Next Article in Journal
Synaptic Activity Regulates Mitochondrial Iron Metabolism to Enhance Neuronal Bioenergetics
Previous Article in Journal
Understanding the Role of Oxidative Stress, Neuroinflammation and Abnormal Myelination in Excessive Aggression Associated with Depression: Recent Input from Mechanistic Studies
Previous Article in Special Issue
MUC1 Tissue Expression and Its Soluble Form CA15-3 Identify a Clear Cell Renal Cell Carcinoma with Distinct Metabolic Profile and Poor Clinical Outcome
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Emerging Hallmarks of Metabolic Reprogramming in Prostate Cancer

by
Francesco Lasorsa
1,
Nicola Antonio di Meo
1,
Monica Rutigliano
1,
Matteo Ferro
2,
Daniela Terracciano
3,
Octavian Sabin Tataru
4,
Michele Battaglia
1,
Pasquale Ditonno
1 and
Giuseppe Lucarelli
1,*
1
Urology, Andrology and Kidney Transplantation Unit, Department of Precision and Regenerative Medicine and Ionian Area, University of Bari “Aldo Moro”, 70124 Bari, Italy
2
Division of Urology, European Institute of Oncology, IRCCS, 20141 Milan, Italy
3
Department of Translational Medical Sciences, University of Naples “Federico II”, 80131 Naples, Italy
4
The Institution Organizing University Doctoral Studies (I.O.S.U.D.), George Emil Palade University of Medicine, Pharmacy, Sciences and Technology, 540142 Târgu Mureș, Romania
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(2), 910; https://doi.org/10.3390/ijms24020910
Submission received: 11 December 2022 / Revised: 30 December 2022 / Accepted: 1 January 2023 / Published: 4 January 2023
(This article belongs to the Special Issue Novel Insights into Molecular Biology of Urological Cancers 2.0)

Abstract

:
Prostate cancer (PCa) is the most common male malignancy and the fifth leading cause of cancer death in men worldwide. Prostate cancer cells are characterized by a hybrid glycolytic/oxidative phosphorylation phenotype determined by androgen receptor signaling. An increased lipogenesis and cholesterogenesis have been described in PCa cells. Many studies have shown that enzymes involved in these pathways are overexpressed in PCa. Glutamine becomes an essential amino acid for PCa cells, and its metabolism is thought to become an attractive therapeutic target. A crosstalk between cancer and stromal cells occurs in the tumor microenvironment because of the release of different cytokines and growth factors and due to changes in the extracellular matrix. A deeper insight into the metabolic changes may be obtained by a multi-omic approach integrating genomics, transcriptomics, metabolomics, lipidomics, and radiomics data.

1. Introduction

Prostate cancer (PCa) is the most common male malignancy and the fifth leading cause of cancer death in men worldwide [1,2]. PCa may be suspected based on digital rectal examination (DRE) and/or prostate-specific antigen (PSA) levels; however, PSA levels may increase in many non-malignant clinical conditions. An early diagnosis of PCa reduces the cost for disease management and improves treatment efficacy and patients’ quality of life. In this scenario, the discovery of novel biomarkers that may increase the sensitivity and specificity for PCa diagnosis and prognosis (i.e., Prostate Cancer Antigen 3, the TMPRSS2-ERG gene fusion, Spondin 2, and circulating tumor cells) is of utmost importance [3,4,5,6]. Many studies have already marked the association between metabolic disorders and an increased risk of PCa, as well as for other urologic tumors [7,8]. Obesity is a well-known risk factor for PCa and an increased BMI is associated with aggressive disease. Obesity triggers a low-grade inflammation (metaflammation), which is thought to be a milestone for the pathophysiology of several chronic diseases as for cancer. As for diabetes, it was demonstrated that TNF and TLR signaling reduce insulin secretion and effects. This crosstalk between metabolism and the immune system has been shown to be evolutionarily conserved. Further future studies will be necessary to unravel molecular mechanisms linking aberrant metabolism and cancer [9,10]. Nevertheless, several previous studies have already stated that obese patients diagnosed with prostate cancer experience worse outcomes. They are at an increased risk of disease recurrence, exacerbated treatment-related adverse effects, the development of obesity-related comorbidities, earlier metastatic disease, and higher prostate cancer mortality. The increased inflammatory environment together with metabolic irregularities are commonly postulated, although physiological mechanisms linking obesity with patients’ poor prognosis remain unclear [11,12,13,14,15,16,17,18,19]. A recent study explored the metabolic and genetic signatures of obese individuals. The authors analysed metabolome perturbations in association with BMI and they found out that metabolite signatures (mBMI) better stratify patients’ health risk [20]. Metabolomics is referred to as the analysis of metabolic products in cells, tissues, organs, and the organism, which may be used to identify new biomarkers for cancer diagnosis and management [21]. The multi-omics approach including the study of cancer cell metabolism represents a powerful strategy for a better comprehension of cancer progression [22].

2. Metabolism of Normal and Tumor Prostate Cells

Prostate cells are characterized by an increased production of citrate and polyamines, which are physiological components of prostatic fluid. The high amount of citrate produced by prostate cells depends on a specific metabolic pathway not described in other human cells. In eukaryotic cells, glucose is transformed via glycolysis into pyruvate and then oxidized in acetyl-CoA in the mitochondria. Eventually, acetyl-CoA reacts with oxaloacetate, producing citrate, which enters the Krebs cycle. In prostate cells, zinc ion inhibits the mitochondrial enzyme aconitase ACO2 (first step of citrate oxidation), thus blocking the Krebs cycle and leading to citrate accumulation [23]. Citrate is then secreted in the semen where it regulates pH, ions homeostasis, and the coagulation/de-coagulation process. Citrate has high affinity for zinc, magnesium, and calcium, and it is the main regulator of ionized calcium concentration in semen. Variations in citrate secretion by prostate epithelial cells alter male fertility since calcium concentration affects sperm count, motility, morphology, and volume [24,25,26,27]. Zinc ion is a cofactor of several cell enzymes with structural, catalytic, and regulatory functions. Two different families of transporters have been identified: the ZIP (Zrt/IRT-like proteins) family that regulates zinc inflow from the extracellular zone, and the ZnT family (Zn transporter) that controls zinc outflow from the cell and ion redistribution into the mitochondria and lysosomes. These proteins are significantly downregulated in PCa [28,29,30,31,32,33,34,35,36]. As a result of the loss of zinc accumulation, a higher energy balance is obtained (via the Krebs cycle). As well as this role, zinc is involved in other biological processes such as cell division, intracellular signaling, apoptosis, cell invasion, and migration. Induced zinc accumulation in PCa cells leads to a marked inhibition of cell growth. The inhibitory effects was associated with increased levels of p21 [37]. Recent studies have demonstrated the influence of zinc in the expression of genes related to tumor growth, angiogenesis, and metastasis through the nuclear transcriptional factor NF-kB. The treatment of PCa with zinc was associated with a reduced expression of tumorigenic cytokines (VEGF, IL-6, IL-8 and MMP-9). Conversely, zinc depletion led to the increased expression of some genes (PKB/Akt, Mdm2 phosphorylation) and reduced levels of p53 and p21 [38,39]. Zinc accumulation has been associated with the mitochondrial release of cytochrome C, thus leading to the activation of caspase-9 and caspase-3 and the cleavage of poly(ADP-ribose)polymerase (intrinsic apoptosis pathway) [40,41]. Moreover, PCa cells pretreated with zinc showed a reduced expression of intercellular adhesion molecules (i.e., ICAM-1) and inactivation of Aminopeptidase N, with an impaired ability to invade the extracellular matrix [42,43,44].

3. Glucose Metabolism Reprogramming

Prostate cancer cells are characterized by a hybrid glycolytic/oxidative phosphorylation (OXPHOS) phenotype determined by androgen receptor (AR) signaling that may stimulate the AMPK-PGC1α cascade [45,46,47]. AMPK regulates the use of carbohydrates, lipids, and amino acids as energy sources, whereas peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC1α) is a transcriptional coactivator playing a key role in mitochondrial biogenesis and function. The FGFR1 pathway involving lactate dehydrogenase (LDH) isoenzymes is one of the mechanisms responsible for the shift from OXPHOS to aerobic glycolysis [48]. The LDHA isoenzyme preferentially converts pyruvate to lactate, which can eventually be oxidized back to pyruvate by LDHB. Phosphorylated LDH isoform A and reduced expression of LDH isoform B (because of methylation of the gene promoter) may lead to the Warburg effect. At the early stages of cancer development, PCa cells are mainly independent of glucose since they use fructose as an alternative energy source: glucose transporter GLUT-1 is less expressed than fructose transporter (GLUT-5) [49]. Bader et al. showed that the mitochondrial pyruvate carrier (MPC) was transcriptionally regulated by AR. This carrier transports pyruvate into mitochondria and links cytosolic with mitochondrial metabolism, and it is increased in primary PCa and is associated with poor clinical outcomes [50]. Another important glucose pathway is the pentose phosphate pathway (PPP), which allows cells to obtain precursors for nucleotide synthesis and NADPH by using glucose-6-phosphate. In PCa cells, glucose-6-phosphate dehydrogenase (G6PDH), the rate-limiting reaction in PPP, is upregulated, and it was shown that AR signaling increases G6PDH, NADPH, and ribose synthesis [51,52,53]. Two genes (PRKAB1 and PFKFB4) were described by Ros et al. to be important for PCa cell survival [54]. PRKAB1 encodes for a regulatory subunit of AMP-activated kinase (AMPK), which turns off ATP-consuming pathways. LKB1 (an AMPK upstream kinase) expression was significantly reduced in high-grade PIN lesions and completely lost in adenocarcinomas [55]. The pivotal role of the LKB1-AMPK axis in controlling oncogenic processes depends on the interaction with the PI3K, mTOR, and MAPK pathways [56,57]. PFKFB4 encodes for 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 4 (an isoform of the glycolytic enzyme phosphofructokinase 2). Higher PFKFB4 mRNA levels were found in metastatic PCa compared with localized tumors because of the role played by this enzyme in controlling glycolysis and antioxidant production.

4. Lipid Metabolism Reprogramming

Lipids may be used as an energy source by the fatty acid (FA) β-oxidation in the mitochondria. Fatty acid translocase (FAT/CD36) is a main FA transporter. CD36 is frequently gained or amplificated in prostate cancer; this feature has been associated with patients’ poor prognosis. In addition, CD36 functions as a receptor capable of activating SRC family kinases, mitogen-activated protein kinases, and reactive oxygen species pathways. CD36 may recognise different ligands such as oxidized low-density lipoproteins, β-amyloid peptide, staphylococcus aureus-derived microbial diacylglycerides and lipoteichoic acid, and mycoplasma macrophage-activating lipopeptide-2. This feature may explain the oncogenic potential of CD36. It has been noted that blocking CD36 reduces fatty acid uptake from the tumor microenvironment, and reduces lipid biosynthesis and the oncogenic lipid signaling pathways, thus limiting cancer growth [58,59,60]. Fatty acid binding proteins (FABPs) regulate the intracellular trafficking of fatty acids. Metastatic progression of different cancers (including PCa) may also depend on FABPs. A cluster of FABP members (FABP4, FABP5, FABP8, FABP9, and FABP12) maps on the chromosome 8q21 region, which is commonly amplified in PCa metastases. It has been suggested that FABPs transfer FA to PPAR β/δ and γ. Downstream effectors of PPARs may enhance cancer epithelial–mesenchymal transition (EMT), angiogenesis, migration, and invasion [61,62,63,64]. A CoA-group is added to FA resulting in acyl-CoA, which will be oxidized into acetyl-CoA in the mitochondria. Eventually, acetyl-CoA will enter the TCA cycle. Fatty acids are part of monoacylglycerols, diacylglycerols (DAG), triacylglycerols (triglycerides), sterol esters, and membrane phospholipids. Cholesterol synthesis demands acetyl-CoA to begin; the rate-limiting reaction is catalyzed by HMG-CoA reductase. As well as fatty acids’ de novo synthesis, they are also derived from adipose tissue lipolysis or breakdown of triglycerides contained in circulating chylomicrons and lipoproteins. An increased lipogenesis and cholesterogenesis have been described in PCa cells (Figure 1) [65].
Many studies have shown that the enzymes involved in these pathways are overexpressed in PCa. ATP-citrate lyase (ACLY), acetyl-CoA carboxylase (ACC), fatty acid synthase (FASN), stearoyl-CoA desaturase-1 (SCD1), 3-hydroxy-3-methyl-glutaryl-CoA reductase (HMGCR), and squalene epoxidase (SQLE) are overexpressed in cancer cells and are under the transcriptional control of the AR. Sterol regulatory element-binding proteins (SREBPs- three isoforms 1a, 1c and 2) are transcription factors that bind sterol regulatory elements (SRE) in the promoter region of genes involved in fatty acid, cholesterol, and lipid metabolism (HMG-CoA synthase, HMG-CoA reductase, FASN, SCD1, and LDLR) [66]. To be activated, a SREBP demands the cleavage from the endoplasmic reticulum (ER) membrane in a two-step process in the Golgi apparatus by the SREBP cleavage-activating protein (SCAP). SREBP2 isoform regulates cholesterol metabolism by inducing the mevalonate pathway, while SREBP1 activates fatty acid synthesis [67]. The SREBP activity depends on the sterol intracellular concentration since the SCAP-SREBP complex is retained at the rough ER when cholesterol is available. The molecular mechanism of this sterol-sensing system is still unclear. Other alternative pathways have been described with SREBP activation such as TNF-α and mTORC [68,69]. The loss of PTEN (a frequently observed genomic anomaly) and the upregulation of PI3K/AKT/mTOR lead to the activation of SREBPs. In addition, SREBP1 has been described as being involved in PCa cell proliferation, migration, and invasion by activating lipogenesis and through an increased production of reactive oxygen species and NADPH oxidase 5 expression. AR activates the expression of SCAP, and SREBP1 increases the expression of the AR. So, the AR-SREBP1 axis forms a self-regulating loop to keep a continuous gene expression [70]. In castration-resistant cancer cell lines, decreased tumor growth, increased apoptosis, altered lipidome, decreased lipid storage, and AR and AR-V7 expression have been observed after inhibiting FASN activity, thus highlighting the role that this enzyme may play in PCa progression [71]. The aberrant accumulation of esterified cholesterol has been observed in lipid droplets within high-grade and metastatic human PCa, but is not detectable in normal prostate cells. The abundance of cholesterol is related to the upregulated producing pathways and uptake from the circulation, whereas sterol efflux is downregulated (ATP binding cassette transporter A1 or G1 are cell transport proteins) [72]. Statins (inhibitors of HMG-CoA reductase) reduce PCa cells’ growth, invasion, and migration, and induce apoptosis. Increased expression of HMG-CoA reductase has been associated with poor prognosis; in addition, a higher expression of this enzyme has been noted in enzalutamide-resistant PCa cells, and its knockout can restore enzalutamide sensitivity. The combination of simvastatin and enzalutamide decreased PCa cell growth in both in vivo and in vitro models. On the other hand, simvastatin alone or in combination with enzalutamide lowered AR expression in enzalutamide-resistant cancer cells [73,74,75,76,77].
Lipidomic analysis revealed that monounsaturated fatty acid and polyunsaturated fatty acid levels overcome free saturated fatty acids. Overexpression of Δ-Enoyl-CoA Delta Isomerase 1 (EC1), a key enzyme of β-oxidation, has been associated with the biochemical recurrence of PCa [78]. In turn, the inhibition of this enzyme reduced cell growth. Furthermore, 2,4 dienoyl-CoA reductase (DECR1, involved in polyunsaturated fatty acyl-CoA oxidation) is more expressed in castration-resistant tumors. Knockdown of DECR1 has been shown to reduce proliferation, migration, and treatment resistance [79].
High levels of phosphocholine, phosphoethanolamine, and glycerophospholipids have been observed in PCa; these compounds are consistent with membrane remodeling and cellular proliferation processes [80]. The “cholinic phenotype” is referred to as the overexpression of choline kinase alpha (CHKA) and phosphocholine (PCho) levels in different types of cancers, including PCa. Two separate genes (CHKA and CHKB) encode for three isoforms, CHKA-1, CHKA-2, and CHKB, which are active in homodimeric, heterodimeric, and oligomeric forms. High levels of CHKA-1/2 have been observed in PCa and in several human malignancies (breast, lung, ovarian, endometrial, colorectal, bladder cancers, osteosarcoma, and T-cell lymphoma). PCho is the main source for phosphatidylcholine (PC-Kennedy pathway), which is the main phospholipid in eukaryotic membranes with other associated functions such as cholesterol transport support, a substrate to produce second messengers, and a cofactor for several enzymes [81,82]. CHK is also involved in sphingomyelin synthesis, another essential membrane phospholipid [83]. The working hypothesis is that AR activity and PCa growth may be negatively influenced by reducing CHKA levels. Blocking this kinase may allow cancer cells to activate alternative pathways for phospholipids synthesis that have toxic effects and may lead to cell destruction. Different effective CHKA inhibitors have been developed in clinical trials, but they soon manifested elevated toxicity.
Sphingomyelinases degrade sphingomyelin (SM) into ceramide and phosphocholine. Sphingomyelin synthases (SGMSs) use phosphatidylcholine and ceramide to form SM and diacylglycerol (DAG) [84]. High levels of SM ensure cell survival, proliferation, migration, and inflammation, whereas high levels of ceramide provoke cell cycle arrest. SGMS and ceramide kinase (CERK) regulate sphingolipid homeostasis. In cancer cells, SGMS activity keeps low levels of ceramide; moreover, ceramide 1-phosphate (produced by CERK-dependent phosphorylation of ceramide) plays an important role in cell proliferation and migration and in cancer aggressiveness [85]. Metastasis-associated lung adenocarcinoma transcript-1 (MALAT1) is a long non-coding RNA highly expressed in various malignancies such as bladder, lung, and prostate cancer. It has been hypothesized that MALAT1 interferes with gene expression at a post-transcriptional level mediated by micro-RNAs. It suppresses miR-140 in PCa cells [86]. MALAT1 depletion induces metabolic reprogramming of cancer cells toward a more glycolytic phenotype. Reduced energy balance does not support cell growth adequately. It has been described that CHKA and CERK are targets of MALAT1 as well as MALAT1 modulating CHKA expression, PCho, and glutathione content in PCa cells [87,88]. Since choline synthesis is a crucial event for cancer cells, MALAT1 targeting downregulates CHKA and CERK activity; this effect seems to overlap the activity of CHKA inhibitors. Heat shock proteins (HSPs), which are upregulated in PCa, act as chaperones by binding the ligand-binding domain (LBD) of AR, promoting its stability, folding, and activation. It has been shown that CHKA acts as a co-chaperone reinforcing AR signaling by binding to its LBD. Hence, a feed-forward AR-CHKA signaling loop exists. This reinforces the evidence of CHKA as a marker of tumor progression and a potential target for PCa [89].
Other metabolic intermediates of de novo lipogenesis, whose concentrations are significantly increased in PCa cells, are diacylglycerol (DAG), phosphatidic acid (PA), 6-cholesteryl ester (CE), sphingosine 1-phosphate (S1P), and lysophosphatidic acid (LPA) [90].

5. Amino Acid Metabolism Reprogramming

Glutamine represents the main source for the Krebs cycle and lipogenesis intermediates. It enters the cells through the solute carrier (SLC) group of transporters or via micropinocytosis from the surrounding microenvironment [91,92,93]. Glutamine is transformed into glutamate by glutaminase GLS1 then turned into α-ketoglutarate (α-KG) by glutamate dehydrogenase (GDH) or glutamic oxaloacetic transaminase (GOT). Glutamine is incorporated into the TCA cycle, and it is a building block for glutathione synthesis. α-KG serves as a cofactor for Fe (II)-α-KG dioxygenase, which contributes to DNA demethylation and for Jumonji-domain-containing histone demethylases as well [94,95,96]. An increased expression of glutamine transporter ASCT2 (SLC1A5) and of glutaminase GLS1 has been observed in PCa cells, so glutamine becomes an essential amino acid. These events have been associated with AR, MYC, and mTOR pathways’ activation [97]. Moreover, glutamine functions as a nitrogen donor for amino acids, nucleotides, and other important metabolites for tumor growth. Because of its essential role in prostate cancer progression, glutamine metabolism is thought to become an attractive therapeutic target [98,99]. Glutamate may also be obtained by cancer cells by N-acetyl-aspartyl-glutamate (NAAG) when other ways are limited. Recent studies have explored the relationship between plasma NAAG concentrations and tumor sizes for different cancer, whereas glutamate levels seem to correlate with the Gleason score [100,101]. Tryptophan (TRP) accumulates in cancer compared to normal prostate tissue. Indoleamine-2,3-dioxygenase (IDO1) and tryptophan 2,3-dioxygenase (TDO) mediate the conversion of TRP into KYN. IDO1 expression is under IFN-γ and TNF-α control. The role of KYN seems to be mediated by the interaction with the cytoplasmic aryl hydrocarbon receptor (AhR) in immune and cancer cells. After binding with KYN, AhR moves into the nucleus promoting target genes involved in tumor cell migration and immune evasion. The TRP/KYN pathway has already been described in both renal and bladder cancers [102,103]. Sarcosine is an N-methyl derivative of glycine whose expression increases during the progression from benign tissue, through localization, to metastatic cancer. Glycine-N-methyltransferase (GNMT) catalyzes the transfer of the methyl group from S-adenosyl methionine (SAM) to glycine. Recently, higher expression of GNMT have been described in PCa cells compared to normal tissues, and higher enzyme levels have been associated with lower disease-free survival rates [104]. A previous study showed that serum sarcosine had a higher predictive value than PSA in patients with PSA <4 ng/mL. At the same time, low-/intermediate-/high-grade cancers were positively associated with sarcosine levels [105,106].

6. Metabolic Crosstalk in Prostate Cancer Microenvironment

A crosstalk between cancer and stromal cells occurs in the tumor microenvironment because of the release of different cytokines and growth factors and due to changes in the extracellular matrix (Figure 2).
This remodulation of the crosstalk ensures not only cancer cells’ proliferation in a hostile environment but also their ability to infiltrate the ECM and metastasize. Cancer-associated fibroblasts (CAFs) and tumor-associated macrophages (TAMs) actively interact with cancer cells. As well as their ability to induce epithelial–mesenchymal transition and stem-like characteristics in PCa, CAFs induce reciprocal metabolic reprogramming. Indeed, glucose transporter GLUT1 is more expressed in CAFs. Once CAFs differentiate into myofibroblasts, they can secrete lactate and pyruvate (through monocarboxylate transporter-MCT4). These metabolites can be taken up by epithelial cancer cells (through the MCT1 transporter) and then incorporated into the Krebs cycle. Within cancer cells, lactate activates the SIRT1/PGC-1α axis, increasing mitochondrial mass and activity. Mitochondrial transfer between CAFs and cancer cells has recently been discovered [107]. This interaction makes cancer cells independent from glucose consumption while becoming dependent on lactate consumption for anabolic pathways and cell proliferation. This metabolic reprogramming between CAFs and PCa cells has been referred to as the “Reverse Warburg Effect” by Pavlides et al. In addition, CAFs provide high levels of glutamine by activated oncogenic Ras [108,109,110]. Adipocytes represent another class of stromal cells with a well-known crosstalk with PCa. Obesity and high visceral fat may increase the risk of PCa progression to metastatic disease probably because of the excess in dietary fatty acids, alterations in the insulin-IGF-1 axis, and higher levels of pro-inflammatory cytokines [111]. Several mechanisms have been suggested to explain the obesity–cancer association such as the presence of cancer-associated adipocytes, adipokines, obesity-related inflammatory cytokine production, and alteration in sex hormone metabolism. Phenotypical changes have been described in tumor-surrounding adipocytes: as well as the downregulation of adipocyte markers and reduced lipid storage, the overproduction of pro-inflammatory cytokines and extracellular matrix-related molecules has been noted [112]. Cell co-culture studies demonstrated that PCa cells store fatty acid from surrounding adipocytes. Indeed, increased expression of fatty acid transporters such as CD36 and FATP5 has been described in some PCa cell lines (PC-3 and LNCaP cells). In addition, in tumor cells, free fatty acids induce oxidative stress through NADPH oxidase 5 (NOX5); the increased ROS production contributes to the tumor cell invasion by activating the HIF1-MMP14 pathway. The increased thickness of periprostatic adipose tissue (PPAT) has been associated with PCa progression and the presence of high-grade disease. Contact with PPAT may promote progression in PCa with extraprostatic extension, and a decreased biochemical recurrence-free survival has been speculated after radical prostatectomy when compared to tumors not invading outside the prostate capsule [113].
Altuna-Coy et al. observed changes in linoleic acid metabolism. It may be transformed into 9- and 13-hydroxyoctadecadienoic acid (9- and 13-HODE) and 9- and 13-oxo-octadecadienoic acid (9- and 13-oxoODE). 13-oxoODE has been stated to be an endogenous ligand of PPARG with anti-inflammatory properties. They found reduced concentrations of 13-oxoODE/9-oxoODE in high-risk PPAT and reduced expression of PPARG and ADIPOQ genes. Reduced expression of PPARG (inhibitor of NF-kB) may explain the increased levels of pro-inflammatory cytokines (IL-6, TNF-α, IL-1B) in the aggressive PCa-related PPAT [114]. Finley et al. noted that adipokines’ concentrations in PPAT were more elevated than serum levels. In particular high levels of IL-6 were found in PPAT together with the increased phosphorylation of STAT3 in high-grade tumors [115]. Moreover, one of the chemokines involved in PCa progression in a mouse model is CXCL12/SDF-1. Saha et al. demonstrated that CXCL12 and its receptors CXCR4 and CXCR7 activate many oncogenic signaling pathways including STAT3, NFkB, and MAPK [116]. Laurent et al. described the role of CCL7, another chemokine secreted by adipocytes in PPAT. Interacting with CCR3 (overexpressed in PCa cells), CCL7 diffuses from the PPAT to the peripheral zone of the prostate, promoting cancer extraprostatic extension and local dissemination [117]. Leptin, a cytokine secreted by adipocytes, promotes MCT4 and its chaperone CD147 expression in cancer cells [118]. According to the density and location of tumor-infiltrating CD8+ T cells, and the presence of immunosuppressive Foxp3+ regulatory T cells and CD11b+ myeloid-derived suppressor cells (MDSC), cancers may be classified as immunologically “hot” or “cold”. Prostate cancers are known to be immunologically cold because of low cytotoxic T-cell infiltration. Tumor-associated macrophages (TAM) regulate angiogenesis, and tumor cells’ proliferation and dissemination. TAM polarization from the M1 phenotype toward the M2 phenotype relates with poor prognosis since they release anti-inflammatory cytokines (IL-10 and TGF-β), thus blocking cytotoxic CD8+ T-cell activity [119,120].

7. Metabolic Gene Alterations

The most frequently observed genomic alteration is PTEN (phosphatase and tensin homolog tumor suppressor) loss on chromosome 10 [121]. PTEN is a direct antagonist of PI3K, which has several important downstream effector molecules, including Ser/Thr protein kinase AKT/PKB, which finally activates the mammalian target of rapamycin (mTOR) [122]. Another possible genetic aberration is the gain of MYC. According to their genetics, two metabolic phenotypes have been described: phospho-AKThigh/MYClow versus phospho-AKTlow/MYChigh. AKT signaling activation has been associated with enhanced aerobic glycolysis (Warburg effect), the pentose phosphate pathway, and fructose metabolism; on the other hand, MYC overexpression has been related to the increased expression of glutaminase, dysregulated lipid metabolism, and reduced expression of some glycolytic enzymes [123,124,125]. However, both metabolic phenotypes have been associated with the increased expression of fatty acid synthase (FASN) [126]. Even if the clinical implications are still unknown, the characterization of these cellular subtypes may have diagnostic and therapeutic roles: 18F-FDG PET has a better diagnostic performance in AKThigh/MYClow PCa. SOX2 is a transcriptional factor with a pivotal role in stem cells’ pluripotency. Its expression has been described in different human malignancies including prostate cancer. First, SOX2 promotes mitochondrial biogenesis and enhances glycolysis, oxidative phosphorylation, purines, pyrimidines, amino acids’ metabolism, and the pentose phosphate pathway. At the same time, SOX2 has different gene targets not typically observed in human embryonic stem cells (i.e., BCL2, EZH2, FGFR3, FOXA1, KRAS, MET, etc.). Nevertheless, its mechanism of action remains unclear. It has been suggested that SOX2 may be involved in driving metastatic survival and growth, and in enabling easier adaptation to the new metastatic microenvironment. Furthermore, it has been shown to be associated with resistance to AR antagonist enzalutamide [127].
Ongoing insights are investigating autophagy and its crucial role in cancer biology. It is a homeostatic process for degradation through the lysosomal system of either older proteins or dysfunctional cytoplasmic organelles [128]. It contributes to the restoration of energy balance, especially during energy deprivation. In addition, autophagy aims to regulate cellular mass, to distribute organelles properly, and to remove harmful compounds. Depending on the cellular context, it may play either a protective or a detrimental role for PCa survival. mTORC plays a pivotal role in autophagy regulation [129]. Different genes related to this process involved in PCa carcinogenesis have been identified upstream of mTORC activity. STK11 encodes for LKB1, a tumor suppressor serine–threonine kinase whose main effector is AMPK. It has been described that LKB1 expression reduces throughout PCa progression from normal to neoplastic tissue [130,131]. Calcium/calmodulin-dependent protein kinase kinase 2 (CaMKK2) is another essential element for autophagic processes, which has been shown to be a target gene of AR [132]. Tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) also triggers autophagy in PCa cells [133]. Autophagy may be thought of as a response mechanism to stressors for PCa including chemotherapy and castration therapy. In turn, autophagy itself may promote castration resistance. Cell death may be enhanced under anticancer treatment by inhibiting autophagic pathways. Further studies could elucidate the role of autophagy modulators to improve therapeutic strategies.

8. Conclusions

Blood, urine, semen, and tissue samples continue to be investigated for the identification of novel biomarkers that might become eligible for PCa diagnosis with a better performance than PSA. At the same time, new markers might be used for selecting candidates for active surveillance. To achieve these goals, algorithms of machine learning (ML), a branch of artificial intelligence (AI), are currently being developed and trained. Deep learning (DL) is a recent field of ML that has already shown superior problem-solving capabilities by using neural networks [134]. In the near future, by integrating different metabolomic datasets and improving AI techniques, new biomarkers for earlier cancer diagnosis might be identified. Prostate cancer cells face important reprogramming of different metabolic pathways. PCa cells and tumor microenvironment (TME) components are known to affect their metabolism mutually. A better comprehension of the relationship between cancer cells and TME might help novel effective therapies to be designed, which might overcome castration resistance disease. In association with standard therapies, targeting these altered metabolic pathways may represent future tools to improve prostate cancer treatments. Drugs targeting specific metabolic pathways are listed in Table 1. A deeper insight into the metabolic changes may be obtained by a multi-omic approach integrating genomics, transcriptomics, metabolomics, lipidomics, and radiomics data [135,136].

Author Contributions

Conceptualization, F.L. and G.L.; literature collection and preparation, F.L., N.A.d.M., M.R., M.F., D.T., O.S.T., M.B., P.D. and G.L.; writing—original draft preparation, F.L.; writing—review and editing, G.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer Statistics, 2022. CA Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef]
  2. 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]
  3. Walsh, A.L.; Tuzova, A.V.; Bolton, E.M.; Lynch, T.H.; Perry, A.S. Long Noncoding RNAs and Prostate Carcinogenesis: The Missing “Linc”? Trends Mol. Med. 2014, 20, 428–436. [Google Scholar] [CrossRef]
  4. Stephan, C.; Ralla, B.; Jung, K. Prostate-Specific Antigen and Other Serum and Urine Markers in Prostate Cancer. Biochim. Biophys. Acta 2014, 1846, 99–112. [Google Scholar] [CrossRef]
  5. Gasi Tandefelt, D.; Boormans, J.; Hermans, K.; Trapman, J. ETS Fusion Genes in Prostate Cancer. Endocr. Relat. Cancer 2014, 21, R143–R152. [Google Scholar] [CrossRef]
  6. Lucarelli, G.; Rutigliano, M.; Bettocchi, C.; Palazzo, S.; Vavallo, A.; Galleggiante, V.; Trabucco, S.; Di Clemente, D.; Selvaggi, F.P.; Battaglia, M.; et al. Spondin-2, a Secreted Extracellular Matrix Protein, Is a Novel Diagnostic Biomarker for Prostate Cancer. J. Urol. 2013, 190, 2271–2277. [Google Scholar] [CrossRef]
  7. Lucarelli, G.; Rutigliano, M.; Galleggiante, V.; Giglio, A.; Palazzo, S.; Ferro, M.; Simone, C.; Bettocchi, C.; Battaglia, M.; Ditonno, P. Metabolomic Profiling for the Identification of Novel Diagnostic Markers in Prostate Cancer. Expert. Rev. Mol. Diagn. 2015, 15, 1211–1224. [Google Scholar] [CrossRef]
  8. Lucarelli, G.; Loizzo, D.; Ferro, M.; Rutigliano, M.; Vartolomei, M.D.; Cantiello, F.; Buonerba, C.; Di Lorenzo, G.; Terracciano, D.; De Cobelli, O.; et al. Metabolomic Profiling for the Identification of Novel Diagnostic Markers and Therapeutic Targets in Prostate Cancer: An Update. Expert. Rev. Mol. Diagn. 2019, 19, 377–387. [Google Scholar] [CrossRef]
  9. Hotamisligil, G.S. Inflammation, Metaflammation and Immunometabolic Disorders. Nature 2017, 542, 177–185. [Google Scholar] [CrossRef]
  10. Subramanian, M.; Wojtusciszyn, A.; Favre, L.; Boughorbel, S.; Shan, J.; Letaief, K.B.; Pitteloud, N.; Chouchane, L. Precision Medicine in the Era of Artificial Intelligence: Implications in Chronic Disease Management. J. Transl. Med. 2020, 18, 472. [Google Scholar] [CrossRef]
  11. Boehm, K.; Sun, M.; Larcher, A.; Blanc-Lapierre, A.; Schiffmann, J.; Graefen, M.; Sosa, J.; Saad, F.; Parent, M.-É.; Karakiewicz, P.I. Waist Circumference, Waist-Hip Ratio, Body Mass Index, and Prostate Cancer Risk: Results from the North-American Case-Control Study Prostate Cancer & Environment Study. Urol. Oncol. 2015, 33, 494.e1–494.e7. [Google Scholar] [CrossRef]
  12. Dickerman, B.A.; Ahearn, T.U.; Giovannucci, E.; Stampfer, M.J.; Nguyen, P.L.; Mucci, L.A.; Wilson, K.M. Weight Change, Obesity and Risk of Prostate Cancer Progression among Men with Clinically Localized Prostate Cancer. Int. J. Cancer 2017, 141, 933–944. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Kasper, J.S.; Liu, Y.; Giovannucci, E. Diabetes Mellitus and Risk of Prostate Cancer in the Health Professionals Follow-up Study. Int. J. Cancer 2009, 124, 1398–1403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Vavallo, A.; Simone, S.; Lucarelli, G.; Rutigliano, M.; Galleggiante, V.; Grandaliano, G.; Gesualdo, L.; Campagna, M.; Cariello, M.; Ranieri, E.; et al. Pre-Existing Type 2 Diabetes Mellitus Is an Independent Risk Factor for Mortality and Progression in Patients with Renal Cell Carcinoma. Medicine 2014, 93, e183. [Google Scholar] [CrossRef]
  15. Breda, A.; Lucarelli, G.; Luccarelli, G.; Rodriguez-Faba, O.; Guirado, L.; Facundo, C.; Bettocchi, C.; Gesualdo, L.; Castellano, G.; Grandaliano, G.; et al. Clinical and Pathological Outcomes of Renal Cell Carcinoma (RCC) in Native Kidneys of Patients with End-Stage Renal Disease: A Long-Term Comparative Retrospective Study with RCC Diagnosed in the General Population. World J. Urol. 2015, 33, 1–7. [Google Scholar] [CrossRef]
  16. Ferro, M.; Vartolomei, M.D.; Russo, G.I.; Cantiello, F.; Farhan, A.R.A.; Terracciano, D.; Cimmino, A.; Di Stasi, S.; Musi, G.; Hurle, R.; et al. An Increased Body Mass Index Is Associated with a Worse Prognosis in Patients Administered BCG Immunotherapy for T1 Bladder Cancer. World J. Urol. 2019, 37, 507–514. [Google Scholar] [CrossRef]
  17. Wright, M.E.; Chang, S.-C.; Schatzkin, A.; Albanes, D.; Kipnis, V.; Mouw, T.; Hurwitz, P.; Hollenbeck, A.; Leitzmann, M.F. Prospective Study of Adiposity and Weight Change in Relation to Prostate Cancer Incidence and Mortality. Cancer 2007, 109, 675–684. [Google Scholar] [CrossRef]
  18. Freedland, S.J.; Platz, E.A. Obesity and Prostate Cancer: Making Sense out of Apparently Conflicting Data. Epidemiol. Rev. 2007, 29, 88–97. [Google Scholar] [CrossRef] [Green Version]
  19. Ferro, M.; Terracciano, D.; Buonerba, C.; Lucarelli, G.; Bottero, D.; Perdonà, S.; Autorino, R.; Serino, A.; Cantiello, F.; Damiano, R.; et al. The Emerging Role of Obesity, Diet and Lipid Metabolism in Prostate Cancer. Future Oncol. 2017, 13, 285–293. [Google Scholar] [CrossRef] [Green Version]
  20. Cirulli, E.T.; Guo, L.; Leon Swisher, C.; Shah, N.; Huang, L.; Napier, L.A.; Kirkness, E.F.; Spector, T.D.; Caskey, C.T.; Thorens, B.; et al. Profound Perturbation of the Metabolome in Obesity Is Associated with Health Risk. Cell Metab. 2019, 29, 488–500.e2. [Google Scholar] [CrossRef]
  21. Spratlin, J.L.; Serkova, N.J.; Eckhardt, S.G. Clinical Applications of Metabolomics in Oncology: A Review. Clin. Cancer Res. 2009, 15, 431–440. [Google Scholar] [CrossRef] [Green Version]
  22. Gómez-Cebrián, N.; Poveda, J.L.; Pineda-Lucena, A.; Puchades-Carrasco, L. Metabolic Phenotyping in Prostate Cancer Using Multi-Omics Approaches. Cancers 2022, 14, 596. [Google Scholar] [CrossRef]
  23. Singh, K.K.; Desouki, M.M.; Franklin, R.B.; Costello, L.C. Mitochondrial Aconitase and Citrate Metabolism in Malignant and Nonmalignant Human Prostate Tissues. Mol. Cancer 2006, 5, 14. [Google Scholar] [CrossRef] [Green Version]
  24. Frégeau-Proulx, L.; Lacouture, A.; Berthiaume, L.; Weidmann, C.; Harvey, M.; Gonthier, K.; Pelletier, J.-F.; Neveu, B.; Jobin, C.; Bastien, D.; et al. Multiple Metabolic Pathways Fuel the Truncated Tricarboxylic Acid Cycle of the Prostate to Sustain Constant Citrate Production and Secretion. Mol. Metab. 2022, 62, 101516. [Google Scholar] [CrossRef]
  25. Owen, D.H.; Katz, D.F. A Review of the Physical and Chemical Properties of Human Semen and the Formulation of a Semen Simulant. J. Androl. 2005, 26, 459–469. [Google Scholar] [CrossRef] [Green Version]
  26. Beigi Harchegani, A.; Irandoost, A.; Mirnamniha, M.; Rahmani, H.; Tahmasbpour, E. Possible Mechanisms for The Effects of Calcium Deficiency on Male Infertility. Int. J. Fertil. Steril. 2019, 12, 267–272. [Google Scholar] [CrossRef]
  27. Shemshaki, G.; Murthy, A.S.N.; Malini, S.S. Assessment and Establishment of Correlation between Reactive Oxidation Species, Citric Acid, and Fructose Level in Infertile Male Individuals: A Machine-Learning Approach. J. Hum. Reprod. Sci. 2021, 14, 129–136. [Google Scholar] [CrossRef]
  28. Sawant Dessai, A.; Dominguez, M.P.; Chen, U.I.; Hasper, J.; Prechtl, C.; Yu, C.; Katsuta, E.; Dai, T.; Zhu, B.; Jung, S.Y.; et al. Transcriptional Repression of SIRT3 Potentiates Mitochondrial Aconitase Activation to Drive Aggressive Prostate Cancer to the Bone. Cancer Res. 2021, 81, 50–63. [Google Scholar] [CrossRef]
  29. Costello, L.C.; Franklin, R.B.; Feng, P. Mitochondrial Function, Zinc, and Intermediary Metabolism Relationships in Normal Prostate and Prostate Cancer. Mitochondrion 2005, 5, 143–153. [Google Scholar] [CrossRef] [Green Version]
  30. Costello, L.C.; Franklin, R.B. The Clinical Relevance of the Metabolism of Prostate Cancer; Zinc and Tumor Suppression: Connecting the Dots. Mol. Cancer 2006, 5, 17. [Google Scholar] [CrossRef]
  31. Costello, L.C.; Franklin, R.B. Zinc Is Decreased in Prostate Cancer: An Established Relationship of Prostate Cancer! J. Biol. Inorg. Chem. 2011, 16, 3–8. [Google Scholar] [CrossRef] [PubMed]
  32. Gaither, L.A.; Eide, D.J. The Human ZIP1 Transporter Mediates Zinc Uptake in Human K562 Erythroleukemia Cells. J. Biol. Chem. 2001, 276, 22258–22264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Gaither, L.A.; Eide, D.J. Functional Expression of the Human HZIP2 Zinc Transporter. J. Biol. Chem. 2000, 275, 5560–5564. [Google Scholar] [CrossRef] [Green Version]
  34. Costello, L.C.; Liu, Y.; Zou, J.; Franklin, R.B. Evidence for a Zinc Uptake Transporter in Human Prostate Cancer Cells Which Is Regulated by Prolactin and Testosterone. J. Biol. Chem. 1999, 274, 17499–17504. [Google Scholar] [CrossRef] [Green Version]
  35. Desouki, M.M.; Geradts, J.; Milon, B.; Franklin, R.B.; Costello, L.C. HZip2 and HZip3 Zinc Transporters Are down Regulated in Human Prostate Adenocarcinomatous Glands. Mol. Cancer 2007, 6, 37. [Google Scholar] [CrossRef] [Green Version]
  36. Kolenko, V.; Teper, E.; Kutikov, A.; Uzzo, R. Zinc and Zinc Transporters in Prostate Carcinogenesis. Nat. Rev. Urol. 2013, 10, 219–226. [Google Scholar] [CrossRef] [Green Version]
  37. Liang, J.Y.; Liu, Y.Y.; Zou, J.; Franklin, R.B.; Costello, L.C.; Feng, P. Inhibitory Effect of Zinc on Human Prostatic Carcinoma Cell Growth. Prostate 1999, 40, 200–207. [Google Scholar] [CrossRef]
  38. Golovine, K.; Uzzo, R.G.; Makhov, P.; Crispen, P.L.; Kunkle, D.; Kolenko, V.M. Depletion of Intracellular Zinc Increases Expression of Tumorigenic Cytokines VEGF, IL-6 and IL-8 in Prostate Cancer Cells via NF-KappaB-Dependent Pathway. Prostate 2008, 68, 1443–1449. [Google Scholar] [CrossRef] [Green Version]
  39. Han, C.-T.; Schoene, N.W.; Lei, K.Y. Influence of Zinc Deficiency on Akt-Mdm2-P53 and Akt-P21 Signaling Axes in Normal and Malignant Human Prostate Cells. Am. J. Physiol. Cell. Physiol. 2009, 297, C1188–C1199. [Google Scholar] [CrossRef]
  40. Feng, P.; Li, T.-L.; Guan, Z.-X.; Franklin, R.B.; Costello, L.C. Direct Effect of Zinc on Mitochondrial Apoptogenesis in Prostate Cells. Prostate 2002, 52, 311–318. [Google Scholar] [CrossRef]
  41. Feng, P.; Liang, J.Y.; Li, T.L.; Guan, Z.X.; Zou, J.; Franklin, R.; Costello, L.C. Zinc Induces Mitochondria Apoptogenesis in Prostate Cells. Mol. Urol. 2000, 4, 31–36. [Google Scholar] [PubMed]
  42. Uzzo, R.G.; Crispen, P.L.; Golovine, K.; Makhov, P.; Horwitz, E.M.; Kolenko, V.M. Diverse Effects of Zinc on NF-KappaB and AP-1 Transcription Factors: Implications for Prostate Cancer Progression. Carcinogenesis 2006, 27, 1980–1990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Ishii, K.; Otsuka, T.; Iguchi, K.; Usui, S.; Yamamoto, H.; Sugimura, Y.; Yoshikawa, K.; Hayward, S.W.; Hirano, K. Evidence That the Prostate-Specific Antigen (PSA)/Zn2+ Axis May Play a Role in Human Prostate Cancer Cell Invasion. Cancer Lett. 2004, 207, 79–87. [Google Scholar] [CrossRef] [PubMed]
  44. Wickström, M.; Larsson, R.; Nygren, P.; Gullbo, J. Aminopeptidase N (CD13) as a Target for Cancer Chemotherapy. Cancer Sci. 2011, 102, 501–508. [Google Scholar] [CrossRef]
  45. Peitzsch, C.; Gorodetska, I.; Klusa, D.; Shi, Q.; Alves, T.C.; Pantel, K.; Dubrovska, A. Metabolic Regulation of Prostate Cancer Heterogeneity and Plasticity. Semin. Cancer Biol. 2022, 82, 94–119. [Google Scholar] [CrossRef]
  46. Moon, J.-S.; Jin, W.-J.; Kwak, J.-H.; Kim, H.-J.; Yun, M.-J.; Kim, J.-W.; Park, S.W.; Kim, K.-S. Androgen Stimulates Glycolysis for de Novo Lipid Synthesis by Increasing the Activities of Hexokinase 2 and 6-Phosphofructo-2-Kinase/Fructose-2,6-Bisphosphatase 2 in Prostate Cancer Cells. BioChem. J. 2011, 433, 225–233. [Google Scholar] [CrossRef] [Green Version]
  47. Lee, Y.G.; Nam, Y.; Shin, K.J.; Yoon, S.; Park, W.S.; Joung, J.Y.; Seo, J.K.; Jang, J.; Lee, S.; Nam, D.; et al. Androgen-Induced Expression of DRP1 Regulates Mitochondrial Metabolic Reprogramming in Prostate Cancer. Cancer Lett. 2020, 471, 72–87. [Google Scholar] [CrossRef]
  48. Liu, J.; Chen, G.; Liu, Z.; Liu, S.; Cai, Z.; You, P.; Ke, Y.; Lai, L.; Huang, Y.; Gao, H.; et al. Aberrant FGFR Tyrosine Kinase Signaling Enhances the Warburg Effect by Reprogramming LDH Isoform Expression and Activity in Prostate Cancer. Cancer Res. 2018, 78, 4459–4470. [Google Scholar] [CrossRef] [Green Version]
  49. Reinicke, K.; Sotomayor, P.; Cisterna, P.; Delgado, C.; Nualart, F.; Godoy, A. Cellular Distribution of Glut-1 and Glut-5 in Benign and Malignant Human Prostate Tissue. J. Cell Biochem. 2012, 113, 553–562. [Google Scholar] [CrossRef]
  50. Bader, D.A.; Hartig, S.M.; Putluri, V.; Foley, C.; Hamilton, M.P.; Smith, E.A.; Saha, P.K.; Panigrahi, A.; Walker, C.; Zong, L.; et al. Mitochondrial Pyruvate Import Is a Metabolic Vulnerability in Androgen Receptor-Driven Prostate Cancer. Nat. Metab. 2019, 1, 70–85. [Google Scholar] [CrossRef]
  51. Lucarelli, G.; Galleggiante, V.; Rutigliano, M.; Sanguedolce, F.; Cagiano, S.; Bufo, P.; Lastilla, G.; Maiorano, E.; Ribatti, D.; Giglio, A.; et al. Metabolomic Profile of Glycolysis and the Pentose Phosphate Pathway Identifies the Central Role of Glucose-6-Phosphate Dehydrogenase in Clear Cell-Renal Cell Carcinoma. Oncotarget 2015, 6, 13371–13386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Tsouko, E.; Khan, A.S.; White, M.A.; Han, J.J.; Shi, Y.; Merchant, F.A.; Sharpe, M.A.; Xin, L.; Frigo, D.E. Regulation of the Pentose Phosphate Pathway by an Androgen Receptor-MTOR-Mediated Mechanism and Its Role in Prostate Cancer Cell Growth. Oncogenesis 2014, 3, e103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Riganti, C.; Gazzano, E.; Polimeni, M.; Aldieri, E.; Ghigo, D. The Pentose Phosphate Pathway: An Antioxidant Defense and a Crossroad in Tumor Cell Fate. Free Radic. Biol. Med. 2012, 53, 421–436. [Google Scholar] [CrossRef]
  54. Ros, S.; Santos, C.R.; Moco, S.; Baenke, F.; Kelly, G.; Howell, M.; Zamboni, N.; Schulze, A. Functional Metabolic Screen Identifies 6-Phosphofructo-2-Kinase/Fructose-2,6-Biphosphatase 4 as an Important Regulator of Prostate Cancer Cell Survival. Cancer Discov. 2012, 2, 328–343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Pearson, H.B.; McCarthy, A.; Collins, C.M.P.; Ashworth, A.; Clarke, A.R. Lkb1 Deficiency Causes Prostate Neoplasia in the Mouse. Cancer Res. 2008, 68, 2223–2232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Tennakoon, J.B.; Shi, Y.; Han, J.J.; Tsouko, E.; White, M.A.; Burns, A.R.; Zhang, A.; Xia, X.; Ilkayeva, O.R.; Xin, L.; et al. Androgens Regulate Prostate Cancer Cell Growth via an AMPK-PGC-1α-Mediated Metabolic Switch. Oncogene 2014, 33, 5251–5261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Grossi, V.; Lucarelli, G.; Forte, G.; Peserico, A.; Matrone, A.; Germani, A.; Rutigliano, M.; Stella, A.; Bagnulo, R.; Loconte, D.; et al. Loss of STK11 Expression Is an Early Event in Prostate Carcinogenesis and Predicts Therapeutic Response to Targeted Therapy against MAPK/P38. Autophagy 2015, 11, 2102–2113. [Google Scholar] [CrossRef] [Green Version]
  58. Watt, M.J.; Clark, A.K.; Selth, L.A.; Haynes, V.R.; Lister, N.; Rebello, R.; Porter, L.H.; Niranjan, B.; Whitby, S.T.; Lo, J.; et al. Suppressing Fatty Acid Uptake Has Therapeutic Effects in Preclinical Models of Prostate Cancer. Sci. Transl. Med. 2019, 11, eaau5758. [Google Scholar] [CrossRef]
  59. Hoebe, K.; Georgel, P.; Rutschmann, S.; Du, X.; Mudd, S.; Crozat, K.; Sovath, S.; Shamel, L.; Hartung, T.; Zähringer, U.; et al. CD36 Is a Sensor of Diacylglycerides. Nature 2005, 433, 523–527. [Google Scholar] [CrossRef]
  60. Stewart, C.R.; Stuart, L.M.; Wilkinson, K.; van Gils, J.M.; Deng, J.; Halle, A.; Rayner, K.J.; Boyer, L.; Zhong, R.; Frazier, W.A.; et al. CD36 Ligands Promote Sterile Inflammation through Assembly of a Toll-like Receptor 4 and 6 Heterodimer. Nat. Immunol. 2010, 11, 155–161. [Google Scholar] [CrossRef]
  61. Liu, R.-Z.; Godbout, R. An Amplified Fatty Acid-Binding Protein Gene Cluster in Prostate Cancer: Emerging Roles in Lipid Metabolism and Metastasis. Cancers 2020, 12, 3823. [Google Scholar] [CrossRef] [PubMed]
  62. Amiri, M.; Yousefnia, S.; Seyed Forootan, F.; Peymani, M.; Ghaedi, K.; Nasr Esfahani, M.H. Diverse Roles of Fatty Acid Binding Proteins (FABPs) in Development and Pathogenesis of Cancers. Gene 2018, 676, 171–183. [Google Scholar] [CrossRef] [PubMed]
  63. Uehara, H.; Takahashi, T.; Oha, M.; Ogawa, H.; Izumi, K. Exogenous Fatty Acid Binding Protein 4 Promotes Human Prostate Cancer Cell Progression. Int. J. Cancer 2014, 135, 2558–2568. [Google Scholar] [CrossRef] [PubMed]
  64. Liu, R.-Z.; Choi, W.-S.; Jain, S.; Dinakaran, D.; Xu, X.; Han, W.H.; Yang, X.-H.; Glubrecht, D.D.; Moore, R.B.; Lemieux, H.; et al. The FABP12/PPARγ Pathway Promotes Metastatic Transformation by Inducing Epithelial-to-Mesenchymal Transition and Lipid-Derived Energy Production in Prostate Cancer Cells. Mol. Oncol. 2020, 14, 3100–3120. [Google Scholar] [CrossRef]
  65. Siltari, A.; Syvälä, H.; Lou, Y.-R.; Gao, Y.; Murtola, T.J. Role of Lipids and Lipid Metabolism in Prostate Cancer Progression and the Tumor’s Immune Environment. Cancers 2022, 14, 4293. [Google Scholar] [CrossRef]
  66. Shimano, H. Sterol Regulatory Element-Binding Proteins (SREBPs): Transcriptional Regulators of Lipid Synthetic Genes. Progress Lipid Res. 2001, 40, 439–452. [Google Scholar] [CrossRef]
  67. Krycer, J.R.; Kristiana, I.; Brown, A.J. Cholesterol Homeostasis in Two Commonly Used Human Prostate Cancer Cell-Lines, LNCaP and PC-3. PLoS ONE 2009, 4, e8496. [Google Scholar] [CrossRef] [Green Version]
  68. Eid, W.; Dauner, K.; Courtney, K.C.; Gagnon, A.; Parks, R.J.; Sorisky, A.; Zha, X. MTORC1 Activates SREBP-2 by Suppressing Cholesterol Trafficking to Lysosomes in Mammalian Cells. Proc. Natl. Acad. Sci. USA 2017, 114, 7999–8004. [Google Scholar] [CrossRef] [Green Version]
  69. Lawler, J.F.; Yin, M.; Diehl, A.M.; Roberts, E.; Chatterjee, S. Tumor Necrosis Factor-Alpha Stimulates the Maturation of Sterol Regulatory Element Binding Protein-1 in Human Hepatocytes through the Action of Neutral Sphingomyelinase. J. Biol. Chem. 1998, 273, 5053–5059. [Google Scholar] [CrossRef] [Green Version]
  70. Huang, W.-C.; Li, X.; Liu, J.; Lin, J.; Chung, L.W.K. Activation of Androgen Receptor, Lipogenesis, and Oxidative Stress Converged by SREBP-1 Is Responsible for Regulating Growth and Progression of Prostate Cancer Cells. Mol. Cancer Res. 2012, 10, 133–142. [Google Scholar] [CrossRef]
  71. Zadra, G.; Ribeiro, C.F.; Chetta, P.; Ho, Y.; Cacciatore, S.; Gao, X.; Syamala, S.; Bango, C.; Photopoulos, C.; Huang, Y.; et al. Inhibition of de Novo Lipogenesis Targets Androgen Receptor Signaling in Castration-Resistant Prostate Cancer. Proc. Natl. Acad. Sci. USA 2019, 116, 631–640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Wu, X.; Daniels, G.; Lee, P.; Monaco, M.E. Lipid Metabolism in Prostate Cancer. Am. J. Clin. Exp. Urol. 2014, 2, 111–120. [Google Scholar] [PubMed]
  73. Murtola, T.J.; Syvälä, H.; Pennanen, P.; Bläuer, M.; Solakivi, T.; Ylikomi, T.; Tammela, T.L.J. The Importance of LDL and Cholesterol Metabolism for Prostate Epithelial Cell Growth. PLoS ONE 2012, 7, e39445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Murtola, T.J.; Syvälä, H.; Pennanen, P.; Bläuer, M.; Solakivi, T.; Ylikomi, T.; Tammela, T.L.J. Comparative Effects of High and Low-Dose Simvastatin on Prostate Epithelial Cells: The Role of LDL. Eur. J. Pharmacol. 2011, 673, 96–100. [Google Scholar] [CrossRef]
  75. Hoque, A.; Chen, H.; Xu, X.-C. Statin Induces Apoptosis and Cell Growth Arrest in Prostate Cancer Cells. Cancer Epidemiol. Biomarkers Prev. 2008, 17, 88–94. [Google Scholar] [CrossRef] [Green Version]
  76. Murtola, T.J.; Pennanen, P.; Syvälä, H.; Bläuer, M.; Ylikomi, T.; Tammela, T.L.J. Effects of Simvastatin, Acetylsalicylic Acid, and Rosiglitazone on Proliferation of Normal and Cancerous Prostate Epithelial Cells at Therapeutic Concentrations. Prostate 2009, 69, 1017–1023. [Google Scholar] [CrossRef]
  77. Kochuparambil, S.T.; Al-Husein, B.; Goc, A.; Soliman, S.; Somanath, P.R. Anticancer Efficacy of Simvastatin on Prostate Cancer Cells and Tumor Xenografts Is Associated with Inhibition of Akt and Reduced Prostate-Specific Antigen Expression. J. Pharmacol. Exp. Ther. 2011, 336, 496–505. [Google Scholar] [CrossRef] [Green Version]
  78. Bramhecha, Y.M.; Guérard, K.-P.; Audet-Walsh, É.; Rouzbeh, S.; Kassem, O.; Pernet, E.; Scarlata, E.; Hamel, L.; Brimo, F.; Divangahi, M.; et al. Fatty Acid Oxidation Enzyme Δ3, Δ2-Enoyl-CoA Isomerase 1 (ECI1) Drives Aggressive Tumor Phenotype and Predicts Poor Clinical Outcome in Prostate Cancer Patients. Oncogene 2022, 41, 2798–2810. [Google Scholar] [CrossRef]
  79. Nassar, Z.D.; Mah, C.Y.; Dehairs, J.; Burvenich, I.J.; Irani, S.; Centenera, M.M.; Helm, M.; Shrestha, R.K.; Moldovan, M.; Don, A.S.; et al. Human DECR1 Is an Androgen-Repressed Survival Factor That Regulates PUFA Oxidation to Protect Prostate Tumor Cells from Ferroptosis. Elife 2020, 9, e54166. [Google Scholar] [CrossRef]
  80. Swanson, M.G.; Keshari, K.R.; Tabatabai, Z.L.; Simko, J.P.; Shinohara, K.; Carroll, P.R.; Zektzer, A.S.; Kurhanewicz, J. Quantification of Choline- and Ethanolamine-Containing Metabolites in Human Prostate Tissues Using 1H HR-MAS Total Correlation Spectroscopy. Magn. Reson. Med. 2008, 60, 33–40. [Google Scholar] [CrossRef]
  81. Chen, X.; Qiu, H.; Wang, C.; Yuan, Y.; Tickner, J.; Xu, J.; Zou, J. Molecular Structure and Differential Function of Choline Kinases CHKα and CHKβ in Musculoskeletal System and Cancer. Cytokine Growth Factor Rev. 2017, 33, 65–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Glunde, K.; Bhujwalla, Z.M.; Ronen, S.M. Choline Metabolism in Malignant Transformation. Nat. Rev. Cancer 2011, 11, 835–848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Janardhan, S.; Srivani, P.; Sastry, G.N. Choline Kinase: An Important Target for Cancer. Curr. Med. Chem. 2006, 13, 1169–1186. [Google Scholar] [CrossRef] [PubMed]
  84. Adada, M.; Luberto, C.; Canals, D. Inhibitors of the Sphingomyelin Cycle: Sphingomyelin Synthases and Sphingomyelinases. Chem. Phys. Lipids 2016, 197, 45–59. [Google Scholar] [CrossRef]
  85. Taniguchi, M.; Okazaki, T. Role of Ceramide/Sphingomyelin (SM) Balance Regulated through “SM Cycle” in Cancer. Cell Signal. 2021, 87, 110119. [Google Scholar] [CrossRef]
  86. Hao, T.; Wang, Z.; Yang, J.; Zhang, Y.; Shang, Y.; Sun, J. MALAT1 Knockdown Inhibits Prostate Cancer Progression by Regulating MiR-140/BIRC6 Axis. Biomed. Pharm. 2020, 123, 109666. [Google Scholar] [CrossRef]
  87. De Martino, S.; Iorio, E.; Cencioni, C.; Aiello, A.; Spallotta, F.; Chirico, M.; Pisanu, M.E.; Grassi, C.; Pontecorvi, A.; Gaetano, C.; et al. MALAT1 as a Regulator of the Androgen-Dependent Choline Kinase A Gene in the Metabolic Rewiring of Prostate Cancer. Cancers 2022, 14, 2902. [Google Scholar] [CrossRef]
  88. Gordon, A.D.; Biswas, S.; Feng, B.; Chakrabarti, S. MALAT1: A Regulator of Inflammatory Cytokines in Diabetic Complications. Endocrinol. Diab. Metab. 2018, 1, e00010. [Google Scholar] [CrossRef]
  89. Asim, M.; Massie, C.E.; Orafidiya, F.; Pértega-Gomes, N.; Warren, A.Y.; Esmaeili, M.; Selth, L.A.; Zecchini, H.I.; Luko, K.; Qureshi, A.; et al. Choline Kinase Alpha as an Androgen Receptor Chaperone and Prostate Cancer Therapeutic Target. J. Natl. Cancer Inst. 2016, 108, djv371. [Google Scholar] [CrossRef] [Green Version]
  90. Li, J.; Ren, S.; Piao, H.-L.; Wang, F.; Yin, P.; Xu, C.; Lu, X.; Ye, G.; Shao, Y.; Yan, M.; et al. Integration of Lipidomics and Transcriptomics Unravels Aberrant Lipid Metabolism and Defines Cholesteryl Oleate as Potential Biomarker of Prostate Cancer. Sci. Rep. 2016, 6, 20984. [Google Scholar] [CrossRef]
  91. White, M.A.; Lin, C.; Rajapakshe, K.; Dong, J.; Shi, Y.; Tsouko, E.; Mukhopadhyay, R.; Jasso, D.; Dawood, W.; Coarfa, C.; et al. Glutamine Transporters Are Targets of Multiple Oncogenic Signaling Pathways in Prostate Cancer. Mol. Cancer Res. 2017, 15, 1017–1028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Scalise, M.; Pochini, L.; Galluccio, M.; Console, L.; Indiveri, C. Glutamine Transport and Mitochondrial Metabolism in Cancer Cell Growth. Front Oncol. 2017, 7, 306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Scalise, M.; Console, L.; Rovella, F.; Galluccio, M.; Pochini, L.; Indiveri, C. Membrane Transporters for Amino Acids as Players of Cancer Metabolic Rewiring. Cells 2020, 9, 2028. [Google Scholar] [CrossRef]
  94. Fedeles, B.I.; Singh, V.; Delaney, J.C.; Li, D.; Essigmann, J.M. The AlkB Family of Fe(II)/α-Ketoglutarate-Dependent Dioxygenases: Repairing Nucleic Acid Alkylation Damage and Beyond. J. Biol. Chem. 2015, 290, 20734–20742. [Google Scholar] [CrossRef] [Green Version]
  95. Carey, B.W.; Finley, L.W.S.; Cross, J.R.; Allis, C.D.; Thompson, C.B. Intracellular α-Ketoglutarate Maintains the Pluripotency of Embryonic Stem Cells. Nature 2015, 518, 413–416. [Google Scholar] [CrossRef] [Green Version]
  96. Kim, G.; Jang, S.K.; Kim, Y.J.; Jin, H.O.; Bae, S.; Hong, J.; Park, I.C.; Lee, J.H. Inhibition of Glutamine Uptake Resensitizes Paclitaxel Resistance in SKOV3-TR Ovarian Cancer Cell via mTORC1/S6K Signaling Pathway. Int. J. Mol. Sci. 2022, 23, 8761. [Google Scholar] [CrossRef]
  97. Wang, Q.; Hardie, R.-A.; Hoy, A.J.; van Geldermalsen, M.; Gao, D.; Fazli, L.; Sadowski, M.C.; Balaban, S.; Schreuder, M.; Nagarajah, R.; et al. Targeting ASCT2-Mediated Glutamine Uptake Blocks Prostate Cancer Growth and Tumour Development. J. Pathol. 2015, 236, 278–289. [Google Scholar] [CrossRef] [Green Version]
  98. Pan, T.; Gao, L.; Wu, G.; Shen, G.; Xie, S.; Wen, H.; Yang, J.; Zhou, Y.; Tu, Z.; Qian, W. Elevated Expression of Glutaminase Confers Glucose Utilization via Glutaminolysis in Prostate Cancer. BioChem. Biophys. Res. Commun. 2015, 456, 452–458. [Google Scholar] [CrossRef]
  99. Mukha, A.; Kahya, U.; Dubrovska, A. Targeting Glutamine Metabolism and Autophagy: The Combination for Prostate Cancer Radiosensitization. Autophagy 2021, 17, 3879–3881. [Google Scholar] [CrossRef]
  100. Nguyen, T.; Kirsch, B.J.; Asaka, R.; Nabi, K.; Quinones, A.; Tan, J.; Antonio, M.J.; Camelo, F.; Li, T.; Nguyen, S.; et al. Uncovering the Role of N-Acetyl-Aspartyl-Glutamate as a Glutamate Reservoir in Cancer. Cell Rep. 2019, 27, 491–501.e6. [Google Scholar] [CrossRef]
  101. Koochekpour, S.; Majumdar, S.; Azabdaftari, G.; Attwood, K.; Scioneaux, R.; Subramani, D.; Manhardt, C.; Lorusso, G.D.; Willard, S.S.; Thompson, H.; et al. Serum Glutamate Levels Correlate with Gleason Score and Glutamate Blockade Decreases Proliferation, Migration, and Invasion and Induces Apoptosis in Prostate Cancer Cells. Clin. Cancer Res. 2012, 18, 5888–5901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Penney, K.L.; Tyekucheva, S.; Rosenthal, J.; El Fandy, H.; Carelli, R.; Borgstein, S.; Zadra, G.; Fanelli, G.N.; Stefanizzi, L.; Giunchi, F.; et al. Metabolomics of Prostate Cancer Gleason Score in Tumor Tissue and Serum. Mol. Cancer Res. 2021, 19, 475–484. [Google Scholar] [CrossRef] [PubMed]
  103. Lucarelli, G.; Rutigliano, M.; Ferro, M.; Giglio, A.; Intini, A.; Triggiano, F.; Palazzo, S.; Gigante, M.; Castellano, G.; Ranieri, E.; et al. Activation of the Kynurenine Pathway Predicts Poor Outcome in Patients with Clear Cell Renal Cell Carcinoma. Urol. Oncol. 2017, 35, 461.e15–461.e27. [Google Scholar] [CrossRef] [PubMed]
  104. Song, Y.H.; Shiota, M.; Kuroiwa, K.; Naito, S.; Oda, Y. The Important Role of Glycine N-Methyltransferase in the Carcinogenesis and Progression of Prostate Cancer. Mod. Pathol. 2011, 24, 1272–1280. [Google Scholar] [CrossRef] [Green Version]
  105. Lucarelli, G.; Fanelli, M.; Larocca, A.M.V.; Germinario, C.A.; Rutigliano, M.; Vavallo, A.; Selvaggi, F.P.; Bettocchi, C.; Battaglia, M.; Ditonno, P. Serum Sarcosine Increases the Accuracy of Prostate Cancer Detection in Patients with Total Serum PSA Less than 4.0 Ng/Ml. Prostate 2012, 72, 1611–1621. [Google Scholar] [CrossRef] [PubMed]
  106. Lucarelli, G.; Ditonno, P.; Bettocchi, C.; Spilotros, M.; Rutigliano, M.; Vavallo, A.; Galleggiante, V.; Fanelli, M.; Larocca, A.M.V.; Germinario, C.A.; et al. Serum Sarcosine Is a Risk Factor for Progression and Survival in Patients with Metastatic Castration-Resistant Prostate Cancer. Future Oncol. 2013, 9, 899–907. [Google Scholar] [CrossRef]
  107. Ippolito, L.; Morandi, A.; Taddei, M.L.; Parri, M.; Comito, G.; Iscaro, A.; Raspollini, M.R.; Magherini, F.; Rapizzi, E.; Masquelier, J.; et al. Cancer-Associated Fibroblasts Promote Prostate Cancer Malignancy via Metabolic Rewiring and Mitochondrial Transfer. Oncogene 2019, 38, 5339–5355. [Google Scholar] [CrossRef]
  108. Pavlides, S.; Whitaker-Menezes, D.; Castello-Cros, R.; Flomenberg, N.; Witkiewicz, A.K.; Frank, P.G.; Casimiro, M.C.; Wang, C.; Fortina, P.; Addya, S.; et al. The Reverse Warburg Effect: Aerobic Glycolysis in Cancer Associated Fibroblasts and the Tumor Stroma. Cell Cycle 2009, 8, 3984–4001. [Google Scholar] [CrossRef] [Green Version]
  109. Fiaschi, T.; Marini, A.; Giannoni, E.; Taddei, M.L.; Gandellini, P.; De Donatis, A.; Lanciotti, M.; Serni, S.; Cirri, P.; Chiarugi, P. Reciprocal Metabolic Reprogramming through Lactate Shuttle Coordinately Influences Tumor-Stroma Interplay. Cancer Res. 2012, 72, 5130–5140. [Google Scholar] [CrossRef] [Green Version]
  110. Mishra, R.; Haldar, S.; Placencio, V.; Madhav, A.; Rohena-Rivera, K.; Agarwal, P.; Duong, F.; Angara, B.; Tripathi, M.; Liu, Z.; et al. Stromal Epigenetic Alterations Drive Metabolic and Neuroendocrine Prostate Cancer Reprogramming. J. Clin. Investig. 2018, 128, 4472–4484. [Google Scholar] [CrossRef]
  111. Keto, C.J.; Aronson, W.J.; Terris, M.K.; Presti, J.C.; Kane, C.J.; Amling, C.L.; Freedland, S.J. Obesity Is Associated with Castration-Resistant Disease and Metastasis in Men Treated with Androgen Deprivation Therapy after Radical Prostatectomy: Results from the SEARCH Database. BJU Int. 2012, 110, 492–498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Laurent, V.; Toulet, A.; Attané, C.; Milhas, D.; Dauvillier, S.; Zaidi, F.; Clement, E.; Cinato, M.; Le Gonidec, S.; Guérard, A.; et al. Periprostatic Adipose Tissue Favors Prostate Cancer Cell Invasion in an Obesity-Dependent Manner: Role of Oxidative Stress. Mol. Cancer Res. 2019, 17, 821–835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Kapoor, J.; Namdarian, B.; Pedersen, J.; Hovens, C.; Moon, D.; Peters, J.; Costello, A.J.; Ruljancich, P.; Corcoran, N.M. Extraprostatic Extension into Periprostatic Fat Is a More Important Determinant of Prostate Cancer Recurrence than an Invasive Phenotype. J. Urol. 2013, 190, 2061–2066. [Google Scholar] [CrossRef] [PubMed]
  114. Altuna-Coy, A.; Ruiz-Plazas, X.; Sánchez-Martin, S.; Ascaso-Til, H.; Prados-Saavedra, M.; Alves-Santiago, M.; Bernal-Escoté, X.; Segarra-Tomás, J.; Chacón, M.R. The Lipidomic Profile of the Tumoral Periprostatic Adipose Tissue Reveals Alterations in Tumor Cell’s Metabolic Crosstalk. BMC Med. 2022, 20, 255. [Google Scholar] [CrossRef] [PubMed]
  115. Finley, D.S.; Calvert, V.S.; Inokuchi, J.; Lau, A.; Narula, N.; Petricoin, E.F.; Zaldivar, F.; Santos, R.; Tyson, D.R.; Ornstein, D.K. Periprostatic Adipose Tissue as a Modulator of Prostate Cancer Aggressiveness. J. Urol. 2009, 182, 1621–1627. [Google Scholar] [CrossRef] [PubMed]
  116. Saha, A.; Ahn, S.; Blando, J.; Su, F.; Kolonin, M.G.; DiGiovanni, J. Proinflammatory CXCL12-CXCR4/CXCR7 Signaling Axis Drives Myc-Induced Prostate Cancer in Obese Mice. Cancer Res. 2017, 77, 5158–5168. [Google Scholar] [CrossRef] [Green Version]
  117. Laurent, V.; Guérard, A.; Mazerolles, C.; Le Gonidec, S.; Toulet, A.; Nieto, L.; Zaidi, F.; Majed, B.; Garandeau, D.; Socrier, Y.; et al. Periprostatic adipocytes act as a driving force for prostate cancer progression in obesity. Nat. Commun. 2016, 7, 10230. [Google Scholar] [CrossRef] [Green Version]
  118. Calgani, A.; Delle Monache, S.; Cesare, P.; Vicentini, C.; Bologna, M.; Angelucci, A. Leptin Contributes to Long-Term Stabilization of HIF-1α in Cancer Cells Subjected to Oxygen Limiting Conditions. Cancer Lett. 2016, 376, 1–9. [Google Scholar] [CrossRef]
  119. Genard, G.; Lucas, S.; Michiels, C. Reprogramming of Tumor-Associated Macrophages with Anticancer Therapies: Radiotherapy versus Chemo- and Immunotherapies. Front Immunol. 2017, 8, 828. [Google Scholar] [CrossRef] [Green Version]
  120. Krstic, J.; Trivanovic, D.; Jaukovic, A.; Santibanez, J.F.; Bugarski, D. Metabolic Plasticity of Stem Cells and Macrophages in Cancer. Front Immunol. 2017, 8, 939. [Google Scholar] [CrossRef]
  121. Jamaspishvili, T.; Berman, D.M.; Ross, A.E.; Scher, H.I.; De Marzo, A.M.; Squire, J.A.; Lotan, T.L. Clinical Implications of PTEN Loss in Prostate Cancer. Nat. Rev. Urol. 2018, 15, 222–234. [Google Scholar] [CrossRef] [PubMed]
  122. Song, M.S.; Salmena, L.; Pandolfi, P.P. The Functions and Regulation of the PTEN Tumour Suppressor. Nat. Rev. Mol. Cell Biol. 2012, 13, 283–296. [Google Scholar] [CrossRef] [PubMed]
  123. Majumder, P.K.; Febbo, P.G.; Bikoff, R.; Berger, R.; Xue, Q.; McMahon, L.M.; Manola, J.; Brugarolas, J.; McDonnell, T.J.; Golub, T.R.; et al. MTOR Inhibition Reverses Akt-Dependent Prostate Intraepithelial Neoplasia through Regulation of Apoptotic and HIF-1-Dependent Pathways. Nat. Med. 2004, 10, 594–601. [Google Scholar] [CrossRef] [PubMed]
  124. Wise, D.R.; DeBerardinis, R.J.; Mancuso, A.; Sayed, N.; Zhang, X.-Y.; Pfeiffer, H.K.; Nissim, I.; Daikhin, E.; Yudkoff, M.; McMahon, S.B.; et al. Myc Regulates a Transcriptional Program That Stimulates Mitochondrial Glutaminolysis and Leads to Glutamine Addiction. Proc. Natl. Acad. Sci. USA 2008, 105, 18782–18787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Gao, P.; Tchernyshyov, I.; Chang, T.-C.; Lee, Y.-S.; Kita, K.; Ochi, T.; Zeller, K.I.; De Marzo, A.M.; Van Eyk, J.E.; Mendell, J.T.; et al. C-Myc Suppression of MiR-23a/b Enhances Mitochondrial Glutaminase Expression and Glutamine Metabolism. Nature 2009, 458, 762–765. [Google Scholar] [CrossRef] [Green Version]
  126. Jones, S.F.; Infante, J.R. Molecular Pathways: Fatty Acid Synthase. Clin. Cancer Res. 2015, 21, 5434–5438. [Google Scholar] [CrossRef] [Green Version]
  127. de Wet, L.; Williams, A.; Gillard, M.; Kregel, S.; Lamperis, S.; Gutgesell, L.C.; Vellky, J.E.; Brown, R.; Conger, K.; Paner, G.P.; et al. SOX2 Mediates Metabolic Reprogramming of Prostate Cancer Cells. Oncogene 2022, 41, 1190–1202. [Google Scholar] [CrossRef]
  128. Loizzo, D.; Pandolfo, S.D.; Rogers, D.; Cerrato, C.; di Meo, N.A.; Autorino, R.; Mirone, V.; Ferro, M.; Porta, C.; Stella, A.; et al. Novel Insights into Autophagy and Prostate Cancer: A Comprehensive Review. Int. J. Mol. Sci. 2022, 23, 3826. [Google Scholar] [CrossRef]
  129. Guo, J.Y.; White, E. Autophagy, Metabolism, and Cancer. Cold Spring Harb. Symp. Quant. Biol. 2016, 81, 73–78. [Google Scholar] [CrossRef]
  130. Chiacchiera, F.; Simone, C. The AMPK-FoxO3A Axis as a Target for Cancer Treatment. Cell Cycle Georget. Tex 2010, 9, 1091–1096. [Google Scholar] [CrossRef]
  131. Chiacchiera, F.; Matrone, A.; Ferrari, E.; Ingravallo, G.; Lo Sasso, G.; Murzilli, S.; Petruzzelli, M.; Salvatore, L.; Moschetta, A.; Simone, C. P38alpha Blockade Inhibits Colorectal Cancer Growth in Vivo by Inducing a Switch from HIF1alpha- to FoxO-Dependent Transcription. Cell Death Differ. 2009, 16, 1203–1214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Lin, C.; Blessing, A.M.; Pulliam, T.L.; Shi, Y.; Wilkenfeld, S.R.; Han, J.J.; Murray, M.M.; Pham, A.H.; Duong, K.; Brun, S.N.; et al. Inhibition of CAMKK2 Impairs Autophagy and Castration-Resistant Prostate Cancer via Suppression of AMPK-ULK1 Signaling. Oncogene 2021, 40, 1690–1705. [Google Scholar] [CrossRef] [PubMed]
  133. Nazim, U.M.; Yin, H.; Park, S.-Y. Neferine Treatment Enhances the TRAIL-induced Apoptosis of Human Prostate Cancer Cells via Autophagic Flux and the JNK Pathway. Int. J. Oncol. 2020, 56, 1152–1161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Castaldo, R.; Cavaliere, C.; Soricelli, A.; Salvatore, M.; Pecchia, L.; Franzese, M. Radiomic and Genomic Machine Learning Method Performance for Prostate Cancer Diagnosis: Systematic Literature Review. J. Med. Internet. Res. 2021, 23, e22394. [Google Scholar] [CrossRef]
  135. Ferro, M.; de Cobelli, O.; Musi, G.; Del Giudice, F.; Carrieri, G.; Busetto, G.M.; Falagario, U.G.; Sciarra, A.; Maggi, M.; Crocetto, F.; et al. Radiomics in Prostate Cancer: An up-to-Date Review. Ther. Adv. Urol 2022, 14, 17562872221109020. [Google Scholar] [CrossRef]
  136. Ferro, M.; de Cobelli, O.; Vartolomei, M.D.; Lucarelli, G.; Crocetto, F.; Barone, B.; Sciarra, A.; Del Giudice, F.; Muto, M.; Maggi, M.; et al. Prostate Cancer Radiogenomics-From Imaging to Molecular Characterization. Int. J. Mol. Sci. 2021, 22, 9971. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Oncoprints of fatty acid (FA) metabolism genes in the cancer genome atlas (TCGA, PanCancer Atlas) PCa patient cohort (PRAD). About 88% of the tumors from 489 patients in PRAD showed altered gene copy number or expression of FA metabolism-related genes. ACLY: ATP citrate lyase; ACACA: acetyl-CoA carboxylase alpha (ACC); FASN: fatty acid synthase; ACCS2: acyl-CoA synthetase short-chain family member 2; MLYCD: malonyl-CoA decarboxylase; SCD: stearoyl-CoA desaturase; SREBF1: sterol regulatory element-binding transcription factor 1 (SPREBP1); CPT1A: carnitine palmitoyltransferase 1A; PPARA: peroxisome proliferator-activated receptor alpha; GPAM: glycerol-3-phosphate acyltransferase; AGPAT: 1-acylglycerol-3-phosphate O-acyltransferase; LPCAT1: lysophosphatidylcholine acyltransferase 1; LPIN1: lipin 1; DGAT1: diacylglycerol O-acyltransferase 1; PNPLA2: patatin-like phospholipase domain containing 2; LIPE: lipase, hormone-sensitive; MGLL: monoglyceride lipase.
Figure 1. Oncoprints of fatty acid (FA) metabolism genes in the cancer genome atlas (TCGA, PanCancer Atlas) PCa patient cohort (PRAD). About 88% of the tumors from 489 patients in PRAD showed altered gene copy number or expression of FA metabolism-related genes. ACLY: ATP citrate lyase; ACACA: acetyl-CoA carboxylase alpha (ACC); FASN: fatty acid synthase; ACCS2: acyl-CoA synthetase short-chain family member 2; MLYCD: malonyl-CoA decarboxylase; SCD: stearoyl-CoA desaturase; SREBF1: sterol regulatory element-binding transcription factor 1 (SPREBP1); CPT1A: carnitine palmitoyltransferase 1A; PPARA: peroxisome proliferator-activated receptor alpha; GPAM: glycerol-3-phosphate acyltransferase; AGPAT: 1-acylglycerol-3-phosphate O-acyltransferase; LPCAT1: lysophosphatidylcholine acyltransferase 1; LPIN1: lipin 1; DGAT1: diacylglycerol O-acyltransferase 1; PNPLA2: patatin-like phospholipase domain containing 2; LIPE: lipase, hormone-sensitive; MGLL: monoglyceride lipase.
Ijms 24 00910 g001
Figure 2. Overview of prostate cancer cell metabolism and its crosstalk with stromal components. TAF: tumor-associated fibroblast; MCT4: monocarboxylate transporter 4; MCT1: monocarboxylate transporter 1; GLUT: glucose transporter; G6P: glucose 6-phosphate; PPP: pentose phosphate pathway; α-KG: α-ketoglutarate; OAA: oxaloacetate; Ac-CoA: acyl-CoA; ASCT2 (SLC1A5): neutral amino acid transporter; ATP: adenosine triphosphate.
Figure 2. Overview of prostate cancer cell metabolism and its crosstalk with stromal components. TAF: tumor-associated fibroblast; MCT4: monocarboxylate transporter 4; MCT1: monocarboxylate transporter 1; GLUT: glucose transporter; G6P: glucose 6-phosphate; PPP: pentose phosphate pathway; α-KG: α-ketoglutarate; OAA: oxaloacetate; Ac-CoA: acyl-CoA; ASCT2 (SLC1A5): neutral amino acid transporter; ATP: adenosine triphosphate.
Ijms 24 00910 g002
Table 1. Cancer drugs targeting metabolic pathways that are approved, or under investigation in clinical trials.
Table 1. Cancer drugs targeting metabolic pathways that are approved, or under investigation in clinical trials.
Metabolic PathwayTargetDrug
GlycolysisGLUT1WZB117, BAY-876
Hexokinase2-Deoxyglucose
Pyruvate Kinase M2 (PKM2)TEPP-46
Lactate dehydrogenase A (LDHA)Quinoline, 3-sulfonamides, FX11, PSTMB
Monocarboxylate transporter 1 (MCT1)AZD3965
Amino acid metabolismGlutaminase 1 (GLS1)CB-839, IPN60090
ASCT2 (SLC1A5)GPNA
Multiple targetsJHU-083, DRP-104
Phosphoglycerate dehydrogenase (PHGDH)CBR-5884, NCT-503
Indoleamine-2,3-dioxygenase-1 (IDO1)Epacadostat, indoximod
Circulating asparagineL-Asparaginase
Large neutral amino acid transporter (LAT1)JPH203
Mitochondrial metabolismPyruvate dehydrogenase (PDH), α-ketoglutarate dehydrogenaseCPI-613
Electron transport chain complex 1Metformin, IACS-010759, IM156
Lipid metabolismATP-citrate lyase (ACLY)SB-204990
Acetyl-CoA carboxylase (ACC)Soraphen-A
fatty acid synthase (FASN)TVB-2640
Enzymes mutated in cancerMutant isocitrate dehydrogenase 1 (IDH1)Ivosidenib, IDH305, BAY1436032, LY3410738, FT-2102
Mutant isocitrate dehydrogenase 2 (IDH2)Enasidenib
Nucleic acid synthesisThymidylate synthase (TS)5-Fluorouracil, capecitabine, pemetrexed, raltitrexed
Dihydrofolate reductase (DHFR)Methotrexate, pemetrexed
Glycinamide ribonucleotide formyltransferase (GARFT)Pemetrexed
Dihydroorotate dehydrogenase (DHODH)Brequinar, leflunomide
Ribonucleotide reductase (RNR)Gemcitabine, clofarabine, fludarabine, cladribine, cytarabine
5-Phosphoribosyl-1-pyro-phosphatase (PRPP) amidotransferaseMercaptopurine, thioguanine
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

Lasorsa, F.; di Meo, N.A.; Rutigliano, M.; Ferro, M.; Terracciano, D.; Tataru, O.S.; Battaglia, M.; Ditonno, P.; Lucarelli, G. Emerging Hallmarks of Metabolic Reprogramming in Prostate Cancer. Int. J. Mol. Sci. 2023, 24, 910. https://doi.org/10.3390/ijms24020910

AMA Style

Lasorsa F, di Meo NA, Rutigliano M, Ferro M, Terracciano D, Tataru OS, Battaglia M, Ditonno P, Lucarelli G. Emerging Hallmarks of Metabolic Reprogramming in Prostate Cancer. International Journal of Molecular Sciences. 2023; 24(2):910. https://doi.org/10.3390/ijms24020910

Chicago/Turabian Style

Lasorsa, Francesco, Nicola Antonio di Meo, Monica Rutigliano, Matteo Ferro, Daniela Terracciano, Octavian Sabin Tataru, Michele Battaglia, Pasquale Ditonno, and Giuseppe Lucarelli. 2023. "Emerging Hallmarks of Metabolic Reprogramming in Prostate Cancer" International Journal of Molecular Sciences 24, no. 2: 910. https://doi.org/10.3390/ijms24020910

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