Next Article in Journal
Effect of Temperature on the Development of Stages of Spermatogenesis and the Functionality of Sertoli Cells In Vitro
Next Article in Special Issue
A Step Forward for the Treatment of Localized Prostate Cancer Using Gold Nanoparticles Combined with Laser Irradiation
Previous Article in Journal
pOsHAK1:OsSUT1 Promotes Sugar Transport and Enhances Drought Tolerance in Rice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Prostate Cancer and the Mevalonate Pathway

by
Patricia Guerrero-Ochoa
1,
Sergio Rodríguez-Zapater
2,
Alberto Anel
3,
Luis Mariano Esteban
4,*,
Alejandro Camón-Fernández
1,
Raquel Espilez-Ortiz
1,5,6,
María Jesús Gil-Sanz
1,5 and
Ángel Borque-Fernando
1,4,5,6,*
1
Health Research Institute of Aragon Foundation, 50009 Zaragoza, Spain
2
Minimally Invasive Research Group (GITMI), Faculty of Veterinary Medicine, University of Zaragoza, 50009 Zaragoza, Spain
3
Department of Biochemistry and Molecular and Cellular Biology, Faculty of Sciences, University of Zaragoza, 50009 Zaragoza, Spain
4
Department of Applied Mathematics, Escuela Universitaria Politécnica de La Almunia, Institute for Biocomputation and Physic of Complex Systems, Universidad de Zaragoza, 50100 La Almunia de Doña Godina, Spain
5
Department of Urology, Miguel Servet University Hospital, 50009 Zaragoza, Spain
6
Area of Urology, Department of Surgery, Faculty of Medicine, University of Zaragoza, 50009 Zaragoza, Spain
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(4), 2152; https://doi.org/10.3390/ijms25042152
Submission received: 10 January 2024 / Revised: 4 February 2024 / Accepted: 7 February 2024 / Published: 10 February 2024

Abstract

:
Antineoplastic therapies for prostate cancer (PCa) have traditionally centered around the androgen receptor (AR) pathway, which has demonstrated a significant role in oncogenesis. Nevertheless, it is becoming progressively apparent that therapeutic strategies must diversify their focus due to the emergence of resistance mechanisms that the tumor employs when subjected to monomolecular treatments. This review illustrates how the dysregulation of the lipid metabolic pathway constitutes a survival strategy adopted by tumors to evade eradication efforts. Integrating this aspect into oncological management could prove valuable in combating PCa.

1. Introduction

The centrality of AR in orchestrating metabolic reprogramming for the growth and advancement of PCa has been established in recent research [1]. Moreover, it has been demonstrated that the interplay between AR and the serum response element (SRE) gene regulatory sequence, along with other non-traditional pathways, assumes crucial significance in the context of castration-resistant PCa (CRPC). Significantly, a substantial body of research is currently focusing on this particular domain. In this context, cholesterol is one of the lipids that has great relevance due to its role in the structure of the cell membrane as well as in signaling pathways linked to the AR pathway, not only due to its precursor role in the generation of androgens, which are the main AR ligands currently used as a target for therapies. While fundamental evidence underscores the profound influence of lipids and, specifically, cholesterol on tumors, the clinical sphere remains marked by conflicting outcomes. Due to this strong evidence, several clinical trials using statins have been launched, and these trials have confirmed the important roles of cholesterol and drugs designed to lower its serum levels in the treatment of PCa.
The current research initiatives are prominently centered around delineating the lipidome, transcriptomics, and metabolomics, entailing a comprehensive quantitative assessment of lipids and metabolites through the application of liquid or gas chromatography coupled with mass spectrometry. These efforts are directed towards exploring their implications across various disease spectrums, encompassing aging and cancer, as highlighted by recent findings [2]. Over the past decade, compelling evidence has emerged, showcasing distinct lipid species originating from diverse biological sources that distinguish PCa patients from controls [3,4,5].

2. The Mevalonate Pathway

De novo cholesterol generation is facilitated by the utilization of glucose, glutamine, and acetate as nutrients which contribute to the production of acetyl-coenzyme A (acetyl-CoA), a primary substrate for the mevalonate (MVA) pathway [6,7,8]. However, the major source of acetyl-CoA originates from fatty acids (FAs) [4], whose synthesis, uptake, β-oxidation, and accumulation are upregulated to support the enhanced cellular proliferation and heightened demands for plasma membrane synthesis and energy production in cancer cells [9]. Intracellular lipid accumulation in lipid droplets and tumor-associated adipocytes facilitates survival under both normoxic and hypoxic conditions [10,11]. Lipid droplets play a crucial role in membrane remodeling and composition through the trafficking of fatty acids among different lipid classes stored in diverse lipid pools and organelles [12]. Notably, the prominent fatty acid transporter CD36 is frequently gained or amplified in PCa, correlating with poor prognoses [13].
Acetyl-CoA emerges as the product of enzymatic activity orchestrated by acetyl-CoA synthase short-chain family member (ACSS), ATP-citrate lyase (ACLY), and the pyruvate dehydrogenase complex (PDC) [14]. Notably, these metabolic enzymes translocate to the nucleus, collaborating with diverse transcriptional factors and regulators to modulate gene expression through the acetylation of histones [15,16,17]. The orchestration of DNA methylation, histone posttranslational modifications (PTMs), and noncoding RNAs (ncRNAs) significantly impacts gene expression in PCa, contributing to tumorigenesis and its progression [18]. Within the context of PCa, a nuclear subunit of PDC, pyruvate dehydrogenase A1 (PDHA1), contributes to the expression of sterol regulatory element-binding protein (SREBP)-dependent genes involved in lipid synthesis in both mitochondria and nuclei, thereby sustaining tumorigenesis [19]. Consequently, acetyl-CoA serves as the pivotal metabolite initiating the MVA pathway, which also influences histone acetylation, representing a crucial mechanism for non-mutational epigenetic reprogramming. Moreover, acetyl-CoA functions as the primary source of nicotinamide adenine dinucleotide phosphate (NADPH), the principal antioxidant in cancer cells, facilitating the management of elevated levels of reactive oxygen species (ROS) typically generated to maintain signaling and various cellular processes within tumors [20,21,22,23].
The condensation of three molecules of acetyl-CoA by 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) synthase sequentially generates HMG-CoA, which is then converted into MVA through the action of HMG-CoA reductase (HMG-CR). The subsequent conversion of MVA into isopentenyl diphosphate (IPP) leads to the production of vital metabolites such as farnesyl pyrophosphate (FPP), geranyl pyrophosphate (GPP), heme A, and coenzyme Q10 (ubiquinone) [24,25]. The metabolization of FPP generates lanosterol, which is further metabolized into cholesterol through an enzymatic pathway [26]. Cholesterol serves as the precursor for various crucial compounds, such as vitamin D, bile salts, oxysterols, and steroid hormones, with the latter being known to initiate and drive the progression of PCa [27].
Notably, MVA metabolites have been shown to regulate conventional and non-conventional T lymphocytes (LT) at various levels. Specifically, γδ LT recognizes lipid antigens independent of major histocompatibility complex (MHC) class I presentation, originating from the MVA pathway, with IPP being the most relevant, presented by the ubiquitous protein butyrophilin 3A1 (BTN3A1, also known as CD277), which binds to phosphorylated antigens [28]. Enzymes derived from the MVA pathway are also responsible for post-translational modifications in RAS and Rho families, essential for their functional regulation, influencing various cancer-relevant pathways, including cytoskeletal organization, vesicle trafficking, gene expression, cell signaling, cell cycle, motility, and cell survival, supporting tumor initiation, growth, metastasis, and therapy resistance [29]. The inhibition of the MVA pathway promotes cancer cell death by activating the intrinsic apoptosis pathway [30].
During the transformation of PCa cells, there is a progressive reactivation of mitochondrial oxidative phosphorylation (OXPHOS) alongside increased glucose metabolism and decreased citrate production [31,32]. Androgen signalling reprograms global cellular metabolic pathways, including aerobic glycolysis, mitochondrial respiration, and de novo lipogenesis, to meet the metabolic and biosynthetic demands of PCa cells [33,34,35]. Long-term androgen treatment also increases mitochondrial biogenesis and activity [35].
The crucial genes involved in the initiation of the MVA pathway enzymes are chiefly under the regulatory control of the SREBP family. SREBP1a and SREBP1c are transcribed from alternative start sites in the SREBPF1 gene, with the former regulating the expression of MVA and fatty acid metabolism genes and the latter predominantly governing the expression of fatty acid metabolism genes. SREBP2, transcribed from the SREBPF2 gene, functions as the principal transcription factor for genes associated with the MVA pathway [26,36,37]. All these considerations are shown in a schematic way in Figure 1.

3. Steroidogenesis and the Androgen Receptor Pathway

Under normal physiological conditions, SREBP2, in the presence of cholesterol and oxysterols, is sequestered in the endoplasmic reticulum (ER), remaining bound to the SREBP cleavage-activating protein (SCAP) complex, along with insulin-induced genes (INSIGs). Upon sterol deprivation, INSIGs undergo ubiquitination and subsequent degradation. SREBP2, along with SCAP, is then transported to the Golgi apparatus for cleavage by site-1 protease (SP1) and site-2 protease (SP2) before translocating to the nucleus to bind to SRE on the DNA [38,39,40].
In contrast, SREBP1 is not strongly regulated by sterol levels; rather, it is inhibited by polyunsaturated fatty acids (PUFAs) at multiple levels, including amino acids and eIF-2α kinase. It can be induced by insulin, high-glucose conditions, liver X receptors, atypical protein kinase C (PKC) isoforms, and fatty acids [40,41]. Notably, oxysterols, derivatives of cholesterol, can activate liver X receptors, which function as regulators of lipid flux and activators of SREBP1c [27].

4. Interrelation between Prostate Cancer, Androgen Receptor, and Cholesterol

4.1. The Role of Cholesterol in Plasma Membrane Lipid Rafts

The dynamic functionality of the plasma membrane is underpinned by specialized regions rich in cholesterol and sphingolipids, known as “Lipid Rafts” (LRs), which have the capacity to modulate the accessibility of proteins to regulatory or effector molecules. Cholesterol constitutes approximately 30% of the lipids in the plasma membrane and plays a pivotal role in closely packing the acyl chains of phospholipids, thereby promoting phase separation and stabilizing the formation of LRs [42]. Cancer cells exhibit higher levels of intracellular cholesterol and LRs compared to their normal counterparts [43]. For instance, in PCa, cell lines like PC-3 have been found to possess LRs at levels five times greater than those in normal cells [44].
Lipid rafts play a crucial role in sustaining proliferative signaling by hosting receptors organized within a structure referred to as CASMER, short for “cluster of apoptotic signaling molecule-enriched rafts,” where death receptors are concentrated [45,46,47]. This implies that both types of signals, those promoting cell survival and those leading to cell death, are interconnected in a dynamic manner. CASMER communicate with other subcellular structures to facilitate the transmission of apoptotic signals or induce signals related to cell survival, proliferation, inflammation, cancer growth, and metastasis. In the context of mTOR complex 1 (mTORC1) signaling, molecules like insulin-like growth factor (IGFs) and the epidermal growth factor (EGF) are involved [48].
Cholesterol is necessary to maintain lipid raft structure since the removal of cholesterol renders these structures non-functional and leads to the dissociation of proteins from the rafts [47,49]. Edelfosine, an ether-phospholipid with a high affinity for cholesterol, has emerged as an antitumor drug that acts through lipid rafts, promoting cell death by co-clustering raft-associated Fas/CD95 receptors and inducing apoptosis in a wide range of tumor cells [48]. PCa cells induce senescence in a subset of immunosuppressive neutrophils by secreting apolipoprotein E (APOE), which binds to the Triggering receptor expressed in myeloid cells 2 (TREM2). Senescence arrest cell proliferation in a stable manner but cells remain metabolically active, resist apoptosis and induce tolerogenic state in T cells. APOE belongs to a family of lipid-binding proteins that serve as major transporters of systemic cholesterol and triglycerides. The APOE/TREM2 axis correlates with a poor prognosis in PCa patients. Specifically eliminating senescent immunosuppressive neutrophils using histone deacetylase inhibitors enhances the effectiveness of standard therapeutic approaches [50].

4.2. Growth Factor Receptors Reside Inside Lipid Rafts

In the context of PCa, elevated cholesterol levels promote disease progression by inducing the epithelial-to-mesenchymal transition (EMT) through the activation of the extracellular-regulated protein kinases 1/2 (ERK1/2) pathway, mediated by the EGFR, and the accumulation of adipocyte plasma membrane-associated proteins (APMAPs) in LRs. Notably, high cholesterol levels in LRs inhibit EGFR internalization. Consistently, both APMAP mRNA and protein levels are elevated in clinical PCa samples [51]. HER2, also called EGFR 2, has been detected on the cell surface of quiescent (G0) PCa cells, which often exhibit resistance to chemotherapy and are associated with recurrence following dormancy [52].
Another member of the EGFR family, HER3, has been found to be overexpressed in human CRPC samples, correlating with enhanced proliferation and survival. Notably, these effects can be counteracted by a neutralizing anti-neuregulin 1 (NRG1), which serves as the ligand for HER3 [53]. NRG1 is often generated by fibroblasts. In addition to these receptors, G protein-coupled receptors (GPCRs), particularly the G protein-coupled estrogen receptor (GPER), are detected in the cytoplasm of basal epithelial cells, as well as in Gleason patterns 2 or 3, but not in luminal secretory epithelial cells. Various mechanisms associated with GPCRs have been reported in PCa, either alone or in conjunction with estrogen receptors (ERs) [54,55,56].
Similarly, lipid rafts (LRs) are significantly enriched in cancer stem cells (CSCs) compared to non-stem cancer cells, and key CSC markers, such as CD24, CD44, and CD133, are located within these lipid rafts, underscoring their pivotal role in CSCs. LRs play a crucial function in facilitating CSC self-renewal, the EMT, drug resistance, and the CSC niche [57,58]. In the context of PCa, CSCs exhibit heightened expression levels of CD44 and CD133 [59]. In vitro assays have demonstrated that cholesterol depletion disrupts lipid rafts, inhibits cancer cell migration, and abolishes β-catenin activation and the CSC phenotype [58], emphasizing the involvement of cholesterol in regulating phenotypic plasticity.

4.3. Interaction of Cholesterol with Androgen Receptor Signaling, the Critical Route of Prostate Cancer

AR expression has been detected in nearly all primary and metastatic PCa cases, regardless of stage or grade, and is sustained in the majority of androgen-independent PCa and CRPC. Intense nuclear AR staining in CRPC bone metastases is associated with a worse clinical outcome [60,61]. Some CRPC cases that emerge in patients are linked to the hyperactivation of the androgen signaling pathway due to various factors, including AR gene amplification, AR mutations, the expression of constitutively active AR splice variants (with AR-V7 being the most common variant found in 75% of metastatic PCa cases), intratumoral androgen synthesis, the altered expression and function of AR coactivators, and the signaling crosstalk with other oncogenic pathways [1,62,63]. AR mutations have been identified in 60% of metastatic tumors [64,65] and 1% of primary PCa cases [65]. Functionally, these mutations enable antiandrogens to act as AR agonists and activate AR through estrogens, progesterone, dehydroepiandrosterone (DHEA), androstenediol, and glucocorticoids [66,67,68]. Studies in both patients and animal models have demonstrated that certain point mutations convert AR antagonists into potent agonists [69]. AR-V7 is an aberrantly spliced mRNA isoform of AR, leading to a protein lacking the C-terminal ligand-binding domain but retaining the transcriptionally active N-terminal domain. Despite its inability to bind ligands, AR-V7 remains constitutively active in a ligand-independent manner and can drive the growth of CRPC [62]. Also, APUC genes have been described as important predictors of clinical response in CRPC, belonging to a family characterized by androgen production, uptake, and conversion, based on genomic analyses of patient germlines [70].
Cholesterol derived from lipid rafts in PCa cell lines has been demonstrated to interact with the N-terminal domain of testosterone-activated AR and form complexes with transient receptor potential melastatin 8 (TRPM8). This interaction inhibits the function of TRPM8, a non-selective cation channel involved in the calcium signaling of tumor and stromal cells and has been proposed as a molecular target in cancer development and progression [71,72,73]. This interaction significantly enhances the migratory capacity of PCa cells, promoting invasion and metastasis. Activated AR has a dose-dependent regulatory effect on the function of the TRPM8 channel [74], in addition to its known ability to induce the transcription of TRPM8 through binding to androgen response elements (AREs) [75]. Likewise, prostate-specific antigen (PSA) has also been observed to modulate TRPM8, facilitating its accumulation in the plasma membrane, similar to the function described for AR but mediated by a G protein-coupled membrane receptor and AKT signaling associated with activated AR [76,77]. However, various findings suggest that additional factors might modulate the cell response to the TRPM8 channel [72]. It is plausible that these factors collectively influence the TRPM8 channel response in PCa [73,78].
SREBP1 has been implicated in PCa cell proliferation, migration, and invasion. Studies have revealed that an AR/mTOR complex promotes SREBF1 expression and activity, primarily due to the constitutive activation of mTOR resulting from the loss of function of PTEN [79]. Moreover, the AR-SREBP1 axis has been identified, forming a self-regulating interaction between the activation of SCAP by AR and SREBP1, which subsequently increases the expression of AR [80,81]. Androgens are recognized as key activators of SREBPs under both physiological conditions and in steroid-regulated cancers [82]. One form of regulation occurs through the stimulation of SCAP expression [83], leading to a switch in the isoform expression of INSIG [84] and the upregulation of several enzymes involved in fatty acid and cholesterol synthesis under the transcriptional control of AR [13]. Perturbations in cholesterol metabolism have been predominantly identified in high-grade tumors [3]. Correspondingly, the inhibition of fatty acid synthase (FASN) enhances the immune response to castration-resistant PCa cell line tumors and decreases the expression and transcriptional activity of AR and the AR-variant splicing 7, resulting in a better response to hormone treatment [85]. The secretion of cholesterol into the intracellular space is regulated by ATP-binding cassette (ABC) transporters, and data have shown the tissue-dependent regulation of ABC transporters by estrogens, progesterone, and androgens [86].

4.4. The Androgen Receptor–Microbiome–Diet Axis in Prostate Cancer

The gut microbiome is linked with diet; also, the immune response is connected to the modulation of immune function, resulting in a different response to immune checkpoint therapy and serving as a source of testosterone. Men with CRPC have an increased abundance of gut bacteria with androgenic functions. Men with high-risk PCa share a specific gut microbial profile. Most importantly, a high-fat diet causes gut dysbiosis and gut bacterial metabolites such as short-chain fatty acids, propionate, and acetate, which enter systemic circulation, resulting in the promotion of PCa growth; this is called the “gut-prostate axis” [87,88]. Humans can utilize absorbed acetate and propionate as substrates for lipid, glucose, and cholesterol metabolism [89]. Androgen deprivation in mice and humans promotes the expansion of defined commensal microbiota capable of converting androgen precursors into active androgens. The depletion of these microbiota and their replacement with Prevotella stercorea control the growth of PCa [90]. Clostridium scindens from fecal human microbiota can convert glucocorticoids into androgens [91], and similarly, Ruminococcus gnavus metabolize pregnenolone to androgenic steroids [90].
In 2016, the International Agency for Research on Cancer (IARC) recommended the avoidance of weight gain based on a body of evidence from around 1000 studies linking obesity to at least 13 types of cancer. This correlation is often linked to higher levels of lipids in the bloodstream [92]. Obesity and altered lipids are associated with the induction of vasculature formation and inflammation through the ETS transcription factor ERG, a gene that encodes key regulators of embryonic development, cell proliferation, angiogenesis, inflammation, and apoptosis. In PCa, the protein ERG contains an ETS DNA-binding domain and a PNT (pointed) domain which is implicated in the self-association of chimeric oncoproteins like TMPSSR2-ERG and NDRG1-ERG [3,93,94]. Individual metabolites correlate with Gleason scores and impact the translocation of ERG, highlighting altered fatty acid oxidation. Specifically, cholesterol has been shown to be negatively correlated with Gleason score [3]. In fact, metabolomic differences between Gleason 4 + 3 and Gleason 3 + 4 have been described, suggesting that lipids could be considered markers of PCa aggression [95], which is related to the identification of metabolic signatures of PCa [3,5], similar to other new personalized signatures developed in early PCa based on multimodal biomarkers [96].

4.5. The Role of the cGAS–Sting Pathway in Prostate Cancer

The cGAS–STING pathway plays an important role in the immune defense against both infections and cancer. Upon the activation of the STING protein, various signaling cascades are initiated, ultimately leading to the activation of interferon regulatory factor 3 (IRF3), nuclear factor-kappa B (NF-κB), and the expression of interferon (IFN) and autophagy [97]. Acting as a vital component of the genomic stability system within the cytosol and endolysosomal compartments, the cyclic GMP-AMP synthase (cGAS) detects double-stranded DNA (dsDNA) and functions as an adenosine monophosphate (AMP) synthase [98].
In the context of PCa, which is often characterized as immunologically “cold” due to limited responses to checkpoint inhibitory therapy, the significance of the STING pathway is underscored. STING activation leads to the upregulation of genes associated with antigen presentation, Th1 chemokine signaling, interferon response, and the expression of programmed cell death protein 1 (PD-L1). These functions of the STING pathway imply its potential relevance in augmenting the immunogenicity of PCa and enhancing the efficacy of immune-based therapies. Consistent with these observations, the enhanced expression of enhancer of zeste homolog 2 (EZH2), a negative regulator of the STING pathway, has been linked to the heightened intratumoral trafficking of activated CD8+ lymphocytes (LT), an increase in M1 tumor-associated macrophages (TAMs), and the reversal of resistance to PD-1 checkpoint therapy [99]. These findings are in consonance with Speckle-type POZ protein (SPOP), another modulator of dsDNA, which is mutated in 15% of PCa patients. SPOP is crucial for regulating the balance between immunosuppressive and antitumor activities downstream of STING. It can shift the pathway toward antitumor cGAS–STING–IFN-β signaling [100]. Also, in vivo assays suggest that the MVA pathway and IFN signaling pathway are part of a metabolic circuit, sensing each other [101].
Recent research efforts have focused on utilizing the cGAS–STING pathway as a therapeutic target for the treatment of PCa, and it has been demonstrated that ionizing radiation can activate this pathway and enhance the response of CD8+ lymphocytes [102,103]. On the other hand, the cGAS–STING pathway has also been linked to promoting tumor proliferation. Propionibacterium acnes, a common microorganism in normal prostate tissues and PCa, can activate the cGAS–STING pathway and induce the expression of interferons, which subsequently promotes the growth of PCa [98].

4.6. Function of Stroma Elements in Prostate Cancer

Cancer-associated fibroblasts (CAFs) represent the predominant stromal cell type in the tumor microenvironment (TME) [104,105], and their abundance is notably pronounced in PCa [106]. These cells, primarily derived from tissue-resident fibroblasts and other local cells, including tumoral epithelial PCa cells [104,105,107,108,109,110], exert a significant influence on tumorigenesis. Notably, only grafts containing initiated prostate epithelium and CAFs generate tumors [111,112]. CAFs induce angiogenesis, immunosuppression, metastatic progression, and resistance to therapy. However, in certain contexts, CAFs have been linked to exerting antitumoral activity during the early stages of tumorigenesis [113]. This finding aligns with recent research indicating rapid epigenetic remodeling within hours of tumor-specific CD8+ T lymphocytes after sustained exposure to tumor antigens, resulting in a persistent dysfunctional state despite robust activation and proliferation [114].
New evidence suggests that tumor cells can circumvent the constraints posed by androgen deprivation therapy (ADT) through the acquisition of glucosamine from CAFs. This process leads to an elevation in O-GlcNAc levels and, ultimately, triggers the induction of 3β-Hydroxysteroid dehydrogenase-1 (3βHSD1), the crucial enzyme responsible for the conversion of dehydroepiandrosterone (DHEA) to the highly potent androgen dihydrotestosterone (DHT) [115]. Additionally, mutations occurring in the AR’s ligand-binding domain cause the expanded responsiveness of the AR signaling pathway to nonandrogenic steroid molecules and antiandrogens, along with the expression of constitutively active AR splice variants, alterations in the function of AR coactivators, and intricate crosstalk with other oncogenic pathways [1,116,117]. In this context, cholesterol plays a role in regulating the expression of AR cofactors. Elevated levels of SRC-1, SRC-2, SRC-3, and PCAF gene and protein expression are observed concurrently with heightened AR gene and protein expression. This interplay exerts a notable influence on the initiation and advancement of PCa in a murine model of CRPC [117]. In agreement with this, the complexity of the AR is further highlighted by the observation that CRPC is associated with AR overexpression due to clonal selective pressure in the androgen-depleted environment, leading to gene upregulation. Notably, androgen-independent cells require 80% less androgens for growth [118], and approximately 30% of these cells exhibit high-level amplifications [119]. Furthermore, AR mutations have been identified to increase sensitivity to androgens by four orders of magnitude lower than that required for androgen-dependent PCa cells [120]. This hyperactivation of the androgen signaling pathway leads to CRPC.
It has been observed that, in PCa, there is a bidirectional relationship between the signaling pathway of AR and the TAMs. AR activation has been described as essential for the ability to induce tumor cell death by TAMs, which is why ADT would be counterproductive to the local innate immune response [121]. However, when TAMs become pro-tumor, they can transfer cholesterol to tumor cells in vitro by increasing androgen production and AR activation, suggesting the need to limit cholesterol transfer, the MVA pathway, or TAMs as adjuvant therapy in the context of PCa [122].

4.7. Mitochondria in the Metabolic Deregulation of Prostate Cancer

Metabolic alterations induced by oncogenesis in PCa are sustained through mitochondrial activity and the involvement of dynamin-related protein 1 (DRP1), a mediator of mitochondrial fission. Elevated DRP1 levels in both androgen-sensitive and CRPC cell lines are associated with AR signaling, leading to an enhancement of pyruvate transport into the mitochondria. This, in turn, promotes oxidative phosphorylation (OXPHOS) and lipogenesis, which are hallmark characteristics of PCa [1]. Similarly, the upregulation of the pyruvate kinase M2 isoform (PKM2) is critical for controlling lipid homeostasis. The regulation of nuclear respiratory factor 1 (NRF1), responsible for the transcription of the endoplasmic reticulum (ER) transmembrane protein 33 (TMEM33), is a key element of this mechanism. As mentioned earlier, the SREBP proteins are localized in the ER, and TMEM33 recruits the RNF5 enzyme to facilitate SCAP degradation and activate the MVA pathway. These findings are associated with the observation that PKM2 is predominantly expressed in PCa cells as opposed to normal prostate cells. The downregulation of PKM2 is linked to a reduction in tumor growth, while the global knockout of PKM2 accelerates the growth of allografted tumors [123]. A part of these considerations is shown in a schematic way in Figure 2.

4.8. The Telomerase Complex Is Related to Androgen Receptor Signaling

Telomeres are the DNA-protein structures that cap the ends of linear chromosomes and protect them from fusing end to end [124]. The telomerase is a protein complex (HSP90, hTERC, diskerin, TEP1, p53, hTERT), the canonical function of which is to protect against telomere shortening [125]. The non-canonical functions of the telomerase contribute to cancer progression through apoptosis resistance; an increase in mitochondrial membrane potential and a decrease in ROS; gene expression; and signal transduction [125].
In PCa cells, telomere stability may be modulated via an allosteric mechanism in-volving the AR. The AR interacts with telomeric proteins, specifically TIN2 from the shelterin complex, and exhibits the capability to disrupt these interactions in AR-positive PCa cell lines when treated with the AR antagonist bicalutamide [124]. Additionally, there is evidence suggesting a concentration-dependent impact of androgens on human telomerase reverse transcriptase (hTERT) expression. Under physiological androgen levels, characterized by low androgen levels, hTERT expression is induced. Conversely, at high supraphysiological androgen levels, hTERT expression is suppressed by the inhibitor of growth 1 (ING1) and 2 (ING2) [126]. Recently, it has been postulated through bioinformatic studies that the AR, zinc finger proteins, and telomeres modulate the global gene expression pattern during PCa progression [127]. Conflicting findings about the correlation of telomerase expression and clinicopathological indicators may be consistent with noise caused during processing of tumor tissues by surrounding normal tissue expressing little or no telomerase activity. [128,129].
The cholesterol-derived hormone estrogen, specifically 17β-estradiol (E2), has the capacity to stimulate the transcription of the human telomerase reverse transcriptase (hTERT) gene in PCa cells. This induction results in an increase in telomerase activity within a short period. Notably, estrogen receptors (ERs) collocate with the hTERT promoter in PCa cells, indicating a direct interaction between estrogen signaling and the regulation of telomerase activity in PCa [130]. The significance of these findings lies in the association of estrogens with dihydrotestosterone biosynthesis through the “backdoor pathway.” This pathway, which is not only implicated in the human testis and adrenal glands but also in the classical pathway for androgen synthesis, has been linked to tumor resistance to castration [131].

5. Statins as Adjuvant Therapy in Prostate Cancer

Statins, widely used for reducing the de novo production of cholesterol in the blood, were first developed from Penicillium citrinum in 1976 [132]. They exert their effect by competitively inhibiting HMG-CR (see Figure 1), leading to reductions in serum total cholesterol, low-density lipoprotein (LDL) cholesterol, apolipoprotein-B (Apo-B), and triglycerides, thereby aiding in the prevention of cardiovascular disease (CVD) [133]. Their remarkably high affinity for HMG-CR, approximately 1000 times higher than the endogenous substrate, results in the induction of LDL cholesterol receptors and subsequent reductions in blood levels of this type of cholesterol. Some statins, such as simvastatin, lovastatin, and pravastatin, are derived from fungal sources, while others are synthetic compounds [7].
The ability of statins to produce intracellular effects is closely linked to their chemical structure, with lipophilic statins having a higher capacity for diffusion into the intracellular space [134,135]. This heightened intracellular diffusion capability is associated with their ability to induce apoptosis [136] and may also contribute to certain adverse effects [137]. Statins play a critical role in modulating cholesterol levels in lipid raft-dependent functions within cancer cells and tumors. In the context of PCa cell lines, the increase in intracellular cholesterol in response to the activation of PI3K can be countered by the administration of statins [138,139,140].
In vitro and in vivo experiments indicate that statins can decrease cellular plasticity by promoting a mesenchymal-like cell state. This transition enhances the ability of metastatic seeding on the one hand but suppresses the reorganization of cells into secondary tumors on the other. In essence, the influence of statins leads to an augmentation of the EMT while inhibiting the vital and dynamic reversal in cellular phenotype known as the mesenchymal-to-epithelial transition (MET), which is crucial for the colonization of distant organs. This overall process, termed epithelial-mesenchymal plasticity (EMP), characterized by a combination of cell states displaying both partial epithelial and mesenchymal characteristics, is notably curbed. In the case of PCa, the utilization of rosuvastatin has shown a reduction in cellular plasticity [141,142], whereas lovastatin or simvastatin has been linked to the deactivation of RhoA, triggering cancer cell apoptosis and inducing cell cycle arrest in the G1 phase [143].
Bisphosphonates are other compounds that act on the MVA pathway; however, they inhibit farnesyl-pyrophosphate synthase, an enzyme downstream HMGCR in this pathway [144]. Bisphophonates can bind to mineral bone surfaces [145] and are used in oncology for the management of skeletal complications and/or bone metastases. In addition, some bisphosphonates have been shown to exhibit direct anticancer properties [146,147], indicating that targeting the MVA pathway could have beneficial effects, at least in certain cancer types affecting bone tissue.

5.1. Adverse Effects Induced by Statins

The detrimental effects associated with high-dose statin use, known as statin-associated side effects (SASs), lead to approximately 50% of patients discontinuing treatment within a year [148,149,150]. The methylation-induced downregulation of ABCG1 due to statins has been linked to an increased risk of developing type 2 diabetes [151]. Various studies suggest that 9–12% of statin recipients face an elevated risk of type 2 diabetes [152]. However, a recent meta-analysis has indicated that the actual increase in risk equates to one case per 100–200 patients treated [153]. Despite these potential side effects, the benefits of statin use outweigh the adverse outcomes, making it highly recommended for individuals with cardiovascular risk.
Furthermore, other less explored symptoms include neurological, neurocognitive, hepatotoxicity, nephrotoxicity, gastrointestinal, urological–genital, and reproductive effects. Statins disrupt the Akt pathway, leading to mitochondrial dysfunction, resulting in insufficient energy production for highly energy-demanding tissues such as muscles. This gives rise to what is known as statin-associated mitochondrial symptoms (SAMSs). SAMSs are the primary reason for treatment discontinuation, and they can manifest as weakness, muscle pain, fatigue, myalgia, tendon pain, night muscle cramps, and, in extreme cases, rhabdomyolysis [154,155]. These adverse effects are closely linked to several key mechanisms. Firstly, the reduction in circulating lipoproteins, which transport a substantial portion (74%) of CoQ10, leads to the depletion of CoQ10 [155,156]. CoQ10 plays an essential role in the functioning of mitochondrial electrical complexes. Consequently, this depletion contributes to mitochondrial depletion and triggers oxidative stress. These effects are further compounded by the downregulation of peroxisome proliferator-activated gamma coactivator receptor (PGC)-1α and PGC-1β [157].
Moreover, the inhibition of the electron transport chain and the disruption of Ca2+ metabolism through increased mitochondrial Ca2+ efflux via the Na+-Ca2+ exchanger [158] and, secondarily, the exit of Ca2+ from the sarcoplasmic reticulum mediated by ryanodine receptors [159] result in the activation of calpain. Calpain is responsible for the translocation of BAX to the outer mitochondrial membrane, leading to the release of cytochrome C into the cytosol and the subsequent activation of the intrinsic pathway of apoptosis [160].

5.2. Clinical Evidence of the Antitumoral Effects of Statins

The mevalonate pathway has been found to play a crucial role in the development of PCa by producing molecules that are either directly or indirectly related to the function of AR, which is the molecule at the center of this pathology. Cholesterol precedes the generation of steroid hormones in the mevalonate pathway and has been found to have a very relevant structural role in the formation of LRs, which are the fundamental axis of the cell’s functioning because they coin the receptors that are the starting point for a cellular response to an external stimulus. Cholesterol directly or indirectly regulates several signaling pathways that are interconnected with the AR signaling pathway.
Clinical data about the use of statins as antitumoral agents are contradictory, but the level of evidence of this contribution must be evaluated before arriving to any solid conclusions.
In 2018, Murtola and colleagues detected a 14% reduction in the tumor proliferation index Ki-67 in patient samples after a 28-day intervention with 80 mg of atorvastatin [161]. A subsequent study demonstrated the positive impact of 80 mg of fluvastatin on cell proliferation, with detectable concentrations in prostate tissue and increased apoptosis in 33 patients who had undergone prostatectomy. Fluvastatin, after 4–12 weeks of use, attained adequate blood levels consistent with histopathological evidence supporting the antitumor role of statins [162,163]. Although these data manage to harmonize the contribution of statins at the cellular level with consistent pharmacological parameters, due to the small sample size, analyses of these results remain inconclusive, emphasizing the need to establish a more robust connection between fundamental scientific evidence and clinical applications [164].
Prognostic indicators are needed to help urologists establish the most appropriate therapy for PCa patients, and we have previously mentioned some of them, like hTERT expression determined in prostate tissue, which is positively correlated with various clinicopathological features, such as Gleason score, tumor differentiation, serum levels of PSA, recurrence and worse survival following radical prostatectomy [128,129]. Patients with tumors with little infiltration of TAMs and CAFs exhibit significantly improved recurrence-free survival (RFS) rates and more favorable outcomes from hormonal therapy compared to those with high infiltration levels. In addition, TAMs and CAFs correlate with serum PSA levels, Gleason score and T stage [165]. CAFs have also been proposed as PCa markers. It has been observed that both HMGCS2 and aldoketo reductase 1 member C3 (AKR1C3) increase their levels, particularly in CRPC cell lines, in cocultures with CAFs. Therapy with simvastatin and AKR1C3 inhibitors decreased the growth of PCa lines, and this effect also extended to CRPC lines and enzalutamide resistant [166,167]. These enzymes showed a significant elevation and are correlated with the Gleason score in human PCa specimens [168].
However, a limitation has been identified in preclinical studies on the anticancer efficacy of statins: the need to achieve concentrations in the micromolar range to effectively inhibit cell proliferation. These concentrations are currently unattainable in clinical practice without inducing significant side effects [169]. Our approach involves the use of statins as a complement to other well-established treatments in PCa and not as a single therapy to inhibit cell proliferation. To reduce the availability of tumor cells for immune escape through the activation of the RA signaling pathway, it is crucial to reach therapeutic levels similar to those required to prevent cardiovascular disease [170].
Dickerman and colleagues emphasize the significance of proper randomization in understanding discrepancies. They found that statin therapy did not have a significant impact on the incidence of PCa in an emulating study that analyzed the electronic health records of 733,804 adults over a decade [171]. They did not analyze the potential benefits of using statins to mitigate the severity of PCa. However, this study does not diminish the strong evidence that has been generated from well-randomized prospective trials that have shown positive results related to combining statins with other therapies.
An initial analysis from the REDUCE database revealed that the use of statins did not increase or diminish the development of PCa [172]. However, five years later, the accumulated data revealed that high total serum cholesterol and high HDL cholesterol were associated with an increased risk of high-grade PCa [173]. In accordance with this, a meta-analysis found that statin use was associated with a 20% lower risk of advanced PCa based on data from 15 cohort and 12 case–control studies [174] and findings from two large prospective studies with substantial follow-up support, the Health Professionals Follow-up Study (24 years) and the Atherosclerosis Risk in Communities (ARIC) study (20 years) [175,176]. In the same way, recent robust findings show that statins could be useful in the advanced stages of PCa when combined with AR axis-targeted therapies (ARATs) for men with high-risk hormone-sensitive PCa (HSPC) or CRPC after radiotherapy. This prospective analysis, which took place over almost 7 years, determined that, in two months, statins improved overall survival (OS) and PCa-specific survival [177].
These substantial findings support other less consistent findings, like thos derived from a post hoc analysis in patients previously treated with docetaxel [178] or a meta-analysis from three phase III studies evaluating enzalutamide: AFFIRM, PREVAIL, and PROSPER; a multivariate analysis suggested that statins are associated with improved OS, but significant limitations meant that the results were confounded by the baseline cardiovascular risk and original indication of statins [179]. These findings agree with those of other studies, specifically one in which a 27% reduction in the risk of overall mortality and a 35% lower risk of specific mortality were established in statin users among men with PCa receiving ADT or ARATs in 25 cohorts; however, these studies had great heterogeneity, which prevented an adequate comparison, despite studies of quality, sensitivity, and strategies to reduce biases [180].
An analysis of a Finnish cohort of 14,424 men with PCa conducted over the course of 18 years reported a decreased risk of starting androgen deprivation therapy (ADT) and PCa-related death in patients who initiated statin use 5 years before diagnosis [181]. Similarly, a large population-based electronic database in the United Kingdom was used to evaluate 11,772 patients over 4.4 years, and it was found that a decreased risk of PCa-specific mortality is associated with statin use, with a stronger association being observed when statins were used before diagnosis [182].
The secondary role of statins in the prevention of PCa was examined in a systematic meta-analysis which included prospective, retrospective, and randomized controlled trials. This study found that the combination of statin use and androgen deprivation therapy (ADT) was associated with a significant reduction in the risks of mortality of all-causes and PCa-specific mortality. Similarly, the combination of statins and radical prostatectomy improved cancer-specific mortality. However, the authors admitted that differences in the time, dose, and duration of statin treatment caused heterogeneity, and the most effective type of statin was not established [183].
One study that did not achieve findings that support the use of statins as a complementary therapy in PCa, despite involving prospective, randomized, and double-blind trials, is the PRO-STAT, a study that involved 181 patients and the use of suboptimal doses of atorvastatin (20 mg/day) after RP [184]. However, this study faced criticism due to the relatively low doses of atorvastatin used, which were not effective in sufficiently reducing cholesterol levels [162]. A cohort study about the relation between statin use and the aggressiveness of PCa (higher Gleason score) established that the benefits are dose-dependent, meaning that the use of statins may continue for a relatively longer time (>11 months) [185].
Certainly, some strong evidence provided by extensive prospective studies supports the use of statins due to their antitumor properties and well-underlined benefits when used in combination with RP, ADT, and ARATs. In addition to clinical data, in vitro results have highlighted the diverse mechanisms linking the MVA pathway and other metabolic pathways to the AR pathway and PCa. These mechanisms encompass castration resistance, metastasis, and the response to different therapeutic approaches. These intricate processes not only involve the epithelial components but also underscore the critical role of stromal elements and microbiota in the progression of PCa, and recently, this has been supported by congruent evidence from metabolomic studies.

Funding

The APC was funded by Group in Applied Engineering and Data Science (Government of Aragon, grant number T69_23R). A.B.-F./L.M.E. were supported by Group in Applied Engineering and Data Science (Government of Aragon, grant number T69_23R). S.R.-Z. was supported by a Margarita Salas Fellowship for the Requalification of the Spanish university system 2021–2023 from the University of Zaragoza and funded by the European Union-NextGenerationEU.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. 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] [PubMed]
  2. Hornburg, D.; Wu, S.; Moqri, M.; Zhou, X.; Contrepois, K.; Bararpour, N.; Traber, G.M.; Su, B.; Metwally, A.A.; Avina, M.; et al. Dynamic lipidome alterations associated with human health, disease and ageing. Nat. Metab. 2023, 5, 1578–1594. [Google Scholar] [CrossRef] [PubMed]
  3. Meller, S.; Meyer, H.A.; Bethan, B.; Dietrich, D.; Maldonado, S.G.; Lein, M.; Montani, M.; Reszka, R.; Schatz, P.; Peter, E.; et al. Integration of tissue metabolomics, transcriptomics and immunohistochemistry reveals ERG-and gleason score-specific metabolomic alterations in prostate cancer. Oncotarget 2015, 7, 1421. [Google Scholar] [CrossRef]
  4. Sagini, K.; Urbanelli, L.; Buratta, S.; Emiliani, C.; Llorente, A. Lipid Biomarkers in Liquid Biopsies: Novel Opportunities for Cancer Diagnosis. Pharmaceutics 2023, 15, 437. [Google Scholar] [CrossRef]
  5. Yu, C.; Niu, L.; Li, L.; Li, T.; Duan, L.; He, Z.; Zhao, Y.; Zou, L.; Wu, X.; Luo, C. Identification of the metabolic signatures of prostate cancer by mass spectrometry-based plasma and urine metabolomics analysis. Prostate 2021, 81, 1320–1328. [Google Scholar] [CrossRef] [PubMed]
  6. Comerford, S.A.; Huang, Z.; Du, X.; Wang, Y.; Cai, L.; Witkiewicz, A.K.; Walters, H.; Tantawy, M.N.; Fu, A.; Manning, H.C.; et al. Acetate Dependence of Tumors. Cell 2014, 159, 1591. [Google Scholar] [CrossRef]
  7. Jiang, W.; Hu, J.W.; He, X.R.; Jin, W.L.; He, X.Y. Statins: A repurposed drug to fight cancer. J. Exp. Clin. Cancer Res. 2021, 40, 241. [Google Scholar] [CrossRef]
  8. Schug, Z.T.; Peck, B.; Jones, D.T.; Zhang, Q.; Grosskurth, S.; Alam, I.S.; Goodwin, L.M.; Smethurst, E.; Mason, S.; Blyth, K.; et al. Acetyl-CoA Synthetase 2 Promotes Acetate Utilization and Maintains Cancer Cell Growth under Metabolic Stress. Cancer Cell 2015, 27, 57. [Google Scholar] [CrossRef]
  9. Broadfield, L.A.; Pane, A.A.; Talebi, A.; Swinnen, J.V.; Fendt, S.M. Lipid metabolism in cancer: New perspectives and emerging mechanisms. Developmental Cell. Cell Press 2021, 56, 1363–1393. [Google Scholar]
  10. Koizume, S.; Miyagi, Y. Lipid Droplets: A Key Cellular Organelle Associated with Cancer Cell Survival under Normoxia and Hypoxia. Int. J. Mol. Sci. 2016, 17, 1430. [Google Scholar] [CrossRef]
  11. Ladanyi, A.; Mukherjee, A.; Kenny, H.A.; Johnson, A.; Mitra, A.K.; Sundaresan, S.; Nieman, K.M.; Pascual, G.; Benitah, S.A.; Montag, A.; et al. Adipocyte-induced CD36 expression drives ovarian cancer progression and metastasis. Oncogene 2018, 37, 2285–2301. [Google Scholar] [CrossRef]
  12. Danielli, M.; Perne, L.; Jarc Jovičić, E.; Petan, T. Lipid droplets and polyunsaturated fatty acid trafficking: Balancing life and death. Front. Cell Dev. Biol. 2023, 11, 1104725. [Google Scholar] [CrossRef]
  13. 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. [Google Scholar] [CrossRef]
  14. Sivanand, S.; Viney, I.; Wellen, K.E. Spatiotemporal Control of Acetyl-CoA Metabolism in Chromatin Regulation. Trends Biochem. Sci. 2018, 43, 61–74. [Google Scholar] [CrossRef]
  15. Boon, R.; Silveira, G.G.; Mostoslavsky, R. Nuclear metabolism and the regulation of the epigenome. Nat. Metab. 2020, 2, 1190–1203. [Google Scholar] [CrossRef]
  16. Guertin, D.A.; Wellen, K.E. Acetyl-CoA metabolism in cancer. Nat. Rev. Cancer 2023, 23, 156–172. [Google Scholar] [CrossRef]
  17. Xu, D.; Shao, F.; Bian, X.; Meng, Y.; Liang, T.; Lu, Z. The Evolving Landscape of Noncanonical Functions of Metabolic Enzymes in Cancer and Other Pathologies. Cell Metab. 2021, 33, 33–50. [Google Scholar] [CrossRef]
  18. Logotheti, S.; Papadaki, E.; Zolota, V.; Logothetis, C.; Vrahatis, A.G.; Soundararajan, R.; Tzelepi, V. Lineage Plasticity and Stemness Phenotypes in Prostate Cancer: Harnessing the Power of Integrated “Omics” Approaches to Explore Measurable Metrics. Cancers 2023, 15, 4357. [Google Scholar] [CrossRef]
  19. Chen, J.; Guccini, I.; Mitri DDi Brina, D.; Revandkar, A.; Sarti, M.; Pasquini, E.; Alajati, A.; Pinton, S.; Losa, M.; Civenni, G.; et al. Compartmentalized activities of the pyruvate dehydrogenase complex sustain lipogenesis in prostate cancer. Nat. Genet. 2018, 50, 219–228. [Google Scholar] [CrossRef]
  20. Chen, Z.; Tian, R.; She, Z.; Cai, J.; Li, H. Role of oxidative stress in the pathogenesis of nonalcoholic fatty liver disease. Free Radic. Biol. Med. 2020, 152, 116–141. [Google Scholar] [CrossRef]
  21. Chiarugi, A.; Dölle, C.; Felici, R.; Ziegler, M. The NAD metabolome—A key determinant of cancer cell biology. Nat. Rev. Cancer 2012, 12, 741–752. [Google Scholar] [CrossRef]
  22. Xiao, W.; Wang, R.S.; Handy, D.E.; Loscalzo, J. NAD(H) and NADP(H) Redox Couples and Cellular Energy Metabolism. Antioxid. Redox. Signal. 2018, 28, 251–272. [Google Scholar] [CrossRef]
  23. Ying, W. NAD+/NADH and NADP+/NADPH in cellular functions and cell death: Regulation and biological consequences. Antioxid. Redox Signal. 2008, 10, 179–206. [Google Scholar] [CrossRef]
  24. Golomb, B.A.; Evans, M.A. Statin Adverse Effects: A Review of the Literature and Evidence for a Mitochondrial Mechanism. Am. J. Cardiovasc. Drugs 2008, 8, 373. [Google Scholar] [CrossRef]
  25. Stancu, C.; Sima, A. Statins: Mechanism of action and effects. J. Cell Mol. Med. 2001, 5, 378–387. [Google Scholar] [CrossRef]
  26. Xue, L.; Qi, H.; Zhang, H.; Ding, L.; Huang, Q.; Zhao, D.; Wu, B.J.; Li, X. Targeting SREBP-2-Regulated Mevalonate Metabolism for Cancer Therapy. Fron. Oncol. 2020, 10, 1510. [Google Scholar] [CrossRef] [PubMed]
  27. Wang, B.; Tontonoz, P. Liver X receptors in lipid signalling and membrane homeostasis. Nat. Rev. Endocrinol. 2018, 14, 452–463. [Google Scholar] [CrossRef] [PubMed]
  28. Thurnher, M.; Gruenbacher, G. T lymphocyte regulation by mevalonate metabolism. Sci. Signal. 2015, 8, re4. [Google Scholar] [CrossRef] [PubMed]
  29. Crosas-Molist, E.; Samain, R.; Kohlhammer, L.; Orgaz, J.L.; George, S.L.; Maiques, O.; Barcelo, J.; Sanz-Moreno, V. RHO GTPase signaling in cancer progression and dissemination. Physiol. Rev. 2022, 102, 455–510. [Google Scholar] [CrossRef] [PubMed]
  30. 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]
  31. Sharma, N.L.; Massie, C.E.; Ramos-Montoya, A.; Zecchini, V.; Scott, H.E.; Lamb, A.D.; MacArthur, S.; Stark, R.; Warren, A.Y.; Mills, I.G.; et al. The Androgen Receptor Induces a Distinct Transcriptional Program in Castration-Resistant Prostate Cancer in Man. Cancer Cell 2013, 23, 35–47. [Google Scholar] [CrossRef]
  32. Bader, D.A.; McGuire, S.E. Tumour metabolism and its unique properties in prostate adenocarcinoma. Nat. Rev. Urol. 2020, 17, 214–231. [Google Scholar] [CrossRef]
  33. Audet-Walsh, É.; Yee, T.; McGuirk, S.; Vernier, M.; Ouellet, C.; St-Pierre, J.; Giguère, V. Androgen-dependent repression of ERRγ reprograms metabolism in prostate cancer. Cancer Res. 2017, 77, 378–389. [Google Scholar] [CrossRef] [PubMed]
  34. Massie, C.E.; Lynch, A.; Ramos-Montoya, A.; Boren, J.; Stark, R.; Fazli, L.; Warren, A.; Scott, H.; Madhu, B.; Sharma, N.; et al. The androgen receptor fuels prostate cancer by regulating central metabolism and biosynthesis. EMBO J. 2011, 30, 2719–2733. [Google Scholar] [CrossRef]
  35. 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]
  36. Jeon, T.I.; Osborne, T.F. SREBPs: Metabolic integrators in physiology and metabolism. Trends Endocrinol. Metab. 2012, 23, 65–72. [Google Scholar] [CrossRef] [PubMed]
  37. Mullen, P.J.; Yu, R.; Longo, J.; Archer, M.C.; Penn, L.Z. The interplay between cell signalling and the mevalonate pathway in cancer. Nat. Rev. Cancer 2016, 16, 718–731. [Google Scholar] [CrossRef] [PubMed]
  38. Clendening, J.W.; Penn, L.Z. Targeting tumor cell metabolism with statins. Oncogene 2012, 31, 4967–4978. [Google Scholar] [CrossRef]
  39. Sharpe, L.J.; Brown, A.J. Controlling cholesterol synthesis beyond 3-hydroxy-3-methylglutaryl-CoA reductase (HMGCR). J. Biol. Chem. 2013, 288, 18707–18715. [Google Scholar] [CrossRef]
  40. Shimano, H.; Sato, R. SREBP-regulated lipid metabolism: Convergent physiology-divergent pathophysiology. Nat. Rev. Endocrinol. 2017, 13, 710–730. [Google Scholar] [CrossRef]
  41. Cheng, C.; Ru, P.; Geng, F.; Liu, J.; Yoo, J.Y.; Wu, X.; Cheng, X.; Euthine, V.; Hu, P.; Guo, J.Y.; et al. Glucose-Mediated N-glycosylation of SCAP Is Essential for SREBP-1 Activation and Tumor Growth. Cancer Cell 2015, 28, 569–581. [Google Scholar] [CrossRef]
  42. Maxfield, F.R.; Van Meer, G. Cholesterol, the central lipid of mammalian cells. Curr. Opin. Cell Biol. 2010, 22, 422–429. [Google Scholar] [CrossRef] [PubMed]
  43. Dambal, S.; Alfaqih, M.; Sanders, S.; Maravilla, E.; Ramirez-Torres, A.; Galvan, G.C.; Reis-Sobreiro, M.; Rotinen, M.; Driver, L.M.; Behrove, M.S.; et al. 27-Hydroxycholesterol Impairs Plasma Membrane Lipid Raft Signaling as Evidenced by Inhibition of IL6–JAK–STAT3 Signaling in Prostate Cancer Cells. Mol. Cancer Res. 2020, 18, 671–684. [Google Scholar] [CrossRef] [PubMed]
  44. Li, Y.C.; Park, M.J.; Ye, S.K.; Kim, C.W.; Kim, Y.N. Elevated levels of cholesterol-rich lipid rafts in cancer cells are correlated with apoptosis sensitivity induced by cholesterol-depleting agents. Am. J. Pathol. 2006, 168, 1107–1118. [Google Scholar] [CrossRef] [PubMed]
  45. Head, B.P.; Patel, H.H.; Insel, P.A. Interaction of membrane/lipid rafts with the cytoskeleton: Impact on signaling and function: Membrane/lipid rafts, mediators of cytoskeletal arrangement and cell signaling. Biochim. Et Biophys. Acta Biomembr. 2014, 1838, 532–545. [Google Scholar] [CrossRef] [PubMed]
  46. Mollinedo, F.; Gajate, C. Lipid rafts as major platforms for signaling regulation in cancer. Adv. Biol. Regul. 2015, 57, 130–146. [Google Scholar] [CrossRef] [PubMed]
  47. Mollinedo, F.; Gajate, C. Lipid rafts as signaling hubs in cancer cell survival/death and invasion: Implications in tumor progression and therapy. J. Lipid Res. 2020, 61, 611–635. [Google Scholar] [CrossRef] [PubMed]
  48. Mollinedo, F.; Gajate, C. Clusters of apoptotic signaling molecule-enriched rafts, CASMERs: Membrane platforms for protein assembly in Fas/CD95 signaling and targets in cancer therapy. Biochem. Soc. Trans. 2022, 50, 1105–1118. [Google Scholar] [CrossRef] [PubMed]
  49. Simons, K.; Ehehalt, R. Cholesterol, lipid rafts, and disease. J. Clin. Investig. 2002, 110, 597–603. [Google Scholar] [CrossRef] [PubMed]
  50. Bancaro, N.; Calì, B.; Troiani, M.; Elia, A.R.; Arzola, R.A.; Attanasio, G.; Lai, P.; Crespo, M.; Gurel, B.; Pereira, R.; et al. Apolipoprotein E induces pathogenic senescent-like myeloid cells in prostate cancer. Cancer Cell 2023, 41, 602–619.e11. [Google Scholar] [CrossRef]
  51. Jiang, S.; Wang, X.; Song, D.; Liu, X.J.; Gu, Y.; Xu, Z.; Wang, X.; Zhang, X.; Ye, Q.; Tong, Z.; et al. Cholesterol induces epithelial-to-mesenchymal transition of prostate cancer cells by suppressing degradation of EGFR through APMAP. Cancer Res. 2019, 79, 3063–3075. [Google Scholar] [CrossRef]
  52. Yumoto, K.; Rashid, J.; Ibrahim, K.G.; Zielske, S.P.; Wang, Y.; Omi, M.; Decker, A.M.; Jung, Y.; Sun, D.; Remmer, H.A.; et al. HER2 as a potential therapeutic target on quiescent prostate cancer cells. Transl. Oncol. 2023, 31, 101642. [Google Scholar] [CrossRef]
  53. Gil, V.; Miranda, S.; Riisnaes, R.; Gurel, B.; D’Ambrosio, M.; Vasciaveo, A.; Crespo, M.; Ferreira, A.; Brina, D.; Troiani, M.; et al. HER3 Is an Actionable Target in Advanced Prostate Cancer. Cancer Res. 2021, 81, 6207–6218. [Google Scholar] [CrossRef]
  54. Muhammad, A.; Forcados, G.E.; Yusuf, A.P.; Abubakar, M.B.; Sadiq, I.Z.; Elhussin, I.; Siddique, M.A.T.; Aminu, S.; Suleiman, R.B.; Abubakar, Y.S.; et al. Comparative G-Protein-Coupled Estrogen Receptor (GPER) Systems in Diabetic and Cancer Conditions: A Review. Molecules 2022, 27, 8943. [Google Scholar] [CrossRef]
  55. Ramírez-de-Arellano, A.; Pereira-Suárez, A.L.; Rico-Fuentes, C.; López-Pulido, E.I.; Villegas-Pineda, J.C.; Sierra-Diaz, E. Distribution and Effects of Estrogen Receptors in Prostate Cancer: Associated Molecular Mechanisms. Front. Endocrinol. 2022, 12, 811578. [Google Scholar] [CrossRef] [PubMed]
  56. Rago, V.; Romeo, F.; Giordano, F.; Maggiolini, M.; Carpino, A. Identification of the estrogen receptor GPER in neoplastic and non-neoplastic human testes. Reprod. Biol. Endocrinol. 2011, 9, 135. [Google Scholar] [CrossRef] [PubMed]
  57. Hermetet, F.; Buffière, A.; Aznague, A.; Pais de Barros, J.P.; Bastie, J.N.; Delva, L.; Quéré, R. High-fat diet disturbs lipid raft/TGF-β signaling-mediated maintenance of hematopoietic stem cells in mouse bone marrow. Nat. Commun. 2019, 10, 523. [Google Scholar] [CrossRef] [PubMed]
  58. Zhang, S.; Zhu, N.; Li, H.F.; Gu, J.; Zhang, C.J.; Liao, D.F.; Qin, L. The lipid rafts in cancer stem cell: A target to eradicate cancer. Stem Cell Res. Ther. 2022, 13, 432. [Google Scholar] [CrossRef] [PubMed]
  59. Collins, A.T.; Berry, P.A.; Hyde, C.; Stower, M.J.; Maitland, N.J. Prospective identification of tumorigenic prostate cancer stem cells. Cancer Res. 2005, 65, 10946–10951. [Google Scholar] [CrossRef] [PubMed]
  60. Aurilio, G.; Cimadamore, A.; Mazzucchelli, R.; Lopez-Beltran, A.; Verri, E.; Scarpelli, M.; Massari, F.; Cheng, L.; Santoni, M.; Montironi, R. Androgen Receptor Signaling Pathway in Prostate Cancer: From Genetics to Clinical Applications. Cells 2020, 9, 2653. [Google Scholar] [CrossRef] [PubMed]
  61. Crnalic, S.; Hörnberg, E.; Wikström, P.; Lerner, U.H.; Tieva, Å.; Svensson, O.; Widmark, A.; Bergh, A. Nuclear androgen receptor staining in bone metastases is related to a poor outcome in prostate cancer patients. Endocr. Relat. Cancer 2010, 17, 885–895. [Google Scholar] [CrossRef]
  62. Antonarakis, E.S.; Lu, C.; Luber, B.; Wang, H.; Chen, Y.; Zhu, Y.; Silberstein, J.L.; Taylor, M.N.; Maughan, B.L.; Denmeade, S.R.; et al. Clinical Significance of Androgen Receptor Splice Variant-7 mRNA Detection in Circulating Tumor Cells of Men with Metastatic Castration-Resistant Prostate Cancer Treated With First-and Second-Line Abiraterone and Enzalutamide. J. Clin. Oncol. 2017, 35, 2149–2156. [Google Scholar] [CrossRef]
  63. Armstrong, A.J.; Halabi, S.; Luo, J.; Nanus, D.M.; Giannakakou, P.; Szmulewitz, R.Z.; Danila, D.C.; Healy, P.; Anand, M.; Rothwell, C.J.; et al. Prospective Multicenter Validation of Androgen Receptor Splice Variant 7 and Hormone Therapy Resistance in High-Risk Castration-Resistant Prostate Cancer: The PROPHECY Study. J. Clin. Oncol. 2019, 37, 1120–1129. [Google Scholar] [CrossRef]
  64. Grasso, C.S.; Wu, Y.M.; Robinson, D.R.; Cao, X.; Dhanasekaran, S.M.; Khan, A.P.; Quist, M.J.; Jing, X.; Lonigro, R.J.; Brenner, J.C.; et al. The mutational landscape of lethal castration-resistant prostate cancer. Nature 2012, 487, 239–243. [Google Scholar] [CrossRef]
  65. Calon, A.; Tauriello, D.V.F.; Batlle, E. TGF-beta in CAF-mediated tumor growth and metastasis. Semin. Cancer Biol. 2014, 25, 15–22. [Google Scholar] [CrossRef] [PubMed]
  66. Eto, M.; Shiota, M.; Akamatsu, S.; Tsukahara, S.; Nagakawa, S.; Matsumoto, T. Androgen receptor mutations for precision medicine in prostate cancer. Endocr Relat. Cancer 2022, 29, R143–R155. [Google Scholar]
  67. Jacob, A.; Raj, R.; Allison, D.B.; Myint, Z.W. Androgen receptor signaling in prostate cancer and therapeutic strategies. Cancers 2021, 13, 5417. [Google Scholar] [CrossRef] [PubMed]
  68. Jernberg, E.; Bergh, A.; Wikström, P. Clinical relevance of androgen receptor alterations in prostate cancer. Endocr. Connect. 2017, 6, R146–R161. [Google Scholar] [CrossRef] [PubMed]
  69. Joseph, J.D.; Lu, N.; Qian, J.; Sensintaffar, J.; Shao, G.; Brigham, D.; Moon, M.; Maneval, E.C.; Chen, I.; Darimont, B.; et al. A Clinically Relevant Androgen Receptor Mutation Confers Resistance to Second-Generation Antiandrogens Enzalutamide and ARN-509. Cancer Discov. 2013, 3, 1020–1029. [Google Scholar] [CrossRef] [PubMed]
  70. McSweeney, S.; Bergom, H.E.; Prizment, A.; Halabi, S.; Sharifi, N.; Ryan, C.; Hwang, J. Regulatory genes in the androgen production, uptake and conversion (APUC) pathway in advanced prostate cancer. Endocr. Oncol. 2022, 2, R51–R64. [Google Scholar] [CrossRef]
  71. Andersen, H.H.; Olsen, R.V.; Møller, H.G.; Eskelund, P.W.; Gazerani, P.; Arendt-Nielsen, L. A review of topical high-concentration L-menthol as a translational model of cold allodynia and hyperalgesia. Eur. J. Pain 2014, 18, 315–325. [Google Scholar] [CrossRef]
  72. Pethő, Z.; Najder, K.; Bulk, E.; Schwab, A. Mechanosensitive ion channels push cancer progression. Cell Calcium. 2019, 80, 79–90. [Google Scholar] [CrossRef]
  73. Liu, G.Y.; Sabatini, D.M. mTOR at the nexus of nutrition, growth, ageing and disease. Nat. Rev. Mol. Cell Biol. 2020, 21, 183–203. [Google Scholar] [CrossRef] [PubMed]
  74. Grolez, G.P.; Gordiendko, D.V.; Clarisse, M.; Hammadi, M.; Desruelles, E.; Fromont, G.; Prevarskaya, N.; Slomianny, C.; Gkika, D. TRPM8-androgen receptor association within lipid rafts promotes prostate cancer cell migration. Cell Death Dis. 2019, 10, 652. [Google Scholar] [CrossRef] [PubMed]
  75. Bidaux, G.; Roudbaraki, M.; Merle, C.; Crépin, A.; Delcourt, P.; Slomianny, C.; Thebault, S.; Bonnal, J.L.; Benahmed, M.; Cabon, F.; et al. Evidence for specific TRPM8 expression in human prostate secretory epithelial cells: Functional androgen receptor requirement. Endocr. Relat. Cancer 2005, 12, 367–382. [Google Scholar] [CrossRef] [PubMed]
  76. Gkika, D.; Flourakis, M.; Lemonnier, L.; Prevarskaya, N. PSA reduces prostate cancer cell motility by stimulating TRPM8 activity and plasma membrane expression. Oncogene 2010, 29, 4611–4616. [Google Scholar] [CrossRef] [PubMed]
  77. Freeman, M.R.; Cinar, B.; Kim, J.; Mukhopadhyay, N.K.; Di Vizio, D.; Adam, R.M.; Solomon, K.R. Transit of hormonal and EGF receptor-dependent signals through cholesterol-rich membranes. Steroids 2007, 72, 210–217. [Google Scholar] [CrossRef] [PubMed]
  78. Licitra, F.; Giovannelli, P.; Di Donato, M.; Monaco, A.; Galasso, G.; Migliaccio, A.; Castoria, G. New Insights and Emerging Therapeutic Approaches in Prostate Cancer. Front. Endocrinol. 2022, 13, 840787. [Google Scholar] [CrossRef] [PubMed]
  79. Audet-Walsh, E.; Vernier, M.; Yee, T.; Laflamme, C.; Li, S.; Chen, Y.; Giguere, V. SREBF1 activity is regulated by an AR/mTOR nuclear axis in prostate cancer. Mol. Cancer Res. 2018, 16, 1396–1405. [Google Scholar] [CrossRef]
  80. Huang, W.C.; Zhau, H.E.; Chung, L.W.K. Androgen receptor survival signaling is blocked by anti-β2- microglobulin monoclonal antibody via a MAPK/lipogenic pathway in human prostate cancer cells. J. Biol. Chem. 2010, 285, 7947–7956. [Google Scholar] [CrossRef]
  81. 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]
  82. Chen, M.; Chen, L.M.; Chai, K.X. Androgen regulation of prostasin gene expression is mediated by sterol-regulatory element-binding proteins and SLUG. Prostate 2006, 66, 911–920. [Google Scholar] [CrossRef] [PubMed]
  83. Heemers, H.; Verrijdt, G.; Organe, S.; Claessens, F.; Heyns, W.; Verhoeven, G.; Swinnen, J.V. Identification of an Androgen Response Element in Intron 8 of the Sterol Regulatory Element-binding Protein Cleavage-activating Protein Gene Allowing Direct Regulation by the Androgen Receptor. J. Biol. Chem. 2004, 279, 30880–30887. [Google Scholar] [CrossRef]
  84. Heemers, H.; Maes, B.; Foufelle, F.; Heyns, W.; Verhoeven, G.; Swinnen, J.V. Androgens Stimulate Lipogenic Gene Expression in Prostate Cancer Cells by Activation of the Sterol Regulatory Element-Binding Protein Cleavage Activating Protein/Sterol Regulatory Element-Binding Protein Pathway. Mol. Endocrinol. 2001, 15, 1817–1828. [Google Scholar] [CrossRef] [PubMed]
  85. Zadra, G.; Ribeiro, C.F.; Chetta, P.; Ho, Y.; Cacciatore, S.; Gao, X.; Syamala, S.; Bango, C.; Photopoulos, C.; Huang, Y. 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]
  86. Mineiro, R.; Santos, C.; Gonçalves, I.; Lemos, M.; Cavaco, J.E.B.; Quintela, T. Regulation of ABC transporters by sex steroids may explain differences in drug resistance between sexes. J. Physiol. Biochem. 2023, 79, 467–487. [Google Scholar] [CrossRef] [PubMed]
  87. Fujita, K.; Matsushita, M.; Banno, E.; De Velasco, M.A.; Hatano, K.; Nonomura, N.; Uemura, H. Gut microbiome and prostate cancer. Int. J. Urol. 2022, 29, 793–798. [Google Scholar] [CrossRef] [PubMed]
  88. Javier-DesLoges, J.; McKay, R.R.; Swafford, A.D.; Sepich-Poore, G.D.; Knight, R.; Parsons, J.K. The microbiome and prostate cancer. Prostate Cancer Prostatic Dis. 2022, 25, 159–164. [Google Scholar] [CrossRef] [PubMed]
  89. Fujita, K.; Matsushita, M.; De Velasco, M.A.; Hatano, K.; Minami, T.; Nonomura, N.; Uemura, H. The Gut-Prostate Axis: A New Perspective of Prostate Cancer Biology through the Gut Microbiome. Cancers 2023, 15, 1375. [Google Scholar] [CrossRef]
  90. Pernigoni, N.; Zagato, E.; Calcinotto, A.; Troiani, M.; Pereira Mestre, R.; Calì, B.; Attanasio, G.; Troisi, J.; Minini, M.; Mosole, S.; et al. Commensal bacteria promote endocrine resistance in prostate cancer through androgen biosynthesis. Science 2021, 374, 216–224. [Google Scholar] [CrossRef]
  91. Ridlon, J.M.; Ikegawa, S.; Alves, J.M.P.; Zhou, B.; Kobayashi, A.; Iida, T.; Mitamura, K.; Tanabe, G.; Serrano, M.; De Guzman, A.; et al. Clostridium scindens: A human gut microbe with a high potential to convert glucocorticoids into androgens. J. Lipid Res. 2013, 54, 2437–2449. [Google Scholar] [CrossRef]
  92. International Agency for Research on Cancer. Absence of excess body fatness. In IARC Handbooks of Cancer Prevention; IARC: Lyon, France, 2021; Volume 16, pp. 1–645. [Google Scholar]
  93. Miettinen, M.; Wang, Z.F.; Paetau, A.; Tan, S.H.; Dobi, A.; Srivastava, S.; Sesterhenn, I. ERG transcription factor as an immunohistochemical marker for vascular endothelial tumors and prostatic carcinoma. Am. J. Surg. Pathol. 2011, 35, 432–441. [Google Scholar] [CrossRef] [PubMed]
  94. Svensson, M.A.; Perner, S.; Ohlson, A.L.; Day, J.R.; Groskopf, J.; Kirsten, R.; Sollie, T.; Helenius, G.; Andersson, S.-O.; Demichelis, F.; et al. A Comparative Study of ERG Status Assessment on DNA, mRNA, and Protein Levels Using Unique Samples from a Swedish Biopsy Cohort. Appl. Immunohistochem. Mol. Morphol. 2014, 22, 136–141. [Google Scholar] [CrossRef]
  95. Randall, E.C.; Zadra, G.; Chetta, P.; Lopez, B.G.C.; Syamala, S.; Basu, S.S.; Agar, J.N.; Loda, M.; Tempany, C.M.; Fennessy, F.M.; et al. Molecular Characterization of Prostate Cancer with Associated Gleason Score Using Mass Spectrometry Imaging. Mol. Cancer Res. 2019, 17, 1155–1165. [Google Scholar] [CrossRef]
  96. Peng, C.K.; Wang, S.C.; Chen, S.L.; Sung, W.W. Multimodal Biomarkers That Predict the Presence of Gleason Pattern 4: Potential Impact for Active Surveillance. Letter. J. Urol. 2023, 210, 580–581. [Google Scholar] [CrossRef] [PubMed]
  97. Yuma, S.; Lia, M.; Fanga, Y.; Chen, Z.J. TBK1 recruitment to STING activates both IRF3 and NF-κB that mediate immune defense against tumors and viral infections. Proc. Natl. Acad. Sci. USA 2021, 118, e2100225118. [Google Scholar] [CrossRef] [PubMed]
  98. Zheng, W.; Feng, D.; Xiong, X.; Liao, X.; Wang, S.; Xu, H.; Le, W.; Wei, Q.; Yang, L. The Role of cGAS-STING in Age-Related Diseases from Mechanisms to Therapies. Aging Dis. 2023, 14, 1145. [Google Scholar] [CrossRef] [PubMed]
  99. Morel, K.L.; Sheahan, A.V.; Burkhart, D.L.; Baca, S.C.; Boufaied, N.; Liu, Y.; Qiu, X.; Cañadas, I.; Roehle, K.; Heckler, M.; et al. EZH2 inhibition activates a dsRNA–STING–interferon stress axis that potentiates response to PD-1 checkpoint blockade in prostate cancer. Nat. Cancer 2021, 2, 444–456. [Google Scholar] [CrossRef] [PubMed]
  100. Geng, C.; Zhang, M.C.; Manyam, G.C.; Vykoukal, J.V.; Fahrmann, J.F.; Peng, S.; Wu, C.; Park, S.; Kondraganti, S.; Wang, D.; et al. SPOP mutations target STING1 signaling in prostate cancer and create therapeutic vulnerabilities to PARP inhibitor–induced growth suppression. Clin. Cancer Res. 2023, 29, 4464–4478. [Google Scholar] [CrossRef]
  101. York, A.G.; Williams, K.J.; Argus, J.P.; Zhou, Q.D.; Brar, G.; Vergnes, L.; Gray, E.E.; Zhen, A.; Wu, N.C.; Yamada, D.H.; et al. Limiting Cholesterol Biosynthetic Flux Spontaneously Engages Type i IFN Signaling. Cell 2015, 163, 1716–1729. [Google Scholar] [CrossRef]
  102. Storozynsky, Q.; Hitt, M.M. The impact of radiation-induced dna damage on cgas-sting-mediated immune responses to cancer. Int. J. Mol. Sci. 2020, 21, 8877. [Google Scholar] [CrossRef] [PubMed]
  103. Haughey, C.M.; Mukherjee, D.; Steele, R.E.; Popple, A.; Dura-Perez, L.; Pickard, A.; Patel, M.; Jain, S.; Mullan, P.B.; Williams, R.; et al. Investigating radiotherapy response in a novel syngeneic model of prostate cancer. Cancers 2020, 12, 2804. [Google Scholar] [CrossRef] [PubMed]
  104. Kalluri, R. The biology and function of fibroblasts in cancer. Nat. Rev. Cancer 2016, 16, 582–598. [Google Scholar] [CrossRef]
  105. Shi, Y.; Du, L.; Lin, L.; Wang, Y. Tumour-associated mesenchymal stem/stromal cells: Emerging therapeutic targets. Nat. Rev. Drug Discov. 2017, 16, 35–52. [Google Scholar] [CrossRef] [PubMed]
  106. Yamauchi, M.; Barker, T.H.; Gibbons, D.L.; Kurie, J.M. The fibrotic tumor stroma. J. Clin. Investig. 2018, 128, 16–25. [Google Scholar] [CrossRef] [PubMed]
  107. Santi, A.; Kugeratski, F.G.; Zanivan, S. Cancer Associated Fibroblasts: The Architects of Stroma Remodeling. Proteomics 2018, 18, 1700167. [Google Scholar] [CrossRef] [PubMed]
  108. Öhlund, D.; Elyada, E.; Tuveson, D. Fibroblast heterogeneity in the cancer wound. J. Exp. Med. 2014, 211, 1503–1523. [Google Scholar] [CrossRef]
  109. Ahel, J.; Hudorović, N.; Vičić-Hudorović, V.; Nikles, H. TGF-beta in the natural history of prostate cancer. Acta Clin. Croat. 2019, 58, 128–138. [Google Scholar] [CrossRef]
  110. Mo, F.; Lin, D.; Takhar, M.; Ramnarine, V.R.; Dong, X.; Bell, R.H.; Volik, S.V.; Wang, K.; Xue, H.; Wang, Y.; et al. Stromal Gene Expression is Predictive for Metastatic Primary Prostate Cancer. Eur. Urol. 2018, 73, 524–532. [Google Scholar] [CrossRef]
  111. Franco, O.E.; Jiang, M.; Strand, D.W.; Peacock, J.; Fernandez, S.; Jackson, R.S.; Revelo, M.P.; Bhowmick, N.A.; Hayward, S.W. Altered TGF-β Signaling in a Subpopulation of Human Stromal Cells Promotes Prostatic Carcinogenesis. Cancer Res. 2011, 71, 1272–1281. [Google Scholar] [CrossRef]
  112. Taylor, R.A.; Toivanen, R.; Frydenberg, M.; Pedersen, J.; Harewood, L.; Australian Prostate Cancer Bioresource; Collins, A.T.; Maitland, N.J.; Risbridger, G.P. Human Epithelial Basal Cells Are Cells of Origin of Prostate Cancer, Independent of CD133 Status. Stem Cells 2012, 30, 1087–1096. [Google Scholar] [CrossRef]
  113. LeBleu, V.S.; Kalluri, R. A peek into cancer-associated fibroblasts: Origins, functions and translational impact. DMM Dis. Models Mech. 2018, 11, dmm029447. [Google Scholar] [CrossRef]
  114. Rudloff, M.W.; Zumbo, P.; Favret, N.R.; Roetman, J.J.; Detrés Román, C.R.; Erwin, M.M.; Murray, K.A.; Jonnakuti, S.T.; Dündar, F.; Betel, D.; et al. Hallmarks of CD8+ T cell dysfunction are established within hours of tumor antigen encounter before cell division. Nat. Immunol. 2023, 24, 1527–1539. [Google Scholar] [CrossRef] [PubMed]
  115. Cui, D.; Li, J.; Zhu, Z.; Berk, M.; Hardaway, A.; McManus, J.; Chung, Y.M.; Alyamani, M.; Valle, S.; Tiwari, R.; et al. Cancer-associated fibroblast-secreted glucosamine alters the androgen biosynthesis program in prostate cancer via HSD3B1 upregulation. J. Clin. Investig. 2023, 133, e161913. [Google Scholar] [CrossRef]
  116. Pienta, K.J.; Bradley, D. Mechanisms underlying the development of androgen-independent prostate cancer. Clin. Cancer Res. 2006, 12, 1665–1671. [Google Scholar] [CrossRef]
  117. Pimenta, R.; Camargo, J.A.; Candido, P.; Ghazarian, V.; Gonçalves, G.L.; Guimarães, V.R.; Romão, P.; Chiovatto, C.; Mioshi, C.M.; dos Santos, G.A.; et al. Cholesterol Triggers Nuclear Co-Association of Androgen Receptor, p160 Steroid Coactivators, and p300/CBP-Associated Factor Leading to Androgenic Axis Transactivation in Castration-Resistant Prostate Cancer. Cell. Physiol. Biochem. 2022, 56, 1–15. [Google Scholar]
  118. Chen, C.D.; Welsbie, D.S.; Tran, C.; Baek, S.H.; Chen, R.; Vessella, R.; Rosenfeld, M.G.; Sawyers, C.L. Molecular determinants of resistance to antiandrogen therapy. Nat. Med. 2004, 10, 33–39. [Google Scholar] [CrossRef] [PubMed]
  119. Attard, G.; Reid, A.H.M.; A’Hern, R.; Parker, C.; Oommen, N.B.; Folkerd, E.; Messiou, C.; Molife, L.R.; Maier, G.; Thompson, E.; et al. Selective Inhibition of CYP17 With Abiraterone Acetate Is Highly Active in the Treatment of Castration-Resistant Prostate Cancer. J. Clin. Oncol. 2009, 27, 3742–3748. [Google Scholar] [CrossRef] [PubMed]
  120. Gregory, C.W.; Johnson, R.T.; Mohler, J.L.; French, F.S.; Wilson, E.M. Androgen receptor stabilization in recurrent prostate cancer is associated with hypersensitivity to low androgen. Cancer Res. 2001, 61, 2892–2898. [Google Scholar] [PubMed]
  121. Zaalberg, A.; Minnee, E.; Mayayo-Peralta, I.; Schuurman, K.; Gregoricchio, S.; van Schaik, T.; Hoekman, L.; Li, D.; Corey, E.; Janssen, H.; et al. A genome-wide CRISPR screen in human prostate cancer cells reveals drivers of macrophage mediated cell killing and positions AR as a tumor-intrinsic immunomodulator. bioRxiv 2023. [Google Scholar] [CrossRef]
  122. El-Kenawi, A.; Dominguez-Viqueira, W.; Liu, M.; Awasthi, S.; Abraham-Miranda, J.; Keske, A.; Steiner, K.L.K.; Noel, L.; Serna, A.N.; Dhillon, J.; et al. Macrophage-derived cholesterol contributes to therapeutic resistance in prostate cancer. Cancer Res. 2021, 81, 5477–5490. [Google Scholar] [CrossRef] [PubMed]
  123. Liu, F.; Ma, M.; Gao, A.; Ma, F.; Ma, G.; Liu, P.; Jia, C.; Wang, Y.; Donahue, K.; Zhang, S.; et al. PKM2-TMEM33 axis regulates lipid homeostasis in cancer cells by controlling SCAP stability. EMBO J. 2021, 40, e108065. [Google Scholar] [CrossRef] [PubMed]
  124. Kim, S.H.; Richardson, M.; Chinnakannu, K.; Bai, V.U.; Menon, M.; Barrack, E.R.; Reddy, G.P.V. Androgen receptor interacts with telomeric proteins in prostate cancer cells. J. Biol. Chem. 2010, 285, 10472–10476. [Google Scholar] [CrossRef]
  125. Taheri, M.; Ghafouri-Fard, S.; Najafi, S.; Kallenbach, J.; Keramatfar, E.; Atri Roozbahani, G.; Heidari Horestani, M.; Hussen, B.M.; Baniahmad, A. Hormonal regulation of telomerase activity and hTERT expression in steroid-regulated tissues and cancer. Cancer Cell Int. 2022, 22, 258. [Google Scholar] [CrossRef]
  126. Bartsch, S.; Mirzakhani, K.; Neubert, L.; Stenzel, A.; Ehsani, M.; Esmaeili, M.; Pungsrinont, T.; Kacal, M.; Mohammad, S.; Rasa, M.; et al. Antithetic hTERT Regulation by Androgens in Prostate Cancer Cells: hTERT Inhibition Is Mediated by the ING1 and ING2 Tumor Suppressors. Cancers 2021, 13, 4025. [Google Scholar] [CrossRef]
  127. Arantes dos Santos, G.; Viana, N.I.; Pimenta, R.; Reis, S.T.; Ramos Moreira Leite, K.; Srougi, M. Hypothesis: The triad androgen receptor, zinc finger proteins and telomeres modulates the global gene expression pattern during prostate cancer progression. Med. Hypotheses 2021, 150, 110566. [Google Scholar] [CrossRef]
  128. Gasinska, A.; Jaszczynski, J.; Rychlik, U.; Łuczynska, E.; Pogodzinski, M.; Palaczynski, M. Prognostic Significance of Serum PSA Level and Telomerase, VEGF and GLUT-1 Protein Expression for the Biochemical Recurrence in Prostate Cancer Patients after Radical Prostatectomy. Pathol. Oncol. Res. 2020, 26, 1049–1056. [Google Scholar] [CrossRef]
  129. Athanassiadou, P.; Bantis, A.; Gonidi, M.; Liossi, A.; Aggelonidou, E.; Petrakakou, E.; Giannopoulos, A. Telomerase expression as a marker in prostate cancer: Correlation to clinicopathologic predictors. J. Exp. Clin. Cancer Res. 2003, 22, 613–618. [Google Scholar]
  130. Nanni, S.; Narducci, M.; Della Pietra, L.; Moretti, F.; Grasselli, A.; De Carli, P.; Sacchi, A.; Pontecorvi, A.; Farsetti, A. Signaling through estrogen receptors modulates telomerase activity in human prostate cancer. J. Clin. Investig. 2002, 110, 219–227. [Google Scholar] [CrossRef]
  131. Marti, N.; Galván, J.A.; Pandey, A.V.; Trippel, M.; Tapia, C.; Müller, M.; Perren, A.; Flück, C.E. Genes and proteins of the alternative steroid backdoor pathway for dihydrotestosterone synthesis are expressed in the human ovary and seem enhanced in the polycystic ovary syndrome. Mol. Cell Endocrinol. 2017, 441, 116–123. [Google Scholar] [CrossRef]
  132. Endo, A.; Kuroda, M.; Tsujita, Y. ML-236A, ML-236B, and ML-236C, new inhibitors of cholesterogenesis produced by Penicillium citrinium. J. Antibiot. 1976, 29, 1346–1348. [Google Scholar] [CrossRef]
  133. Mollazadeh, H.; Tavana, E.; Fanni, G.; Bo, S.; Banach, M.; Pirro, M.; von Haehling, S.; Jamialahmadi, T.; Sahebkar, A. Effects of statins on mitochondrial pathways. J. Cachexia Sarcopenia Muscle 2021, 12, 237–251. [Google Scholar] [CrossRef]
  134. Dulak, J.; Jozkowicz, A. Anti-Angiogenic and Anti-Inflammatory Effects of Statins: Relevance to Anti-Cancer Therapy. Curr. Cancer Drug Targets 2005, 5, 579–594. [Google Scholar] [CrossRef]
  135. Hamelin, B.A.; Turgeon, J. Hydrophilicity/ lipophilicity: Relevance for the pharmacology and clinical effects of HMG-CoA reductase inhibitors. Trends Pharmacol. Sci. 1998, 19, 26–37. [Google Scholar] [CrossRef]
  136. Kato, S.; Smalley, S.; Sadarangani, A.; Chen-Lin, K.; Oliva, B.; Brañes, J.; Carvajal, J.; Gejman, R.; Owen, G.I.; Cuello, M. Lipophilic but not hydrophilic statins selectively induce cell death in gynaecological cancers expressing high levels of HMGCoA reductase. J. Cell Mol. Med. 2010, 14, 1180. [Google Scholar]
  137. Bytyçi, I.; Bajraktari, G.; Bhatt, D.L.; Morgan, C.J.; Ahmed, A.; Aronow, W.S.; Banach, M. Hydrophilic vs lipophilic statins in coronary artery disease: A meta-analysis of randomized controlled trials. J. Clin. Lipidol. 2017, 11, 624–637. [Google Scholar] [CrossRef]
  138. Yue, S.; Li, J.; Lee, S.Y.; Lee, H.J.; Shao, T.; Song, B.; Cheng, L.; Masterson, T.A.; Liu, X.; Ratliff, T.L. Cholesteryl ester accumulation induced by PTEN loss and PI3K/AKT activation underlies human prostate cancer aggressiveness. Cell Metab. 2014, 19, 393–406. [Google Scholar] [CrossRef]
  139. Zhuang, L.; Kim, J.; Adam, R.M.; Solomon, K.R.; Freeman, M.R. Cholesterol targeting alters lipid raft composition and cell survival in prostate cancer cells and xenografts. J. Clin. Invest. 2005, 115, 959–968. [Google Scholar] [CrossRef] [PubMed]
  140. Allott, E.H.; Ebot, E.M.; Stopsack, K.H.; Gonzalez-Feliciano, A.G.; Markt, S.C.; Wilson, K.M.; Ahearn, T.U.; Gerke, T.A.; Downer, M.K.; Rider, J.R.; et al. Statin use is associated with lower risk of PTEN-null and lethal prostate cancer. Clin. Cancer Res. 2020, 26, 1086–1093. [Google Scholar] [CrossRef] [PubMed]
  141. Dorsch, M.; Kowalczyk, M.; Planque, M.; Heilmann, G.; Urban, S.; Dujardin, P.; Forster, J.; Ueffing, K.; Nothdurft, S.; Oeck, S.; et al. Statins affect cancer cell plasticity with distinct consequences for tumor progression and metastasis. Cell Rep. 2021, 37, 110056. [Google Scholar] [CrossRef] [PubMed]
  142. Deezagi, A.; Safari, N. Rosuvastatin inhibit spheroid formation and epithelial–mesenchymal transition (EMT) in prostate cancer PC-3 cell line. Mol. Biol. Rep. 2020, 47, 8727–8737. [Google Scholar] [CrossRef] [PubMed]
  143. Hoque, A.; Chen, H.; Xu, X.C. Statin Induces Apoptosis and Cell Growth Arrest in Prostate Cancer Cells. Cancer Epidemiol. Biomark. Prev. 2008, 17, 88–94. [Google Scholar] [CrossRef] [PubMed]
  144. Kobayashi, Y.; Kashima, H.; Rahmanto, Y.S.; Banno, K.; Yu, Y.; Matoba, Y.; Watanabe, K.; Iijima, M.; Takeda, T.; Kunitomi, H.; et al. Drug repositioning of mevalonate pathway inhibitors as antitumor agents for ovarian cancer. Oncotarget 2017, 8, 72147. [Google Scholar] [CrossRef]
  145. Russell, R.G.G.; Watts, N.B.; Ebetino, F.H.; Rogers, M.J. Mechanisms of action of bisphosphonates: Similarities and differences and their potential influence on clinical efficacy. Osteoporos. Int. 2008, 19, 733–759. [Google Scholar] [CrossRef]
  146. Rogers, M.J.; Crockett, J.C.; Coxon, F.P.; Mönkkönen, J. Biochemical and molecular mechanisms of action of bisphosphonates. Bone 2011, 49, 34–41. [Google Scholar] [CrossRef] [PubMed]
  147. Clézardin, P. Bisphosphonates’ antitumor activity: An unravelled side of a multifaceted drug class. Bone 2011, 48, 71–79. [Google Scholar] [CrossRef]
  148. Mann, D.M.; Woodward, M.; Muntner, P.; Falzon, L.; Kronish, I. Predictors of nonadherence to statins: A systematic review and meta-analysis. Ann. Pharmacother. 2010, 44, 1410–1421. [Google Scholar] [CrossRef]
  149. Lemstra, M.; Blackburn, D.; Crawley, A.; Fung, R. Proportion and risk indicators of nonadherence to statin therapy: A meta-analysis. Can. J. Cardiol. 2012, 28, 574–580. [Google Scholar] [CrossRef]
  150. Hirota, T.; Fujita, Y.; Ieiri, I. An updated review of pharmacokinetic drug interactions and pharmacogenetics of statins. Expert Opin. Drug Metab. Toxicol. 2020, 16, 809–822. [Google Scholar] [CrossRef]
  151. Ochoa-Rosales, C.; Portilla-Fernandez, E.; Nano, J.; Wilson, R.; Lehne, B.; Mishra, P.P.; Gao, X.; Ghanbari, M.; Rueda-Ochoa, O.L.; Juvinao-Quintero, D.; et al. Epigenetic Link Between Statin Therapy and Type 2 Diabetes. Diabetes Care 2020, 43, 875–884. [Google Scholar] [CrossRef]
  152. Qu, H.; Guo, M.; Chai, H.; Wang, W.T.; Ga, Z.Y.; Shi, D.Z. Effects of coenzyme Q10 on statin-induced myopathy: An updated meta-analysis of randomized controlled trials. J. Am. Heart Assoc. 2018, 7, e009835. [Google Scholar] [CrossRef]
  153. Sattar, N. Statins and diabetes: What are the connections? Best Pract. Res. Clin. Endocrinol. Metab. 2023, 37, 101749. [Google Scholar] [CrossRef]
  154. Ward, N.C.; Watts, G.F.; Eckel, R.H. Statin Toxicity: Mechanistic Insights and Clinical Implications. Circ. Res. 2019, 124, 328–350. [Google Scholar] [CrossRef] [PubMed]
  155. Zaleski, A.L.; Taylor, B.A.; Thompson, P.D. Coenzyme Q10 as Treatment for Statin-Associated Muscle Symptoms—A Good Idea, but…. Adv. Nutr. 2018, 9, 519S–523S. [Google Scholar] [CrossRef]
  156. Marcoff, L.; Thompson, P.D. The Role of Coenzyme Q10 in Statin-Associated Myopathy: A Systematic Review. J. Am. Coll. Cardiol. 2007, 49, 2231–2237. [Google Scholar] [CrossRef]
  157. Vaughan, R.A.; Garcia-Smith, R.; Bisoffi, M.; Conn, C.A.; Trujillo, K.A. Ubiquinol rescues simvastatin-suppression of mitochondrial content, function and metabolism: Implications for statin-induced rhabdomyolysis. Eur. J. Pharmacol. 2013, 711, 1–9. [Google Scholar] [CrossRef]
  158. Sirvent, P.; Mercier, J.; Vassort, G.; Lacampagne, A. Simvastatin triggers mitochondria-induced Ca2+ signaling alteration in skeletal muscle. Biochem. Biophys. Res. Commun. 2005, 329, 1067–1075. [Google Scholar] [CrossRef]
  159. Herminghaus, A.; Laser, E.; Schulz, J.; Truse, R.; Vollmer, C.; Picker, I.B.O. Pravastatin and Gemfibrozil Modulate Differently Hepatic and Colonic Mitochondrial Respiration in Tissue Homogenates from Healthy Rats. Cells 2019, 8, 983. [Google Scholar] [CrossRef] [PubMed]
  160. Sacher, J.; Weigl, L.; Werner, M.; Szegedi, C.; Hohenegger, M. Delineation of myotoxicity induced by 3-hydroxy-3-methylglutaryl CoA reductase inhibitors in human skeletal muscle cells. J. Pharmacol. Exp. Ther. 2005, 314, 1032–1041. [Google Scholar] [CrossRef]
  161. Murtola, T.J.; Syvälä, H.; Tolonen, T.; Helminen, M.; Riikonen, J.; Koskimäki, J.; Pakarainen, T.; Kaipia, A.; Isotalo, T.; Kujala, P.; et al. Atorvastatin Versus Placebo for Prostate Cancer Before Radical Prostatectomy—A Randomized, Double-blind, Placebo-controlled Clinical Trial. Eur. Urol. 2018, 74, 697–701. [Google Scholar] [CrossRef] [PubMed]
  162. Murtola, T.J.; Siltari, A. Statins for prostate cancer: When and how much? Clin. Cancer Res. 2021, 27, 4947–4949. [Google Scholar] [CrossRef] [PubMed]
  163. Longo, J.; Hamilton, R.J.; Masoomian, M.; Khurram, N.; Branchard, E.; Mullen, P.J.; Elbaz, M.; Hersey, K.; Chadwick, D.; Ghai, S.; et al. A pilot window-of-opportunity study of preoperative fluvastatin in localized prostate cancer. Prostate Cancer Prostatic Dis. 2020, 23, 630–637. [Google Scholar] [CrossRef] [PubMed]
  164. Molendijk, J.; Robinson, H.; Djuric, Z.; Hill, M.M. Lipid mechanisms in hallmarks of cancer. Mol. Omics 2020, 16, 6–18. [Google Scholar] [CrossRef] [PubMed]
  165. Nonomura, N.; Takayama, H.; Nakayama, M.; Nakai, Y.; Kawashima, A.; Mukai, M.; Nagahara, A.; Aozasa, K.; Tsujimura, A. Infiltration of tumour-associated macrophages in prostate biopsy specimens is predictive of disease progression after hormonal therapy for prostate cancer. BJU Int. 2011, 107, 1918–1922. [Google Scholar] [CrossRef]
  166. Eder, T.; Weber, A.; Neuwirt, H.; Grünbacher, G.; Ploner, C.; Klocker, H.; Sampson, N.; Eder, I. Cancer-Associated Fibroblasts Modify the Response of Prostate Cancer Cells to Androgen and Anti-Androgens in Three-Dimensional Spheroid Culture. Int. J. Mol. Sci. 2016, 17, 1458. [Google Scholar] [CrossRef] [PubMed]
  167. Neuwirt, H.; Bouchal, J.; Kharaishvili, G.; Ploner, C.; Jöhrer, K.; Pitterl, F.; Weber, A.; Klocker, H.; Eder, I.E. Cancer-associated fibroblasts promote prostate tumor growth and progression through upregulation of cholesterol and steroid biosynthesis. Cell Commun. Signal. 2020, 18, 11. [Google Scholar] [CrossRef]
  168. Owen, J.S.; Clayton, A.; Pearson, H.B. Cancer-Associated Fibroblast Heterogeneity, Activation and Function: Implications for Prostate Cancer. Biomolecules 2023, 13, 67. [Google Scholar] [CrossRef]
  169. Barbalata, C.I.; Tefas, L.R.; Achim, M.; Tomuta, I.; Porfire, A.S. Statins in risk-reduction and treatment of cancer. World J. Clin. Oncol. 2020, 11, 573–588. [Google Scholar] [CrossRef]
  170. Orozco-Beltrán, D.; Brotons Cuixart, C.; Banegas Banegas, J.R.; Gil Guillén, V.F.; Cebrián Cuenca, A.M.; Martín Rioboó, E.; Jordá Baldó, A.; Vicuña, J.; Navarro Pérez, J. Cardiovascular preventive recommendations. PAPPS 2022 thematic updates. Working groups of the PAPPS. Aten Primaria 2022, 54, 102444. [Google Scholar] [CrossRef]
  171. Dickerman, B.A.; García-Albéniz, X.; Logan, R.W.; Denaxas, S.; Hernán, M.A. Avoidable flaws in observational analyses: An application to statins and cancer. Nat. Med. 2019, 25, 1601–1606. [Google Scholar] [CrossRef]
  172. Freedland, S.J.; Hamilton, R.J.; Gerber, L.; Banez, L.L.; Moreira, D.M.; Andriole, G.L.; Rittmaster, R.S. Statin use and risk of prostate cancer and high-grade prostate cancer: Results from the REDUCE study. Prostate Cancer Prostatic Dis. 2013, 16, 254–259. [Google Scholar] [CrossRef]
  173. Jamnagerwalla, J.; Howard, L.E.; Allott, E.H.; Vidal, A.C.; Moreira, D.M.; Castro-Santamaria, R.; Freeman, M.R.; Freedland, S.J. Serum cholesterol and risk of high-grade prostate cancer: Results from the REDUCE study. Prostate Cancer Prostatic Dis. 2018, 21, 252–259. [Google Scholar] [CrossRef]
  174. Bansal, D.; Undela, K.; D’cruz, S.; Schifano, F. Statin Use and Risk of Prostate Cancer: A Meta-Analysis of Observational Studies. PLoS ONE 2012, 7, e46691. [Google Scholar] [CrossRef] [PubMed]
  175. Allott, E.H.; Arab, L.; Su, L.J.; Farnan, L.; Fontham, E.T.H.; Mohler, J.L.; Bensen, J.T.; Steck, S.E. Saturated fat intake and prostate cancer aggressiveness: Results from the population-based North Carolina-Louisiana Prostate Cancer Project. Prostate Cancer Prostatic Dis. 2017, 20, 48–54. [Google Scholar] [CrossRef]
  176. Mondul, A.M.; Joshu, C.E.; Barber, J.R.; Prizment, A.E.; Bhavsar, N.A.; Selvin, E.; Folsom, A.R.; Platz, E.A. Longer-term lipid-lowering drug use and risk of incident and fatal prostate cancer in black and white men in the ARIC Study HHS Public Access. Cancer Prev. Res. 2018, 11, 779–788. [Google Scholar] [CrossRef]
  177. Hamilton, R. Making sense of the statin-prostate cancer relationship: Is it time for a randomized controlled trial? Eur. Urol. Focus 2017, 3, 221–222. [Google Scholar] [CrossRef]
  178. Wilson, B.E.; Armstrong, A.J.; de Bono, J.; Sternberg, C.N.; Ryan, C.J.; Scher, H.I.; Smith, M.R.; Rathkopf, D.; Logothetis, C.J.; Chi, K.N. Effects of metformin and statins on outcomes in men with castration-resistant metastatic prostate cancer: Secondary analysis of COU-AA-301 and COU-AA-302. Eur. J. Cancer 2022, 170, 296–304. [Google Scholar] [CrossRef]
  179. Joshua, A.M.; Armstrong, A.; Crumbaker, M.; Scher, H.I.; de Bono, J.; Tombal, B.; Hussain, M.; Sternberg, C.N.; Gillessen, S.; Carles, J.; et al. Statin and metformin use and outcomes in patients with castration-resistant prostate cancer treated with enzalutamide: A meta-analysis of AFFIRM, PREVAIL and PROSPER. Eur. J. Cancer 2022, 170, 285–295. [Google Scholar] [CrossRef]
  180. Jayalath, V.H.; Lajkosz, K.; Fleshner, N.E.; Hamilton, R.J.; Jenkins, D.J.A. The effect of lowering cholesterol through diet on serum prostate-specific antigen levels: A secondary analysis of clinical trials. Can. Urol. Assoc. J. 2022, 16, 279–282. [Google Scholar] [CrossRef] [PubMed]
  181. Joentausta, R.M.; Rannikko, A.; Murtola, T.J. Prostate cancer survival among statin users after prostatectomy in a Finnish nationwide cohort. Prostate 2019, 79, 583–591. [Google Scholar] [CrossRef] [PubMed]
  182. Yu, O.; Eberg, M.; Benayoun, S.; Aprikian, A.; Batist, G.; Suissa, S.; Azoulay, L. Use of statins and the risk of death in patients with prostate cancer. J. Clin. Oncol. 2014, 32, 5–11. [Google Scholar] [CrossRef]
  183. Hou, Y.C.; Shao, Y.H. The Effects of Statins on Prostate Cancer Patients Receiving Androgen Deprivation Therapy or Definitive Therapy: A Systematic Review and Meta-Analysis. Pharmaceuticals 2022, 15, 131. [Google Scholar] [CrossRef] [PubMed]
  184. Jeong, I.G.; Lim, B.; Yun, S.C.; Lim, J.H.; Hong, J.H.; Kim, C.S. Adjuvant low-dose statin use after radical prostatectomy: The PRO-STAT randomized clinical trial. Clin. Cancer Res. 2021, 27, 5004–5011. [Google Scholar] [CrossRef] [PubMed]
  185. Wang, K.; Gerke, T.A.; Chen, X.; Prosperi, M. Association of statin use with risk of Gleason score-specific prostate cancer: A hospital-based cohort study. Cancer Med. 2019, 8, 7399–7407. [Google Scholar] [CrossRef] [PubMed]
Figure 1. MVA pathway and related molecules. LDLR, low-density lipoprotein receptor; GA3P, glyceraldehyde 3-phosphate dehydrogenase; OAA, oxalacetate; Malonyl-CoA, malonyl coenzyme A; Acetyl-CoA, acetyl coenzyme A; Ac-Ac-CoA, aceto-acetyl coenzyme A; HMG-CoA, 3-hidroxi-3-metilglutaril-coenzima A or β-hidroxi-β-metilglutaril-coenzima A; FPP, farnesyl pyrophosphate; GPP, geranyl pyrophosphate; FASN, fatty acid synthase; SREBP, sterol regulatory element-binding protein; HSP90, heat shock protein 90; SPRING, SREBF pathway regulator in Golgi; COPII, Coat Protein Complex II; INSIGS, insulin-induced genes; SCAP, SREBP cleavage-activating protein; AR, androgen receptor; ERα, estrogen receptor α; ARE, androgen response elements; ERE, estrogen response elements; SRE, sterol response elements.
Figure 1. MVA pathway and related molecules. LDLR, low-density lipoprotein receptor; GA3P, glyceraldehyde 3-phosphate dehydrogenase; OAA, oxalacetate; Malonyl-CoA, malonyl coenzyme A; Acetyl-CoA, acetyl coenzyme A; Ac-Ac-CoA, aceto-acetyl coenzyme A; HMG-CoA, 3-hidroxi-3-metilglutaril-coenzima A or β-hidroxi-β-metilglutaril-coenzima A; FPP, farnesyl pyrophosphate; GPP, geranyl pyrophosphate; FASN, fatty acid synthase; SREBP, sterol regulatory element-binding protein; HSP90, heat shock protein 90; SPRING, SREBF pathway regulator in Golgi; COPII, Coat Protein Complex II; INSIGS, insulin-induced genes; SCAP, SREBP cleavage-activating protein; AR, androgen receptor; ERα, estrogen receptor α; ARE, androgen response elements; ERE, estrogen response elements; SRE, sterol response elements.
Ijms 25 02152 g001
Figure 2. Transcriptional regulators of AR and related molecules in PCa. ER, estrogen receptor; TGFBR, tumor growth factor β receptor; IGF1R, insulin-like growth factor 1 receptor; EGFR, epithelial growth factor receptor; TRPM8, transient receptor potential cation channel subfamily M (melastatin) member 8; ABCA1, ATP-binding cassette subfamily A member 1; LDLR, LDL cholesterol receptor; DRP1, dynamin-related protein; DHT, dihydrotestosterone; 5α-R, 5 alpha receptor; DHEA, dehydroepyandrosterone; PTEN, phosphatase and tensin homolog; cGAS, cyclic GMP-AMP synthase; PKM2, pyruvate kinase M2; TMEM33, transmembrane protein 33; STING, stimulator of interferon response cGAMP interactor 1; NRF1, nuclear respiratory factor 1; ARE, androgen response element.
Figure 2. Transcriptional regulators of AR and related molecules in PCa. ER, estrogen receptor; TGFBR, tumor growth factor β receptor; IGF1R, insulin-like growth factor 1 receptor; EGFR, epithelial growth factor receptor; TRPM8, transient receptor potential cation channel subfamily M (melastatin) member 8; ABCA1, ATP-binding cassette subfamily A member 1; LDLR, LDL cholesterol receptor; DRP1, dynamin-related protein; DHT, dihydrotestosterone; 5α-R, 5 alpha receptor; DHEA, dehydroepyandrosterone; PTEN, phosphatase and tensin homolog; cGAS, cyclic GMP-AMP synthase; PKM2, pyruvate kinase M2; TMEM33, transmembrane protein 33; STING, stimulator of interferon response cGAMP interactor 1; NRF1, nuclear respiratory factor 1; ARE, androgen response element.
Ijms 25 02152 g002
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

Guerrero-Ochoa, P.; Rodríguez-Zapater, S.; Anel, A.; Esteban, L.M.; Camón-Fernández, A.; Espilez-Ortiz, R.; Gil-Sanz, M.J.; Borque-Fernando, Á. Prostate Cancer and the Mevalonate Pathway. Int. J. Mol. Sci. 2024, 25, 2152. https://doi.org/10.3390/ijms25042152

AMA Style

Guerrero-Ochoa P, Rodríguez-Zapater S, Anel A, Esteban LM, Camón-Fernández A, Espilez-Ortiz R, Gil-Sanz MJ, Borque-Fernando Á. Prostate Cancer and the Mevalonate Pathway. International Journal of Molecular Sciences. 2024; 25(4):2152. https://doi.org/10.3390/ijms25042152

Chicago/Turabian Style

Guerrero-Ochoa, Patricia, Sergio Rodríguez-Zapater, Alberto Anel, Luis Mariano Esteban, Alejandro Camón-Fernández, Raquel Espilez-Ortiz, María Jesús Gil-Sanz, and Ángel Borque-Fernando. 2024. "Prostate Cancer and the Mevalonate Pathway" International Journal of Molecular Sciences 25, no. 4: 2152. https://doi.org/10.3390/ijms25042152

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