,194]. Autophagy during cancer progression is modulated thorough lipid metabolism.

3.1. Enzymes in Lipid Metabolism Can Regulate Autophagy in Different Cancers

3.1.1. Fatty Acid Synthase (FASN)

FASN is an important cytoplasmic enzyme involved in fatty acid biosynthesis. The main function of FASN is to synthesize palmitate from acetyl-CoA and malonyl-CoA [195]. FASN expression increases the activation of mTOR, resulting in autophagy inhibition in leukemic blast cells obtained from AML patients [196]. Moreover, FASN can downregulate the expression of TFEB via mTOR. TFEB is phosphorylated by mTOR, leading to TFEB becoming localized in the cytoplasm and, thereby, inhibiting its transcriptional activity [196]. This can negatively affect autophagy induction since TFEB regulates the expression of the genes involved in autophagy including lysosomal biogenesis, lysosomal membrane, lysosomal acidification, and other autophagy-related genes [197]. This has been confirmed by the finding that the inhibition of FASN promotes the nuclear translocation of TFEB where it can positively regulate the expression of lysosomal biogenesis genes such as LAMP1 [196]. Additionally, FASN has been found to self-modulate its expression via crosstalk with PI3K/Akt in clinical specimens of human osteosarcomas. It was found that the overexpression of FASN has a positive relationship with the activation of PI3K-Akt in these clinical samples. In contrast, the inhibition of FASN decreased Akt phosphorylation while Akt inhibition similarly inhibited the mRNA and protein expression of FASN in vitro [198]. In another study, it was found that FASN overexpression enhanced the cellular respiration (e.g., FAO) that favors cancer cells during CRC. In addition, FASN overexpression was correlated with less MAPK activation and the accumulation of p62, indicating autophagy inhibition, which can favor the survival of CRC HCT116 and SW480 cells [199]. The overall effect of FAs in the regulation of autophagy is summarized in Figure 5.

3.1.2. 3-Hydroxy-3-MethylGlutaryl-CoA Reductase (HMGCR) Inhibitors

HMGCR is the rate-limiting enzyme in the mevalonate pathway that synthesizes different lipids, such as cholesterol, geranylgeranyl pyrophosphate (GGPP), and farnesyl pyrophosphate (FPP) [28,29,200,201]. Statins are a class of medications that inhibit HMGCR and, thus, decrease the de novo cholesterol levels [200,202,203]. Therefore, statins have been commonly used as cholesterol-lowering agents. Statins have been shown to have pleiotropic effects that are independent of their cholesterol-lowering properties [201]. However, inhibition of HMGCR by statins can affect the production of other downstream lipids that are produced in the mevalonate cascade. Changes in these lipids, in turn, can modulate autophagy and, therefore, affect cancer cell survival. It has been shown in human leukemia cells that blocking cholesterol biosynthesis using statins leads to autophagy induction via inhibition of the Akt-mTOR pathway [204]. Interestingly, the addition of cholesterol (but not GGPP or FPP) to cells attenuates the induced autophagy, indicating the cholesterol-dependent inhibition of autophagy in these cancer cells [205].
In HepG2 cells, HMGCR inhibition by statins can induce autophagy through increasing LC3-II expression. Autophagy induction was confirmed after observing that statins further increased LC3-II expression when the cells were also treated with Bafilomycin A1, an autophagy inhibitor [206]. Similar results were observed in atorvastatin-treated PC3 prostate cancer cells, where a 10-fold increase in LC3-II expression was found. This group observed that the addition of GGPP, but not FPP, could reverse the atorvastatin-induced autophagy. This indicated that autophagy induction by atorvastatin was due to the inhibition of GGPP biosynthesis [207]. In human rhabdomyosarcoma cells, the depletion of GGPP with hydrophobic statins induced autophagy [208,209]. It was also shown that simvastatin inhibits the flux of autophagy in alveolar rhabdomyosarcoma cells through changing the acidity of lysosomes, which is different from C2C12 cells [210]. Statins (fluvastatin, lovastatin, simvastatin, and atorvastatin) can trigger the induction of autophagy in PC3 cells via the inhibition of geranylgeranylation. The findings of this study suggested that autophagy induction by statins can potentially have a protective role against prostate cancer progression [207,211]. These findings highlight that lipid metabolites produced by the mevalonate pathway play a crucial role in autophagy regulation upon HMGCR inhibition in different types of cancer. Additionally, it has been shown in previous studies that the combination therapy of statins with other anticancer agents can improve the treatment response in different cancer cell lines (small-cell lung, ovarian and breast cancer, GBM, and acute myeloid leukemia) [29,212,213,214,215]. In agreement with these studies, we have previously shown that simvastatin prevents the fusion of autophagosomes and lysosomes in GBM cell lines; thereby, inhibiting autophagy flux. This autophagy flux inhibition potentiates the temozolomide-induced apoptosis in GBM cancer cells [29]. The impact of cholesterol and the cholesterol biosynthesis pathway in the regulation of autophagy is summarized in Figure 6.

3.1.3. Sphingosine-1-Kinase (SPHK1)

SPHK1 is the enzyme that synthesizes sphingosine-1-phosphate (S1P) from sphingosine. It also plays a crucial role in maintaining the sphingolipid balance within the cell [216]. It has been shown that SPHK1 participates in the regulation of autophagy and in different aspects of malignancy including therapy resistance, cancer development, growth, and metastasis [217,218]. SPHK1 is highly expressed in different types of cancers, such as human astrocytoma, CRC, and prostate cancer. It has also been associated with angiogenesis and resistance to chemotherapy and radiotherapy [219,220,221]. Thus, SPHK1 has been proposed to serve as an oncogene in carcinogenesis [222]. Prior studies have found that SPHK1 plays a role in the biogenesis of autophagosomes by modulating the expression levels of LC3 and, therefore, autophagy in neuroblastoma cells. However, the exact mechanisms by which SPHK1 regulates autophagy in these cancer cells has yet to be elucidated [223]. Another study found that in breast cancer MCF-7 cells, SPHK1 induces autophagy by increasing the formation of LC3-positive autophagosomes, thereby protecting them during nutrient deprivation [224]. SPHK1 also induces autophagy in CRC HT-29 cells via activation of SPHK1/ERK/pERK, which leads to the inhibition of mTOR and, consequently, ULK1 activation. SPHK1 was also shown to increase the expression of ATG5, ULK1, and LC3 in CRC HT-29 cells. Furthermore, autophagy has been linked to the survival of CRC cells during chemotherapy stress and is a marker of an invasive phenotype. Therefore, autophagy inhibitors, such as Bafilomycin A1 and Chloroquine, are suggested as promising interventions for CRC treatment [225]. Moreover, SPHK1 has been shown to induce EMT and, consequently, contribute to invasion and metastasis in HCC HepG2 cells [226]. This study revealed that SPHK1 induced EMT through enhancing lysosomal degradation of E-cadherin, a protein critical in epithelial cell–cell adhesion. Furthermore, this group found that S1P could stimulate the binding of TRAF2 (TNF Receptor Associated Factor 2) to BECLIN1. TRAF2 binding catalyzes the ubiquitination of BECLIN1 and, therefore, the induction of autophagy. Overall, the authors concluded that the SPHK1-E-cadherin-TRAF2-BECLIN1 pathway is responsible for EMT induction and subsequent metastasis in HCC cells. Thus, SPHK1 inhibition is proposed to decrease autophagy activation and is a possible strategy for HCC treatment. SPHK1 also modulates autophagy in human peritoneal mesothelial cells (HPMCs) and plays a role in the progression of gastric cancer peritoneal dissemination (GCPD) [227]. SPHK1 expression was directly related to LC3 expression in 120 human samples from gastric cancer peritoneal tissue. The SPHK1-induced autophagy in HPMCs increased the invasive capability of gastric cancer cells, whereas SPHK1 knockdown inhibited autophagy. This was associated with a poor prognosis and recurrence in these patients. SPHK1 is also involved in the production of ceramide, a prominent lipid metabolite that will be discussed in the following section. Additionally, SPHK1 has an inhibitory effect on mTOR and, therefore, allows for the activation of autophagy independent of Akt [81]. The connection of sphingomyelin, ceramide, and SPHK1 to autophagy is summarized in Figure 7.

3.1.4. Fatty Acid Translocase/Cluster of Differentiation 36 (FAT/CD36)

FAT/CD36 is an enzyme localized within the cell membrane that is responsible for long-chain fatty acid uptake into the cell [228]. A study found that FAT overexpression inhibits autophagy in HCC HepG2 and Huh7 cells [229]. Consistent with these findings, they found that FAT knockdown induces autophagy in these cells as shown by the increased formation of autophagosomes, LC3-II expression, and p62 degradation. FAT modulates autophagy through the phosphorylation of BECLIN1 and ULK1, acting as a negative regulator of autophagy in HCC cells [229]. In this same study, in vivo studies in mice with FAT knockout also showed elevated autophagy. This was further confirmed by the observation that autophagy was decreased after reconstituting FAT expression in FAT knockdown mice. Additionally, FAT knockdown caused an increase in lipophagy levels together with enhanced FA β-oxidation. These findings highlight that modulation of FAT expression could be a promising way to induce lipophagy in HCC cells, thereby mitigating the excessive lipid accumulation (Figure 7).
We have summarized the effect of enzymes involved in lipid metabolism on the different steps of autophagy in Table 2.

3.2. Lipid Metabolites Can Regulate Autophagy in Different Cancers

3.2.1. Phosphatidic Acid (PA)

PA is produced through the breakdown of phosphatidylcholine into PA and choline by Phospholipase D [230]. Although different targets for PA have been reported, the most significant target of PA in cancer cells is likely to be mTOR [231]. PA regulates autophagy by mediating the induction of autophagy by inducing autophagosome curvature and inhibiting mTOR [232]. Furthermore, PA can be produced under nutrient and oxygen deprivation, which is common in human tumor sites [231]. Other metabolites of PA can also regulate autophagy. For example, Phosphatidic Acid Phosphatases can convert PA into diacylglycerol, which has been shown to regulate autophagy [233]. Diacylglycerol induces autophagy through activation of PKC. It also helps dissociate BCL2 from BECLIN1 through activation of JNK and, thus, initiates autophagy induction [233]. Lysophosphatidic acid (LPA) activates mTORC1 and decreases autophagy via the phosphorylation of S6K and ULK in prostate cancer cells (PC3, LNCaP, and Du145 cells) [234]. It has been recently shown that the HS1BP3 protein negatively regulates autophagosome formation in moderately differentiated osteosarcoma cells (moderately differentiated sarcoma) through phospholipase D (PLD) and PA production and the increased localization of PLD1 to ATG16L1-positive membranes [235].

3.2.2. Sphingolipids

Another class of lipids are sphingolipids, which play a prominent role in autophagy regulation as well as cell proliferation and differentiation [236]. Sphingolipids are a very important group of lipids that can regulate a wide range of crucial functions in cancer cells, such as cell growth and proliferation [237]. In particular, two sphingolipids have appeared to significantly regulate autophagy, ceramide and S1P. Ceramide can inhibit Akt, which consequently inhibits mTOR, inducing autophagy. It has also been found that both ceramide and S1P can upregulate the transcription of BECLIN1 and cause autophagy induction [238]. The product of ceramide hydrolysis is sphingosine, which produces S1P via phosphorylation by SPHK1 [224]. Recent studies have shown that SPHK1 is activated under starvation conditions, which consequently forms S1P [224]. Ceramide synthase 2 (CerS2) is one of six mammalian isoforms of ceramide synthase that can synthesize ceramide. It was shown that overexpression of CerS2 causes the accumulation of ceramides, resulting in growth arrest and autophagy induction in SMS-KCNR neuroblastoma and MCF-7 breast cancer cells [237]. Additionally, it has been shown that the treatment of HT-29 CRC cells with ceramide induces autophagy in these cells by elevating the pool of ceramides [238,239,240]. It was shown in MCF-7 breast cancer cells that ceramides increased the expression of BECLIN1 and, therefore, induced autophagy [239]. In addition, it has been shown that short-chain ceramides (C(2)-ceramide and C(6)-ceramide) and tamoxifen-induced stimulation of de novo ceramide synthesis caused the dissociation of BCL2 from BECLIN1 and, therefore, the induction of autophagy in HeLa cells [238].
Sphingosine is generated from the hydrolysis of ceramide by ceramidase followed by its conversion to Sphingosine-1-Phosphate (S1P) by sphingosine kinases. S1P binds to its cell receptors and can modulate various cellular functions including proliferation, migration, and autophagy. It has been shown that elevated levels of intracellular S1P through overexpression of SPHK1 induces autophagy via the inhibition of mTOR in human MCF-7 breast cancer cells [224,241]. Moreover, S1P can act as a survival mechanism in prostate cancer. It was shown that treatment with S1P in PC3 prostate cancer cells can induce autophagy through the inhibition of mTOR [242]. It was concluded that S1P could mediate nutrient deprivation-induced autophagy as a survival mechanism during the early phase of prostate cancer development.

3.2.3. Peroxisome Proliferator-Activating Receptors (PPARs)

PPARs comprise a family of nuclear proteins including PPARα, PPARδ, and PPARγ that serve as transcription factors, regulating the expression of different genes [243]. In order to be functional as a transcription factor, PPARs need to bind to lipid ligands via lipid-binding proteins [244]. These lipid ligands have been reported to be oleic acid, linoleic acids, linolenic acids, prostaglandins, eicosanoids, and oxidized lipids [245,246]. PPARs have been shown to regulate autophagy in different cancers via various pathways such as the regulation of autophagy-related genes [81]. Furthermore, TFEB is also activated following the activation of PPARs, which upregulates the expression of different autophagy-related genes, such as Rab7 and LAMP2 [247]. This finding has been further confirmed by the observation that mouse cells with PPARδ knockout have a significant reduction in autophagy markers [248]. The levels of these lipid ligands change during cancer, which can directly influence PPAR and TFEB activation and, thereby, autophagy during cancer. For this reason, PPARs have been proposed as promising treatment candidates for cancer. It has been found that the treatment of Caco-2, a CRC cell line, with rosiglitazone (a PPARδ agonist) induces autophagy [249]. In addition, PPARγ can activate PTEN, which is an inhibitor of PI3K. Upon PI3K inhibition, PIP3 production is reduced, resulting in mTOR inhibition and the induction of autophagy in CRC cells [249]. PPARγ is active during cancer, and it has been suggested that PPARγ activates autophagy in a self-sufficient manner. For example, PPARγ has been shown to regulate autophagy in MDA-MB-231 breast cancer cells. It was shown that troglitazone (a PPARγ agonist) induces autophagy in these cancer cells as shown by an increase in lysosomal acidification [250]. These findings have instigated new proposals among cancer researchers that PPAR-induced autophagy benefits cancer cells by protecting them from metabolic stress and cancer therapy [250].

3.2.4. Free Fatty Acids (FFAs)

FFAs are byproducts of lipophagy, and they can also regulate autophagy. The excessive accumulation of FFAs can result in cell death (known as lipotoxicity) [251]. Thus, the levels of FFAs are tightly regulated by lipophagy [252]. Palmitic acid is a common dietary FA [253]. A study on HCC cells showed that palmitic acid was able to induce autophagy flux independent of mTOR [254]. Human HCC HepG2 cells treated with palmitic acid showed activation of PKC and increased levels of diacylglycerol. As mentioned above, diacylglycerol can induce autophagy via the activation of PKC and the dissociation of BCL2 from BECLIN1. Furthermore, PKC inhibition significantly reduced autophagy flux in these cells, confirming that palmitic acid induced autophagy flux via PKC [254].

3.2.5. Omega-3 Polyunsaturated Fatty Acids (n-3 PUFAs)

n-3 PUFAs are involved in the modulation of autophagy [255]. Docosahexaenoic acid (DHA) is one of the most studied n-3 PUFAs [256]. DHA can be synthesized in the body in low amounts through the conversion of α-linolenic acid (ALA), which is a shorter chain omega-3 fatty acid [257] by elongase [258,259]. DHA has antitumor effects on various types of cancer through the modulation of autophagy [260]. DHA has been studied in various cancers [261,262,263] and it exists in many phospholipids that form the cell membrane in different tissues [257]. In p53-wild type prostate cancer cell lines, it has been reported that DHA induces autophagy through AMPK activation resulting in the inhibition of mTOR [264]. In contrast, it induces autophagy through mitochondrial ROS production in p53-mutant prostate cancer cells (PC3 and DU145). The generated ROS lead to the inhibition of Akt and mTOR and, thus, autophagy induction [265]. Interestingly, DHA alone or in combination with other chemotherapeutic agents, can exert anticancer effects by boosting the immune response to cancer cells. It was found that DHA-induced autophagy increases the processing of cancer antigens. The subsequent cross-presentation of these cancer antigens to Antigen Presenting Cells (APCs) improved the immune response by ensuring the production of antitumor T cells [266]. DHA can also induce autophagy in CRC cell lines, Caco-2 and SW620 (by regulating the expression of genes involved in the autophagy process, such as p62, LC3 and ATG14) [267], and the HCC cell line, HepG2 (shown by the accumulation of LC3-II and the formation of LC3puncta) [268]. DHA exerts its anticancer effects in other cancer types, such as GBM and cervical cancer [264], by the induction of autophagy through mTOR inactivation.
DHA can be converted to docosahexaenoyl ethanolamine (DHEA) [269]. It has been shown that DHEA can activate autophagy in MCF-7 breast cancer cells [270]. It can increase the expression of PPARγ, which can, in turn, upregulate PTEN. Finally, PTEN inhibits Akt-mTOR, which subsequently induces autophagy. Moreover, in the same cell line, DHEA was able to phosphorylate BCL2, leading to its dissociation from BECLIN1 and, consequently, inducing autophagy [270]. Although these promising findings highlight the potentiality of DHA and DHEA for anticancer treatment interventions, we will need more in vitro and in vivo studies to better examine their use in anticancer therapies.

3.2.6. Cholesterol

Cholesterol is mainly biosynthesized in the endoplasmic reticulum (ER) of hepatocytes. This process requires acetyl-CoA, which is produced primarily during β-oxidation of FAs in the mitochondria [271]. Cholesterol is one of the major products synthesized by the mevalonate cascade, and is necessary for the formation of cellular membranes, modulation of cell membrane fluidity, cellular structures, and the synthesis of hormones and vitamin D [272]. The rate-limiting step in cholesterol biosynthesis is the conversion of 3-hydroxy-3-methylglutaryl-CoA (HMG-CoA) to mevalonate by HMG-CoA Reductase (HMGCR) [273], which is further processed into cholesterol and other lipid molecules, such as GGPP, FPP, and squalene [274,275]. As discussed previously, the production of lipids, including cholesterol, can be inhibited by the cholesterol-lowering agents known as statins. Statins inhibit the rate-limiting enzyme in this pathway, HMGCR [200].
The current knowledge concerning how cholesterol can modulate autophagy in the context of cancer is very limited. One study has shown that the inhibition of cholesterol biosynthesis by statins in blood cancer cells induces autophagy through the inhibition of mTOR signaling [204]. Autophagy was also induced after the treatment of these cells with methyl-β-cyclodextrin, a cholesterol-reducing agent, indicating that autophagy induction by statins was due to cholesterol depletion. The cholesterol–mTORC1 axis regulated autophagy through Golgi membrane protein 1 (GOLM1) and by interacting with LC3 through an LC3-interacting region (LIR) in hepatocellular cancer cells [276]. The cholesterol of the cytosolic membrane lipid raft regulated the autophagic cell death in breast cancer cell lines (MDM-MB231) via the regulation of caspase-8 activity [277]. We have summarized the impact of lipid metabolites in autophagy process in Table 3.

4. Lung Cancer, Metabolism and Autophagy

Lung cancer is a major contributor to the global cancer burden. It is the second most diagnosed cancer in the world, only recently being surpassed by breast cancer in 2020 [278,279]. Globally, lung cancer accounts for 11.4% of all newly diagnosed cancers, representing 1 in 10 new cases [278,279]. It continues to be the number one cause of cancer-related deaths worldwide, accounting for 1 in 5 cancer deaths in 2020 [280]. The death toll from lung cancer exceeds that of breast, prostate, and colorectal cancer combined [278,280]. The five-year survival rate of lung cancer (18.4%) is lower than colorectal (64.5%), breast (89.6%) and prostate (98.2%) cancers [280]. Metastasis, the dissemination of a primary tumor to other parts of the body, is the major cause of morbidity and mortality in lung cancer. Only 16% of cases of lung cancer are diagnosed at an early stage when the cancer is still localized to the lungs [280]. The five-year survival rate is only 5% for those with distant metastases, compared to 56% for those detected at an early stage [280]. Diagnosis at the advanced stage contributes to the substantial mortality in lung cancer, with more than half of people dying of lung cancer within a year of their diagnosis [280].
In the following sections, we will focus on the impact of metabolism and autophagy in the pathogenesis of lung cancer.

4.1. Etiology, Pathogenesis, and Natural History of Lung Cancer

The majority of lung cancers are either non-small cell lung cancer (NSCLC) or small cell lung cancer (SCLC). NSCLC accounts for more than 80% of all lung cancers. Adenocarcinoma and squamous cell cancers are the most common subtypes of NSCLC [281]. Adenocarcinoma is currently the most common histologic subtype worldwide and histopathologically diagnosed through its specific morphology and immunohistochemical profile that includes thyroid transcription factor-1 (TTF-1) and Napsin [282,283,284]. About 5–15% of adenocarcinomas harbor targetable genetic driver mutations including EGFR, ALK, BRAF and KRAS. The signaling proteins in these genes drive the proliferation of the malignant cells [285]. Programmed cell death-ligand 1 (PDLI) is a cell-surface protein expressed in NSCLC patients and is associated with poor prognosis and a therapeutic target for immune checkpoint inhibitors [286,287,288]. Squamous cell carcinoma (LUSC) was previously the most common subtype but is now the second most common [282,289]. The majority of squamous cell carcinomas present as more central lung cancers (i.e., closer to the central bronchial tree). It is often defined by immunohistochemical staining for p40, p63, and CK5 or CK5/6 [282,290].
Although there is a combination of genetic and environmental risk factors for the development of lung cancer, the strongest risk factor is known to be environmental exposure to smoking [291]. Although there are growing populations of non-smoking individuals developing lung cancer, smoking is estimated to be a risk factor in 90% of patients who develop lung cancer [291]. Smoking exposure as a risk factor for developing lung cancer not only applies to people who smoke but also extends to second-hand smoke (involuntary inhalation of smoke by non-smokers) or third-hand smoke (exposure to smoke deposited on surfaces, such as clothing, furniture, etc.) [292,293]. Although the data on third-hand smoke exposure is developing, there is established evidence that second-hand smoke exposure is associated with a higher risk of developing lung cancer. The evidence suggests that there is a dose–response relationship between second-hand smoke exposure and the incidence of lung cancer among never-smokers [292,293]. For example, women who have never smoked had a 27% higher incidence of lung cancer if their spouse smoked [292].
A common established risk factor for lung cancer is exposure to radon [294,295]. Radon is a gas resulting from the decay of radium and uranium [294,295]. Due to its potential presence in soil and groundwater, it can potentially accumulate in homes [294,295]. Asbestos is also a known risk factor for lung cancer and appears to have a synergistic effect with smoking to increase the risk of lung cancer [296,297]. Asbestos exposure can be due to occupational exposure or to exposure in homes or other buildings (i.e., schools, workplaces) [296,297]. Other environmental/occupational exposures associated with an increased risk of lung cancer are chromium, nickel, and arsenic [298]. Pulmonary fibrosis appears to increase the risk of developing lung cancer even after controlling for the effect of smoking [299]. HIV and HPV infection may also increase the risk of developing lung cancer, although it is unclear if this observation is being driven predominantly by smoking exposures [300,301,302,303].
Genetic factors are also important. The most common genetic risk factors are mutations in EGFR (epidermal growth factor receptors) and ALK (anaplastic lymphoma kinase) [304,305]. EGFR mutations are more common in Asian populations [306].
As NSCLC is the most common lung cancer, we will focus this part of our review on this type of lung cancer. We will first explain the genetic mutations of autophagy in this type of the lung cancer. Then, we will discuss how changes in metabolism could affect autophagy in NSCLC.

4.2. Modulation of Autophagy through Genetic Changes in NSCLC

Components of the Rho GTPase signaling pathway are found to be significantly mutated in NSCLC. For example, inactivating mutations in the Rho GTPase activating protein (RhoGap) ARHGAP35 (p190RhoGAP) are found in ~6% of LUSCs, which results in increased RhoA GTPase activity. Autophagy and Rho GTPase signaling show reciprocal regulation [307]. Previously, it was shown that the knockdown of ROCK1 (Rho-associated kinase 1), a downstream effector of RhoA, was found to enhance starvation-induced autophagy [308]. In addition, increased ROCK signaling inhibited autophagy, whereas the inhibition of ROCK resulted in an augmented autophagy flux with an increased size of early and late autophagic structures [309]. The authors suggested that ROCK1 signaling delimits the elongation phase and the eventual size of newly-formed autophagosomes [309]. Subsequently, it was found that ROCK1 promoted starvation-induced autophagy through the phosphorylation of BECN1 (ATG6) to prevent BECN1–BCL2L1 interaction [310]. Sequestosome-1 (p62 SQSTM1), an autophagy receptor, sequesters RhoA for autophagic degradation [311]. In addition, the knockdown of ATG5 results in enhanced RhoA levels and defective cytokinesis, increasing aneuploidy and motility [311,312]. The knockout (KO) of Atg5, Atg7, and Ulk1 in Mouse Embryonic Fibroblasts (MEFs) resulted in increased cell motility with an amoeboid morphology [313]. Tumor cells have two major forms of motility, mesenchymal and amoeboid-type movements, controlled by the Rac1 and RhoA GTPases, respectively [314]. Interaction with p62 SQSTM1 results in degradation of the Rho guanine nucleotide exchange factor ARHGEF2 (GEF-H1) [313]. Atg7 KO cells show increased ARHGEF2 and RhoA activity [313]. Lung adenocarcinomas (LUADs) show mutations in the ARHGEG2 gene (4%); however, the biological consequence of these mutations is not known at this time. Overall, enhanced RhoA activity increases stress-induced autophagy and motility, which, in turn, sets up a negative feedback loop through autophagy-mediated degradation of RhoA and its GTP exchange factor, ARHGEF2. It has been recently reported that circular RNA Rho GTPase activating protein 10 (circARHGAP10) is an important factor for the development of NSCLC [315]. circARHGAP10 has been shown to be upregulated in NSCLC tissues, cells, and serum-derived exosomes. On the other hand, serum-derived exosomes increased circARHGAP10 expression in NSCLC cells [315]. miR-638 is a target of circARHGAP10 and is involved in its effect on proliferation, migration, invasion, and glycolysis in NSCLC cells [315]. Another investigation showed that low expression of ARHGAP6 is linked to increased metabolic activity in NSCLC, which confirms that ARHGAP6 is a potential tumor suppressor in NSCLC [316]. Therefore, it could be concluded that there might be an indirect link between the effect of small Rho GTPase on glycolysis via the regulation of autophagy.
The BCL2L1 gene is focally amplified in ~8% of lung LUSCs and 3% of LUADs. BCL2L1 binds BECLIN1 [317], inhibiting its ability to activate autophagy [318,319]. Therefore, elevated levels of BCL2L1 would be expected to inhibit autophagy induction [319]. In addition, elevated levels of BCL2L1 would also be anti-apoptotic through the binding of pro-apoptotic proteins, such as BAX and BAK [320]. This inhibition of apoptosis would also impart drug resistance to tumor cells [321]. It is still an open question whether BCL2 and BCL2L1 can alter autophagy directly [322]. One study using BAX, BAK1 knockout cells, to prevent apoptosis induction, showed that overexpression of BCL2, MCL1, and BCL2L1 in the knockout cells did not alter autophagy [323,324]. Only when BAK and BAX were present, was apoptosis-induced autophagy activated. This suggested that BCL2-like proteins regulate autophagy indirectly by inhibiting BAX/BAK1-mediated apoptosis [323,324]. A follow up study, from another group, showed that the BH3 peptidomimetic drug ABT-737 induced autophagy in BAX/BAK1 double knockout cells when longer durations and higher doses were used, which correlated with the disruption of BCL2–BECN1 binding [325]. However, a further study from Reljic et al. suggested that higher doses of BH3 peptidomimetics caused off target effects [326]. Using the cBioPortal.org web site and TCGA data, LUAD patients with BCL2L1 amplification showed a significant reduction in progression-free survival but not overall survival (Figure 8A,B). However, the number of patients with amplified BCL2L1 was small (n = 7) and more patients with this genotype should be studied [327,328,329]. Further studies are needed to determine whether the effects of BCL2L1 amplification is due to alterations of apoptosis or autophagy. One of the hallmarks of cancer is the evasion of apoptosis [330,331], which is associated with an increase in invasion and chemo-resistance [332], A part of the decrease in apoptosis in tumor cells is correlated with changes in the balance between the BCL2 family of pro- and anti-apoptotic proteins [333]. Therefore, the process of apoptosis, autophagy, and the metabolic status of cancer cells, including NSCLC tumor cells, are tightly linked to BCL2 family proteins and the genetic changes of this family could have a direct effect on the regulation of tumor cell response to chemotherapy or tumor invasion.
The frequent mutation in PI3 kinase signaling pathway genes in lung tumors results in activation of the mTORC I kinase complex, which, in turn, would inhibit autophagy (for review [334]). This suppression of autophagy by oncogenes or tumor suppressors may cause an acceleration of early-stage tumor formation. In a K-rasG12D-induced mouse model of NSCLC, it was shown that the knockout of Atg5 resulted in increased tumor growth during the early stages compared with autophagy competent mice [335]. The accelerated early oncogenesis was due to increased infiltration of FOXP3+ regulatory T cells, which subverted immunosurveillance [335]. However, the Atg5 knockout mice showed increased overall survival due to the reduced progression of adenomas to adenocarcinomas, which could be overcome by Tp53 loss [335]. Another study using Atg7 knockout mice showed that autophagy was needed for sustained K-ras-induced tumor cell proliferation and progression [205]. The lack of Atg7 resulted in oncocytoma formation due to the accumulation of morphologically abnormal mitochondria [205]. Tp53 loss allowed the progression to adenocarcinomas in this study; however, the acceleration of early oncogenesis was not seen [205]. It was previously shown that wild type TP53 under drug or stress treatments inhibits mTOR1 kinase complex activity and, in turn, activates autophagy [336]. In tumors with TP53 loss, mTOR1 would not be inhibited resulting in reduced stress induced autophagy. Therefore, the inhibition of autophagy may be another tumor suppressor function of TP53 [336].

4.3. Autophagy, Glucose Metabolism, and Lung Cancer

As we have discussed in previous sections, glucose metabolism is tightly linked to autophagy in different types of cancers. In this section, we focus the discussion on the impact of glucose metabolism and autophagy in lung cancer.
LKB1 is a master tumor suppressor that limits cell growth through repression of the mammalian target of rapamycin complex 1 (mTORC1) and activation of AMPK to control metabolism. Therefore, the loss of LKB1 leads to uncontrolled mTORC1 activation, activation of HIF-1α and, consequently, angiogenesis and the Warburg effect [337,338,339]. Co-mutations in the KRAS proto-oncogene and the LKB1 tumor suppressor gene frequently occur in hyper-metabolic and aggressive human lung adenocarcinoma tumors. mTORC1 is hyperactive in LKB1/KRAS mutant lung adenocarcinoma cancer cells and, thus, targeting this pathway by mTOR inhibitors might represent an effective way of treating lung adenocarcinoma [340,341]. In 2018, researchers were able to use the coupling of [18F]-2-fluoro-2-deoxy-D-glucose positron emission tomography (18F-FDG PET)/computed tomography (CT) imaging to quantitative immunohistochemistry (qIHC) as an effective method to measure metabolic changes and therapeutic responses caused by a selective catalytic mTOR kinase inhibitor in mouse models of lung cancer. Their analysis showed that the mTOR inhibitor MLN0128 could suppress tumor growth and glycolysis in genetically engineered mouse models (GEMMs) of lung cancer with KRAS and LKB1 co-mutations, as shown by reduced 18F-FDG consumption [342]. Interestingly, in addition to the AMPK-mTOR signaling pathway, SIRT3/ HIF-1α was significantly upregulated in KRAS mutant human lung adenocarcinoma lines (H460, A549, and H358) [343]. Lian-Xiang Luo demonstrated that the simultaneous use of an antitumor (Honokiol) and an autophagy inhibitor, 3-methyladenine (3-MA) triggered pro-death autophagy through activating mitochondrial SIRT3, suppressing HIF-1α expression, and disrupting oncogenic KRAS-mediated RAF/MEK/ERK and PI3K/AKT/mTOR signaling pathways [343]. The last step of glycolysis is the conversion of PEP to pyruvate and ATP, which is catalyzed by PK. In KRAS mutant cells, PKM2 is hyperactivated and acts as a histone kinase in the nucleus and upregulates the expression of c-Myc. Their overexpression affects metabolic functions including the enhanced expression of glucose transporter GLUT1, the activity of glycolytic enzymes (HK2, PFK, enolase1), and LDHA, with the overproduction of lactic acid and non-metabolic pathways (upregulation of autophagy). Even though PKM2 activates the mTOR pathway resulting in a block in autophagy in KRAS mutants. Autophagy is hyperactive in KRAS mutant cancer cells [344]. M Morita used KRAS in PKM2 mutant lung cancer mouse models (KrasG12D or KrasG12V) and in SCLC cells [345]. In addition to KRAS, oncogenic promoters such as the EGFR support glycolytic metabolism in lung tumors. EGFR signaling activates the class I PI3K-AKT-mTORC1 pathway. AKT and mTORC1 phosphorylate BECLIN1 on serine residues, leading to downregulation of VPS34 activity and autophagy. In contrast, ULK1 and AMPK stimulate autophagy by phosphorylating BECLIN1 and promoting the formation of the active BECLIN1–VPS34 complex. mTORC1 also suppresses autophagy initiation by phosphorylating and inhibiting ULK1. These EGFR-dependent events are blocked by tyrosine kinase inhibitors (TKIs) [346]. Therefore, it can be concluded that autophagy plays an important role in KRAS and EGFR mutant lung tumors through the small Rho GTPase and PI3-AKT-mTORC1 pathway.
M Ye et al. showed that overexpression of GLUT1 and MCT-4 led to a significant increase in glycolysis and autophagy in ER6 cells (erlotinib-resistant sub-line of HCC827) compared to EGFR-TKI sensitive HCC827 lung adenocarcinoma cell lines. They demonstrated that the combination of an AKT inhibitor (MK2206), to suppress AKT phosphorylation or autophagy, with glucose deprivation and a GLUT1 specific inhibitor (STF-31) could tackle the resistance to erlotinib in NSCLC [347]. Other compounds that target the EGFR are flavonoids (natural products with variable phenolic structures that reduce the risk of cancer). A recent study revealed that Scandenolone (a natural isoflavonoid) directly attaches to the ATP-binding site of EGFR to impede the phosphorylation of the AKT/ERK signaling pathway, which is essential for sensitizing SK-MEL-28 melanoma cell lines to intrinsic apoptosis and blocking autophagy flux. In fact, exposure to Scandenolone inhibits the AKT/ERK pathway, leading to an increase in Bad, one of the BH3-only proteins, and a reduction in Bcl-2, an anti-apoptotic protein, disrupting not only the inhibitory interactions between BECLIN1 and Bcl-2 but also the activation of the lysosomal protease Cathepsin B, which induces the induction of autophagy in these cells [347]. Another study examined whether adding an additional isopentenyl group to the Scandenolone structure would improve its anticancer effect. This new antitumor compound, named Final-2, could not only suppress cancer cell growth by promoting apoptosis and blocking autophagy but also by inhibiting the expression of the oncogenes KRAS, c-MYC, and glucose transporters (GLUTs), such as GLUT1, GLUT3, and GLUT4, which resulted in the downregulation of glucose metabolism in the human lung cancer cells A549 [348].
It has been shown that under stress conditions (hypoxia, glucose deprivation, low pH, ROS accumulation, and Ca2+ homeostasis perturbation), lung cancer cells experience ER stress and activate GRP78/BiP-related unfolded protein response (UPR) signaling. Upon ER stress, GRP78/BiP can trigger autophagy to improve the survival of cancer cells through three different UPR pathways including eukaryotic translation initiation factor 2 alpha kinase 3/PKR-like ER kinase (EIF2AK3/PERK), endoplasmic reticulum to nucleus signaling 1/inositol-requiring enzyme 1 (ERN1/IRE1), and activating transcription factor 6 (ATF6), which target the ATG16L1 complex, MAPK/JNK and X-box binding protein 1 (XBP1), and the AKT-mTOR pathway, respectively [349,350,351,352]. It has been reported that overexpression of GRP78 is often seen in advanced-stage lung cancer, which is associated with a poor prognosis [353]. A recent study has reported that the inhibition of ATPase activity of GRP78 by a thiazole benzenesulfonamid (HA15) provokes ER stress-mediated UPR and induces pro-apoptotic autophagy in the lung cancer cell line A549 [354]. Glucose starvation induces ER stress-related GPR78 in NSCLC cells, leading to autophagy as a cyto-protective process. In this regard, the microtubule cytoskeleton is involved in autophagy signaling and βIII-Tubulin is one of the β-tubulin isotypes overexpressed in glucose-starved NSCLC cells. On the other hand, GRP78 interacts with βIII-tubulin; however, the functional consequences of this association are unknown. Tubulins, and in particular βIII-tubulin, associate with a variety of glycolytic enzymes including PK, phosphofructokinase, aldolase, hexokinase, GAPDH, and LDH. These observations suggest an inherent link between the microtubule cytoskeleton, glycolysis, autophagy, and the initiation of ER stress responses [355]. βIII-Tubulin enables rapid GPR78/Akt activation in response to glucose starvation and decreases the reliance of cells on glycolytic metabolism to maintain cell survival and proliferation. Amelia L. Parker, in 2016, indicated that suppression of high βIII-tubulin expression with 4-phenylbutyric acid and N-acetylglucosamine ameliorates ER stress and reduces autophagy by delaying the association of GRP78 with Akt in glucose-starved NSCLC cells [356]. The major stressors influencing autophagy flux are Ca2+ homeostasis disturbances and glucose starvation. Thus, voltage and ligand-gated calcium channels (such as T-type, L-type, and IP3-R), and GLUTs, which are often upregulated in cancer, are also targets for autophagy regulation. Both Ca2+ and glucose influx perturbations might be signals for UPR and ER stress activation, leading to the ER chaperones such as GRP78/BiP stimulating autophagy through three different molecular pathways, such as ERN1/IRE1-JNK-Bcl-2, EIF2AK3/PERK-eIF2a-ATF4, or ATF6-XBP1-ATG. In addition, both starvation and Ca2+ perturbations may lead to activation of ER stress-mediated autophagy in cancer cells via distinct Ca2+ channel activation, CaMKKβ, which mediates AMPK-dependent inhibition of mTORC1 or by activating the IP3-R-BECLIN1-Bcl-2 pathway. Therefore, the development of Ca2+ channel and GLUT blockers would be a novel cancer therapy. Hong-Kun Rim et al. revealed that co-blocking the T-type Ca2+ channel and GLUT with KYS05047 and KYS05090 synthesized compounds can induce autophagy and apoptosis in human lung adenocarcinoma A549 cells and xenografts resulting in elevated ROS generation and reduced intracellular Ca2+ levels and glucose uptake [352,357,358,359].
Glutamine, the blood’s most abundant amino acid, is metabolized through glutaminolysis, catalyzed by glutaminase (GLS) and leucine-dependent glutamate dehydrogenase (GDH), to produce αKG for the Krebs cycle, nucleotides, and other amino acids. The expression level of GLS is regulated by the oncogene c-MYC, and its upregulation is a hallmark of tumor growth. In fact, enhancing glutaminolytic αKG production stimulates lysosomal translocation and the activation of mTORC1. mTORC1 (master switch for cell proliferation and autophagy) is regulated by glutaminolysis [360]. Recently, RC Bruntz et al., by using [13C6]-glucose and [13C5,15N2]-glutamine as tracers, showed that Selenite inhibited glutaminolysis by suppressing GLS1 expression and inducing excess ROS production, which mediates autophagy and eventual cell death in NSCLC A549 cells [361]. The upregulated glutaminolysis is a feature of cancer cell metabolism, activated by extracellular lactate signaling. Under oxygenated conditions in the tumor environment, lactate enters the cells via MCT1, stabilizing hypoxia-inducible factor-2α (HIF-2α), and then transactivating c-MYC. This, in turn, activates the expression of glutamine transporters (LAT1 and ASCT2) and glutaminase 1 (GLS1), resulting in an increased proportion of glutamine absorption by tumor cells and triggering OXPHOS [362,363]. Forming an acidic tumor microenvironment is a common event in lung cancer due to the production of lactate by the Warburg effect. For lung cancer cells to survive in such a situation in vivo or in vitro, autophagy is induced by the acidic pH through ER stress-related GRP78. Glucose starvation and the inhibition of MCT-1 with a-cyano-4-hydroxycinnamate (CHC) in mice injected with hypoxic Lewis lung carcinoma (LLC) cells resulted in decreased tumor growth [364]. The knockdown of GRP78 by siRNA reduced autophagy activation under acidic conditions in the human lung cancer cell lines (A549 and NCI-H226) and enhanced acid-induced apoptosis [365]. Glucose depletion (GD) induces necrotic cell death in A549 lung carcinoma cells, activating an inflammatory response due to the release of high-mobility group box 1 (HMGB1). SC Lim et al. revealed that sodium salicylate, an active metabolite of aspirin, could inhibit HMGB1 release, Cu/Zn superoxide dismutase1 (SOD1) release, and ROS production in A549 cells cultured in GD medium, which, in turn, led to the anti-inflammatory effect. Moreover, aspirin could activate AMPK, inhibit mTOR to induce autophagic cell death, and suppress tumor development [366,367,368]. Glutamate oxaloacetate transaminase 1 (GOT1), by catalyzing the conversion of aspartate to oxaloacetate (OAA), provides the metabolites necessary for gluconeogenesis and, thus, GOT1 is involved in the maintenance of redox homeostasis and the regulation of glycolytic metabolism in normal cells [369], while the inhibition of GOT1 leads to increased glucose utilization and lactate production, resulting in a decrease in cellular pH [370]. X Zhou in 2018 demonstrated that in GOT1-null 143B osteosarcoma cells or GOT1 siRNA knockdown in A549 lung cancer cells resulted in the elevated expression of autophagy-related genes (ATG5 and Beclin1) and increased secretion rate of lactate compared with wild type cells [370]. Placing these cells in a glucose-free culture medium caused a reduction in cell viability [370].
Nickel compounds are mainly associated with metabolic changes in sino-nasal and lung cancers. YT Kang et al., in 2017, documented that hexokinase 2 (HK2), involved in the first step of glycolysis, plays a significant role in Nickel-induced autophagy in lung bronchial epithelial cells (BEAS-2B). They determined that the inhibition of HK2 by the antihyperglycemic drug metformin and the knockdown of lipocalin 2 (LCN2) by shRNA mitigated NiCl2-mediated autophagy and induced apoptosis in BEAS-2B cells [371]. VK Chaudhri et al., by evaluating two cell groups including primary human lung tumor CAFs and “normal” fibroblasts (NF) isolated from nonneoplastic lung tissue located at least 5 cm away from the tumor, determined that metabolic alterations are not limited to cancer cells. They demonstrated that the reduction in the amount of glucose in the medium from 25 to 5 mmol/L also increased dipeptides and autophagy in NFs [372].
We have summarized the role of various glycolytic enzymes in relation to autophagy in lung cancer in Table 4.

4.4. Lipid Metabolism Regulation of Autophagy in Lung Cancer

4.4.1. HMGCR Inhibition

Statins (inhibitors of HMGCR) are widely used for lowering serum cholesterol. They also have several pleiotropic effects including their potential effects in different types of cancers [201,381]. A recent xenograph study in nude mice by Yang et al. reported that the use of the statin fluvastatin could significantly inhibit the metastasis of lung adenocarcinoma cells to bone. The anti-metastatic effect of fluvastatin was shown to be due to the induction of autophagy in lung adenocarcinoma cells. The chemical inhibition of autophagy by 3-methyadenine (3-MA) or Bafilomycin A1 (Baf A1) and the genetic inhibition of autophagy by knocking down Atg5 or Atg7 hampered the suppressive effects of fluvastatin on bone metastasis of lung cancer cells. These findings highlight the importance of the mevalonate pathway inhibition by statins (through HMGCR inhibition) on the regulation of autophagy and how this leads to the suppression of bone metastasis of lung cancer cells [382].
Another statin that has been extensively studied in cancer is lovastatin. It has been reported that lovastatin has inhibitory effects on the growth and proliferation of NSCLC adenocarcinoma lung cancer cells A549 [383]. In addition, it increases the sensitivity of A549 lung cancer cells to ionizing radiation and inhibits their colony forming ability. Interestingly, these effects by lovastatin were reversed when mevalonate was used in the treatment as well [383]. This indicates that the effects of lovastatin were dependent on its ability to inhibit the mevalonate pathway. In this same study, the treatment of A549 cells with lovastatin resulted in the induction of cell death and induced AMPK phosphorylation, an upstream regulator of autophagy, by directly activating ULK1 through phosphorylation [383]. Collectively, these results show that the lovastatin-induced cell death could be due to AMPK-mediated lethal autophagy. These results emphasize the crucial role that the mevalonate pathway can play on the regulation of autophagy, which, consequently, impacts lung cancer cells’ proliferation and clonogenicity and improves their sensitivity to ionizing radiation [382].

4.4.2. SPHK1

The expression of SPHK1 is significantly increased in NSCLC patients. In a study by Zhu et al. [384], it was found that SPHK1 expression is highly elevated in A549 cells. They showed that SPHK1 overexpression is enhanced in A549 cells while SPHK1 knockdown prevented the migration and invasion of these cells. In fact, the expression of E-cadherin was reduced, and the expression of SNAIL was increased in overexpressed cells. More importantly, SPHK1 induced AKT activation. Since the mTOR-AKT pathway is a master regulator of autophagy, it is safe to say that SPHK1 promotes the invasion and migration of lung cancer cells simultaneously with (and possibly through) autophagy inhibition [384]. This evidence highlights that the targeting of SPHK1 is a promising candidate for NSCLC treatment. Consistent with these findings, another research group evaluated the expression levels of SPHK1 in tissues obtained from 176 NSCLC patients using immunohistochemistry (IHC) [385]. The expression of SPHK1 was correlated with clinicopathological factors and it was shown that the patients with a high expression of SPHK1 that were treated with adjuvant platinum-based chemotherapy had a significantly lower survival rate. This indicates that even performing IHC on tissues from NSCLC patients can be a potential predictive tool for the survival of NSCLC patients.
In addition, another study used SPHK1 expression levels to predict survival and progression in NSCLC patients. Findings from this study revealed that SPHK1 expression was significantly increased in 218 tissue samples from NSCLC patients [386]. This was associated with an inhibition of doxorubicin- or docetaxel-induced apoptosis and the induction of anti-apoptotic factors. Conversely, the chemical inhibition and knockdown of SPHK1 remarkably increased NSCLC cells’ sensitivity to apoptosis triggered by anticancer agents both in vitro and in vivo. Regarding the association of SPHK1 and autophagy, it was also shown that high SPHK1 expression induced the activation of the PI3K/AKT pathway, which, as mentioned earlier, can inhibit autophagy via mTOR signaling. Interestingly, inhibition of the PI3K-AKT pathway failed to induce anti-apoptotic effects by SPHK1; therefore, highlighting this pathway and its downstream pathway (autophagy) as a mediator of SPHK1 effects on NSCLC cells. In agreement with these findings, it was found that SK1-I, a competitive SPHK1 inhibitor, induces apoptosis in cancer cells and suppresses the metastasis of cancer cells into lymph nodes [387]. As mentioned in previous sections, phosphoinositides can modulate autophagy. In addition, any deregulation in the function of PI3K leads to autophagy deregulation because PI3K is part of the PI3K-AKT-mTOR pathway [79].

4.4.3. PA

Phospholipase D (PLD) is the enzyme that hydrolyses phosphatidylcholine (PC) to PA and choline [388]. The expression and activity of a single nucleotide polymorphism (SNP) in the Phospholipase D1 (PLD1) gene (SNP A2698C) is significantly increased in a variety of human cancers [389]. It has been shown that PLD1 expression is significantly increased in lung surgical specimens obtained from NSCLC patients compared to normal lung tissues [390]. Moreover, PLD1 has been associated with the development and aggression of lung cancer [386].
As discussed in earlier sections, PA is involved in (1) the formation of autophagosome curvature, and (2) the inhibition of mTOR and, therefore, the induction of autophagy [81,391]. It can be suggested that the higher expression of PLD1 in NSCLC leads to the higher production of PA and, consequently, higher rates of autophagy induction in NSCLC cells [389]. In addition, it was reported some time ago that upon activation (through bradykinin and sphingosine 1), PLD1 is involved in the PKC pathway in NSCLC A549 cells [392,393], which can regulate autophagy. This further suggests the regulation of autophagy by lipid metabolism in NSCLC.
Lipin-1 is the enzyme that converts PA to diacylglycerol (DAG) [394]. Lipin-1 expression is significantly increased in NSCLC cell lines and patient tissues. It has recently been shown that the elevated expression of Lipin-1 is associated with poor prognosis of NSCLC patients. Moreover, Lipin-1 suppression greatly reduces the growth and viability of NSCLC cells while it has almost no significant effects on non-cancerous cells in the lung [395]. In addition, the same research group demonstrated the loss of Lipin-1 sensitized NSCLC cells to cisplatin. Since PA is involved in the formation of autophagosome curvature [388], the disruption of Lipin-1 can potentially help accumulate the PA levels and consequently boost autophagosome formation during autophagy activation. This can potentially reveal the impact of PA in autophagy regulation during NSCLC progression. These findings suggest that targeting Lipin-1 (and, therefore, the conversion rate of PA to DAG) in combination with other NSCLC anticancer agents could be a promising treatment intervention for lung cancer [395].
Interestingly, it has been observed that the levels of Interleukin-8 (IL-8 or CXCL8), which is a pro-inflammatory chemokine, are significantly elevated in different cancers including NSCLC. This increased IL-8 level has been associated with the enhanced activity of PLD and the activation of AKT/mTOR, which can directly regulate autophagy [389]. Furthermore, IL-8 activates PLD, which consequently converts PC to PA. These findings similarly show the involvement of PDL (and hence PA) in the regulation of autophagy in lung cancer cells.
Due to their importance in cancer development and progression, targeting PDL1 and its signaling molecules has been suggested as a treatment strategy in different cancers with an increased PLD1 activity and expression [388].

4.4.4. Peroxisome Proliferator-Activating Receptors (PPARs)

As mentioned previously, PPARs are ligand-dependent nuclear transcription factors. In order to be functional as transcription factors, PPARs need to bind to lipid ligands (e.g., oleic acid, linoleic acids, linolenic acids, prostaglandins, eicosanoids, and oxidized lipids) through lipid-binding proteins [244,245,246]. A study [396] showed that the treatment of A549 NSCLC cells with troglitazone (an activator of PPARγ) induced the conversion of LC3-I to LC3-II and increased the degradation of p62, indicating activation of autophagy flux while inducing apoptosis in A549 cells. Consistently, the activation of PPARγ and, subsequently, autophagy activation was inhibited when the cells were treated with a PPARγ antagonist, GW9662. This highlights troglitazone as a potential candidate for the combination therapy of apoptosis-resistant NSCLC. Moreover, another study on human NSCLC cell lines, HCC827 and H1650, demonstrated that PPARγ agonists sensitize PTEN-deficient lung cancer cells to EGFR TKIs via autophagy induction [397]. One of the mechanisms of developing resistance in lung cancer is through PTEN loss, which contributes to the resistance to EGFR TKIs [398]. It was shown that PPARγ agonists increased the expression of PTEN and, consequently, inhibited the PI3K-AKT signaling pathway, which will activate autophagy through mTOR inhibition. This results in significant sensitivity to gefitinib (an EGFR TKI). The mechanism of this sensitivity to EGFR-TKIs was found to be the induction of autophagy. In fact, the findings by this group showed that PPARγ agonists induced higher levels of LC3-II accumulation and the degradation of p62 in NSCLC cells. In addition, genetic silencing of ATG5 eliminated this potentiation effect by PPARγ agonists. Therefore, the combined use of PPARγ agonists and EGFR TKIs have recently been suggested for the treatment of lung cancer patients. In an animal study [399], it was shown that the treatment of mice with pioglitazone (synthetic PPARγ ligand) remarkably prevented lung tumor formation induced by 4-(methylnitrosamino)-l-(3-pyridyl)-lbutanone (NNK). In addition, it was found that endogenous PPARγ was decreased well before lung tumor formation, indicating a role for the endogenous PPARγ molecular pathway in lung cancer development. These findings show that the potential increase in the activity of PPARγ using its ligands not only could be inhibitory for lung tumor formation but also beneficial for the treatment of lung cancer. Moreover, assessment of the activity levels of PPARγ may be useful as biomarkers for lung cancer.

4.4.5. Omega-3 Polyunsaturated Fatty Acids (n-3 PUFAs) (Docosahexaenoic Acid (DHA), Docosahexaenoyl Ethanolamine (DHEA), EPA)

Free lipids can also regulate autophagy during lung cancer, although only a handful of studies have investigated this relationship. n-3 PUFAs and, in particular, DHA or EPA, have been defined as important preventive lipids against lung cancer development [400,401]. They can exert their effects through different ways including cell surface receptor functions, membrane fluidity, and the enhancement of oxidative stress, all of which might induce autophagy [401]. One study [402] demonstrated that both DHA and EPA caused the formation of autophagosomes in a concentration-dependent fashion in the NSCLC A549 cell line. This was confirmed by the finding that treatment of A549 cells with the autophagy inhibitor, 3-MA, reversed the formation of autophagosomes. This autophagy inhibition also prevented apoptosis (reduced activation of caspase-3/7) and increased cell viability, showing that DHA and EPA-induced autophagy acts to enhance the apoptotic cell death of A549 cells. It has also been reported that the treatment of A549 cells with DHA or EPA induced autophagy in a dose- and time-dependent manner [262]. The results from this study also showed that the anti-tumor effects of DHA and EPA were associated with the activation of autophagy.

5. Conclusions and Future Directions

Autophagy plays an important role at different stages of a tumor’s development, progression, and metastasis. During the early stages of tumor development, it prevents tumorigenesis by the removal of damaged mitochondria and misfolded proteins, while it supports tumor progression in the later stages of tumor formation. Autophagy also regulates tumor metastasis both positively or negatively; however, it is dependent on the type of tumor. As we have discussed in our article, both glucose and lipid metabolism can regulate autophagy via different mechanisms to compensate for the high ATP demand of cancer cells, as summarized in Figure 9. The enzymes of glycolysis can directly interact with the autophagy machinery (HK2) [107] to change their activity or expression levels or to directly induce changes in autophagy in tumor cells (PFKFB4) as an adaption to cellular stress conditions [124,125]. On the other hand, the major adaptations of lipid metabolism observed in cancer are generally associated with changing the autophagy pathway in favor of the increased aggressiveness of cancer cells. Therefore, targeting lipid metabolism or glycolysis is an attractive therapeutic strategy.
The interconnection between lipid/carbohydrate metabolism and autophagy and also the role of these metabolic pathways in the regulation of autophagy could be harnessed to develop new clinical interventions against cancer. There have been multiple approaches, including the development of small molecule inhibitors of enzymes that catalyze ceramide catabolism, synthetic ceramide analogs, inhibitors of sphingosine kinase, and antagonists of the S1P receptor [403,404]. Other approaches also exist; for instance (a) reactivation of the SMase, SPL, and S1P phosphatase genes that are suppressed in cancer cells; (b) targeting and hence inhibiting the dihydroceramide desaturase; and (c) using small molecules to activate SMase enzymes [405]. A study by Hernandez-Tiedra et al. [406] showed that the treatment of U87 glioma cells with Delta 9-tetrahydrocannabinol (THC, the main active component of Cannabis sativa) enhances the ratio of dihydroceramide:ceramide in the endoplasmic reticulum of these cells. This ratio change has an impact on the autophagosomes and autophagolysosomes leading to the permeabilization of lysosomal membranes, and the release of hydrolases and cathepsins. Subsequently, this will result in the activation of apoptosis in these cells. Acid ceramidase (AC) is a lysosomal hydrolase that catalyzes the conversion of ceramide into fatty acid and sphingosine. AC has been reported to regulate the levels of molecules that have pro- or anti-apoptotic effects [407,408]. Tumors with high levels of AC expression and, thus, lower ceramide levels, demonstrate higher protection from stress stimuli such as chemotherapy. An innovative approach to avoid and circumvent cancer resistance to single and combined chemotherapy, known as multidrug resistance (MDR), is, thus, the combined use of chemotherapeutic agents and AC inhibitors to better disarm cancer cells [409]. As an example, targeting lipid metabolism to control autophagy in cancer has been widely investigated. One of the important regulators of fatty acid composition in the cell is stearoyl-CoA desaturase 1 (SCD1), which is involved in monounsaturated fatty acid (MUFA) synthesis from saturated fatty acid (SFA) [410]. Furthermore, a connection between SCD1 and autophagy has been reported. In fact, ULK1 inhibition reduces SCD1 expression in liver cells, increasing the SFA/MUFA ratio and, subsequently, leading to cell death [411]. In addition, the autophagy of lipid molecules (lipophagy) provides the necessary fatty acids that are essential for appropriate mitochondrial respiration [412]. Given the lipid metabolism and autophagy crosstalk, and the key role of SCD1 in lipid metabolism, SCD1 is a promising target for developing novel anti-cancer treatments. SCD1 inhibitors, including A939572, CAY10566, MF-438 and CVT-11127, significantly decrease the proliferation rate and induce apoptosis in different cancer types including lung cancer [413,414,415]. That said, due to the adverse effects, most of the studies have not progressed into clinical trials after the pre-clinical studies. Furthermore, due to the SCD1 and autophagy interconnection, new combination therapies with both SCD1 inhibitors and autophagy inducers/inhibitors have been suggested [416,417]. It is, however, worth mentioning that the most effective cancer therapy should be designed based on the biological context; for example, the presence of mutations, tissue type, cancer heterogeneity, etc.
On the other hand, the alteration of various mitophagy modulators has been shown to play a role in tumor progression (for review [418]). LUAD lung cancer patients with a BNIP3L deletion did not show significant differences in overall or progression-free survival. BNIP3L acts as an adaptor molecule that targets mitochondria for turnover and recruits components of the mitophagy machinery to the mitochondria [419]. Another study found that 26.7% of lung tumor samples showed reduced expression of BNIP3L but no rearrangement or mutation of the gene was seen [420]. The closely related BNIP3 had been shown to act like a tumor suppressor in triple negative breast cancers [421,422]. When deficient, it results in increased metastasis and an increase in damaged mitochondrial mass, resulting in the upregulation of HIF1α and reactive oxygen species (ROS) levels [421,422].
The PRKN (PARK2 or PARKIN) gene is a site of frequent loss of heterozygosity in NSCLCs [423]. Transfection of the PRKN gene into a deficient lung cancer cell line reduced in vivo tumor growth but had no effect on in vitro cell growth [423]. Lung cancer mutated forms of PRKN show impaired mitophagy and reduced cell death after mitophagy induction [424]. PINK1 phosphorylation derepresses the PRKN E3 ubiquitin ligase resulting in the ubiquitination of mitochondrial substrates, committing mitochondria to mitophagic turnover (for review [425]). LUSC patients with a PRKN mutation or homozygous deletions (7%) showed significantly reduced progression-free survival. Why LUSCs and not LUADs show this reduced progression-free survival when mitophagy genes are deleted is not known at this time but further study is warranted. In breast cancer, PRKN targets HIF-1α for ubiquitination, controlling its stability and resulting in metastasis suppression [426]. Hence, more work is needed to determine the contributions of mitophagy and HIF-1α stability to PRKN’s tumor suppressor function.
Therefore, a greater understanding of the specific roles played by lipid and glycolytic metabolites and/or the enzymes involved in lipid metabolism and glycolysis in cancer, and understanding the role of mitophagy in cancer, is still needed. Our current knowledge of the role that these pathways can play in the regulation of autophagy in different cancers is very much limited to a few studies. Future studies will, undoubtedly, provide us with promising effective candidates for the prevention and treatment of different cancers.

Author Contributions

J.A. prepared the draft for the lipid metabolism sections 1249 and helped S.G. in the final edit of the paper. M.K. prepared the draft for the glucose metabolism sections and prepared Figure 2, Figure 3, Figure 4, Figure 5, Figure 6 and Figure 7 in the draft. N.S. prepared the 1251 draft for the lactate pathway and prepared the Figure 9 draft. S.L. prepared the 1252 draft of the general autophagy section and finalized Figure 1 and Figure 9. A.R. supervised the 1253 draft of the preparation of the lipid metabolism sections. B.K. and N.A. prepared the 1254 draft of the clinical lung cancer section. F.M. helped in the general me-1255 draft catabolism section. M.R.M. drafted the genetics of autophagy in the lung cancer section and helped in the preparation of the final 1256 draft. S.G. supervised the whole team. de-1257 signed the paper and the final proof of the paper. All authors have read and agreed to the published version of the manuscript.

Funding

S.G., B.K., N.A. and M.M. are supported by the CancerCare Manitoba Operating grant (763117252) funded by CancerCare Manitoba Foundation. J.A. is supported by a Vanier CIHR Ph.D. studentship. Navjit Singh was supported by the Max Rady College of Medicine’s BSc (Med) program.

Acknowledgments

Javad Alizadeh was supported by a Vanier Ph.D. studentship. Navjit Singh was supported by the Max Rady College of Medicine’s BSc (Med) program.

Conflicts of Interest

The authors have no conflict of interest to declare.

References

  1. Esmaeili, Y.; Yarjanli, Z.; Pakniya, F.; Bidram, E.; Los, M.J.; Eshraghi, M.; Klionsky, D.J.; Ghavami, S.; Zarrabi, A. Targeting autophagy, oxidative stress, and ER stress for neurodegenerative disease treatment. J. Control. Release 2022, 345, 147–175. [Google Scholar] [CrossRef]
  2. Shojaei, S.; Suresh, M.; Klionsky, D.J.; Labouta, H.I.; Ghavami, S. Autophagy and SARS-CoV-2 infection: Apossible smart targeting of the autophagy pathway. Virulence 2020, 11, 805–810. [Google Scholar] [CrossRef]
  3. Dalvand, A.; da Silva Rosa, S.C.; Ghavami, S.; Marzban, H. Potential role of TGFBeta and autophagy in early crebellum development. Biochem. Biophys. Rep. 2022, 32, 101358. [Google Scholar] [CrossRef] [PubMed]
  4. Mizushima, N.; Komatsu, M. Autophagy: Renovation of cells and tissues. Cell 2011, 147, 728–741. [Google Scholar] [CrossRef] [Green Version]
  5. Ghavami, S.; Gupta, S.; Ambrose, E.; Hnatowich, M.; Freed, D.H.; Dixon, I.M. Autophagy and heart disease: Implications for cardiac ischemia-reperfusion damage. Curr. Mol. Med. 2014, 14, 616–629. [Google Scholar] [CrossRef] [PubMed]
  6. Ghavami, S.; Zamani, M.; Ahmadi, M.; Erfani, M.; Dastghaib, S.; Darbandi, M.; Darbandi, S.; Vakili, O.; Siri, M.; Grabarek, B.O.; et al. Epigenetic regulation of autophagy in gastrointestinal cancers. Biochim. Biophys. Acta Mol. Basis Dis. 2022, 1868, 166512. [Google Scholar] [CrossRef]
  7. Li, W.W.; Li, J.; Bao, J.K. Microautophagy: Lesser-known self-eating. Cell. Mol. Life Sci. CMLS 2012, 69, 1125–1136. [Google Scholar] [CrossRef]
  8. Cuervo, A.M.; Wong, E. Chaperone-mediated autophagy: Roles in disease and aging. Cell Res. 2014, 24, 92–104. [Google Scholar] [CrossRef] [Green Version]
  9. Mokarram, P.; Albokashy, M.; Zarghooni, M.; Moosavi, M.A.; Sepehri, Z.; Chen, Q.M.; Hudecki, A.; Sargazi, A.; Alizadeh, J.; Moghadam, A.R.; et al. New frontiers in the treatment of colorectal cancer: Autophagy and the unfolded protein response as promising targets. Autophagy 2017, 13, 781–819. [Google Scholar] [CrossRef] [PubMed]
  10. Seo, A.Y.; Lau, P.W.; Feliciano, D.; Sengupta, P.; Gros, M.A.L.; Cinquin, B.; Larabell, C.A.; Lippincott-Schwartz, J. AMPK and vacuole-associated Atg14p orchestrate mu-lipophagy for energy production and long-term survival under glucose starvation. eLife 2017, 6, e21690. [Google Scholar] [CrossRef]
  11. Martelli, A.; Omrani, M.; Zarghooni, M.; Citi, V.; Brogi, S.; Calderone, V.; Sureda, A.; Lorzadeh, S.; da Silva Rosa, S.C.; Grabarek, B.O.; et al. New Visions on Natural Products and Cancer Therapy: Autophagy and Related Regulatory Pathways. Cancers 2022, 14, 5839. [Google Scholar] [CrossRef]
  12. Jung, S.; Choe, S.; Woo, H.; Jeong, H.; An, H.K.; Moon, H.; Ryu, H.Y.; Yeo, B.K.; Lee, Y.W.; Choi, H.; et al. Autophagic death of neural stem cells mediates chronic stress-induced decline of adult hippocampal neurogenesis and cognitive deficits. Autophagy 2020, 16, 512–530. [Google Scholar] [CrossRef] [Green Version]
  13. Zhang, G.; He, J.; Ye, X.; Zhu, J.; Hu, X.; Shen, M.; Ma, Y.; Mao, Z.; Song, H.; Chen, F. β-Thujaplicin induces autophagic cell death, apoptosis, and cell cycle arrest through ROS-mediated Akt and p38/ERK MAPK signaling in human hepatocellular carcinoma. Cell Death Dis. 2019, 10, 255. [Google Scholar] [CrossRef] [Green Version]
  14. Ruolan, W.; Liangjiao, C.; Longquan, S. The mTOR/ULK1 signaling pathway mediates the autophagy-promoting and osteogenic effects of dicalcium silicate nanoparticles. J. Nanobiotechnol. 2020, 18, 119. [Google Scholar] [CrossRef]
  15. Guevara-Cruz, M.; Godinez-Salas, E.T.; Sanchez-Tapia, M.; Torres-Villalobos, G.; Pichardo-Ontiveros, E.; Guizar-Heredia, R.; Arteaga-Sanchez, L.; Gamba, G.; Mojica-Espinosa, R.; Schcolnik-Cabrera, A.; et al. Genistein stimulates insulin sensitivity through gut microbiota reshaping and skeletal muscle AMPK activation in obese subjects. BMJ Open Diabetes Res. Care 2020, 8, e000948. [Google Scholar] [CrossRef] [Green Version]
  16. Rothschild, J.A.; Islam, H.; Bishop, D.J.; Kilding, A.E.; Stewart, T.; Plews, D.J. Factors Influencing AMPK Activation During Cycling Exercise: A Pooled Analysis and Meta-Regression. Sport. Med. 2022, 52, 1273–1294. [Google Scholar] [CrossRef]
  17. Goldsmith, J.; Levine, B.; Debnath, J. Autophagy and cancer metabolism. Methods Enzymol. 2014, 542, 25–57. [Google Scholar]
  18. Liu, Y.; Shoji-Kawata, S.; Sumpter, R., Jr.; Wei, Y.; Ginet, V.; Zhang, L.; Posner, B.; Tran, K.A.; Green, D.R.; Xavier, R.J.; et al. Autosis is a Na+, K+-ATPase-regulated form of cell death triggered by autophagy-962 inducing peptides, starvation, and hypoxia-ischemia. Proc. Natl. Acad. Sci. USA 2013, 110, 964. [Google Scholar] [CrossRef] [Green Version]
  19. Baghaei, K.; Mazhari, S.; Tokhanbigli, S.; Parsamanesh, G.; Alavifard, H.; Schaafsma, D.; Ghavami, S. Therapeutic potential of targeting regulatory mechanisms of hepatic stellate cell activation in liver fibrosis. Drug Discov. Today 2022, 27, 1044–1061. [Google Scholar] [CrossRef]
  20. Eshraghi, M.; Ahmadi, M.; Afshar, S.; Lorzadeh, S.; Adlimoghaddam, A.; Rezvani Jalal, N.; West, R.; Dastghaib, S.; Igder, S.; Torshizi, S.R.N.; et al. Enhancing autophagy in Alzheimer’s disease through drug repositioning. Pharmacol. Ther. 2022, 237, 108171. [Google Scholar] [CrossRef]
  21. Habibzadeh, P.; Dastsooz, H.; Eshraghi, M.; Los, M.J.; Klionsky, D.J.; Ghavami, S. Autophagy: The Potential Link between SARS-CoV-2 and Cancer. Cancers 2021, 13, 5721. [Google Scholar] [CrossRef] [PubMed]
  22. Scherz-Shouval, R.; Shvets, E.; Fass, E.; Shorer, H.; Gil, L.; Elazar, Z. Reactive oxygen species are essential for autophagy and specifically regulate the activity of Atg4. EMBO J. 2019, 38, e101812. [Google Scholar] [CrossRef] [PubMed]
  23. Barzegar Behrooz, A.; Talaie, Z.; Jusheghani, F.; Los, M.J.; Klonisch, T.; Ghavami, S. Wnt and PI3K/Akt/mTOR Survival Pathways as Therapeutic Targets in Glioblastoma. Int. J. Mol. Sci. 2022, 23, 1353. [Google Scholar] [CrossRef]
  24. Sharifzad, F.; Ghavami, S.; Verdi, J.; Mardpour, S.; Mollapour Sisakht, M.; Azizi, Z.; Taghikhani, A.; Los, M.J.; Fakharian, E.; Ebrahimi, M.; et al. Glioblastoma cancer stem cell biology: Potential theranostic targets. Drug Resist. Updates 2019, 42, 35–45. [Google Scholar] [CrossRef]
  25. Alizadeh, J.; Lorzadeh, S.; Ghavami, S. Autophagy and cancer metastasis: A Trojan horse. J. Investig. Med. 2021, 69, 1145–1147. [Google Scholar] [CrossRef] [PubMed]
  26. Brun, P.; Tarricone, E.; Di Stefano, A.; Mattiuzzo, E.; Mehrbod, P.; Ghavami, S.; Leonardi, A. The regulatory activity of autophagy in conjunctival fibroblasts and its possible role in vernal keratoconjunctivitis. J. Allergy Clin. Immunol. 2020, 146, 1210–1213.e1219. [Google Scholar] [CrossRef]
  27. Samiei, E.; Seyfoori, A.; Toyota, B.; Ghavami, S.; Akbari, M. Investigating Programmed Cell Death and Tumor Invasion in a Three-Dimensional (3D) Microfluidic Model of Glioblastoma. Int. J. Mol. Sci. 2020, 21, 3162. [Google Scholar] [CrossRef]
  28. Shojaei, S.; Alizadeh, J.; Thliveris, J.; Koleini, N.; Kardami, E.; Hatch, G.M.; Xu, F.; Hombach-Klonisch, S.; Klonisch, T.; Ghavami, S. Statins: A new approach to combat temozolomide chemoresistance in glioblastoma. J. Investig. Med. 2018, 66, 1083. [Google Scholar] [CrossRef]
  29. Shojaei, S.; Koleini, N.; Samiei, E.; Aghaei, M.; Cole, L.K.; Alizadeh, J.; Islam, M.I.; Vosoughi, A.R.; Albokashy, M.; Butterfield, Y.; et al. Simvastatin increases temozolomide-induced cell death by targeting the fusion of autophagosomes and lysosomes. FEBS J. 2020, 287, 1005–1034. [Google Scholar] [CrossRef]
  30. Kawalec, P.; Martens, M.D.; Field, J.T.; Mughal, W.; Caymo, A.M.; Chapman, D.; Xiang, B.; Ghavami, S.; Dolinsky, V.W.; Gordon, J.W. Differential impact of doxorubicin dose on cell death and autophagy pathways during acute cardiotoxicity. Toxicol. Appl. Pharmacol. 2022, 453, 116210. [Google Scholar] [CrossRef]
  31. Dastghaib, S.; Shojaei, S.; Mostafavi-Pour, Z.; Sharma, P.; Patterson, J.B.; Samali, A.; Mokarram, P.; Ghavami, S. Simvastatin Induces Unfolded Protein Response and Enhances Temozolomide-Induced Cell Death in Glioblastoma Cells. Cells 2020, 9, 2339. [Google Scholar] [CrossRef]
  32. Stefanek, E.; Samiei, E.; Kavoosi, M.; Esmaeillou, M.; Roustai Geraylow, K.; Emami, A.; Ashrafizadeh, M.; Perrin, D.; Gordon, J.W.; Akbari, M.; et al. A bioengineering method for modeling alveolar Rhabdomyosarcoma and assessing chemotherapy responses. MethodsX 2021, 8, 101473. [Google Scholar] [CrossRef] [PubMed]
  33. Moghadam, A.R.; da Silva Rosa, S.C.; Samiei, E.; Alizadeh, J.; Field, J.; Kawalec, P.; Thliveris, J.; Akbari, M.; Ghavami, S.; Gordon, J.W. Autophagy modulates temozolomide-induced cell death in alveolar Rhabdomyosarcoma cells. Cell Death Discov. 2018, 4, 52. [Google Scholar] [CrossRef] [Green Version]
  34. Li, J.; Zhang, Y.; Wang, L.; Li, M.; Yang, J.; Chen, P.; Zhu, J.; Li, X.; Zeng, Z.; Li, G.; et al. FOXA1 prevents nutrients deprivation induced autophagic cell death through inducing loss of imprinting of IGF2 in lung adenocarcinoma. Cell Death Dis. 2022, 13, 711. [Google Scholar] [CrossRef] [PubMed]
  35. Han, S.; Li, X.; Wang, K.; Zhu, D.; Meng, B.; Liu, J.; Liang, X.; Jin, Y.; Liu, X.; Wen, Q.; et al. PURPL represses autophagic cell death to promote cutaneous melanoma by modulating ULK1 phosphorylation. Cell Death Dis. 2021, 12, 1070. [Google Scholar] [CrossRef] [PubMed]
  36. Long, J.S.; Kania, E.; McEwan, D.G.; Barthet, V.J.A.; Brucoli, M.; Ladds, M.; Nossing, C.; Ryan, K.M. ATG7 is a haploinsufficient repressor of tumor progression and promoter of metastasis. Proc. Natl. Acad. Sci. USA 2022, 119, e2113465119. [Google Scholar] [CrossRef] [PubMed]
  37. Bermudez, O.; Marchetti, S.; Pages, G.; Gimond, C. Post-translational regulation of the ERK phosphatase DUSP6/MKP3 by the mTOR pathway. Oncogene 2008, 27, 3685–3691. [Google Scholar] [CrossRef] [Green Version]
  38. Alizadeh, J.; Glogowska, A.; Thliveris, J.; Kalantari, F.; Shojaei, S.; Hombach-Klonisch, S.; Klonisch, T.; Ghavami, S. Autophagy modulates transforming growth factor beta 1 induced epithelial to mesenchymal transition in non-small cell lung cancer cells. Biochim. Et Biophys. Acta (BBA) Mol. Cell Res. 2018, 1865, 749–768. [Google Scholar] [CrossRef]
  39. Alizadeh, J.; Shojaei, S.; Sepanjnia, A.; Hashemi, M.; Eftekharpour, E.; Ghavami, S. Simultaneous Detection of Autophagy and Epithelial to Mesenchymal Transition in the Non-small Cell Lung Cancer Cells. Methods Mol. Biol. 2019, 1854, 87–103. [Google Scholar] [CrossRef]
  40. Gundamaraju, R.; Lu, W.; Paul, M.K.; Jha, N.K.; Gupta, P.K.; Ojha, S.; Chattopadhyay, I.; Rao, P.V.; Ghavami, S. Autophagy and EMT in cancer and metastasis: Who controls whom? Biochim. Biophys. Acta Mol. Basis Dis. 2022, 1868, 166431. [Google Scholar] [CrossRef]
  41. Fell, D.; Cornish-Bowden, A. Understanding the Control of Metabolism; Portland Press: London, UK, 1997. [Google Scholar]
  42. Newsholme, E.A.; Start, C. Regulation in Metabolism; John Wiley and Sons: Hoboken, NJ, USA, 1973. [Google Scholar]
  43. Morrison, A.J. Cancer cell metabolism connects epigenetic modifications to transcriptional regulation. FEBS J. 2022, 289, 1302–1314. [Google Scholar] [CrossRef]
  44. Wang, X.; An, P.; Gu, Z.; Luo, Y.; Luo, J. Mitochondrial Metal Ion Transport in Cell Metabolism and Disease. Int. J. Mol. Sci. 2021, 22, 7525. [Google Scholar] [CrossRef] [PubMed]
  45. Sekar, M.; Thirumurugan, K. Autophagy: A molecular switch to regulate adipogenesis and lipolysis. Mol. Cell. Biochem. 2022, 477, 727–742. [Google Scholar] [CrossRef]
  46. Watanabe, S.; Tsujino, S. Applications of Medium-Chain Triglycerides in Foods. Front. Nutr. 2022, 9, 802805. [Google Scholar] [CrossRef] [PubMed]
  47. Vance, J.E.; Vance, D.E. Biochemistry of Lipids, Lipoproteins and Membranes; Elsevier: Amsterdam, The Netherlands, 2008. [Google Scholar]
  48. Solinas, G.; Borén, J.; Dulloo, A.G. De novo lipogenesis in metabolic homeostasis: More friend than foe? Mol. Metab. 2015, 4, 367–377. [Google Scholar] [CrossRef] [PubMed]
  49. Ameer, F.; Scandiuzzi, L.; Hasnain, S.; Kalbacher, H.; Zaidi, N. De novo lipogenesis in health and disease. Metabolism 2014, 63, 895–902. [Google Scholar] [CrossRef]
  50. Wang, Y.; Yu, W.; Li, S.; Guo, D.; He, J.; Wang, Y. Acetyl-CoA Carboxylases and Diseases. Front. Oncol. 2022, 12, 836058. [Google Scholar] [CrossRef]
  51. Berg, J.; Tymoczko, J.; Stryer, L.; Gatto, G. Chapter 17: The citric acid cycle. In Biochemistry; WH Freeman and Company: New York, NY, USA, 2002. [Google Scholar]
  52. Cases, S.; Smith, S.J.; Zheng, Y.-W.; Myers, H.M.; Lear, S.R.; Sande, E.; Novak, S.; Collins, C.; Welch, C.B.; Lusis, A.J. Identification of a gene encoding an acyl CoA: Diacylglycerol acyltransferase, a key enzyme in triacylglycerol synthesis. Proc. Natl. Acad. Sci. USA 1998, 95, 13018–13023. [Google Scholar] [CrossRef] [Green Version]
  53. Yen, C.-L.E.; Stone, S.J.; Koliwad, S.; Harris, C.; Farese, R.V., Jr. Thematic review series: Glycerolipids. DGAT enzymes and triacylglycerol biosynthesis. J. Lipid Res. 2008, 49, 2283–2301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Fillmore, N.; Alrob, O.A.; Lopaschuk, G.D. Fatty Acid beta-Oxidation; AOCS Lipid library: Urbana, IL, USA, 2011. [Google Scholar]
  55. Calder, P.C. Omega-3 fatty acids and metabolic partitioning of fatty acids within the liver in the context of nonalcoholic fatty liver disease. Curr. Opin. Clin. Nutr. Metab. Care 2022, 25, 248–255. [Google Scholar] [CrossRef]
  56. Gao, Z.; Chen, X. Fatty Acid beta-Oxidation in Kidney Diseases: Perspectives on Pathophysiological Mechanisms and Therapeutic Opportunities. Front. Pharmacol. 2022, 13, 805281. [Google Scholar] [CrossRef]
  57. Coleman, R.A.; Lee, D.P. Enzymes of triacylglycerol synthesis and their regulation. Prog. Lipid Res. 2004, 43, 134–176. [Google Scholar] [CrossRef]
  58. Aguado, B.; Campbell, R.D. Characterization of a human lysophosphatidic acid acyltransferase that is encoded by a gene located in the class III region of the human major histocompatibility complex. J. Biol. Chem. 1998, 273, 4096–4105. [Google Scholar] [CrossRef] [Green Version]
  59. Shi, Y.; Cheng, D. Beyond triglyceride synthesis: The dynamic functional roles of MGAT and DGAT enzymes in energy metabolism. Am. J. Physiol. Endocrinol. Metab. 2009, 297, E10–E18. [Google Scholar] [CrossRef] [Green Version]
  60. Anne, L. The Heterogeneity of Cancer Metabolism; Springer: Berlin/Heidelberg, Germany, 2021. [Google Scholar]
  61. El Sayed, S.M. Biochemical Origin of the Warburg Effect in Light of 15 Years of Research Experience: A Novel Evidence-Based View (An Expert Opinion Article). OncoTargets Ther. 2023, 16, 143–155. [Google Scholar] [CrossRef]
  62. Vaupel, P.; Schmidberger, H.; Mayer, A. The Warburg effect: Essential part of metabolic reprogramming and central contributor to cancer progression. Int. J. Radiat. Biol. 2019, 95, 912–919. [Google Scholar] [CrossRef] [PubMed]
  63. Gambardella, J.; Fiordelisi, A.; Cerasuolo, F.A.; Buonaiuto, A.; Avvisato, R.; Viti, A.; Sommella, E.; Merciai, F.; Salviati, E.; Campiglia, P.; et al. Experimental evidence and clinical implications of Warburg effect in the skeletal muscle of Fabry disease. iScience 2023, 26, 106074. [Google Scholar] [CrossRef] [PubMed]
  64. Liberti, M.V.; Locasale, J.W. The Warburg Effect: How Does it Benefit Cancer Cells? Trends Biochem. Sci. 2016, 41, 211–218. [Google Scholar] [CrossRef] [Green Version]
  65. San-Millán, I.; Brooks, G.A. Reexamining cancer metabolism: Lactate production for carcinogenesis could be the purpose and explanation of the Warburg Effect. Carcinogenesis 2017, 38, 119–133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Huang, R.; Zong, X. Aberrant cancer metabolism in epithelial-mesenchymal transition and cancer metastasis: Mechanisms in cancer progression. Crit. Rev. Oncol./Hematol. 2017, 115, 13–22. [Google Scholar] [CrossRef]
  67. Sanders, F.W.; Griffin, J.L. De novo lipogenesis in the liver in health and disease: More than just a shunting yard for glucose. Biol. Rev. Camb. Philos. Soc. 2016, 91, 452–468. [Google Scholar] [CrossRef] [Green Version]
  68. Han, H.-S.; Kang, G.; Kim, J.S.; Choi, B.H.; Koo, S.-H. Regulation of glucose metabolism from a liver-centric perspective. Exp. Mol. Med. 2016, 48, e218. [Google Scholar] [CrossRef] [Green Version]
  69. Xia, C.; Li, M.; Ran, G.; Wang, X.; Lu, Z.; Li, T.; Tang, X.; Zhang, Z.; He, Q. Redox-responsive nanoassembly restrained myeloid-derived suppressor cells recruitment through autophagy-involved lactate dehydrogenase A silencing for enhanced cancer immunochemotherapy. J. Control. Release 2021, 335, 557–574. [Google Scholar] [CrossRef] [PubMed]
  70. Gwangwa, M.V.; Joubert, A.M.; Visagie, M.H. Crosstalk between the Warburg effect, redox regulation and autophagy induction in tumourigenesis. Cell. Mol. Biol. Lett. 2018, 23, 20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Huang, T.; Song, X.; Yang, Y.; Wan, X.; Alvarez, A.A.; Sastry, N.; Feng, H.; Hu, B.; Cheng, S.Y. Autophagy and Hallmarks of Cancer. Crit. Rev. Oncog. 2018, 23, 247–267. [Google Scholar] [CrossRef]
  72. Peng, X.; Gong, F.; Chen, Y.; Jiang, Y.; Liu, J.; Yu, M.; Zhang, S.; Wang, M.; Xiao, G.; Liao, H. Autophagy promotes paclitaxel resistance of cervical cancer cells: Involvement of Warburg effect activated hypoxia-induced factor 1-α-mediated signaling. Cell Death Dis. 2014, 5, e1367. [Google Scholar] [CrossRef] [Green Version]
  73. Guido, C.; Whitaker-Menezes, D.; Capparelli, C.; Balliet, R.; Lin, Z.; Pestell, R.G.; Howell, A.; Aquila, S.; Andò, S.; Martinez-Outschoorn, U. Metabolic reprogramming of cancer-associated fibroblasts by TGF-β drives tumor growth: Connecting TGF-β signaling with “Warburg-like” cancer metabolism and L-lactate production. Cell Cycle 2012, 11, 3019–3035. [Google Scholar] [CrossRef] [Green Version]
  74. Ward, P.S.; Thompson, C.B. Metabolic reprogramming: A cancer hallmark even warburg did not anticipate. Cancer Cell 2012, 21, 297–308. [Google Scholar] [CrossRef] [Green Version]
  75. Luo, X.; Cheng, C.; Tan, Z.; Li, N.; Tang, M.; Yang, L.; Cao, Y. Emerging roles of lipid metabolism in cancer metastasis. Mol. Cancer 2017, 16, 76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Yang, M.; Zhang, Y.; Ren, J. Autophagic Regulation of Lipid Homeostasis in Cardiometabolic Syndrome. Front. Cardiovasc. Med. 2018, 5, 38. [Google Scholar] [CrossRef] [Green Version]
  77. Xie, Y.; Li, J.; Kang, R.; Tang, D. Interplay Between Lipid Metabolism and Autophagy. Front. Cell Dev. Biol. 2020, 8, 431. [Google Scholar] [CrossRef] [PubMed]
  78. Jaishy, B.; Abel, E.D. Lipids, lysosomes, and autophagy. J. Lipid Res. 2016, 57, 1619–1635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Dall’Armi, C.; Devereaux, K.A.; Di Paolo, G. The Role of Lipids in the Control of Autophagy. Curr. Biol. 2013, 23, R33–R45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Su, T.; Li, X.; Yang, M.; Shao, Q.; Zhao, Y.; Ma, C.; Wang, P. Autophagy: An Intracellular Degradation Pathway Regulating Plant Survival and Stress Response. Front. Plant Sci. 2020, 11, 164. [Google Scholar] [CrossRef] [Green Version]
  81. Soto-Avellaneda, A.; Morrison, B.E. Signaling and other functions of lipids in autophagy: A review. Lipids Health Dis. 2020, 19, 214. [Google Scholar] [CrossRef] [PubMed]
  82. Zhang, S.; Peng, X.; Yang, S.; Li, X.; Huang, M.; Wei, S.; Liu, J.; He, G.; Zheng, H.; Yang, L.; et al. The regulation, function, and role of lipophagy, a form of selective autophagy, in metabolic disorders. Cell Death Dis. 2022, 13, 132. [Google Scholar] [CrossRef]
  83. Paik, S.; Jo, E.-K. An Interplay Between Autophagy and Immunometabolism for Host Defense Against Mycobacterial Infection. Front. Immunol. 2020, 11, 603951. [Google Scholar] [CrossRef]
  84. Lahiri, V.; Hawkins, W.D.; Klionsky, D.J. Watch What You (Self-) Eat: Autophagic Mechanisms that Modulate Metabolism. Cell Metab. 2019, 29, 803–826. [Google Scholar] [CrossRef] [Green Version]
  85. Zong, W.X.; Rabinowitz, J.D.; White, E. Mitochondria and Cancer. Mol. Cell 2016, 61, 667–676. [Google Scholar] [CrossRef] [Green Version]
  86. Wallace, D.C. Mitochondria and cancer. Nat. Rev. Cancer 2012, 12, 685–698. [Google Scholar] [CrossRef] [Green Version]
  87. Missiroli, S.; Perrone, M.; Genovese, I.; Pinton, P.; Giorgi, C. Cancer metabolism and mitochondria: Finding novel mechanisms to fight tumours. EBioMedicine 2020, 59, 102943. [Google Scholar] [CrossRef]
  88. Shan, W.; Zhou, Y.; Tam, K.Y. The development of small-molecule inhibitors targeting hexokinase 2. Drug Discov. Today 2022, 27, 2574–2585. [Google Scholar] [CrossRef] [PubMed]
  89. Du, D.; Liu, C.; Qin, M.; Zhang, X.; Xi, T.; Yuan, S.; Hao, H.; Xiong, J. Metabolic dysregulation and emerging therapeutical targets for hepatocellular carcinoma. Acta Pharm. Sin. B 2022, 12, 558–580. [Google Scholar] [CrossRef] [PubMed]
  90. Halestrap, A.P.; Pereira, G.C.; Pasdois, P. The role of hexokinase in cardioprotection—Mechanism and potential for translation. Br. J. Pharmacol. 2015, 172, 2085–2100. [Google Scholar] [CrossRef] [Green Version]
  91. DeWaal, D.; Nogueira, V.; Terry, A.R.; Patra, K.C.; Jeon, S.-M.; Guzman, G.; Au, J.; Long, C.P.; Antoniewicz, M.R.; Hay, N. Hexokinase-2 depletion inhibits glycolysis and induces oxidative phosphorylation in hepatocellular carcinoma and sensitizes to metformin. Nat. Commun. 2018, 9, 1–14. [Google Scholar]
  92. Palmieri, D.; Fitzgerald, D.; Shreeve, S.M.; Hua, E.; Bronder, J.L.; Weil, R.J.; Davis, S.; Stark, A.M.; Merino, M.J.; Kurek, R. Analyses of resected human brain metastases of breast cancer reveal the association between up-regulation of hexokinase 2 and poor prognosis. Mol. Cancer Res. 2009, 7, 1438–1445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Wu, J.; Zhang, X.; Wang, Y.; Sun, Q.; Chen, M.; Liu, S.; Zou, X. Licochalcone A suppresses hexokinase 2-mediated tumor glycolysis in gastric cancer via downregulation of the Akt signaling pathway. Oncol. Rep. 2018, 39, 1181–1190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Xu, D.; Jin, J.; Yu, H.; Zhao, Z.; Ma, D.; Zhang, C.; Jiang, H. Chrysin inhibited tumor glycolysis and induced apoptosis in hepatocellular carcinoma by targeting hexokinase-2. J. Exp. Clin. Cancer Res. 2017, 36, 1–11. [Google Scholar] [CrossRef] [Green Version]
  95. Chen, J.; Zhang, S.; Li, Y.; Tang, Z.; Kong, W. Hexokinase 2 overexpression promotes the proliferation and survival of laryngeal squamous cell carcinoma. Tumor Biol. 2014, 35, 3743–3753. [Google Scholar] [CrossRef]
  96. Yoshino, H.; Enokida, H.; Itesako, T.; Kojima, S.; Kinoshita, T.; Tatarano, S.; Chiyomaru, T.; Nakagawa, M.; Seki, N. Tumor-suppressive micro RNA-143/145 cluster targets hexokinase-2 in renal cell carcinoma. Cancer Sci. 2013, 104, 1567–1574. [Google Scholar] [CrossRef]
  97. Arundhathi, J.R.D.; Mathur, S.R.; Gogia, A.; Deo, S.V.S.; Mohapatra, P.; Prasad, C.P. Metabolic changes in triple negative breast cancer-focus on aerobic glycolysis. Mol. Biol. Rep. 2021, 48, 4733–4745. [Google Scholar] [CrossRef] [PubMed]
  98. Mathupala, S.; Ko, Y.A.; Pedersen, P. Hexokinase II: Cancer’s double-edged sword acting as both facilitator and gatekeeper of malignancy when bound to mitochondria. Oncogene 2006, 25, 4777–4786. [Google Scholar] [CrossRef] [Green Version]
  99. Fan, K.; Fan, Z.; Cheng, H.; Huang, Q.; Yang, C.; Jin, K.; Luo, G.; Yu, X.; Liu, C. Hexokinase 2 dimerization and interaction with voltage-dependent anion channel promoted resistance to cell apoptosis induced by gemcitabine in pancreatic cancer. Cancer Med. 2019, 8, 5903–5915. [Google Scholar] [CrossRef] [PubMed]
  100. Gu, J.J.; Singh, A.; Xue, K.; Mavis, C.; Barth, M.; Yanamadala, V.; Lenz, P.; Grau, M.; Lenz, G.; Czuczman, M.S. Up-regulation of hexokinase II contributes to rituximab-chemotherapy resistance and is a clinically relevant target for therapeutic development. Oncotarget 2018, 9, 4020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Suh, D.H.; Kim, M.A.; Kim, H.; Kim, M.-K.; Kim, H.S.; Chung, H.H.; Kim, Y.-B.; Song, Y.S. Association of overexpression of hexokinase II with chemoresistance in epithelial ovarian cancer. Clin. Exp. Med. 2014, 14, 345–353. [Google Scholar] [CrossRef]
  102. Varghese, E.; Samuel, S.M.; Liskova, A.; Samec, M.; Kubatka, P.; Busselberg, D. Targeting Glucose Metabolism to Overcome Resistance to Anticancer Chemotherapy in Breast Cancer. Cancers 2020, 12, 2252. [Google Scholar] [CrossRef]
  103. Patra, K.C.; Wang, Q.; Bhaskar, P.T.; Miller, L.; Wang, Z.; Wheaton, W.; Chandel, N.; Laakso, M.; Muller, W.J.; Allen, E.L. Hexokinase 2 is required for tumor initiation and maintenance and its systemic deletion is therapeutic in mouse models of cancer. Cancer Cell 2013, 24, 213–228. [Google Scholar] [CrossRef] [Green Version]
  104. Gregersen, L.H.; Jacobsen, A.; Frankel, L.B.; Wen, J.; Krogh, A.; Lund, A.H. MicroRNA-143 down-regulates Hexokinase 2 in colon cancer cells. BMC Cancer 2012, 12, 1–10. [Google Scholar] [CrossRef] [Green Version]
  105. Jiang, S.; Zhang, L.F.; Zhang, H.W.; Hu, S.; Lu, M.H.; Liang, S.; Li, B.; Li, Y.; Li, D.; Wang, E.D. A novel miR-155/miR-143 cascade controls glycolysis by regulating hexokinase 2 in breast cancer cells. EMBO J. 2012, 31, 1985–1998. [Google Scholar] [CrossRef] [Green Version]
  106. Kojima, S.; Enokida, H.; Yoshino, H.; Itesako, T.; Chiyomaru, T.; Kinoshita, T.; Fuse, M.; Nishikawa, R.; Goto, Y.; Naya, Y. The tumor-suppressive microRNA-143/145 cluster inhibits cell migration and invasion by targeting GOLM1 in prostate cancer. J. Hum. Genet. 2014, 59, 78–87. [Google Scholar] [CrossRef]
  107. Tan, V.P.; Miyamoto, S. HK2/hexokinase-II integrates glycolysis and autophagy to confer cellular protection. Autophagy 2015, 11, 963–964. [Google Scholar] [CrossRef] [Green Version]
  108. Pilkis, S.J.; Claus, T.H.; Kurland, I.J.; Lange, A.J. 6-Phosphofructo-2-kinase/fructose-2,6-bisphosphatase: A metabolic signaling enzyme. Annu. Rev. Biochem. 1995, 64, 799–835. [Google Scholar] [CrossRef]
  109. El-Maghrabi, M.R.; Noto, F.; Wu, N.; Manes, N. 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase: Suiting structure to need, in a family of tissue-specific enzymes. Curr. Opin. Clin. Nutr. Metab. Care 2001, 4, 411–418. [Google Scholar] [CrossRef] [PubMed]
  110. Okar, D.A.; Manzano, A.; Navarro-Sabate, A.; Riera, L.; Bartrons, R.; Lange, A.J. PFK-2/FBPase-2: Maker and breaker of the essential biofactor fructose-2,6-bisphosphate. Trends Biochem. Sci. 2001, 26, 30–35. [Google Scholar] [CrossRef]
  111. O’Neal, J.; Clem, A.; Reynolds, L.; Dougherty, S.; Imbert-Fernandez, Y.; Telang, S.; Chesney, J.; Clem, B.F. Inhibition of 6-phosphofructo-2-kinase (PFKFB3) suppresses glucose metabolism and the growth of HER2+ breast cancer. Breast Cancer Res. Treat. 2016, 160, 29–40. [Google Scholar] [CrossRef] [PubMed]
  112. Imbert-Fernandez, Y.; Clem, B.F.; O’Neal, J.; Kerr, D.A.; Spaulding, R.; Lanceta, L.; Clem, A.L.; Telang, S.; Chesney, J. Estradiol stimulates glucose metabolism via 6-phosphofructo-2-kinase (PFKFB3). J. Biol. Chem. 2014, 289, 9440–9448. [Google Scholar] [CrossRef] [Green Version]
  113. Novellasdemunt, L.; Obach, M.; Millán-Ariño, L.; Manzano, A.; Ventura, F.; Rosa, J.L.; Jordan, A.; Navarro-Sabate, À.; Bartrons, R. Progestins activate 6-phosphofructo-2-kinase/fructose-2, 6-bisphosphatase 3 (PFKFB3) in breast cancer cells. Biochem. J. 2012, 442, 345–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Dasgupta, S.; Rajapakshe, K.; Zhu, B.; Nikolai, B.C.; Yi, P.; Putluri, N.; Choi, J.M.; Jung, S.Y.; Coarfa, C.; Westbrook, T.F. Metabolic enzyme PFKFB4 activates transcriptional coactivator SRC-3 to drive breast cancer. Nature 2018, 556, 249–254. [Google Scholar] [CrossRef] [PubMed]
  115. Gao, R.; Liu, Y.; Li, D.; Xun, J.; Zhou, W.; Wang, P.; Liu, C.; Li, X.; Shen, W.; Su, W. PFKFB4 promotes breast cancer metastasis via induction of hyaluronan production in a p38-dependent manner. Cell. Physiol. Biochem. 2018, 50, 2108–2123. [Google Scholar] [CrossRef]
  116. Yao, L.; Wang, L.; Cao, Z.-G.; Hu, X.; Shao, Z.-M. High expression of metabolic enzyme PFKFB4 is associated with poor prognosis of operable breast cancer. Cancer Cell Int. 2019, 19, 1–8. [Google Scholar] [CrossRef] [PubMed]
  117. Gao, R.; Li, D.; Xun, J.; Zhou, W.; Li, J.; Wang, J.; Liu, C.; Li, X.; Shen, W.; Qiao, H. CD44ICD promotes breast cancer stemness via PFKFB4-mediated glucose metabolism. Theranostics 2018, 8, 6248. [Google Scholar] [CrossRef]
  118. Hu, K.Y.; Wang, D.G.; Liu, P.F.; Cao, Y.W.; Wang, Y.H.; Yang, X.C.; Hu, C.X.; Sun, L.J.; Niu, H.T. Targeting of MCT1 and PFKFB3 influences cell proliferation and apoptosis in bladder cancer by altering the tumor microenvironment. Oncol. Rep. 2016, 36, 945–951. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Yun, S.J.; Jo, S.-W.; Ha, Y.-S.; Lee, O.-J.; Kim, W.T.; Kim, Y.-J.; Lee, S.-C.; Kim, W.-J. (Eds.) PFKFB4 as a prognostic marker in non-muscle-invasive bladder cancer. In Urologic Oncology: Seminars and Original Investigations; Elsevier: Amsterdam, The Netherlands, 2012. [Google Scholar]
  120. Bobarykina, A.Y.; Minchenko, D.O.; Opentanova, I.L.; Moenner, M.; Caro, J.; Esumi, H.; Minchenko, O.H. Hypoxic regulation of PFKFB-3 and PFKFB-4 gene expression in gastric and pancreatic cancer cell lines and expression of PFKFB genes in gastric cancers. Acta Biochim. Pol. 2006, 53, 789–799. [Google Scholar] [CrossRef] [Green Version]
  121. Minchenko, O.H.; Tsuchihara, K.; Minchenko, D.O.; Bikfalvi, A.; Esumi, H. Mechanisms of regulation of PFKFB expression in pancreatic and gastric cancer cells. World J. Gastroenterol. 2014, 20, 13705. [Google Scholar] [CrossRef] [PubMed]
  122. Bolaños, J.P. Adapting glycolysis to cancer cell proliferation: The MAPK pathway focuses on PFKFB3. Biochem. J. 2013, 452, e7–e9. [Google Scholar] [CrossRef] [Green Version]
  123. Klarer, A.C.; O’Neal, J.; Imbert-Fernandez, Y.; Clem, A.; Ellis, S.R.; Clark, J.; Clem, B.; Chesney, J.; Telang, S. Inhibition of 6-phosphofructo-2-kinase (PFKFB3) induces autophagy as a survival mechanism. Cancer Metab. 2014, 2, 1–14. [Google Scholar] [CrossRef] [Green Version]
  124. Strohecker, A.M.; Joshi, S.; Possemato, R.; Abraham, R.T.; Sabatini, D.M.; White, E. Identification of 6-phosphofructo-2-kinase/fructose-2, 6-bisphosphatase as a novel autophagy regulator by high content shRNA screening. Oncogene 2015, 34, 5662–5676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Yang, Z.; Goronzy, J.J.; Weyand, C.M. The glycolytic enzyme PFKFB3/phosphofructokinase regulates autophagy. Autophagy 2014, 10, 382–383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Ripple, M.O.; Wilding, G. Alteration of glyceraldehyde-3-phosphate dehydrogenase activity and messenger RNA content by androgen in human prostate carcinoma cells. Cancer Res. 1995, 55, 4234–4236. [Google Scholar]
  127. Marcucci, F.; Rumio, C. Tumor Cell Glycolysis-At the Crossroad of Epithelial-Mesenchymal Transition and Autophagy. Cells 2022, 11, 1041. [Google Scholar] [CrossRef]
  128. Butera, G.; Mullappilly, N.; Masetto, F.; Palmieri, M.; Scupoli, M.T.; Pacchiana, R.; Donadelli, M. Regulation of Autophagy by Nuclear GAPDH and Its Aggregates in Cancer and Neurodegenerative Disorders. Int. J. Mol. Sci. 2019, 20, 2062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Chang, C.; Su, H.; Zhang, D.; Wang, Y.; Shen, Q.; Liu, B.; Huang, R.; Zhou, T.; Peng, C.; Wong, C.C. AMPK-dependent phosphorylation of GAPDH triggers Sirt1 activation and is necessary for autophagy upon glucose starvation. Mol. Cell 2015, 60, 930–940. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Tokunaga, K.; Nakamura, Y.; Sakata, K.; Fujimori, K.; Ohkubo, M.; Sawada, K.; Sakiyama, S. Enhanced expression of a glyceraldehyde-3-phosphate dehydrogenase gene in human lung cancers. Cancer Res. 1987, 47, 5616–5619. [Google Scholar] [PubMed]
  131. Phadke, M.; Krynetskaia, N.; Mishra, A.; Krynetskiy, E. Accelerated cellular senescence phenotype of GAPDH-depleted human lung carcinoma cells. Biochem. Biophys. Res. Commun. 2011, 411, 409–415. [Google Scholar] [CrossRef] [Green Version]
  132. Huang, Q.; Lan, F.; Zheng, Z.; Xie, F.; Han, J.; Dong, L.; Xie, Y.; Zheng, F. Akt2 kinase suppresses glyceraldehyde-3-phosphate dehydrogenase (GAPDH)-mediated apoptosis in ovarian cancer cells via phosphorylating GAPDH at threonine 237 and decreasing its nuclear translocation. J. Biol. Chem. 2011, 286, 42211–42220. [Google Scholar] [CrossRef] [Green Version]
  133. Harada, N.; Yasunaga, R.; Higashimura, Y.; Yamaji, R.; Fujimoto, K.; Moss, J.; Inui, H.; Nakano, Y. Glyceraldehyde-3-phosphate dehydrogenase enhances transcriptional activity of androgen receptor in prostate cancer cells. J. Biol. Chem. 2007, 282, 22651–22661. [Google Scholar] [CrossRef] [Green Version]
  134. Correa, C.; Bertollo, C.; Zouain, C.; Goes, A. Glyceraldehyde-3-phosphate dehydrogenase as a surface associated antigen on human breast cancer cell lines MACL-1 and MGSO-3. Oncol. Rep. 2010, 24, 677–685. [Google Scholar]
  135. Pacchiana, R.; Mullappilly, N.; Pinto, A.; Bova, S.; Forciniti, S.; Cullia, G.; Dalla Pozza, E.; Bottani, E.; Decimo, I.; Dando, I.; et al. 3-Bromo-Isoxazoline Derivatives Inhibit GAPDH Enzyme in PDAC Cells Triggering Autophagy and Apoptotic Cell Death. Cancers 2022, 14, 3153. [Google Scholar] [CrossRef]
  136. Dando, I.; Pacchiana, R.; Dalla Pozza, E.; Cataldo, I.; Bruno, S.; Conti, P.; Cordani, M.; Grimaldi, A.; Butera, G.; Caraglia, M. UCP2 inhibition induces ROS/Akt/mTOR axis: Role of GAPDH nuclear translocation in genipin/everolimus anticancer synergism. Free Radic. Biol. Med. 2017, 113, 176–189. [Google Scholar] [CrossRef]
  137. Wang, S.; Jiang, B.; Zhang, T.; Liu, L.; Wang, Y.; Wang, Y.; Chen, X.; Lin, H.; Zhou, L.; Xia, Y. Insulin and mTOR pathway regulate HDAC3-mediated deacetylation and activation of PGK1. PLoS Biol. 2015, 13, e1002243. [Google Scholar]
  138. Rojas-Pirela, M.; Andrade-Alviarez, D.; Rojas, V.; Kemmerling, U.; Caceres, A.J.; Michels, P.A.; Concepcion, J.L.; Quinones, W. Phosphoglycerate kinase: Structural aspects and functions, with special emphasis on the enzyme from Kinetoplastea. Open Biol. 2020, 10, 200302. [Google Scholar] [CrossRef] [PubMed]
  139. Qian, X.; Li, X.; Cai, Q.; Zhang, C.; Yu, Q.; Jiang, Y.; Lee, J.-H.; Hawke, D.; Wang, Y.; Xia, Y. Phosphoglycerate kinase 1 phosphorylates BECLIN1 to induce autophagy. Mol. Cell 2017, 65, 917–931.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Zhang, T.; Wang, Y.; Yu, H.; Zhang, T.; Guo, L.; Xu, J.; Wei, X.; Wang, N.; Wu, Y.; Wang, X.; et al. PGK1 represses autophagy-mediated cell death to promote the proliferation of liver cancer cells by phosphorylating PRAS40. Cell Death Dis. 2022, 13, 68. [Google Scholar] [CrossRef] [PubMed]
  141. Altenberg, B.A.; Greulich, K. Genes of glycolysis are ubiquitously overexpressed in 24 cancer classes. Genomics 2004, 84, 1014–1020. [Google Scholar] [CrossRef] [PubMed]
  142. Hardt, P.; Mazurek, S.; Toepler, M.; Schlierbach, P.; Bretzel, R.; Eigenbrodt, E.; Kloer, H. Faecal tumour M2 pyruvate kinase: A new, sensitive screening tool for colorectal cancer. Br. J. Cancer 2004, 91, 980–984. [Google Scholar] [CrossRef] [Green Version]
  143. Verma, H.; Cholia, R.P.; Kaur, S.; Dhiman, M.; Mantha, A.K. A short review on cross-link between pyruvate kinase (PKM2) and Glioblastoma Multiforme. Metab. Brain Dis. 2021, 36, 751–765. [Google Scholar] [CrossRef]
  144. Dey, P.; Kundu, A.; Sachan, R.; Park, J.; Ahn, M.Y.; Yoon, K.; Lee, J.; Kim, N.D.; Kim, I.S.; Lee, B.M. PKM2 Knockdown Induces Autophagic Cell Death via the AKT/mTOR Pathway in Human Prostate Cancer Cells. Preprints 2019. [Google Scholar] [CrossRef] [Green Version]
  145. Lin, M. Pyruvate Kinase Isoform M2 Influences Autophagy and Related Processes in Hepatocellular Carcinoma Cells. 2018. Available online: https://opencommons.uconn.edu/usp_projects/42/ (accessed on 1 December 2022).
  146. Tamada, M.; Suematsu, M.; Saya, H. Pyruvate kinase M2: Multiple faces for conferring benefits on cancer cells. Clin. Cancer Res. 2012, 18, 5554–5561. [Google Scholar] [CrossRef] [Green Version]
  147. Christofk, H.R.; Vander Heiden, M.G.; Harris, M.H.; Ramanathan, A.; Gerszten, R.E.; Wei, R.; Fleming, M.D.; Schreiber, S.L.; Cantley, L.C. The M2 splice isoform of pyruvate kinase is important for cancer metabolism and tumour growth. Nature 2008, 452, 230–233. [Google Scholar] [CrossRef]
  148. Kumar, Y.; Tapuria, N.; Kirmani, N.; Davidson, B.R. Tumour M2-pyruvate kinase: A gastrointestinal cancer marker. Eur. J. Gastroenterol. Hepatol. 2007, 19, 265–276. [Google Scholar] [CrossRef]
  149. Lv, L.; Li, D.; Zhao, D.; Lin, R.; Chu, Y.; Zhang, H.; Zha, Z.; Liu, Y.; Li, Z.; Xu, Y. Acetylation targets the M2 isoform of pyruvate kinase for degradation through chaperone-mediated autophagy and promotes tumor growth. Mol. Cell 2011, 42, 719–730. [Google Scholar] [CrossRef] [Green Version]
  150. Wang, L.; Yang, L.; Yang, Z.; Tang, Y.; Tao, Y.; Zhan, Q.; Lei, L.; Jing, Y.; Jiang, X.; Jin, H. Glycolytic enzyme PKM2 mediates autophagic activation to promote cell survival in NPM1-mutated leukemia. Int. J. Biol. Sci. 2019, 15, 882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Yu, Z.; Wang, D.; Tang, Y. PKM2 promotes cell metastasis and inhibits autophagy via the JAK/STAT3 pathway in hepatocellular carcinoma. Mol. Cell. Biochem. 2021, 476, 2001–2010. [Google Scholar] [CrossRef] [PubMed]
  152. Dawson, D.M.; Goodfriend, T.L.; Kaplan, N.O.; Kaplan, N.O. Lactic Dehydrogenases: Functions of the Two Types Rates of Synthesis of the Two Major Forms Can Be Correlated with Metabolic Differentiation. Science 1964, 143, 929–933. [Google Scholar] [CrossRef]
  153. Valvona, C.J.; Fillmore, H.L.; Nunn, P.B.; Pilkington, G.J. The regulation and function of lactate dehydrogenase A: Therapeutic potential in brain tumor. Brain Pathol. 2016, 26, 3–17. [Google Scholar] [CrossRef] [Green Version]
  154. Gordon, J.S.; Wood, C.T.; Luc, J.G.Y.; Watson, R.A.; Maynes, E.J.; Choi, J.H.; Morris, R.J.; Massey, H.T.; Throckmorton, A.L.; Tchantchaleishvili, V. Clinical implications of LDH isoenzymes in hemolysis and continuous-flow left ventricular assist device-induced thrombosis. Artif. Organs 2020, 44, 231–238. [Google Scholar] [CrossRef]
  155. Sheng, S.L.; Liu, J.J.; Dai, Y.H.; Sun, X.G.; Xiong, X.P.; Huang, G. Knockdown of lactate dehydrogenase A suppresses tumor growth and metastasis of human hepatocellular carcinoma. FEBS J. 2012, 279, 3898–3910. [Google Scholar] [CrossRef] [PubMed]
  156. Zhou, S.; Min, Z.; Sun, K.; Qu, S.; Zhou, J.; Duan, H.; Liu, H.; Liu, X.; Gong, Z.; Li, D. miR-199a-3p/Sp1/LDHA axis controls aerobic glycolysis in testicular tumor cells. Int. J. Mol. Med. 2018, 42, 2163–2174. [Google Scholar] [CrossRef] [Green Version]
  157. Guddeti, R.K.; Bali, P.; Karyala, P.; Pakala, S.B. MTA1 coregulator regulates LDHA expression and function in breast cancer. Biochem. Biophys. Res. Commun. 2019, 520, 54–59. [Google Scholar] [CrossRef]
  158. Huang, X.; Li, X.; Xie, X.; Ye, F.; Chen, B.; Song, C.; Tang, H.; Xie, X. High expressions of LDHA and AMPK as prognostic biomarkers for breast cancer. Breast 2016, 30, 39–46. [Google Scholar] [CrossRef]
  159. Yang, Y.; Su, D.; Zhao, L.; Zhang, D.; Xu, J.; Wan, J.; Fan, S.; Chen, M. Different effects of LDH-A inhibition by oxamate in non-small cell lung cancer cells. Oncotarget 2014, 5, 11886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Das, C.K.; Parekh, A.; Parida, P.K.; Bhutia, S.K.; Mandal, M. Lactate dehydrogenase A regulates autophagy and tamoxifen resistance in breast cancer. Biochim. Et Biophys. Acta (BBA) Mol. Cell Res. 2019, 1866, 1004–1018. [Google Scholar] [CrossRef] [PubMed]
  161. Brisson, L.; Bański, P.; Sboarina, M.; Dethier, C.; Danhier, P.; Fontenille, M.-J.; Van Hée, V.F.; Vazeille, T.; Tardy, M.; Falces, J. Lactate dehydrogenase B controls lysosome activity and autophagy in cancer. Cancer Cell 2016, 30, 418–431. [Google Scholar] [CrossRef] [Green Version]
  162. Pilon-Thomas, S.; Kodumudi, K.N.; El-Kenawi, A.E.; Russell, S.; Weber, A.M.; Luddy, K.; Damaghi, M.; Wojtkowiak, J.W.; Mulé, J.J.; Ibrahim-Hashim, A. Neutralization of tumor acidity improves antitumor responses to immunotherapy. Cancer Res. 2016, 76, 1381–1390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. de la Cruz-López, K.G.; Castro-Muñoz, L.J.; Reyes-Hernández, D.O.; García-Carrancá, A.; Manzo-Merino, J. Lactate in the regulation of tumor microenvironment and therapeutic approaches. Front. Oncol. 2019, 9, 1143. [Google Scholar] [CrossRef] [Green Version]
  164. Rizwan, A.; Serganova, I.; Khanin, R.; Karabeber, H.; Ni, X.; Thakur, S.; Zakian, K.L.; Blasberg, R.; Koutcher, J.A. Relationships between LDH-A, lactate, and metastases in 4T1 breast tumors. Clin. Cancer Res. 2013, 19, 5158–5169. [Google Scholar] [CrossRef] [Green Version]
  165. Chen, P.; Zuo, H.; Xiong, H.; Kolar, M.J.; Chu, Q.; Saghatelian, A.; Siegwart, D.J.; Wan, Y. Gpr132 sensing of lactate mediates tumor–macrophage interplay to promote breast cancer metastasis. Proc. Natl. Acad. Sci. USA 2017, 114, 580–585. [Google Scholar] [CrossRef] [Green Version]
  166. Végran, F.; Boidot, R.; Michiels, C.; Sonveaux, P.; Feron, O. Lactate influx through the endothelial cell monocarboxylate transporter MCT1 supports an NF-κB/IL-8 pathway that drives tumor angiogenesis. Cancer Res. 2011, 71, 2550–2560. [Google Scholar] [CrossRef] [Green Version]
  167. Sonveaux, P.; Copetti, T.; De Saedeleer, C.J.; Végran, F.; Verrax, J.; Kennedy, K.M.; Moon, E.J.; Dhup, S.; Danhier, P.; Frérart, F. Targeting the lactate transporter MCT1 in endothelial cells inhibits lactate-induced HIF-1 activation and tumor angiogenesis. PLoS ONE 2012, 7, e33418. [Google Scholar] [CrossRef] [Green Version]
  168. Lu, H.; Dalgard, C.L.; Mohyeldin, A.; McFate, T.; Tait, A.S.; Verma, A. Reversible inactivation of HIF-1 prolyl hydroxylases allows cell metabolism to control basal HIF-1. J. Biol. Chem. 2005, 280, 41928–41939. [Google Scholar] [CrossRef] [Green Version]
  169. Martínez-Zaguilán, R.; Seftor, E.A.; Seftor, R.E.; Chu, Y.-W.; Gillies, R.J.; Hendrix, M.J. Acidic pH enhances the invasive behavior of human melanoma cells. Clin. Exp. Metastasis 1996, 14, 176–186. [Google Scholar] [CrossRef]
  170. Calcinotto, A.; Filipazzi, P.; Grioni, M.; Iero, M.; De Milito, A.; Ricupito, A.; Cova, A.; Canese, R.; Jachetti, E.; Rossetti, M. Modulation of microenvironment acidity reverses anergy in human and murine tumor-infiltrating T lymphocytes. Cancer Res. 2012, 72, 2746–2756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Jones, R.; Morris, M. Monocarboxylate transporters: Therapeutic targets and prognostic factors in disease. Clin. Pharmacol. Ther. 2016, 100, 454–463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Garcia, C.K.; Goldstein, J.L.; Pathak, R.K.; Anderson, R.G.; Brown, M.S. Molecular characterization of a membrane transporter for lactate, pyruvate, and other monocarboxylates: Implications for the Cori cycle. Cell 1994, 76, 865–873. [Google Scholar] [CrossRef] [PubMed]
  173. Pinheiro, C.; Albergaria, A.; Paredes, J.; Sousa, B.; Dufloth, R.; Vieira, D.; Schmitt, F.; Baltazar, F. Monocarboxylate transporter 1 is up-regulated in basal-like breast carcinoma. Histopathology 2010, 56, 860–867. [Google Scholar] [CrossRef]
  174. Pinheiro, C.; Longatto-Filho, A.; Ferreira, L.; Pereira, S.M.M.; Etlinger, D.; Moreira, M.A.; Jube, L.F.; Queiroz, G.S.; Schmitt, F.; Baltazar, F. Increasing expression of monocarboxylate transporters 1 and 4 along progression to invasive cervical carcinoma. Int. J. Gynecol. Pathol. 2008, 27, 568–574. [Google Scholar] [CrossRef]
  175. Pinheiro, C.; Penna, V.; Morais-Santos, F.; Abrahão-Machado, L.F.; Ribeiro, G.; Curcelli, E.C.; Olivieri, M.V.; Morini, S.; Valença, I.; Ribeiro, D. Characterization of monocarboxylate transporters (MCTs) expression in soft tissue sarcomas: Distinct prognostic impact of MCT1 sub-cellular localization. J. Transl. Med. 2014, 12, 1–11. [Google Scholar] [CrossRef] [Green Version]
  176. Zhao, Z.; Wu, M.-S.; Zou, C.; Tang, Q.; Lu, J.; Liu, D.; Wu, Y.; Yin, J.; Xie, X.; Shen, J. Downregulation of MCT1 inhibits tumor growth, metastasis and enhances chemotherapeutic efficacy in osteosarcoma through regulation of the NF-κB pathway. Cancer Lett. 2014, 342, 150–158. [Google Scholar] [CrossRef]
  177. Pinheiro, C.; Longatto-Filho, A.; Pereira, S.M.M.; Etlinger, D.; Moreira, M.A.; Jubé, L.F.; Queiroz, G.S.; Schmitt, F.; Baltazar, F. Monocarboxylate transporters 1 and 4 are associated with CD147 in cervical carcinoma. Dis. Mrk. 2009, 26, 97–103. [Google Scholar] [CrossRef]
  178. Jin, P.; Jiang, J.; Xie, N.; Zhou, L.; Huang, Z.; Zhang, L.; Qin, S.; Fu, S.; Peng, L.; Gao, W. MCT1 relieves osimertinib-induced CRC suppression by promoting autophagy through the LKB1/AMPK signaling. Cell Death Dis. 2019, 10, 1–15. [Google Scholar] [CrossRef] [Green Version]
  179. Long, Y.; Gao, Z.; Hu, X.; Xiang, F.; Wu, Z.; Zhang, J.; Han, X.; Yin, L.; Qin, J.; Lan, L. Downregulation of MCT 4 for lactate exchange promotes the cytotoxicity of NK cells in breast carcinoma. Cancer Med. 2018, 7, 4690–4700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Kierans, S.J.; Taylor, C.T. Regulation of glycolysis by the hypoxia-inducible factor (HIF): Implications for cellular physiology. J. Physiol. 2021, 599, 23–37. [Google Scholar] [CrossRef]
  181. Mazure, N.M.; Pouysségur, J. Hypoxia-induced autophagy: Cell death or cell survival? Curr. Opin. Cell Biol. 2010, 22, 177–180. [Google Scholar] [CrossRef] [PubMed]
  182. Noman, M.Z.; Janji, B.; Berchem, G.; Mami-Chouaib, F.; Chouaib, S. Hypoxia-induced autophagy: A new player in cancer immunotherapy? Autophagy 2012, 8, 704–706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Papandreou, I.; Lim, A.; Laderoute, K.; Denko, N. Hypoxia signals autophagy in tumor cells via AMPK activity, independent of HIF-1, BNIP3, and BNIP3L. Cell Death Differ. 2008, 15, 1572–1581. [Google Scholar] [CrossRef]
  184. de Mello, N.P.; Orellana, A.M.; Mazucanti, C.H.; de Morais Lima, G.; Scavone, C.; Kawamoto, E.M. Insulin and autophagy in neurodegeneration. Front. Neurosci. 2019, 13, 491. [Google Scholar] [CrossRef] [Green Version]
  185. Shen, Y.; Liu, Y.; Sun, T.; Yang, W. LincRNA-p21 knockdown enhances radiosensitivity of hypoxic tumor cells by reducing autophagy through HIF-1/Akt/mTOR/P70S6K pathway. Exp. Cell Res. 2017, 358, 188–198. [Google Scholar] [CrossRef]
  186. Talks, K.L.; Turley, H.; Gatter, K.C.; Maxwell, P.H.; Pugh, C.W.; Ratcliffe, P.J.; Harris, A.L. The expression and distribution of the hypoxia-inducible factors HIF-1α and HIF-2α in normal human tissues, cancers, and tumor-associated macrophages. Am. J. Pathol. 2000, 157, 411–421. [Google Scholar] [CrossRef]
  187. Fu, D.; He, C.; Wei, J.; Zhang, Z.; Luo, Y.; Tan, H.; Ren, C. PGK1 is a Potential Survival Biomarker and Invasion Promoter by Regulating the HIF-1α–Mediated Epithelial-Mesenchymal Transition Process in Breast Cancer. Cell. Physiol. Biochem. 2018, 51, 2434–2444. [Google Scholar] [CrossRef] [PubMed]
  188. Sugiyama, T.; Taniguchi, K.; Matsuhashi, N.; Tajirika, T.; Futamura, M.; Takai, T.; Akao, Y.; Yoshida, K. MiR-133b inhibits growth of human gastric cancer cells by silencing pyruvate kinase muscle-splicer polypyrimidine tract-binding protein 1. Cancer Sci. 2016, 107, 1767–1775. [Google Scholar] [CrossRef] [Green Version]
  189. Shi, L.; Yan, H.; An, S.; Shen, M.; Jia, W.; Zhang, R.; Zhao, L.; Huang, G.; Liu, J. SIRT 5-mediated deacetylation of LDHB promotes autophagy and tumorigenesis in colorectal cancer. Mol. Oncol. 2019, 13, 358–375. [Google Scholar] [CrossRef]
  190. Wang, C.W. Lipid droplets, lipophagy, and beyond. Biochim. Et Biophys. Acta 2016, 1861 (8 Pt B), 793–805. [Google Scholar] [CrossRef]
  191. Efeyan, A.; Comb, W.C.; Sabatini, D.M. Nutrient-sensing mechanisms and pathways. Nature 2015, 517, 302–310. [Google Scholar] [CrossRef] [Green Version]
  192. Ward, C.; Martinez-Lopez, N.; Otten, E.G.; Carroll, B.; Maetzel, D.; Singh, R.; Sarkar, S.; Korolchuk, V.I. Autophagy, lipophagy and lysosomal lipid storage disorders. Biochim. Et Biophys. Acta 2016, 1861, 269–284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Schulze, R.J.; Sathyanarayan, A.; Mashek, D.G. Breaking fat: The regulation and mechanisms of lipophagy. Biochim. Biophys. Acta. Mol. Cell Biol. Lipids 2017, 1862 (10 Pt B), 1178–1187. [Google Scholar] [CrossRef]
  194. Martinez-Lopez, N.; Singh, R. Autophagy and Lipid Droplets in the Liver. Annu. Rev. Nutr. 2015, 35, 215–237. [Google Scholar] [CrossRef] [PubMed]
  195. Fafián-Labora, J.; Carpintero-Fernández, P.; Jordan, S.J.D.; Shikh-Bahaei, T.; Abdullah, S.M.; Mahenthiran, M.; Rodríguez-Navarro, J.A.; Niklison-Chirou, M.V.; O’Loghlen, A. FASN activity is important for the initial stages of the induction of senescence. Cell Death Dis. 2019, 10, 318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Humbert, M.; Seiler, K.; Mosimann, S.; Rentsch, V.; McKenna, S.L.; Tschan, M.P. Autophagy-mediated degradation of Fatty Acid Synthase (FASN) facilitates ATRA-induced granulocytic differentiation of acute myeloid leukemia (AML) cells. bioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  197. Palmieri, M.; Impey, S.; Kang, H.; di Ronza, A.; Pelz, C.; Sardiello, M.; Ballabio, A. Characterization of the CLEAR network reveals an integrated control of cellular clearance pathways. Hum. Mol. Genet. 2011, 20, 3852–3866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Wang, H.; Luo, Q.F.; Peng, A.F.; Long, X.H.; Wang, T.F.; Liu, Z.L.; Zhang, G.M.; Zhou, R.P.; Gao, S.; Zhou, Y.; et al. Positive feedback regulation between Akt phosphorylation and fatty acid synthase expression in osteosarcoma. Int. J. Mol. Med. 2014, 33, 633–639. [Google Scholar] [CrossRef] [Green Version]
  199. Zaytseva, Y.Y.; Harris, J.W.; Mitov, M.I.; Kim, J.T.; Butterfield, D.A.; Lee, E.Y.; Weiss, H.L.; Gao, T.; Evers, B.M. Increased expression of fatty acid synthase provides a survival advantage to colorectal cancer cells via upregulation of cellular respiration. Oncotarget 2015, 6, 18891–18904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Alizadeh, J.; Zeki, A.A.; Mirzaei, N.; Tewary, S.; Rezaei Moghadam, A.; Glogowska, A.; Nagakannan, P.; Eftekharpour, E.; Wiechec, E.; Gordon, J.W.; et al. Mevalonate Cascade Inhibition by Simvastatin Induces the Intrinsic Apoptosis Pathway via Depletion of Isoprenoids in Tumor Cells. Sci. Rep. 2017, 7, 44841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Ahmadi, M.; Amiri, S.; Pecic, S.; Machaj, F.; Rosik, J.; Łos, M.J.; Alizadeh, J.; Mahdian, R.; da Silva Rosa, S.C.; Schaafsma, D.; et al. Pleiotropic effects of statins: A focus on cancer. Biochim. Et Biophys. Acta (BBA) Mol. Basis Dis. 2020, 1866, 165968. [Google Scholar] [CrossRef]
  202. Peymani, P.; Dehesh, T.; Aligolighasemabadi, F.; Sadeghdoust, M.; Kotfis, K.; Ahmadi, M.; Mehrbod, P.; Iranpour, P.; Dastghaib, S.; Nasimian, A.; et al. Statins in patients with COVID-19: A retrospective cohort study in Iranian COVID-19 patients. Transl. Med. Commun. 2021, 6, 3. [Google Scholar] [CrossRef]
  203. Ghavami, S.; Sharma, P.; Yeganeh, B.; Ojo, O.O.; Jha, A.; Mutawe, M.M.; Kashani, H.H.; Los, M.J.; Klonisch, T.; Unruh, H.; et al. Airway mesenchymal cell death by mevalonate cascade inhibition: Integration of autophagy, unfolded protein response and apoptosis focusing on Bcl2 family proteins. Biochim. Et Biophys. Acta 2014, 1843, 1259–1271. [Google Scholar] [CrossRef] [Green Version]
  204. Vilimanovich, U.; Bosnjak, M.; Bogdanovic, A.; Markovic, I.; Isakovic, A.; Kravic-Stevovic, T.; Mircic, A.; Trajkovic, V.; Bumbasirevic, V. Statin-mediated inhibition of cholesterol synthesis induces cytoprotective autophagy in human leukemic cells. Eur. J. Pharmacol. 2015, 765, 415–428. [Google Scholar] [CrossRef]
  205. Guo, J.Y.; Karsli-Uzunbas, G.; Mathew, R.; Aisner, S.C.; Kamphorst, J.J.; Strohecker, A.M.; Chen, G.; Price, S.; Lu, W.; Teng, X.; et al. Autophagy suppresses progression of K-ras-induced lung tumors to oncocytomas and maintains lipid homeostasis. Genes Dev. 2013, 27, 1447–1461. [Google Scholar] [CrossRef] [Green Version]
  206. Wang, H.J.; Park, J.Y.; Kwon, O.; Choe, E.Y.; Kim, C.H.; Hur, K.Y.; Lee, M.S.; Yun, M.; Cha, B.S.; Kim, Y.B.; et al. Chronic HMGCR/HMG-CoA reductase inhibitor treatment contributes to dysglycemia by upregulating hepatic gluconeogenesis through autophagy induction. Autophagy 2015, 11, 2089–2101. [Google Scholar] [CrossRef] [Green Version]
  207. Parikh, A.; Childress, C.; Deitrick, K.; Lin, Q.; Rukstalis, D.; Yang, W. Statin-induced autophagy by inhibition of geranylgeranyl biosynthesis in prostate cancer PC3 cells. Prostate 2010, 70, 971–981. [Google Scholar] [CrossRef]
  208. Araki, M.; Motojima, K. Hydrophobic statins induce autophagy in cultured human rhabdomyosarcoma cells. Biochem. Biophys. Res. Commun. 2008, 367, 462–467. [Google Scholar] [CrossRef] [PubMed]
  209. Araki, M.; Maeda, M.; Motojima, K. Hydrophobic statins induce autophagy and cell death in human rhabdomyosarcoma cells by depleting geranylgeranyl diphosphate. Eur. J. Pharmacol. 2012, 674, 95–103. [Google Scholar] [CrossRef] [PubMed]
  210. Emami, A.; Shojaei, S.; da Silva Rosa, S.C.; Aghaei, M.; Samiei, E.; Vosoughi, A.R.; Kalantari, F.; Kawalec, P.; Thliveris, J.; Sharma, P.; et al. Mechanisms of simvastatin myotoxicity: The role of autophagy flux inhibition. Eur. J. Pharmacol. 2019, 862, 172616. [Google Scholar] [CrossRef] [PubMed]
  211. Hou, H.; Zhang, Y.; Huang, Y.; Yi, Q.; Lv, L.; Zhang, T.; Chen, D.; Hao, Q.; Shi, Q. Inhibitors of phosphatidylinositol 3′-kinases promote mitotic cell death in HeLa cells. PLoS ONE 2012, 7, e35665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Khanzada, U.K.; Pardo, O.E.; Meier, C.; Downward, J.; Seckl, M.J.; Arcaro, A. Potent inhibition of small-cell lung cancer cell growth by simvastatin reveals selective functions of Ras isoforms in growth factor signalling. Oncogene 2006, 25, 877–887. [Google Scholar] [CrossRef] [Green Version]
  213. Calabro, A.; Tai, J.; Allen, S.L.; Budman, D.R. In-vitro synergism of m-TOR inhibitors, statins, and classical chemotherapy: Potential implications in acute leukemia. Anti-Cancer Drugs 2008, 19, 705–712. [Google Scholar] [CrossRef]
  214. Cemeus, C.; Zhao, T.T.; Barrett, G.M.; Lorimer, I.A.; Dimitroulakos, J. Lovastatin enhances gefitinib activity in glioblastoma cells irrespective of EGFRvIII and PTEN status. J. Neuro-Oncol. 2008, 90, 9–17. [Google Scholar] [CrossRef]
  215. Martirosyan, A.; Clendening, J.W.; Goard, C.A.; Penn, L.Z. Lovastatin induces apoptosis of ovarian cancer cells and synergizes with doxorubicin: Potential therapeutic relevance. BMC Cancer 2010, 10, 103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Cannavo, A.; Liccardo, D.; Komici, K.; Corbi, G.; de Lucia, C.; Femminella, G.D.; Elia, A.; Bencivenga, L.; Ferrara, N.; Koch, W.J.; et al. Sphingosine Kinases and Sphingosine 1-Phosphate Receptors: Signaling and Actions in the Cardiovascular System. Front. Pharmacol. 2017, 8, 556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Long, J.; Xie, Y.; Yin, J.; Lu, W.; Fang, S. SphK1 promotes tumor cell migration and invasion in colorectal cancer. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2016, 37, 6831–6836. [Google Scholar] [CrossRef] [PubMed]
  218. Almejún, M.B.; Borge, M.; Colado, A.; Elías, E.E.; Podaza, E.; Risnik, D.; De Brasi, C.D.; Stanganelli, C.; Slavutsky, I.; Cabrejo, M.; et al. Sphingosine kinase 1 participates in the activation, proliferation and survival of chronic lymphocytic leukemia cells. Haematologica 2017, 102, e257–e260. [Google Scholar] [CrossRef] [Green Version]
  219. Pchejetski, D.; Doumerc, N.; Golzio, M.; Naymark, M.; Teissié, J.; Kohama, T.; Waxman, J.; Malavaud, B.; Cuvillier, O. Chemosensitizing effects of sphingosine kinase-1 inhibition in prostate cancer cell and animal models. Mol. Cancer Ther. 2008, 7, 1836. [Google Scholar] [CrossRef] [Green Version]
  220. Kawamori, T.; Kaneshiro, T.; Okumura, M.; Maalouf, S.; Uflacker, A.; Bielawski, J.; Hannun, Y.A.; Obeid, L.M. Role for sphingosine kinase 1 in colon carcinogenesis. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2009, 23, 405–414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Akao, Y.; Banno, Y.; Nakagawa, Y.; Hasegawa, N.; Kim, T.J.; Murate, T.; Igarashi, Y.; Nozawa, Y. High expression of sphingosine kinase 1 and S1P receptors in chemotherapy-resistant prostate cancer PC3 cells and their camptothecin-induced up-regulation. Biochem. Biophys. Res. Commun. 2006, 342, 1284–1290. [Google Scholar] [CrossRef] [PubMed]
  222. Bao, M.; Chen, Z.; Xu, Y.; Zhao, Y.; Zha, R.; Huang, S.; Liu, L.; Chen, T.; Li, J.; Tu, H.; et al. Sphingosine kinase 1 promotes tumour cell migration and invasion via the S1P/EDG1 axis in hepatocellular carcinoma. Liver Int. Off. J. Int. Assoc. Study Liver 2012, 32, 331–338. [Google Scholar] [CrossRef] [PubMed]
  223. Moruno Manchon, J.F.; Uzor, N.E.; Finkbeiner, S.; Tsvetkov, A.S. SPHK1/sphingosine kinase 1-mediated autophagy differs between neurons and SH-SY5Y neuroblastoma cells. Autophagy 2016, 12, 1418–1424. [Google Scholar] [CrossRef] [Green Version]
  224. Lavieu, G.; Scarlatti, F.; Sala, G.; Carpentier, S.; Levade, T.; Ghidoni, R.; Botti, J.; Codogno, P. Regulation of autophagy by sphingosine kinase 1 and its role in cell survival during nutrient starvation. J. Biol. Chem. 2006, 281, 8518–8527. [Google Scholar] [CrossRef] [Green Version]
  225. Xu, C.; Zhang, W.; Liu, S.; Wu, W.; Qin, M.; Huang, J. Activation of the SphK1/ERK/p-ERK pathway promotes autophagy in colon cancer cells. Oncol. Lett. 2018, 15, 9719–9724. [Google Scholar] [CrossRef] [Green Version]
  226. Liu, H.; Ma, Y.; He, H.-W.; Zhao, W.-L.; Shao, R.-G. SPHK1 (sphingosine kinase 1) induces epithelial-mesenchymal transition by promoting the autophagy-linked lysosomal degradation of CDH1/E-cadherin in hepatoma cells. Autophagy 2017, 13, 900–913. [Google Scholar] [CrossRef]
  227. Yin, S.; Miao, Z.; Tan, Y.; Wang, P.; Xu, X.; Zhang, C.; Hou, W.; Huang, J.; Xu, H. SPHK1-induced autophagy in peritoneal mesothelial cell enhances gastric cancer peritoneal dissemination. Cancer Med. 2019, 8, 1731–1743. [Google Scholar] [CrossRef]
  228. Jeppesen, J.; Mogensen, M.; Prats, C.; Sahlin, K.; Madsen, K.; Kiens, B. FAT/CD36 is localized in sarcolemma and in vesicle-like structures in subsarcolemma regions but not in mitochondria. J. Lipid Res. 2010, 51, 1504–1512. [Google Scholar] [CrossRef] [Green Version]
  229. Li, Y.; Yang, P.; Zhao, L.; Chen, Y.; Zhang, X.; Zeng, S.; Wei, L.; Varghese, Z.; Moorhead, J.F.; Chen, Y.; et al. CD36 plays a negative role in the regulation of lipophagy in hepatocytes through an AMPK-dependent pathway. J. Lipid Res. 2019, 60, 844–855. [Google Scholar] [CrossRef] [Green Version]
  230. Su, W.; Frohman, M.A. Chapter 144—Phospholipase D. In Handbook of Cell Signaling, 2nd ed.; Bradshaw, R.A., Dennis, E.A., Eds.; Academic Press: San Diego, CA, USA, 2010; pp. 1167–1176. [Google Scholar] [CrossRef]
  231. Foster, D.A. Phosphatidic acid signaling to mTOR: Signals for the survival of human cancer cells. Biochim. Et Biophys. Acta 2009, 1791, 949–955. [Google Scholar] [CrossRef] [Green Version]
  232. Fang, Y.; Vilella-Bach, M.; Bachmann, R.; Flanigan, A.; Chen, J. Phosphatidic acid-mediated mitogenic activation of mTOR signaling. Science 2001, 294, 1942–1945. [Google Scholar] [CrossRef]
  233. Shahnazari, S.; Yen, W.-L.; Birmingham, C.L.; Shiu, J.; Namolovan, A.; Zheng, Y.T.; Nakayama, K.; Klionsky, D.J.; Brumell, J.H. A Diacylglycerol-Dependent Signaling Pathway Contributes to Regulation of Antibacterial Autophagy. Cell Host Microbe 2010, 8, 137–146. [Google Scholar] [CrossRef] [Green Version]
  234. Genc, G.E.; Hipolito, V.E.B.; Botelho, R.J.; Gumuslu, S. Lysophosphatidic acid represses autophagy in prostate carcinoma cells. Biochem. Cell Biol. 2019, 97, 387–396. [Google Scholar] [CrossRef]
  235. Holland, P.; Knaevelsrud, H.; Soreng, K.; Mathai, B.J.; Lystad, A.H.; Pankiv, S.; Bjorndal, G.T.; Schultz, S.W.; Lobert, V.H.; Chan, R.B.; et al. HS1BP3 negatively regulates autophagy by modulation of phosphatidic acid levels. Nat. Commun. 2016, 7, 13889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Carmen, B.; Thierry, L.; Patrice, C. Regulation of Autophagy by Sphingolipids. Anti-Cancer Agents Med. Chem. 2011, 11, 844–853. [Google Scholar] [CrossRef]
  237. Spassieva, S.D.; Mullen, T.D.; Townsend, D.M.; Obeid, L.M. Disruption of ceramide synthesis by CerS2 down-regulation leads to autophagy and the unfolded protein response. Biochem. J. 2009, 424, 273–283. [Google Scholar] [CrossRef] [Green Version]
  238. Pattingre, S.; Bauvy, C.; Carpentier, S.; Levade, T.; Levine, B.; Codogno, P. Role of JNK1-dependent Bcl-2 phosphorylation in ceramide-induced macroautophagy. J. Biol. Chem. 2009, 284, 2719–2728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Scarlatti, F.; Bauvy, C.; Ventruti, A.; Sala, G.; Cluzeaud, F.; Vandewalle, A.; Ghidoni, R.; Codogno, P. Ceramide-mediated macroautophagy involves inhibition of protein kinase B and up-regulation of BECLIN 1. J. Biol. Chem. 2004, 279, 18384–18391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Guenther, G.G.; Peralta, E.R.; Rosales, K.R.; Wong, S.Y.; Siskind, L.J.; Edinger, A.L. Ceramide starves cells to death by downregulating nutrient transporter proteins. Proc. Natl. Acad. Sci. USA 2008, 105, 17402–17407. [Google Scholar] [CrossRef] [Green Version]
  241. Lépine, S.; Allegood, J.C.; Park, M.; Dent, P.; Milstien, S.; Spiegel, S. Sphingosine-1-phosphate phosphohydrolase-1 regulates ER stress-induced autophagy. Cell Death Differ. 2011, 18, 350–361. [Google Scholar] [CrossRef] [PubMed]
  242. Chang, C.L.; Ho, M.C.; Lee, P.H.; Hsu, C.Y.; Huang, W.P.; Lee, H. S1P(5) is required for sphingosine 1-phosphate-induced autophagy in human prostate cancer PC-3 cells. Am. J. Physiol. Cell Physiol. 2009, 297, C451–C458. [Google Scholar] [CrossRef]
  243. Berger, J.; Moller, D.E. The Mechanisms of Action of PPARs. Annu. Rev. Med. 2002, 53, 409–435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Miyata, K.S.; McCaw, S.E.; Marcus, S.L.; Rachubinski, R.A.; Capone, J.P. The peroxisome proliferator-activated receptor interacts with the retinoid X receptor in vivo. Gene 1994, 148, 327–330. [Google Scholar] [CrossRef] [PubMed]
  245. Bull, A.W.; Steffensen, K.R.; Leers, J.; Rafter, J.J. Activation of PPAR γ in colon tumor cell lines by oxidized metabolites of linoleic acid, endogenous ligands for PPAR γ. Carcinogenesis 2003, 24, 1717–1722. [Google Scholar] [CrossRef] [PubMed]
  246. Velkov, T. Interactions between Human Liver Fatty Acid Binding Protein and Peroxisome Proliferator Activated Receptor Selective Drugs. PPAR Res. 2013, 2013, 938401. [Google Scholar] [CrossRef] [Green Version]
  247. Kim, Y.S.; Lee, H.-M.; Kim, J.K.; Yang, C.-S.; Kim, T.S.; Jung, M.; Jin, H.S.; Kim, S.; Jang, J.; Oh, G.T.; et al. PPAR-α Activation Mediates Innate Host Defense through Induction of TFEB and Lipid Catabolism. J. Immunol. 2017, 198, 3283–3295. [Google Scholar] [CrossRef] [Green Version]
  248. Palomer, X.; Capdevila-Busquets, E.; Botteri, G.; Salvadó, L.; Barroso, E.; Davidson, M.M.; Michalik, L.; Wahli, W.; Vázquez-Carrera, M. PPARβ/δ; attenuates palmitate-induced endoplasmic reticulum stress and induces autophagic markers in human cardiac cells. Int. J. Cardiol. 2014, 174, 110–118. [Google Scholar] [CrossRef] [PubMed]
  249. Assumpção, J.A.F.; Magalhães, K.G.; Corrêa, J.R. The role of pparγ and autophagy in ros production, lipid droplets biogenesis and its involvement with colorectal cancer cells modulation. Cancer Cell Int. 2017, 17, 82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Zhou, J.; Zhang, W.; Liang, B.; Casimiro, M.C.; Whitaker-Menezes, D.; Wang, M.; Lisanti, M.P.; Lanza-Jacoby, S.; Pestell, R.G.; Wang, C. PPARγ activation induces autophagy in breast cancer cells. Int. J. Biochem. Cell Biol. 2009, 41, 2334–2342. [Google Scholar] [CrossRef] [Green Version]
  251. Bajaj, M.; Suraamornkul, S.; Kashyap, S.; Cusi, K.; Mandarino, L.; DeFronzo, R.A. Sustained Reduction in Plasma Free Fatty Acid Concentration Improves Insulin Action without Altering Plasma Adipocytokine Levels in Subjects with Strong Family History of Type 2 Diabetes. J. Clin. Endocrinol. Metab. 2004, 89, 4649–4655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Wang, C.-W. Lipid droplet dynamics in budding yeast. Cell. Mol. Life Sci. 2015, 72, 2677–2695. [Google Scholar] [CrossRef]
  253. Carta, G.; Murru, E.; Banni, S.; Manca, C. Palmitic Acid: Physiological Role, Metabolism and Nutritional Implications. Front. Physiol. 2017, 8, 902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Tan, S.H.; Shui, G.; Zhou, J.; Li, J.J.E.; Bay, B.-H.; Wenk, M.R.; Shen, H.-M. Induction of Autophagy by Palmitic Acid via Protein Kinase C-mediated Signaling Pathway Independent of mTOR (Mammalian Target of Rapamycin). J. Biol. Chem. 2012, 287, 14364–14376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Fukui, M.; Kang, K.S.; Okada, K.; Zhu, B.T. EPA, an omega-3 fatty acid, induces apoptosis in human pancreatic cancer cells: Role of ROS accumulation, caspase-8 activation, and autophagy induction. J. Cell. Biochem. 2013, 114, 192–203. [Google Scholar] [CrossRef]
  256. Swanson, D.; Block, R.; Mousa, S.A. Omega-3 fatty acids EPA and DHA: Health benefits throughout life. Adv. Nutr. 2012, 3, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Krauss-Etschmann, S.; Shadid, R.; Campoy, C.; Hoster, E.; Demmelmair, H.; Jiménez, M.; Gil, A.; Rivero, M.; Veszprémi, B.; Decsi, T.; et al. Effects of fish-oil and folate supplementation of pregnant women on maternal and fetal plasma concentrations of docosahexaenoic acid and eicosapentaenoic acid: A European randomized multicenter trial. Am. J. Clin. Nutr. 2007, 85, 1392–1400. [Google Scholar] [CrossRef] [Green Version]
  258. Neff, L.M.; Culiner, J.; Cunningham-Rundles, S.; Seidman, C.; Meehan, D.; Maturi, J.; Wittkowski, K.M.; Levine, B.; Breslow, J.L. Algal docosahexaenoic acid affects plasma lipoprotein particle size distribution in overweight and obese adults. J. Nutr. 2011, 141, 207–213. [Google Scholar] [CrossRef] [Green Version]
  259. Chiu, C.C.; Su, K.P.; Cheng, T.C.; Liu, H.C.; Chang, C.J.; Dewey, M.E.; Stewart, R.; Huang, S.Y. The effects of omega-3 fatty acids monotherapy in Alzheimer’s disease and mild cognitive impairment: A preliminary randomized double-blind placebo-controlled study. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2008, 32, 1538–1544. [Google Scholar] [CrossRef]
  260. Varela-López, A.; Vera-Ramírez, L.; Giampieri, F.; Navarro-Hortal, M.D.; Forbes-Hernández, T.Y.; Battino, M.; Quiles, J.L. The central role of mitochondria in the relationship between dietary lipids and cancer progression. Semin. Cancer Biol. 2021, 73, 86–100. [Google Scholar] [CrossRef]
  261. Shin, S.; Jing, K.; Jeong, S.; Kim, N.; Song, K.-S.; Heo, J.-Y.; Park, J.-H.; Seo, K.-S.; Han, J.; Park, J.-I.; et al. The Omega-3 Polyunsaturated Fatty Acid DHA Induces Simultaneous Apoptosis and Autophagy via Mitochondrial ROS-Mediated Akt-mTOR Signaling in Prostate Cancer Cells Expressing Mutant p53. BioMed Res. Int. 2013, 2013, 568671. [Google Scholar] [CrossRef] [Green Version]
  262. Yao, Q.-H.; Zhang, X.-C.; Fu, T.; Gu, J.-Z.; Wang, L.; Wang, Y.; Lai, Y.-B.; Wang, Y.-Q.; Guo, Y. ω-3 polyunsaturated fatty acids inhibit the proliferation of the lung adenocarcinoma cell line A549 in vitro. Mol. Med. Rep. 2014, 9, 401–406. [Google Scholar] [CrossRef] [Green Version]
  263. Hsu, H.-C.; Chen, C.-Y.; Chiang, C.-H.; Chen, M.-F. Eicosapentaenoic acid attenuated oxidative stress-induced cardiomyoblast apoptosis by activating adaptive autophagy. Eur. J. Nutr. 2014, 53, 541–547. [Google Scholar] [CrossRef] [PubMed]
  264. Jing, K.; Song, K.-S.; Shin, S.; Kim, N.; Jeong, S.; Oh, H.-R.; Park, J.-H.; Seo, K.-S.; Heo, J.-Y.; Han, J.; et al. Docosahexaenoic acid induces autophagy through p53/AMPK/mTOR signaling and promotes apoptosis in human cancer cells harboring wild-type p53. Autophagy 2011, 7, 1348–1358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Tasdemir, E.; Maiuri, M.C.; Galluzzi, L.; Vitale, I.; Djavaheri-Mergny, M.; D’Amelio, M.; Criollo, A.; Morselli, E.; Zhu, C.; Harper, F.; et al. Regulation of autophagy by cytoplasmic p53. Nat. Cell Biol. 2008, 10, 676–687. [Google Scholar] [CrossRef] [Green Version]
  266. D’Eliseo, D.; Di Renzo, L.; Santoni, A.; Velotti, F. Docosahexaenoic acid (DHA) promotes immunogenic apoptosis in human multiple myeloma cells, induces autophagy and inhibits STAT3 in both tumor and dendritic cells. Genes Cancer 2017, 8, 426–437. [Google Scholar] [CrossRef] [Green Version]
  267. Pettersen, K.; Monsen, V.T.; Hakvåg Pettersen, C.H.; Overland, H.B.; Pettersen, G.; Samdal, H.; Tesfahun, A.N.; Lundemo, A.G.; Bjørkøy, G.; Schønberg, S.A. DHA-induced stress response in human colon cancer cells—Focus on oxidative stress and autophagy. Free Radic. Biol. Med. 2016, 90, 158–172. [Google Scholar] [CrossRef] [Green Version]
  268. Kanno, S.-I.; Kurauchi, K.; Tomizawa, A.; Yomogida, S.; Ishikawa, M. Pifithrin-alpha has a p53-independent cytoprotective effect on docosahexaenoic acid-induced cytotoxicity in human hepatocellular carcinoma HepG2 cells. Toxicol. Lett. 2015, 232, 393–402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  269. Omura, Y. Beneficial effects and side effects of DHEA: True anti-aging and age-promoting effects, as well as anti-cancer and cancer-promoting effects of DHEA evaluated from the effects on the normal and cancer cell telomeres and other parameters. Acupunct. Electro-Ther. Res. 2005, 30, 219–261. [Google Scholar] [CrossRef]
  270. Rovito, D.; Giordano, C.; Vizza, D.; Plastina, P.; Barone, I.; Casaburi, I.; Lanzino, M.; De Amicis, F.; Sisci, D.; Mauro, L.; et al. Omega-3 PUFA ethanolamides DHEA and EPEA induce autophagy through PPARγ activation in MCF-7 breast cancer cells. J. Cell. Physiol. 2013, 228, 1314–1322. [Google Scholar] [CrossRef] [PubMed]
  271. Alphonse, P.A.; Jones, P.J. Revisiting Human Cholesterol Synthesis and Absorption: The Reciprocity Paradigm and its Key Regulators. Lipids 2016, 51, 519–536. [Google Scholar] [CrossRef] [PubMed]
  272. Yang, S.-T.; Kreutzberger, A.J.B.; Lee, J.; Kiessling, V.; Tamm, L.K. The role of cholesterol in membrane fusion. Chem. Phys. Lipids 2016, 199, 136–143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Cerqueira, N.M.; Oliveira, E.F.; Gesto, D.S.; Santos-Martins, D.; Moreira, C.; Moorthy, H.N.; Ramos, M.J.; Fernandes, P.A. Cholesterol Biosynthesis: A Mechanistic Overview. Biochemistry 2016, 55, 5483–5506. [Google Scholar] [CrossRef]
  274. Liao, P.; Hemmerlin, A.; Bach, T.J.; Chye, M.L. The potential of the mevalonate pathway for enhanced isoprenoid production. Biotechnol. Adv. 2016, 34, 697–713. [Google Scholar] [CrossRef]
  275. Jiang, S.-Y.; Li, H.; Tang, J.-J.; Wang, J.; Luo, J.; Liu, B.; Wang, J.-K.; Shi, X.-J.; Cui, H.-W.; Tang, J.; et al. Discovery of a potent HMG-CoA reductase degrader that eliminates statin-induced reductase accumulation and lowers cholesterol. Nat. Commun. 2018, 9, 5138. [Google Scholar] [CrossRef] [Green Version]
  276. Shao, W.Q.; Zhu, W.W.; Luo, M.J.; Fan, M.H.; Li, Q.; Wang, S.H.; Lin, Z.F.; Zhao, J.; Zheng, Y.; Dong, Q.Z.; et al. Cholesterol suppresses GOLM1-dependent selective autophagy of RTKs in hepatocellular carcinoma. Cell Rep. 2022, 39, 110712. [Google Scholar] [CrossRef]
  277. Kumar, M.; Irungbam, K.; Kataria, M. Depletion of membrane cholesterol compromised caspase-8 imparts in autophagy induction and inhibition of cell migration in cancer cells. Cancer Cell Int. 2018, 18, 23. [Google Scholar] [CrossRef] [Green Version]
  278. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  279. World Health Organization. Cancer; World Health Organization: Geneva, Switzerland, 2022. [Google Scholar]
  280. American Lung Association. Lung Cancer Fact Sheet; American Lung Association: Chicago, IL, USA, 2020; Available online: https://www.lung.org/lung-health-diseases/lung-disease-lookup/lung-cancer/resource-library/lung-cancer-fact-sheet (accessed on 1 December 2022).
  281. Canadian Cancer Statistics Advisory Committee. Canadian Cancer Statistics 2018; Canadian Cancer Statistics Advisory Committee: Toronto, ON, USA, 2018. [Google Scholar]
  282. WHO Classification of Tumours of the Lung, Pleura, Thymus, and Heart. In WHO Classification of Tumours of the Lung, Pleura, Thymus, and Heart; Travis, W.D.; Brambilla, E.; Burke, A.P.; Marx, A.; Nicholson, A.G. (Eds.) WHO: Lyon, France, 2015. [Google Scholar]
  283. Sakamoto, T.; Furukawa, T.; Pham, H.H.N.; Kuroda, K.; Tabata, K.; Kashima, Y.; Okoshi, E.N.; Morimoto, S.; Bychkov, A.; Fukuoka, J. A collaborative workflow between pathologists and deep learning for the evaluation of tumour cellularity in lung adenocarcinoma. Histopathology 2022, 81, 758–769. [Google Scholar] [CrossRef]
  284. Oktay, E.; Oflazoglu, U.; Varol, Y.; Tanriverdi, O.; Mermur, N.; Arda, H.U.; Demir, L.; Keskin, O.; Ahmadli, T.; Somali, I.; et al. The prognostic role of thyroid transcription factor-1 in lung adenocarcinoma. J. Cancer Res. Ther. 2020, 16, 737–744. [Google Scholar] [CrossRef] [PubMed]
  285. Pao, W.; Girard, N. New driver mutations in non-small-cell lung cancer. Lancet Oncol. 2011, 12, 175–180. [Google Scholar] [CrossRef]
  286. Gennen, K.; Kasmann, L.; Taugner, J.; Eze, C.; Karin, M.; Roengvoraphoj, O.; Neumann, J.; Tufman, A.; Orth, M.; Reu, S.; et al. Prognostic value of PD-L1 expression on tumor cells combined with CD8+ TIL density in patients with locally advanced non-small cell lung cancer treated with concurrent chemoradiotherapy. Radiat. Oncol. 2020, 15, 5. [Google Scholar] [CrossRef] [PubMed]
  287. Yang, H.; Shi, J.; Lin, D.; Li, X.; Zhao, C.; Wang, Q.; Zhang, L.; Jiang, T.; Zhao, S.; Liu, X.; et al. Prognostic value of PD-L1 expression in combination with CD8(+) TILs density in patients with surgically resected non-small cell lung cancer. Cancer Med. 2018, 7, 32–45. [Google Scholar] [CrossRef] [Green Version]
  288. Wang, A.; Wang, H.Y.; Liu, Y.; Zhao, M.C.; Zhang, H.J.; Lu, Z.Y.; Fang, Y.C.; Chen, X.F.; Liu, G.T. The prognostic value of PD-L1 expression for non-small cell lung cancer patients: A meta-analysis. Eur. J. Surg. Oncol. 2015, 41, 450–456. [Google Scholar] [CrossRef] [PubMed]
  289. Daaboul, N.; Nicholas, G.; Laurie, S.A. Algorithm for the treatment of advanced or metastatic squamous non-small-cell lung cancer: An evidence-based overview. Curr. Oncol. 2018, 25 (Suppl. 1), S77–S85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  290. Umakanthan, S.; Bukelo, M.M. Concise genetic profile of lung carcinoma. Postgrad. Med. J. 2021. Online ahead of print. [Google Scholar] [CrossRef]
  291. Alberg, A.J.; Samet, J.M. Epidemiology of lung cancer. Chest 2003, 123 (Suppl. 1), 21S–49S. [Google Scholar] [CrossRef] [Green Version]
  292. Taylor, R.; Najafi, F.; Dobson, A. Meta-analysis of studies of passive smoking and lung cancer: Effects of study type and continent. Int. J. Epidemiol. 2007, 36, 1048–1059. [Google Scholar] [CrossRef]
  293. Kim, C.H.; Lee, Y.C.; Hung, R.J.; McNallan, S.R.; Cote, M.L.; Lim, W.Y.; Chang, S.C.; Kim, J.H.; Ugolini, D.; Chen, Y.; et al. Exposure to secondhand tobacco smoke and lung cancer by histological type: A pooled analysis of the International Lung Cancer Consortium (ILCCO). Int. J. Cancer 2014, 135, 1918–1930. [Google Scholar] [CrossRef] [Green Version]
  294. Darby, S.; Hill, D.; Auvinen, A.; Barros-Dios, J.M.; Baysson, H.; Bochicchio, F.; Deo, H.; Falk, R.; Forastiere, F.; Hakama, M.; et al. Radon in homes and risk of lung cancer: Collaborative analysis of individual data from 13 European case-control studies. BMJ 2005, 330, 223. [Google Scholar] [CrossRef] [Green Version]
  295. Krewski, D.; Lubin, J.H.; Zielinski, J.M.; Alavanja, M.; Catalan, V.S.; Field, R.W.; Klotz, J.B.; Letourneau, E.G.; Lynch, C.F.; Lyon, J.L.; et al. A combined analysis of North American case-control studies of residential radon and lung cancer. J. Toxicol. Environ. Health A 2006, 69, 533–597. [Google Scholar] [CrossRef] [Green Version]
  296. Neuberger, J.S.; Field, R.W. Occupation and lung cancer in nonsmokers. Rev. Environ. Health 2003, 18, 251–267. [Google Scholar] [CrossRef]
  297. van Loon, A.J.; Kant, I.J.; Swaen, G.M.; Goldbohm, R.A.; Kremer, A.M.; van den Brandt, P.A. Occupational exposure to carcinogens and risk of lung cancer: Results from The Netherlands cohort study. Occup. Environ. Med. 1997, 54, 817–824. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  298. Gottschall, E.B. Occupational and environmental thoracic malignancies. J. Thorac. Imaging 2002, 17, 189–197. [Google Scholar] [CrossRef]
  299. Hubbard, R.; Venn, A.; Lewis, S.; Britton, J. Lung cancer and cryptogenic fibrosing alveolitis. A population-based cohort study. Am. J. Respir. Crit. Care Med. 2000, 161, 5–8. [Google Scholar] [CrossRef]
  300. Chen, Y.C.; Chen, J.H.; Richard, K.; Chen, P.Y.; Christiani, D.C. Lung adenocarcinoma and human papillomavirus infection. Cancer 2004, 101, 1428–1436. [Google Scholar] [CrossRef]
  301. Kirk, G.D.; Merlo, C.; O’Driscoll, P.; Mehta, S.H.; Galai, N.; Vlahov, D.; Samet, J.; Engels, E.A. HIV infection is associated with an increased risk for lung cancer, independent of smoking. Clin. Infect. Dis. 2007, 45, 103–110. [Google Scholar] [CrossRef] [Green Version]
  302. Chaturvedi, A.K.; Pfeiffer, R.M.; Chang, L.; Goedert, J.J.; Biggar, R.J.; Engels, E.A. Elevated risk of lung cancer among people with AIDS. AIDS 2007, 21, 207–213. [Google Scholar] [CrossRef] [PubMed]
  303. D’Jaen, G.A.; Pantanowitz, L.; Bower, M.; Buskin, S.; Neil, N.; Greco, E.M.; Cooley, T.P.; Henry, D.; Stem, J.; Dezube, B.J.; et al. Human immunodeficiency virus-associated primary lung cancer in the era of highly active antiretroviral therapy: A multi-institutional collaboration. Clin. Lung Cancer 2010, 11, 396–404. [Google Scholar] [CrossRef] [PubMed]
  304. Dziadziuszko, R.; Peters, S.; Ruf, T.; Cardona, A.; Guerini, E.; Kurtsikidze, N.; Smoljanovic, V.; Planchard, D. Clinical experience and management of adverse events in patients with advanced ALK-positive non-small-cell lung cancer receiving alectinib. ESMO Open 2022, 7, 100612. [Google Scholar] [CrossRef]
  305. Sun, R.; Hou, Z.; Zhang, Y.; Jiang, B. Drug resistance mechanisms and progress in the treatment of EGFR-mutated lung adenocarcinoma. Oncol. Lett. 2022, 24, 408. [Google Scholar] [CrossRef]
  306. Lindeman, N.I.; Cagle, P.T.; Beasley, M.B.; Chitale, D.A.; Dacic, S.; Giaccone, G.; Jenkins, R.B.; Kwiatkowski, D.J.; Saldivar, J.S.; Squire, J.; et al. Molecular testing guideline for selection of lung cancer patients for EGFR and ALK tyrosine kinase inhibitors: Guideline from the College of American Pathologists, International Association for the Study of Lung Cancer, and Association for Molecular Pathology. J. Thorac. Oncol. 2013, 8, 823–859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  307. Dower, C.M.; Wills, C.A.; Frisch, S.M.; Wang, H.G. Mechanisms and context underlying the role of autophagy in cancer metastasis. Autophagy 2018, 14, 1110–1128. [Google Scholar] [CrossRef] [Green Version]
  308. Chan, E.Y.; Kir, S.; Tooze, S.A. siRNA screening of the kinome identifies ULK1 as a multidomain modulator of autophagy. J. Biol. Chem. 2007, 282, 25464–25474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  309. Mleczak, A.; Millar, S.; Tooze, S.A.; Olson, M.F.; Chan, E.Y. Regulation of autophagosome formation by Rho kinase. Cell Signal 2013, 25, 1–11. [Google Scholar] [CrossRef] [PubMed]
  310. Gurkar, A.U.; Chu, K.; Raj, L.; Bouley, R.; Lee, S.-H.; Kim, Y.-B.; Dunn, S.E.; Mandinova, A.; Lee, S.W. Identification of ROCK1 kinase as a critical regulator of BECLIN1-mediated autophagy during metabolic stress. Nat. Commun. 2013, 4, 2189. [Google Scholar] [CrossRef] [Green Version]
  311. Belaid, A.; Cerezo, M.; Chargui, A.; Corcelle-Termeau, E.; Pedeutour, F.; Giuliano, S.; Ilie, M.; Rubera, I.; Tauc, M.; Barale, S.; et al. Autophagy plays a critical role in the degradation of active RHOA, the control of cell cytokinesis, and genomic stability. Cancer Res. 2013, 73, 4311–4322. [Google Scholar] [CrossRef] [Green Version]
  312. Belaid, A.; Ndiaye, P.D.; Cerezo, M.; Cailleteau, L.; Brest, P.; Klionsky, D.J.; Carle, G.F.; Hofman, P.; Mograbi, B. Autophagy and SQSTM1 on the RHOA(d) again: Emerging roles of autophagy in the degradation of signaling proteins. Autophagy 2014, 10, 201–208. [Google Scholar] [CrossRef] [Green Version]
  313. Yoshida, T.; Tsujioka, M.; Honda, S.; Tanaka, M.; Shimizu, S. Autophagy suppresses cell migration by degrading GEF-H1, a RhoA GEF. Oncotarget 2016, 7, 34420–34429. [Google Scholar] [CrossRef]
  314. Sanz-Moreno, V.; Gadea, G.; Ahn, J.; Paterson, H.; Marra, P.; Pinner, S.; Sahai, E.; Marshall, C.J. Rac activation and inactivation control plasticity of tumor cell movement. Cell 2008, 135, 510–523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  315. Fang, K.; Chen, X.; Qiu, F.; Xu, J.; Xiong, H.; Zhang, Z. Serum-Derived Exosomes-Mediated Circular RNA ARHGAP10 Modulates the Progression of Non-Small Cell Lung Cancer Through the miR-638/FAM83F Axis. Cancer Biother. Radiopharm. 2022, 37, 96–110. [Google Scholar] [CrossRef]
  316. Li, P.; Lv, H.; Xu, M.; Zang, B.; Ma, Y. ARHGAP6 Promotes Apoptosis and Inhibits Glycolysis in Lung Adenocarcinoma Through STAT3 Signaling Pathway. Cancer Manag. Res. 2020, 12, 9665–9678. [Google Scholar] [CrossRef] [PubMed]
  317. Maiuri, M.C.; Le Toumelin, G.; Criollo, A.; Rain, J.C.; Gautier, F.; Juin, P.; Tasdemir, E.; Pierron, G.; Troulinaki, K.; Tavernarakis, N.; et al. Functional and physical interaction between Bcl-X(L) and a BH3-like domain in BECLIN1. EMBO J. 2007, 26, 2527–2539. [Google Scholar] [CrossRef]
  318. Erlich, S.; Mizrachy, L.; Segev, O.; Lindenboim, L.; Zmira, O.; Adi-Harel, S.; Hirsch, J.A.; Stein, R.; Pinkas-Kramarski, R. Differential interactions between BECLIN 1 and Bcl-2 family members. Autophagy 2007, 3, 561–568. [Google Scholar] [CrossRef] [Green Version]
  319. Pattingre, S.; Tassa, A.; Qu, X.; Garuti, R.; Liang, X.H.; Mizushima, N.; Packer, M.; Schneider, M.D.; Levine, B. Bcl-2 antiapoptotic proteins inhibit BECLIN 1-dependent autophagy. Cell 2005, 122, 927–939. [Google Scholar] [CrossRef] [Green Version]
  320. Adams, J.M.; Cory, S. The Bcl-2 apoptotic switch in cancer development and therapy. Oncogene 2007, 26, 1324–1337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  321. Amundson, S.A.; Myers, T.G.; Scudiero, D.; Kitada, S.; Reed, J.C.; Fornace, A.J., Jr. An informatics approach identifying markers of chemosensitivity in human cancer cell lines. Cancer Res. 2000, 60, 6101–6110. [Google Scholar]
  322. Lindqvist, L.M.; Simon, A.K.; Baehrecke, E.H. Current questions and possible controversies in autophagy. Cell Death Discov. 2015, 1, 15036. [Google Scholar] [CrossRef] [Green Version]
  323. Lindqvist, L.M.; Heinlein, M.; Huang, D.C.; Vaux, D.L. Prosurvival Bcl-2 family members affect autophagy only indirectly, by inhibiting Bax and Bak. Proc. Natl. Acad. Sci. USA 2014, 111, 8512–8517. [Google Scholar] [CrossRef] [Green Version]
  324. Lindqvist, L.M.; Vaux, D.L. BCL2 and related prosurvival proteins require BAK1 and BAX to affect autophagy. Autophagy 2014, 10, 1474–1475. [Google Scholar] [CrossRef] [Green Version]
  325. Pedro, J.M.B.-S.; Wei, Y.; Sica, V.; Maiuri, M.C.; Zou, Z.; Kroemer, G.; Levine, B. BAX and BAK1 are dispensable for ABT-737-induced dissociation of the BCL2-BECN1 complex and autophagy. Autophagy 2015, 11, 452–459. [Google Scholar] [CrossRef] [Green Version]
  326. Reljic, B.; Conos, S.; Lee, E.F.; Garnier, J.M.; Dong, L.; Lessene, G.; Fairlie, W.D.; Vaux, D.L.; Lindqvist, L.M. BAX-BAK1-independent LC3B lipidation by BH3 mimetics is unrelated to BH3 mimetic activity and has only minimal effects on autophagic flux. Autophagy 2016, 12, 1083–1093. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  327. Cerami, E.; Gao, J.; Dogrusoz, U.; Gross, B.E.; Sumer, S.O.; Aksoy, B.A.; Jacobsen, A.; Byrne, C.J.; Heuer, M.L.; Larsson, E.; et al. The cBio cancer genomics portal: An open platform for exploring multidimensional cancer genomics data. Cancer Discov. 2012, 2, 401–404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  328. Gao, J.; Aksoy, B.A.; Dogrusoz, U.; Dresdner, G.; Gross, B.; Sumer, S.O.; Sun, Y.; Jacobsen, A.; Sinha, R.; Larsson, E.; et al. Integrative analysis of complex cancer genomics and clinical profiles using the cBioPortal. Sci. Signal. 2013, 6, pl1. [Google Scholar] [CrossRef] [Green Version]
  329. Campbell, J.D.; Alexandrov, A.; Kim, J.; Wala, J.; Berger, A.H.; Pedamallu, C.S.; Shukla, S.A.; Guo, G.; Brooks, A.N.; Murray, B.A.; et al. Distinct patterns of somatic genome alterations in lung adenocarcinomas and squamous cell carcinomas. Nat. Genet. 2016, 48, 607–616. [Google Scholar] [CrossRef] [Green Version]
  330. Hanahan, D. Hallmarks of Cancer: New Dimensions. Cancer Discov. 2022, 12, 31–46. [Google Scholar] [CrossRef] [PubMed]
  331. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  332. Hanahan, D.; Weinberg, R.A. The hallmarks of cancer. Cell 2000, 100, 57–70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  333. Hombach-Klonisch, S.; Mehrpour, M.; Shojaei, S.; Harlos, C.; Pitz, M.; Hamai, A.; Siemianowicz, K.; Likus, W.; Wiechec, E.; Toyota, B.D.; et al. Glioblastoma and chemoresistance to alkylating agents: Involvement of apoptosis, autophagy, and unfolded protein response. Pharmacol. Ther. 2018, 184, 13–41. [Google Scholar] [CrossRef]
  334. Laplante, M.; Sabatini, D.M. mTOR signaling in growth control and disease. Cell 2012, 149, 274–293. [Google Scholar] [CrossRef] [Green Version]
  335. Rao, S.; Tortola, L.; Perlot, T.; Wirnsberger, G.; Novatchkova, M.; Nitsch, R.; Sykacek, P.; Frank, L.; Schramek, D.; Komnenovic, V.; et al. A dual role for autophagy in a murine model of lung cancer. Nat. Commun. 2014, 5, 3056. [Google Scholar] [CrossRef] [Green Version]
  336. Feng, Z.; Zhang, H.; Levine, A.J.; Jin, S. The coordinate regulation of the p53 and mTOR pathways in cells. Proc. Natl. Acad. Sci. USA 2005, 102, 8204–8209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  337. Shaw, R.J.; Bardeesy, N.; Manning, B.D.; Lopez, L.; Kosmatka, M.; DePinho, R.A.; Cantley, L.C. The LKB1 tumor suppressor negatively regulates mTOR signaling. Cancer Cell 2004, 6, 91–99. [Google Scholar] [CrossRef] [Green Version]
  338. Shackelford, D.B.; Vasquez, D.S.; Corbeil, J.; Wu, S.; Leblanc, M.; Wu, C.-L.; Vera, D.R.; Shaw, R.J. mTOR and HIF-1α-mediated tumor metabolism in an LKB1 mouse model of Peutz-Jeghers syndrome. Proc. Natl. Acad. Sci. USA 2009, 106, 11137–11142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  339. Faubert, B.; Vincent, E.E.; Griss, T.; Samborska, B.; Izreig, S.; Svensson, R.U.; Mamer, O.A.; Avizonis, D.; Shackelford, D.B.; Shaw, R.J. Loss of the tumor suppressor LKB1 promotes metabolic reprogramming of cancer cells via HIF-1α. Proc. Natl. Acad. Sci. USA 2014, 111, 2554–2559. [Google Scholar] [CrossRef] [Green Version]
  340. Sanchez-Cespedes, M.; Parrella, P.; Esteller, M.; Nomoto, S.; Trink, B.; Engles, J.M.; Westra, W.H.; Herman, J.G.; Sidransky, D. Inactivation of LKB1/STK11 is a common event in adenocarcinomas of the lung. Cancer Res. 2002, 62, 3659–3662. [Google Scholar]
  341. Ding, L.; Getz, G.; Wheeler, D.A.; Mardis, E.R.; McLellan, M.D.; Cibulskis, K.; Sougnez, C.; Greulich, H.; Muzny, D.M.; Morgan, M.B. Somatic mutations affect key pathways in lung adenocarcinoma. Nature 2008, 455, 1069–1075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  342. Momcilovic, M.; Bailey, S.T.; Lee, J.T.; Zamilpa, C.; Jones, A.; Abdelhady, G.; Mansfield, J.; Francis, K.P.; Shackelford, D.B. Utilizing 18F-FDG PET/CT imaging and quantitative histology to measure dynamic changes in the glucose metabolism in mouse models of lung cancer. J. Vis. Exp. JoVE 2018, 137, e57167. [Google Scholar]
  343. Luo, L.-X.; Li, Y.; Liu, Z.-Q.; Fan, X.-X.; Duan, F.-G.; Li, R.-Z.; Yao, X.-J.; Leung, E.L.-H.; Liu, L. Honokiol induces apoptosis, G1 arrest, and autophagy in KRAS mutant lung cancer cells. Front. Pharmacol. 2017, 8, 199. [Google Scholar] [CrossRef] [Green Version]
  344. Kapuy, O.; Makk-Merczel, K.; Szarka, A. Therapeutic Approach of KRAS Mutant Tumours by the Combination of Pharmacologic Ascorbate and Chloroquine. Biomolecules 2021, 11, 652. [Google Scholar] [CrossRef]
  345. Morita, M.; Sato, T.; Nomura, M.; Sakamoto, Y.; Inoue, Y.; Tanaka, R.; Ito, S.; Kurosawa, K.; Yamaguchi, K.; Sugiura, Y.; et al. PKM1 Confers Metabolic Advantages and Promotes Cell-Autonomous Tumor Cell Growth. Cancer Cell 2018, 33, 355–367.e357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  346. Fantin, V.R.; Abraham, R.T. Self-eating limits EGFR-dependent tumor growth. Cell 2013, 154, 1184–1186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  347. Ye, M.; Wang, S.; Wan, T.; Jiang, R.; Qiu, Y.; Pei, L.; Pang, N.; Huang, Y.; Huang, Y.; Zhang, Z. Combined inhibitions of glycolysis and AKT/autophagy can overcome resistance to EGFR-targeted therapy of lung cancer. J. Cancer 2017, 8, 3774. [Google Scholar] [CrossRef] [Green Version]
  348. Chen, Z.; Yang, D.; Jiang, X.; Tian, L.; Sun, J.; Tian, D.; Bai, W. Final-2 targeted glycolysis mediated apoptosis and autophagy in human lung adenocarcinoma cells but failed to inhibit xenograft in nude mice. Food Chem. Toxicol. 2019, 130, 1–11. [Google Scholar] [CrossRef]
  349. Xia, S.; Duan, W.; Liu, W.; Zhang, X.; Wang, Q. GRP78 in lung cancer. J. Transl. Med. 2021, 19, 1–14. [Google Scholar] [CrossRef]
  350. Bello-Perez, M.; Sola, I.; Novoa, B.; Klionsky, D.J.; Falco, A. Canonical and noncanonical autophagy as potential targets for COVID-19. Cells 2020, 9, 1619. [Google Scholar] [CrossRef]
  351. Wang, M.; Wey, S.; Zhang, Y.; Ye, R.; Lee, A.S. Role of the unfolded protein response regulator GRP78/BiP in development, cancer, and neurological disorders. Antioxid. Redox Signal. 2009, 11, 2307–2316. [Google Scholar] [CrossRef] [Green Version]
  352. Kania, E.; Pająk, B.; Orzechowski, A. Calcium homeostasis and ER stress in control of autophagy in cancer cells. BioMed Res. Int. 2015, 2015, 352794. [Google Scholar] [CrossRef] [Green Version]
  353. Ma, X.; Guo, W.; Yang, S.; Zhu, X.; Xiang, J.; Li, H. Serum GRP78 as a tumor marker and its prognostic significance in non-small cell lung cancers: A retrospective study. Dis. Mrk. 2015, 2015, 814670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  354. Wu, J.; Wu, Y.; Lian, X. Targeted inhibition of GRP78 by HA15 promotes apoptosis of lung cancer cells accompanied by ER stress and autophagy. Biol. Open 2020, 9, bio053298. [Google Scholar] [CrossRef]
  355. Parker, A.L.; Kavallaris, M.; McCarroll, J.A. Microtubules and their role in cellular stress in cancer. Front. Oncol. 2014, 4, 153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  356. Parker, A.L.; Turner, N.; McCarroll, J.A.; Kavallaris, M. βIII-Tubulin alters glucose metabolism and stress response signaling to promote cell survival and proliferation in glucose-starved non-small cell lung cancer cells. Carcinogenesis 2016, 37, 787–798. [Google Scholar] [CrossRef] [Green Version]
  357. Rim, H.-K.; Cho, S.; Shin, D.-H.; Chung, K.-S.; Cho, Y.-W.; Choi, J.-H.; Lee, J.Y.; Lee, K.-T. T-type Ca2+ channel blocker, KYS05090 induces autophagy and apoptosis in A549 cells through inhibiting glucose uptake. Molecules 2014, 19, 9864–9875. [Google Scholar] [CrossRef]
  358. Bootman, M.D.; Chehab, T.; Bultynck, G.; Parys, J.B.; Rietdorf, K. The regulation of autophagy by calcium signals: Do we have a consensus? Cell Calcium 2018, 70, 32–46. [Google Scholar] [CrossRef]
  359. Sallán, M.C.; Visa, A.; Shaikh, S.; Nàger, M.; Herreros, J.; Cantí, C. T-type Ca2+ Channels: T for Targetable. Cancer Res. 2018, 78, 603–609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  360. Villar, V.H.; Durán, R.V. Glutamoptosis: A new cell death mechanism inhibited by autophagy during nutritional imbalance. Autophagy 2017, 13, 1078–1079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  361. Bruntz, R.C.; Belshoff, A.C.; Zhang, Y.; Macedo, J.K.; Higashi, R.M.; Lane, A.N.; Fan, T.W.M. Inhibition of anaplerotic glutaminolysis underlies selenite toxicity in human lung cancer. Proteomics 2019, 19, 1800486. [Google Scholar] [CrossRef]
  362. Pérez-Escuredo, J.; Dadhich, R.K.; Dhup, S.; Cacace, A.; Van Hée, V.F.; De Saedeleer, C.J.; Sboarina, M.; Rodriguez, F.; Fontenille, M.-J.; Brisson, L. Lactate promotes glutamine uptake and metabolism in oxidative cancer cells. Cell Cycle 2016, 15, 72–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  363. Jin, L.; Alesi, G.; Kang, S. Glutaminolysis as a target for cancer therapy. Oncogene 2016, 35, 3619–3625. [Google Scholar] [CrossRef] [Green Version]
  364. Sonveaux, P.; Végran, F.; Schroeder, T.; Wergin, M.C.; Verrax, J.; Rabbani, Z.N.; De Saedeleer, C.J.; Kennedy, K.M.; Diepart, C.; Jordan, B.F. Targeting lactate-fueled respiration selectively kills hypoxic tumor cells in mice. J. Clin. Investig. 2008, 118, 3930–3942. [Google Scholar] [CrossRef] [Green Version]
  365. Xie, W.-Y.; Zhou, X.-D.; Li, Q.; Chen, L.-X.; Ran, D.-H. Acid-induced autophagy protects human lung cancer cells from apoptosis by activating ER stress. Exp. Cell Res. 2015, 339, 270–279. [Google Scholar] [CrossRef] [PubMed]
  366. Lim, S.-C.; Kim, S.M.; Choi, J.E.; Kim, C.H.; Duong, H.-Q.; Han, S.I.; Kang, H.S. Sodium salicylate switches glucose depletion-induced necrosis to autophagy and inhibits high mobility group box protein 1 release in A549 lung adenocarcinoma cells. Oncol. Rep. 2008, 19, 1165–1171. [Google Scholar] [CrossRef]
  367. Jiang, M.-J.; Dai, J.-J.; Gu, D.-N.; Huang, Q.; Tian, L. Aspirin in pancreatic cancer: Chemopreventive effects and therapeutic potentials. Biochim. Et Biophys. Acta (BBA) Rev. Cancer 2016, 1866, 163–176. [Google Scholar] [CrossRef]
  368. Din, F.V.; Valanciute, A.; Houde, V.P.; Zibrova, D.; Green, K.A.; Sakamoto, K.; Alessi, D.R.; Dunlop, M.G. Aspirin inhibits mTOR signaling, activates AMP-activated protein kinase, and induces autophagy in colorectal cancer cells. Gastroenterology 2012, 142, 1504–1515.e1503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  369. Perera, R.; Son, J.; Lyssiotis, C.; Ying, H.; Wang, X.; Hua, S.; Ligorio, M.; Perera, R.; Ferrone, C.; Mullarky, E. Glutamine supports pancreatic cancer growth through a KRAS-regulated metabolic pathway. Nature 2013, 496, 101–105. [Google Scholar]
  370. Zhou, X.; Curbo, S.; Li, F.; Krishnan, S.; Karlsson, A. Inhibition of glutamate oxaloacetate transaminase 1 in cancer cell lines results in altered metabolism with increased dependency of glucose. BMC Cancer 2018, 18, 1–14. [Google Scholar] [CrossRef]
  371. Kang, Y.-T.; Hsu, W.-C.; Wu, C.-H.; Hsin, I.-L.; Wu, P.-R.; Yeh, K.-T.; Ko, J.-L. Metformin alleviates nickel-induced autophagy and apoptosis via inhibition of hexokinase-2, activating lipocalin-2, in human bronchial epithelial cells. Oncotarget 2017, 8, 105536. [Google Scholar] [CrossRef] [Green Version]
  372. Chaudhri, V.K.; Salzler, G.G.; Dick, S.A.; Buckman, M.S.; Sordella, R.; Karoly, E.D.; Mohney, R.; Stiles, B.M.; Elemento, O.; Altorki, N.K. Metabolic alterations in lung cancer–associated fibroblasts correlated with increased glycolytic metabolism of the tumor. Mol. Cancer Res. 2013, 11, 579–592. [Google Scholar] [CrossRef] [Green Version]
  373. Wang, H.; Wang, L.; Zhang, Y.; Wang, J.; Deng, Y.; Lin, D. Inhibition of glycolytic enzyme hexokinase II (HK2) suppresses lung tumor growth. Cancer Cell Int. 2016, 16, 1–11. [Google Scholar] [CrossRef] [Green Version]
  374. Xi, F.; Ye, J. Inhibition of lung carcinoma A549 cell growth by knockdown of hexokinase 2 in situ and in vivo. Oncol. Res. Featur. Preclin. Clin. Cancer Ther. 2016, 23, 53–59. [Google Scholar] [CrossRef]
  375. Wei, J.; Ma, Z.; Li, Y.; Zhao, B.; Wang, D.; Jin, Y.; Jin, Y. miR-143 inhibits cell proliferation by targeting autophagy-related 2B in non-small cell lung cancer H1299 cells. Mol. Med. Rep. 2015, 11, 571–576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  376. Lypova, N.; Telang, S.; Chesney, J.; Imbert-Fernandez, Y. Increased 6-phosphofructo-2-kinase/fructose-2, 6-bisphosphatase-3 activity in response to EGFR signaling contributes to non–small cell lung cancer cell survival. J. Biol. Chem. 2019, 294, 10530–10543. [Google Scholar] [CrossRef] [PubMed]
  377. Wang, Q.; Zeng, F.; Sun, Y.; Qiu, Q.; Zhang, J.; Huang, W.; Huang, J.; Huang, X.; Guo, L. Etk interaction with PFKFB4 modulates chemoresistance of small-cell lung cancer by regulating autophagy. Clin. Cancer Res. 2018, 24, 950–962. [Google Scholar] [CrossRef] [Green Version]
  378. Meng, M.-B.; Wang, H.-H.; Guo, W.-H.; Wu, Z.-Q.; Zeng, X.-L.; Zaorsky, N.G.; Shi, H.-S.; Qian, D.; Niu, Z.-M.; Jiang, B. Targeting pyruvate kinase M2 contributes to radiosensitivity of non-small cell lung cancer cells in vitro and in vivo. Cancer Lett. 2015, 356, 985–993. [Google Scholar] [CrossRef] [PubMed]
  379. Zou, Y.-M.; Hu, G.-Y.; Zhao, X.-Q.; Lu, T.; Zhu, F.; Yu, S.-Y.; Xiong, H. Hypoxia-induced autophagy contributes to radioresistance via c-Jun-mediated BECLIN1 expression in lung cancer cells. J. Huazhong Univ. Sci. Technol. [Med. Sci.] 2014, 34, 761–767. [Google Scholar] [CrossRef]
  380. Hsu, H.-L.; Liao, P.-L.; Cheng, Y.-W.; Huang, S.-H.; Wu, C.-H.; Li, C.-H.; Kang, J.-J. Chloramphenicol induces autophagy and inhibits the hypoxia inducible factor-1 alpha pathway in non-small cell lung cancer cells. Int. J. Mol. Sci. 2019, 20, 157. [Google Scholar] [CrossRef] [Green Version]
  381. Nielsen, S.F.; Nordestgaard, B.G.; Bojesen, S.E. Statin use and reduced cancer-related mortality. N. Engl. J. Med. 2012, 367, 1792–1802. [Google Scholar] [CrossRef] [Green Version]
  382. Yang, Z.; Su, Z.; DeWitt, J.P.; Xie, L.; Chen, Y.; Li, X.; Han, L.; Li, D.; Xia, J.; Zhang, Y.; et al. Fluvastatin Prevents Lung Adenocarcinoma Bone Metastasis by Triggering Autophagy. EBioMedicine 2017, 19, 49–59. [Google Scholar] [CrossRef] [Green Version]
  383. Sanli, T.; Liu, C.; Rashid, A.; Hopmans, S.N.; Tsiani, E.; Schultz, C.; Farrell, T.; Singh, G.; Wright, J.; Tsakiridis, T. Lovastatin sensitizes lung cancer cells to ionizing radiation: Modulation of molecular pathways of radioresistance and tumor suppression. J. Thorac. Oncol. 2011, 6, 439–450. [Google Scholar] [CrossRef] [Green Version]
  384. Zhu, L.; Wang, Z.; Lin, Y.; Chen, Z.; Liu, H.; Chen, Y.; Wang, N.; Song, X. Sphingosine kinase 1 enhances the invasion and migration of non-small cell lung cancer cells via the AKT pathway. Oncol. Rep. 2015, 33, 1257–1263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  385. Gachechiladze, M.; Tichý, T.; Kolek, V.; Grygárková, I.; Klein, J.; Mgebrishvili, G.; Kharaishvili, G.; Janíková, M.; Smičková, P.; Cierna, L.; et al. Sphingosine kinase-1 predicts overall survival outcomes in non-small cell lung cancer patients treated with carboplatin and navelbine. Oncol. Lett. 2019, 18, 1259–1266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  386. Song, L.; Xiong, H.; Li, J.; Liao, W.; Wang, L.; Wu, J.; Li, M. Sphingosine kinase-1 enhances resistance to apoptosis through activation of PI3K/Akt/NF-κB pathway in human non-small cell lung cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2011, 17, 1839–1849. [Google Scholar] [CrossRef] [Green Version]
  387. Nagahashi, M.; Ramachandran, S.; Kim, E.Y.; Allegood, J.C.; Rashid, O.M.; Yamada, A.; Zhao, R.; Milstien, S.; Zhou, H.; Spiegel, S.; et al. Sphingosine-1-phosphate produced by sphingosine kinase 1 promotes breast cancer progression by stimulating angiogenesis and lymphangiogenesis. Cancer Res. 2012, 72, 726–735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  388. Cho, J.H.; Hong, S.K.; Kim, E.Y.; Park, S.Y.; Park, C.H.; Kim, J.M.; Kwon, O.J.; Kwon, S.J.; Lee, K.S.; Han, J.S. Overexpression of phospholipase D suppresses taxotere-induced cell death in stomach cancer cells. Biochim. Et Biophys. Acta 2008, 1783, 912–923. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  389. Hsu, Y.L.; Hung, J.Y.; Ko, Y.C.; Hung, C.H.; Huang, M.S.; Kuo, P.L. Phospholipase D signaling pathway is involved in lung cancer-derived IL-8 increased osteoclastogenesis. Carcinogenesis 2010, 31, 587–596. [Google Scholar] [CrossRef]
  390. Ahn, M.-J.; Park, S.-Y.; Kim, W.K.; Cho, J.H.; Chang, B.J.; Kim, D.J.; Ahn, J.S.; Park, K.; Han, J.-S. A Single Nucleotide Polymorphism in the Phospholipase D1 Gene is Associated with Risk of Non-Small Cell Lung Cancer. Int. J. Biomed. Sci. 2012, 8, 121–128. [Google Scholar]
  391. Shatz, O.; Holland, P.; Elazar, Z.; Simonsen, A. Complex Relations Between Phospholipids, Autophagy, and Neutral Lipids. Trends Biochem. Sci. 2016, 41, 907–923. [Google Scholar] [CrossRef]
  392. Meacci, E.; Vasta, V.; Moorman, J.P.; Bobak, D.A.; Bruni, P.; Moss, J.; Vaughan, M. Effect of Rho and ADP-ribosylation Factor GTPases on Phospholipase D Activity in Intact Human Adenocarcinoma A549 Cells. J. Biol. Chem. 1999, 274, 18605–18612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  393. Meacci, E.; Nuti, F.; Catarzi, S.; Vasta, V.; Donati, C.; Bourgoin, S.; Bruni, P.; Moss, J.; Vaughan, M. Activation of Phospholipase D by Bradykinin and Sphingosine 1-Phosphate in A549 Human Lung Adenocarcinoma Cells via Different GTP-Binding Proteins and Protein Kinase C Delta Signaling Pathways. Biochemistry 2003, 42, 284–292. [Google Scholar] [CrossRef]
  394. Granade, M.E.; Harris, T.E. Purification of Lipin and Measurement of Phosphatidic Acid Phosphatase Activity from Liposomes. Methods Enzymol. 2018, 607, 373–388. [Google Scholar] [CrossRef]
  395. Fan, X.; Weng, Y.; Bai, Y.; Wang, Z.; Wang, S.; Zhu, J.; Zhang, F. Lipin-1 determines lung cancer cell survival and chemotherapy sensitivity by regulation of endoplasmic reticulum homeostasis and autophagy. Cancer Med. 2018, 7, 2541–2554. [Google Scholar] [CrossRef]
  396. Nazim, U.M.; Moon, J.-H.; Lee, Y.-J.; Seol, J.-W.; Park, S.-Y. PPARγ activation by troglitazone enhances human lung cancer cells to TRAIL-induced apoptosis via autophagy flux. Oncotarget 2017, 8, 26819–26831. [Google Scholar] [CrossRef] [Green Version]
  397. To, K.K.W.; Wu, W.K.K.; Loong, H.H.F. PPARgamma agonists sensitize PTEN-deficient resistant lung cancer cells to EGFR tyrosine kinase inhibitors by inducing autophagy. Eur. J. Pharmacol. 2018, 823, 19–26. [Google Scholar] [CrossRef]
  398. Gkountakos, A.; Sartori, G.; Falcone, I.; Piro, G.; Ciuffreda, L.; Carbone, C.; Tortora, G.; Scarpa, A.; Bria, E.; Milella, M.; et al. PTEN in Lung Cancer: Dealing with the Problem, Building on New Knowledge and Turning the Game Around. Cancers 2019, 11, 1141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  399. Li, M.Y.; Yuan, H.; Ma, L.T.; Kong, A.W.; Hsin, M.K.; Yip, J.H.; Underwood, M.J.; Chen, G.G. Roles of peroxisome proliferator-activated receptor-alpha and -gamma in the development of non-small cell lung cancer. Am. J. Respir. Cell Mol. Biol. 2010, 43, 674–683. [Google Scholar] [CrossRef] [PubMed]
  400. Freitas, R.D.S.; Campos, M.M. Protective Effects of Omega-3 Fatty Acids in Cancer-Related Complications. Nutrients 2019, 11, 945. [Google Scholar] [CrossRef] [Green Version]
  401. Yang, P.; Cartwright, C.; Chan, D.; Ding, J.; Felix, E.; Pan, Y.; Pang, J.; Rhea, P.; Block, K.; Fischer, S.M.; et al. Anticancer activity of fish oils against human lung cancer is associated with changes in formation of PGE2 and PGE3 and alteration of Akt phosphorylation. Mol. Carcinog 2014, 53, 566–577. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  402. Zajdel, A.; Wilczok, A.; Tarkowski, M. Toxic effects of n-3 polyunsaturated fatty acids in human lung A549 cells. Toxicol. Vitr. Int. J. Publ. Assoc. BIBRA 2015, 30 Pt B, 486–491. [Google Scholar] [CrossRef]
  403. Skácel, J.; Slusher, B.S.; Tsukamoto, T. Small Molecule Inhibitors Targeting Biosynthesis of Ceramide, the Central Hub of the Sphingolipid Network. J. Med. Chem. 2021, 64, 279–297. [Google Scholar] [CrossRef] [PubMed]
  404. Peterschmitt, M.J.; Crawford, N.P.S.; Gaemers, S.J.M.; Ji, A.J.; Sharma, J.; Pham, T.T. Pharmacokinetics, Pharmacodynamics, Safety, and Tolerability of Oral Venglustat in Healthy Volunteers. Clin. Pharmacol. Drug Dev. 2021, 10, 86–98. [Google Scholar] [CrossRef] [PubMed]
  405. Loh, K.C.; Leong, W.-I.; Carlson, M.E.; Oskouian, B.; Kumar, A.; Fyrst, H.; Zhang, M.; Proia, R.L.; Hoffman, E.P.; Saba, J.D. Sphingosine-1-Phosphate Enhances Satellite Cell Activation in Dystrophic Muscles through a S1PR2/STAT3 Signaling Pathway. PLoS ONE 2012, 7, e37218. [Google Scholar] [CrossRef]
  406. Hernández-Tiedra, S.; Fabriàs, G.; Dávila, D.; Salanueva, Í.J.; Casas, J.; Montes, L.R.; Antón, Z.; García-Taboada, E.; Salazar-Roa, M.; Lorente, M.; et al. Dihydroceramide accumulation mediates cytotoxic autophagy of cancer cells via autolysosome destabilization. Autophagy 2016, 12, 2213–2229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  407. Lai, M.; Amato, R.; La Rocca, V.; Bilgin, M.; Freer, G.; Spezia, P.; Quaranta, P.; Piomelli, D.; Pistello, M. Acid ceramidase controls apoptosis and increases autophagy in human melanoma cells treated with doxorubicin. Sci. Rep. 2021, 11, 11221. [Google Scholar] [CrossRef]
  408. Realini, N.; Palese, F.; Pizzirani, D.; Pontis, S.; Basit, A.; Bach, A.; Ganesan, A.; Piomelli, D. Acid Ceramidase in Melanoma: EXPRESSION, LOCALIZATION, AND EFFECTS OF PHARMACOLOGICAL INHIBITION. J. Biol. Chem. 2016, 291, 2422–2434. [Google Scholar] [CrossRef] [Green Version]
  409. Lai, M.; La Rocca, V.; Amato, R.; Freer, G.; Pistello, M. Sphingolipid/Ceramide Pathways and Autophagy in the Onset and Progression of Melanoma: Novel Therapeutic Targets and Opportunities. Int. J. Mol. Sci. 2019, 20, 3436. [Google Scholar] [CrossRef] [Green Version]
  410. Costabile, G.; Della Pepa, G.; Salamone, D.; Luongo, D.; Naviglio, D.; Brancato, V.; Cavaliere, C.; Salvatore, M.; Cipriano, P.; Vitale, M.; et al. Reduction of De Novo Lipogenesis Mediates Beneficial Effects of Isoenergetic Diets on Fatty Liver: Mechanistic Insights from the MEDEA Randomized Clinical Trial. Nutrients 2022, 14, 2178. [Google Scholar] [CrossRef] [PubMed]
  411. Sinha, R.A.; Singh, B.K.; Zhou, J.; Xie, S.; Farah, B.L.; Lesmana, R.; Ohba, K.; Tripathi, M.; Ghosh, S.; Hollenberg, A.N.; et al. Loss of ULK1 increases RPS6KB1-NCOR1 repression of NR1H/LXR-mediated Scd1 transcription and augments lipotoxicity in hepatic cells. Autophagy 2017, 13, 169–186. [Google Scholar] [CrossRef] [Green Version]
  412. Loucif, H.; Dagenais-Lussier, X.; Beji, C.; Cassin, L.; Jrade, H.; Tellitchenko, R.; Routy, J.P.; Olagnier, D.; van Grevenynghe, J. Lipophagy confers a key metabolic advantage that ensures protective CD8A T-cell responses against HIV-1. Autophagy 2021, 17, 3408–3423. [Google Scholar] [CrossRef]
  413. Li, W.; Bai, H.; Liu, S.; Cao, D.; Wu, H.; Shen, K.; Tai, Y.; Yang, J. Targeting stearoyl-CoA desaturase 1 to repress endometrial cancer progression. Oncotarget 2018, 9, 12064–12078. [Google Scholar] [CrossRef] [Green Version]
  414. Chen, L.; Ren, J.; Yang, L.; Li, Y.; Fu, J.; Li, Y.; Tian, Y.; Qiu, F.; Liu, Z.; Qiu, Y. Stearoyl-CoA desaturase-1 mediated cell apoptosis in colorectal cancer by promoting ceramide synthesis. Sci. Rep. 2016, 6, 19665. [Google Scholar] [CrossRef] [Green Version]
  415. Piao, C.; Cui, X.; Zhan, B.; Li, J.; Li, Z.; Li, Z.; Liu, X.; Bi, J.; Zhang, Z.; Kong, C. Inhibition of stearoyl CoA desaturase-1 activity suppresses tumour progression and improves prognosis in human bladder cancer. J. Cell. Mol. Med. 2019, 23, 2064–2076. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  416. Ono, A.; Sano, O.; Kazetani, K.I.; Muraki, T.; Imamura, K.; Sumi, H.; Matsui, J.; Iwata, H. Feedback activation of AMPK-mediated autophagy acceleration is a key resistance mechanism against SCD1 inhibitor-induced cell growth inhibition. PLoS ONE 2017, 12, e0181243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  417. Huang, G.M.; Jiang, Q.H.; Cai, C.; Qu, M.; Shen, W. SCD1 negatively regulates autophagy-induced cell death in human hepatocellular carcinoma through inactivation of the AMPK signaling pathway. Cancer Lett. 2015, 358, 180–190. [Google Scholar] [CrossRef]
  418. Drake, L.E.; Springer, M.Z.; Poole, L.P.; Kim, C.J.; Macleod, K.F. Expanding perspectives on the significance of mitophagy in cancer. Semin. Cancer Biol. 2017, 47, 110–124. [Google Scholar] [CrossRef] [PubMed]
  419. Zhang, J.; Ney, P.A. Role of BNIP3 and NIX in cell death, autophagy, and mitophagy. Cell Death Differ. 2009, 16, 939–946. [Google Scholar] [CrossRef] [Green Version]
  420. Sun, J.L.; He, X.S.; Yu, Y.H.; Chen, Z.C. Expression and structure of BNIP3L in lung cancer. Ai Zheng = Aizheng = Chin. J. Cancer 2004, 23, 8–14. [Google Scholar]
  421. Chourasia, A.H.; Tracy, K.; Frankenberger, C.; Boland, M.L.; Sharifi, M.N.; Drake, L.E.; Sachleben, J.R.; Asara, J.M.; Locasale, J.W.; Karczmar, G.S.; et al. Mitophagy defects arising from BNip3 loss promote mammary tumor progression to metastasis. EMBO Rep. 2015, 16, 1145–1163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  422. Chourasia, A.H.; Macleod, K.F. Tumor suppressor functions of BNIP3 and mitophagy. Autophagy 2015, 11, 1937–1938. [Google Scholar] [CrossRef] [Green Version]
  423. Picchio, M.C.; Martin, E.S.; Cesari, R.; Calin, G.A.; Yendamuri, S.; Kuroki, T.; Pentimalli, F.; Sarti, M.; Yoder, K.; Kaiser, L.R.; et al. Alterations of the tumor suppressor gene Parkin in non-small cell lung cancer. Clin. Cancer Res. 2004, 10, 2720–2724. [Google Scholar] [CrossRef] [Green Version]
  424. Zhang, Z.L.; Wang, N.N.; Ma, Q.L.; Chen, Y.; Yao, L.; Zhang, L.; Li, Q.S.; Shi, M.H.; Wang, H.F.; Ying, Z. Somatic and germline mutations in the tumor suppressor gene PARK2 impair PINK1/Parkin-mediated mitophagy in lung cancer cells. Acta Pharmacol. Sin. 2020, 41, 93–100. [Google Scholar] [CrossRef] [PubMed]
  425. Bernardini, J.P.; Lazarou, M.; Dewson, G. Parkin and mitophagy in cancer. Oncogene 2016, 36, 1315. [Google Scholar] [CrossRef] [PubMed]
  426. Liu, J.; Zhang, C.; Zhao, Y.; Yue, X.; Wu, H.; Huang, S.; Chen, J.; Tomsky, K.; Xie, H.; Khella, C.A.; et al. Parkin targets HIF-1α for ubiquitination and degradation to inhibit breast tumor progression. Nat. Commun. 2017, 8, 1823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Autophagy is a self-eating mechanism occurring in cells. (A) Different steps in the autophagy pathway. Cargos are first engulfed by autophagosomes, which then fuse with lysosomes. Lysosomal enzymes degrade the cargo resulting in the recycling of the cargo’s building blocks. (B) Selective autophagy. Autophagy can selectively target and degrade specific cargo using different receptor proteins.
Figure 1. Autophagy is a self-eating mechanism occurring in cells. (A) Different steps in the autophagy pathway. Cargos are first engulfed by autophagosomes, which then fuse with lysosomes. Lysosomal enzymes degrade the cargo resulting in the recycling of the cargo’s building blocks. (B) Selective autophagy. Autophagy can selectively target and degrade specific cargo using different receptor proteins.
Cancers 15 02195 g001
Figure 2. (A) Glycolysis: This includes 10 consecutive enzymatic reactions in the conversion of glucose into two molecules of pyruvate, which connect with other metabolic pathways. Glucose is phosphorylated to glucose 6-phosphate by the enzyme hexokinase (HK). Glucose 6-phosphate is turned into fructose 6-phosphate by glucose phosphate isomerase (GPI). Phosphofructokinase-1 (PFK) catalyzes the conversion of fructose 6-phosphate and ATP to fructose 1,6-bisphosphate and ADP. 6-phosphofructo2-kinase/fructose 2,6-bisphosphatase (PFKFB), a family of bifunctional enzymes that control the levels of fructose 2,6-bisphosphate (PFKFB3 and PFKFB4 are the two main isoenzymes of this family) (PFK and PFKFB are the two major regulatory enzymes in glycolysis). Subsequently, the aldolase enzyme catalyzes a reversible reaction in which fructose 1,6-bisphosphate converts into glyceraldehyde-3-phosphate (GLAP) and dihydroxyacetone phosphate (DHAP). Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) produces 1,3-bisphosphoglycerate. Next, phosphoglycerate kinase (PGK) forms ATP and 3-phosphoglycerate. Phosphoglycerate mutase (PGM) isomerizes 3-phosphoglycerate into 2-phosphoglycerate. Next, enolase converts 2-phosphoglycerate to phosphoenolpyruvate, which is phosphorylated to form a molecule of pyruvate and a molecule of ATP via pyruvate kinase (PK). Finally, lactate dehydrogenase (LDH) catalyzes the conversion of pyruvate to lactate. (B) Krebs cycle: The TCA cycle starts with the combination of acetyl-CoA, generated from fatty acids, amino acids, or pyruvate oxidation, with oxaloacetate (OAA) to produce citrate. Citrate is converted into isocitrate. The cycle continues with two consecutive oxidative decarboxylations in which isocitrate is converted into the α-ketoglutarate (α-KG) and the succinyl-CoA concomitantly, producing two CO2 and two NADH. Succinyl-CoA coverts into succinate, coupled with the release of ATP. Subsequently, fumarate is formed by the oxidation of succinate. During this step, two hydrogen atoms are transferred to FAD, producing 2FADH. Next, fumarate is converted into malate and further into OAA, which combines with another acetyl-CoA, continuing the cycle [51]. (C) De novo lipogenesis C: In this simplified scheme, pyruvate from glycolysis feeds acetyl-CoA to the TCA cycle. Citrate from the TCA cycle is converted to acetyl-CoA in the cytosol by ATP citrate lyase (ACLY). Acetyl-CoA from the TCA cycle and other sources is subsequently converted to complicated fatty acids by a series of enzymes including ATP-citrate lyase (ACLY), acetyl-CoA carboxylases 1 (ACC1), fatty acid synthase (FASN), and stearoyl-CoA desaturase-1 (SCD1). (D) Fatty acid β-oxidation: Each cycle leads to the formation of acetyl-CoA, nicotinamide adenine dinucleotide (NADH), and flavin adenine dinucleotide (FADH 2). The electron carriers NADH and FADH2 are used by the mitochondrial respiratory chain to generate ATP. (E) Triacylglycerol synthesis: This reaction occurs at the surface of the endoplasmic reticulum (ER) bilayer membrane. The first step in this pathway is the acylation of glycerol-3-phosphate by glycerol-3-phosphate acyltransferase (GPAT), producing lysophosphatidic acid. This is followed by further acylation by LPA acyltransferase (LPAAT) and dephosphorylation by phosphatidic acid phosphorylase (PAP) to yield diacylglycerol (DAG). The final step is converting 1,2-diacylglycerol into triacylglycerol (TAG), which is catalyzed by diacylglycerol acyltransferase (DGAT).
Figure 2. (A) Glycolysis: This includes 10 consecutive enzymatic reactions in the conversion of glucose into two molecules of pyruvate, which connect with other metabolic pathways. Glucose is phosphorylated to glucose 6-phosphate by the enzyme hexokinase (HK). Glucose 6-phosphate is turned into fructose 6-phosphate by glucose phosphate isomerase (GPI). Phosphofructokinase-1 (PFK) catalyzes the conversion of fructose 6-phosphate and ATP to fructose 1,6-bisphosphate and ADP. 6-phosphofructo2-kinase/fructose 2,6-bisphosphatase (PFKFB), a family of bifunctional enzymes that control the levels of fructose 2,6-bisphosphate (PFKFB3 and PFKFB4 are the two main isoenzymes of this family) (PFK and PFKFB are the two major regulatory enzymes in glycolysis). Subsequently, the aldolase enzyme catalyzes a reversible reaction in which fructose 1,6-bisphosphate converts into glyceraldehyde-3-phosphate (GLAP) and dihydroxyacetone phosphate (DHAP). Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) produces 1,3-bisphosphoglycerate. Next, phosphoglycerate kinase (PGK) forms ATP and 3-phosphoglycerate. Phosphoglycerate mutase (PGM) isomerizes 3-phosphoglycerate into 2-phosphoglycerate. Next, enolase converts 2-phosphoglycerate to phosphoenolpyruvate, which is phosphorylated to form a molecule of pyruvate and a molecule of ATP via pyruvate kinase (PK). Finally, lactate dehydrogenase (LDH) catalyzes the conversion of pyruvate to lactate. (B) Krebs cycle: The TCA cycle starts with the combination of acetyl-CoA, generated from fatty acids, amino acids, or pyruvate oxidation, with oxaloacetate (OAA) to produce citrate. Citrate is converted into isocitrate. The cycle continues with two consecutive oxidative decarboxylations in which isocitrate is converted into the α-ketoglutarate (α-KG) and the succinyl-CoA concomitantly, producing two CO2 and two NADH. Succinyl-CoA coverts into succinate, coupled with the release of ATP. Subsequently, fumarate is formed by the oxidation of succinate. During this step, two hydrogen atoms are transferred to FAD, producing 2FADH. Next, fumarate is converted into malate and further into OAA, which combines with another acetyl-CoA, continuing the cycle [51]. (C) De novo lipogenesis C: In this simplified scheme, pyruvate from glycolysis feeds acetyl-CoA to the TCA cycle. Citrate from the TCA cycle is converted to acetyl-CoA in the cytosol by ATP citrate lyase (ACLY). Acetyl-CoA from the TCA cycle and other sources is subsequently converted to complicated fatty acids by a series of enzymes including ATP-citrate lyase (ACLY), acetyl-CoA carboxylases 1 (ACC1), fatty acid synthase (FASN), and stearoyl-CoA desaturase-1 (SCD1). (D) Fatty acid β-oxidation: Each cycle leads to the formation of acetyl-CoA, nicotinamide adenine dinucleotide (NADH), and flavin adenine dinucleotide (FADH 2). The electron carriers NADH and FADH2 are used by the mitochondrial respiratory chain to generate ATP. (E) Triacylglycerol synthesis: This reaction occurs at the surface of the endoplasmic reticulum (ER) bilayer membrane. The first step in this pathway is the acylation of glycerol-3-phosphate by glycerol-3-phosphate acyltransferase (GPAT), producing lysophosphatidic acid. This is followed by further acylation by LPA acyltransferase (LPAAT) and dephosphorylation by phosphatidic acid phosphorylase (PAP) to yield diacylglycerol (DAG). The final step is converting 1,2-diacylglycerol into triacylglycerol (TAG), which is catalyzed by diacylglycerol acyltransferase (DGAT).
Cancers 15 02195 g002
Figure 3. Autophagy is regulated by multiple signaling pathways, which creates a complex interaction system. The glycolysis-related enzymes and glycolysis metabolic intermediates (AF) are involved in the regulation of autophagy, acting as a double-edged sword to sustain cancer cell survival under conditions of energy shortage.
Figure 3. Autophagy is regulated by multiple signaling pathways, which creates a complex interaction system. The glycolysis-related enzymes and glycolysis metabolic intermediates (AF) are involved in the regulation of autophagy, acting as a double-edged sword to sustain cancer cell survival under conditions of energy shortage.
Cancers 15 02195 g003
Figure 4. Hypoxia-induced autophagy (left): Hypoxic stress leads to the accumulation of HIF-1a. BNIP3 and BNIP3L, gene products targeted by HIF, are transcriptionally upregulated and compete with the Bcl-XL and Bcl-2 complex. This competition releases BECLIN1 from the complex, leading to the activation of autophagic machinery through the recruitment of several autophagic proteins including ATG5 and LC3. Acid-induced autophagy (right): Upon MCT1-mediated entry, lactate and NAD+ are converted into pyruvate, NADH, and H+ by LDHB. Then, the protons generated by LDHB can promote lysosomal acidification and autophagy.
Figure 4. Hypoxia-induced autophagy (left): Hypoxic stress leads to the accumulation of HIF-1a. BNIP3 and BNIP3L, gene products targeted by HIF, are transcriptionally upregulated and compete with the Bcl-XL and Bcl-2 complex. This competition releases BECLIN1 from the complex, leading to the activation of autophagic machinery through the recruitment of several autophagic proteins including ATG5 and LC3. Acid-induced autophagy (right): Upon MCT1-mediated entry, lactate and NAD+ are converted into pyruvate, NADH, and H+ by LDHB. Then, the protons generated by LDHB can promote lysosomal acidification and autophagy.
Cancers 15 02195 g004
Figure 5. Crosstalk between lipids and autophagy. Lipids can modulate autophagy through different mechanisms.
Figure 5. Crosstalk between lipids and autophagy. Lipids can modulate autophagy through different mechanisms.
Cancers 15 02195 g005
Figure 6. Mechanism of action of autophagy regulation upon HMGCR inhibition by statins. In addition, the mechanism by which FASN can modulate autophagy during cancer is shown. ACLY: ATP citrate lyase; ACAT: acetyl-CoA acetyltransferase; HMGCS: 3-hydroxy-3-methylglutaryl-CoA synthase; HMGCR: 3-hydroxy-3-methylglutaryl-CoA reductase; ACC: acetyl-CoA carboxylase; FASN: fatty acid synthase; SCD: stearoyl-CoA desaturase; MUFA: monounsaturated fatty acid; PUFA: polyunsaturated fatty acid.
Figure 6. Mechanism of action of autophagy regulation upon HMGCR inhibition by statins. In addition, the mechanism by which FASN can modulate autophagy during cancer is shown. ACLY: ATP citrate lyase; ACAT: acetyl-CoA acetyltransferase; HMGCS: 3-hydroxy-3-methylglutaryl-CoA synthase; HMGCR: 3-hydroxy-3-methylglutaryl-CoA reductase; ACC: acetyl-CoA carboxylase; FASN: fatty acid synthase; SCD: stearoyl-CoA desaturase; MUFA: monounsaturated fatty acid; PUFA: polyunsaturated fatty acid.
Cancers 15 02195 g006
Figure 7. SPHK1 and FAT can regulate autophagy during cancer. A simplified scheme of sphingolipid metabolism and turnover is also shown.
Figure 7. SPHK1 and FAT can regulate autophagy during cancer. A simplified scheme of sphingolipid metabolism and turnover is also shown.
Cancers 15 02195 g007
Figure 8. (A): Overall Survival Kaplan–Meier Estimate of LUAD cancer patients with BCL2L1 gene amplification (red) or without (blue). (B): Disease/Progression-free Kaplan–Meier Estimate in LUAD cancer patients with BCL2L1 gene amplification (red) and without (blue). Graph from Oncomine.org site using TCGA cancer data [327,328,329].
Figure 8. (A): Overall Survival Kaplan–Meier Estimate of LUAD cancer patients with BCL2L1 gene amplification (red) or without (blue). (B): Disease/Progression-free Kaplan–Meier Estimate in LUAD cancer patients with BCL2L1 gene amplification (red) and without (blue). Graph from Oncomine.org site using TCGA cancer data [327,328,329].
Cancers 15 02195 g008
Figure 9. Schematic representation highlighting the metabolic pathways upregulated in cancer. In tumour cells, the uptake of glucose and aerobic glycolysis (the Warburg effect) are upregulated for rapid ATP production. The abundance of lactate that is produced is excreted into the extracellular space/tumour microenvironment. Some pyruvate is shunted into the mitochondrion for OXPHOS. The excess glucose feeds into biosynthesis, such as nucleotide, protein, and lipid synthesis, which supports uncontrolled proliferation. Altered glucose metabolism is linked to altered lipid metabolism in tumour cells, as glucose is the major substrate for lipid synthesis. Both endogenous fatty acid synthesis and fatty acid uptake are upregulated in tumour cells. The fatty acids can be subsequently used for ATP production through β-oxidation, which is also upregulated in tumour cells and supports uncontrolled growth and proliferation. ATP, adenosine triphosphate; CPT1, carnitine palmitoyl transferase 1; ETC, electron transport chain; FAO, fatty acid oxidation; FAS, fatty acid synthesis; GLUT, glucose transporter; MCT4, monocarboxylate transporter 4; PPP, pentose phosphate pathway; TCA cycle, tricarboxylic acid cycle.
Figure 9. Schematic representation highlighting the metabolic pathways upregulated in cancer. In tumour cells, the uptake of glucose and aerobic glycolysis (the Warburg effect) are upregulated for rapid ATP production. The abundance of lactate that is produced is excreted into the extracellular space/tumour microenvironment. Some pyruvate is shunted into the mitochondrion for OXPHOS. The excess glucose feeds into biosynthesis, such as nucleotide, protein, and lipid synthesis, which supports uncontrolled proliferation. Altered glucose metabolism is linked to altered lipid metabolism in tumour cells, as glucose is the major substrate for lipid synthesis. Both endogenous fatty acid synthesis and fatty acid uptake are upregulated in tumour cells. The fatty acids can be subsequently used for ATP production through β-oxidation, which is also upregulated in tumour cells and supports uncontrolled growth and proliferation. ATP, adenosine triphosphate; CPT1, carnitine palmitoyl transferase 1; ETC, electron transport chain; FAO, fatty acid oxidation; FAS, fatty acid synthesis; GLUT, glucose transporter; MCT4, monocarboxylate transporter 4; PPP, pentose phosphate pathway; TCA cycle, tricarboxylic acid cycle.
Cancers 15 02195 g009
Table 1. Glycolysis-related enzymes involved in the regulation of autophagy.
Table 1. Glycolysis-related enzymes involved in the regulation of autophagy.
Enzyme NameEnzyme IsoformType of CancerMoleculesRoles in Glycolysis and AutophagyRef.
PFKFBsPFKFB3Colon
Adenocarcinoma
SiRNA and 3POKnockdown of PFKFB3 induces autophagy.[123]
GAPDH-Pancreatic AdenocarcinomaGenipin and Everolimus/3-MAInduces Caspase-mediated apoptosis and inhibits BECLIN1-mediated autophagy.[136]
PGKPGK1Breast CancershRNAReduces the Warburg effect and reduces invasion.[187]
PKPKM2AMLSiRNAReduces BECLIN1-mediated autophagy.[150]
GAPDH-Gastric CancermiR-133bReduces glycolysis and induces autophagy by switching the PKM isoform expression from PKM2 to PKM1.[188]
LDHLDHABreast CancerSiRNARegulates BECLIN1-mediated autophagy.[160]
PKLDHBColorectal carcinomaGW5074Reduces autophagy by SIRT5 knockout or the inhibition of LDHB acetylation at Lys329.[189]
HIFHIF-1αHuman Liver and Glioma CancershRNA- lincRNA-p21Reduces glycolysis and inhibits autophagy by downregulating the HIF-1 protein.[185]
LDHLDHACervical CancerSiRNA and
2-deoxyD-glucose or koningic acid
Inhibits glycolysis and reduces autophagy.[72]
Lactate-Colorectal CarcinomaOsimertinibInduces autophagy by regulating the expression of MCT1 and activating LKB1/AMPK signaling.[178]
HIFHIF-1-αBreast Cancer7acc1Induces autophagy via the inhibition of MCT4.[179]
HKHK2Colon, breast, prostatemiR-143Induces autophagy and induces proliferation.[91,92,93,94,95,96,97]
Table 2. The effect of enzymes involved in lipid metabolism on different steps of autophagy.
Table 2. The effect of enzymes involved in lipid metabolism on different steps of autophagy.
EnzymeFunctionEffect on AutophagyCancer TypeRef.
Fatty Acid Synthase (FASN)Synthesizes palmitate from acetyl-CoA and malonyl-CoA
(a)
Activation of mTOR and subsequent downregulation of TFEB
InhibitionAcute Myeloid Leukemia[196]
(b)
Activation of PI3K-Akt
InhibitionOsteosarcoma[198]
(c)
MAPK inactivation
InhibitionColorectal Cancer[199]
3-Hydroxy-3-MethylGlutaryl-CoA Reductase (HMGCR)Synthesizes cholesterol, geranylgeranyl pyrophosphate and farnesyl pyrophosphate, etc.
(a)
Cholesterol-dependent inhibition of Akt-mTOR upon HMGCR inhibition by Lovastatin
InductionAcute Lymphocytic Leukemia[204]
(b)
Geranylgeranyl pyrophosphate-dependent upregulation of LC3-II upon HMGCR inhibition by Atorvastatin
InductionProstate Cancer, Rhabdomyosarcoma[207]
Sphingosine-1-Kinase (SPHK1)Synthesizes sphingosine 1 phosphate (S1P) from sphingosine
(a)
Promotes biogenesis of autophagosomes by modulating the expression levels of LC3
InductionNeuroblastoma[223]
(b)
Increases the formation of LC3-positive autophagosomes
InductionBreast Cancer[224]
(c)
Activates SphK1/ERK/p ERK leading to the inactivation of mTOR, and elevated expression of ATG5 and ULK1
InductionColon Cancer[225]
(d)
Enhances lysosomal degradation, increasing the ubiquitination of BECLIN1
InductionHepatocellular Carcinoma[226]
(e)
Upregulation of LC3 expression
InductionGastric Cancer[227]
Fatty Acid Translocase (FAT/CD36)Long-chain fatty acid uptake
(a)
Phosphorylation of BECLIN1 and ULK1
InhibitionHepatocellular Carcinoma[229]
Table 3. Regulation of autophagy by lipid metabolites during cancer.
Table 3. Regulation of autophagy by lipid metabolites during cancer.
Lipid MetaboliteEffect on AutophagyCancer TypeRef.
Phosphatidic AcidInduction of autophagosome curvature, inhibiting mTORInductionBreast Cancer[232]
CeramideInhibition of Akt/mTOR; Upregulation of BECLIN1 expressionInductionBreast Cancer, Neuroblastoma[237,239]
Upregulation of BECLIN1 expressionInductionColon Cancer[239]
Dissociation of BCL2 from BECLIN1InductionCervical Cancer[238]
Sphingosine 1-phosphate (S1P)mTOR inhibitionInductionProstate Cancer[242]
mTOR inhibitionInductionBreast Cancer[224,241]
Peroxisome Proliferator-Activating Receptors (PPARs)PTEN activation; mTOR inhibitionInductionColorectal Cancer[249]
Increase in lysosomal acidificationInductionBreast Cancer[250]
Palmitic AcidActivation of PKC, increasing the diacylglycerol levelsInductionHepatocarcinoma[254]
Docosahexaenoic acid (DHA)AMPK activation resulting in the inhibition of mTORInductionp53-wild type prostate cancer[264]
Producing mitochondrial reactive oxygen species (ROS) resulting in Akt and mTOR inhibitionInductionp53-mutant prostate cancer[264]
Upregulation of LC3 and ATG14 expressionInductionColon Cancer[267]
mTOR inhibitionInductionGlioblastoma[264]
mTOR inhibitionInductionCervix Carcinoma[264]
Docosahexaenoyl ethanolamine (DHEA)Upregulation of PPARγ, which upregulates PTEN and, consequently, inhibits Akt-mTOR; promotes BCL2 phosphorylation and, therefore, its dissociation from BECLIN1InductionBreast Cancer[270]
CholesterolmTOR inhibitionInductionAcute Lymphocytic Leukemia, Chronic Lymphocytic Leukemia[204]
Table 4. The effect of enzymes involved in glycolysis on autophagy in lung cancer.
Table 4. The effect of enzymes involved in glycolysis on autophagy in lung cancer.
Enzyme NameEnzyme IsoformType of CancerMoleculesRoles in Glycolysis and AutophagyRef.
HKHK2NSCLC2-DGInduce autophagy and apoptosis by increasing LC3II and cleaved caspase3 protein levels.[373]
NSCLCSiRNAInhibit cancer cell proliferation and tumorigenesis through decreasing glycolysis and HK2 expression at both the mRNA and protein levels.[374]
NSCLCmiR-143Inhibit glycolysis and autophagy by the knockdown of HK2 and Atg2B.[375]
PFKFBsPFKFB3NSCLCErlotinib drivesReduce the EGF-mediated increase in glycolysis.[376]
PFKFB4SCLCIbrutinibInhibit autophagy by targeting Erk.[377]
GAPDH-NSCLCSiRNAReduce glycolysis and cease tumor cell proliferation.[131]
PKPKM2NSCLCpshRNAInduce autophagy and apoptosis.[378]
LDHLDHALung and Breast CancerWZB117 and ShRNAReduce glycolysis and the induction of apoptosis.[379]
HIFHIF-1NSCLCSP600125Inhibit hypoxia-induced autophagy by targeting the BECLIN1 protein.[379]
NSCLCChloramphenicolInhibit hypoxia-induced autophagy.[380]
Lactate-NSCLCSiRNAReduce pH-stimulated autophagy by the knockdown of the BECLIN1 protein and ER stress marker protein (GRP78).[365]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alizadeh, J.; Kavoosi, M.; Singh, N.; Lorzadeh, S.; Ravandi, A.; Kidane, B.; Ahmed, N.; Mraiche, F.; Mowat, M.R.; Ghavami, S. Regulation of Autophagy via Carbohydrate and Lipid Metabolism in Cancer. Cancers 2023, 15, 2195. https://doi.org/10.3390/cancers15082195

AMA Style

Alizadeh J, Kavoosi M, Singh N, Lorzadeh S, Ravandi A, Kidane B, Ahmed N, Mraiche F, Mowat MR, Ghavami S. Regulation of Autophagy via Carbohydrate and Lipid Metabolism in Cancer. Cancers. 2023; 15(8):2195. https://doi.org/10.3390/cancers15082195

Chicago/Turabian Style

Alizadeh, Javad, Mahboubeh Kavoosi, Navjit Singh, Shahrokh Lorzadeh, Amir Ravandi, Biniam Kidane, Naseer Ahmed, Fatima Mraiche, Michael R. Mowat, and Saeid Ghavami. 2023. "Regulation of Autophagy via Carbohydrate and Lipid Metabolism in Cancer" Cancers 15, no. 8: 2195. https://doi.org/10.3390/cancers15082195

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop