Next Article in Journal
A Clinical Prognostic Model Based on Machine Learning from the Fondazione Italiana Linfomi (FIL) MCL0208 Phase III Trial
Next Article in Special Issue
Metabolic Reprogramming of Ovarian Cancer Spheroids during Adhesion
Previous Article in Journal
The Role of Dielectrophoresis for Cancer Diagnosis and Prognosis
Previous Article in Special Issue
Evaluating Targeted Therapies in Ovarian Cancer Metabolism: Novel Role for PCSK9 and Second Generation mTOR Inhibitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Folate Transport and One-Carbon Metabolism in Targeted Therapies of Epithelial Ovarian Cancer

by
Adrianne Wallace-Povirk
1,2,
Zhanjun Hou
1,2,
Md. Junayed Nayeen
3,
Aleem Gangjee
3 and
Larry H. Matherly
1,2,4,*
1
Molecular Therapeutics Program, Barbara Ann Karmanos Cancer Institute, 421 East Canfield Street, Detroit, MI 48201, USA
2
Department of Oncology, Wayne State University School of Medicine, Detroit, MI 48201, USA
3
Division of Medicinal Chemistry, Graduate School of Pharmaceutical Sciences, Duquesne University, Pittsburgh, PA 15282, USA
4
Department of Pharmacology, Wayne State University School of Medicine, Detroit, MI 48201, USA
*
Author to whom correspondence should be addressed.
Cancers 2022, 14(1), 191; https://doi.org/10.3390/cancers14010191
Submission received: 5 December 2021 / Revised: 23 December 2021 / Accepted: 27 December 2021 / Published: 31 December 2021
(This article belongs to the Special Issue Metabolism in Ovarian Cancer)

Abstract

:

Simple Summary

New therapies are urgently needed for ovarian cancer, the most lethal malignancy in women. To identify new approaches for targeting ovarian cancer, metabolic vulnerabilities must be discovered and strategies for the selective delivery of therapeutic agents must be established. New approaches that are tumor-selective and that facilitate the internalization of novel drugs or provide targets for therapy are being developed for treating ovarian cancer involving folate receptors and the proton-coupled folate transporter. New drugs are being discovered that target key metabolic processes in tumors and neighboring immune cells which contribute to tumor progression. In this review, we describe the remarkable advances in this rapidly evolving area and their extraordinary potential to improve the lives of women diagnosed with this devastating disease.

Abstract

New therapies are urgently needed for epithelial ovarian cancer (EOC), the most lethal gynecologic malignancy. To identify new approaches for targeting EOC, metabolic vulnerabilities must be discovered and strategies for the selective delivery of therapeutic agents must be established. Folate receptor (FR) α and the proton-coupled folate transporter (PCFT) are expressed in the majority of EOCs. FRβ is expressed on tumor-associated macrophages, a major infiltrating immune population in EOC. One-carbon (C1) metabolism is partitioned between the cytosol and mitochondria and is important for the synthesis of nucleotides, amino acids, glutathione, and other critical metabolites. Novel inhibitors are being developed with the potential for therapeutic targeting of tumors via FRs and the PCFT, as well as for inhibiting C1 metabolism. In this review, we summarize these exciting new developments in targeted therapies for both tumors and the tumor microenvironment in EOC.

1. Introduction

Epithelial ovarian cancer (EOC) is the leading cause of death in women diagnosed with gynecological cancers in the USA and accounts for ~90% of ovarian cancers [1]. The high mortality for patients diagnosed with EOC reflects the development of chemoresistance to standard cytotoxic chemotherapy in late-stage disease [2]. Clearly, there is an urgent need for new therapeutic strategies that provide longer disease-free intervals and improve overall survival, particularly for patients with platinum-resistant EOC who have limited treatment options.
In recent years, attention has focused on targeted therapies for EOC. While much attention has involved poly(adenosine diphosphate-ribose) polymerase (PARP) inhibitors in patients with BRCA mutations and homologous recombination deficiency [3,4], of particular interest is the targeting of EOC through the folate receptor (FR) proteins [5,6,7] and the proton-coupled folate transporter (PCFT) [8,9,10]. FRα is expressed in ~85% of EOCs [6,11,12,13], whereas the PCFT is expressed constitutively in EOCs [14]. Further, as the tumor microenvironment (TME) has emerged as an important determinant of disease progression for EOC, it is notable that tumor-associated macrophages (TAMs) express FRβ, thus providing new opportunities to suppress ovarian tumor progression by inhibiting TAMs with FR-targeted therapies [15,16,17,18].
One-carbon (C1) metabolism has remained an important therapeutic target for many cancers since the introduction of methotrexate over 60 years ago and, more recently, pemetrexed [19,20]. This reflects the central role of C1 metabolism in the synthesis of thymidylate, purines, serine, and methionine as well as in supporting biological methylation reactions from S-adenosylmethionine (SAM) [21,22,23,24,25].
With the recognition that C1 metabolism is compartmentalized between the cytosol and mitochondria, recent attention has focused on the therapeutic potential for targeting mitochondrial C1 metabolism in cancer, reflecting its critical role as a source of C1 units, glycine, reducing equivalents (e.g., NAD(P)H), and ATP [21,22,25]. However, the value of targeting C1 metabolism in EOC has yet to be realized, although there is increasing interest in developing targeted antifolates with FR and/or PCFT selectivity and that inhibit essential metabolic pathways or processes with therapeutic impact [8,10,19,21,26,27].
In this review, we provide an overview of C1 metabolism, including the recent major discoveries in this area. We discuss the potential for exploiting the vulnerabilities of EOC in C1 metabolism with novel antifolates and the selective uptake of targeted therapeutics by FRs and the PCFT in the context of current therapies and the unique biology of EOC.

2. Folate Homeostasis, Folate Transport, and One-Carbon Metabolism

2.1. Folate Homeostasis and Transport

Folates are members of the vitamin B9 family and are composed of a bicyclic pteridine ring which in cells can be reversibly reduced by dihydrofolate reductase with one (dihydrofolate) or two (tetrahydrofolate (THF)) reducing equivalents, a p-aminobenzoate, and an L-glutamate. Folates are hydrophilic anionic molecules and have limited capacities to diffuse across cell membranes. Extracellular folates use PCFT (SLC46A1) and RFC (SLC19A1), which are facilitive transporters, as well as FRs, for cellular uptake [6,9,28,29] (Figure 1). Each of these transport systems plays a unique role in the maintenance of C1 homeostasis.
The RFC (SLC19A1) is a folate-organic anion antiporter with optimal activity at pH 7.4 [9], whereas the PCFT (SLC46A1) is a folate-proton symporter that is optimally active at an acidic pH (pH 5.5) [30]. Both RFC and PCFT are members of the major facilitator superfamily of transport proteins [31].
RFC is ubiquitously expressed in tissues and tumors and is the major tissue transporter of folate cofactors (e.g., 5-methyl THF) from the systemic circulation [9,28,29]. RFC is highly expressed in the liver and placenta, with substantial levels in the kidney, lung, bone marrow, intestine, and central nervous system [32]. In mouse tissues, RFC is expressed at the basolateral membrane of renal tubule epithelia, the apical surface of the choroid plexus, hepatocyte membranes, and the apical membrane of cells lining the spinal canal [33]. While RFC is expressed on the apical brush border membrane of the small intestine and colon [33], RFC is not the primary transport system for the absorption of dietary folates in the gastrointestinal (GI) system (see below). Tissue-specific expression of human SLC19A1 gene is regulated by intricate transcriptional and posttranscriptional controls involving as many as six non-coding exons/promoters [9,32]. This likely reflects the unique tissue requirements for reduced folates and may in part relate to the pathophysiology of folate deficiency. Indeed, low levels of RFC are associated with pathophysiological conditions associated with folate deficiency, ranging from fetal abnormalities to cardiovascular disease, neurologic disorders, and cancer [9]. Further establishing the central role of RFC in folate homeostasis and biology, the loss of the SLC19A1 gene in the absence of folate supplementation is embryonic lethal, whereas folate supplementation in the absence of RFC is associated with developmental delays and is accompanied by developmental malformations [34,35]. In ovarian cancer, RFC levels did not correlate with disease stage, histologic grade, or histologic subtypes; however, RFC levels correlated with overall and disease-free survival [36].
PCFT was discovered as the major folate intestinal transporter in that it facilitates the absorption of dietary folates at the acidic pH of the upper GI tract [37]. Variants in the human SLC46A1 gene result in hereditary folate malabsorption due to a loss of PCFT expression or transport activity and are accompanied by severely low levels of systemic and cerebral folates [38]. Although PCFT is also expressed in the choroid plexus, kidney, liver, placenta and spleen [8,29,39], it is not considered to be a major transporter of folates in most normal tissues. Further, even when expressed, PCFT transport activity is very modest in most tissues outside of the upper GI due to a bicarbonate inhibition of transport at a neutral pH [40].
Interestingly, PCFT is highly expressed in a variety of solid tumors, including colorectal adenocarcinoma, malignant pleural mesothelioma, non-small cell lung cancer, and EOC [14,41,42,43]. Conversely, PCFT is expressed at very low levels in leukemias. In at least some cases, low levels of PCFT in human tissues reflect SLC46A1 promoter hypermethylation, as PCFT can be induced by treatment with 5-aza-2′-deoxycytidine [42,44,45]. Given its substantial transport activity at an acidic pH, including those associated with the low pH conditions of the tumor microenvironment [46], there is growing interest in developing targeted therapeutics for cancer with PCFT transport selectivity over that mediated by RFC [10,19,41].
FRs consist of α, β, γ, and δ isoforms encoded from a multigene family (FOLR1, FOLR2, FOLR3, and FOLR4, respectively). FRα, FRβ, and FRδ are tethered to the cell membrane through a glycosylphosphatidylinositol (GPI) linkage, whereas FRγ is not [6,47]. FRα is unique in its tissue specificity in that it is expressed on apical membranes in a small number of healthy epithelial tissues, including those of the female reproductive tract, as well as the kidney, lung, choroid plexus, and placenta [6,12,13,48,49]. FRα serves an important function in embryogenesis [50]. FRβ is expressed in placenta, mature neutrophils, and activated monocytes and macrophages [51,52,53]. FRγ is secreted and is expressed in hematopoietic tissues [54]. FRδ is expressed in the membranes of gametes [47]. (Anti)folate uptake by the membrane-associated FRs involves the high-affinity binding of folate at the cell surface, followed by endocytosis through which the FRs on the cell surface are internalized due to the formation of cytosolic vesicles [55]. These vesicles are acidified to effect dissociation of the bound ligand, which is subsequently released to the cytosol [55]. Compared to facilitated (anti)folate transport by RFC and PCFT, cellular uptake through endocytosis by the GPI-linked FRs is inefficient [6,29].
FRα is expressed on the membrane surface of non-small cell lung cancers, triple negative breast cancers, and kidney, endometrial, and colorectal cancers [6,11,12,13,48,49]. Pancreas, brain, gastric, liver, prostate, and bladder cancers are reported to express FRα in lower levels [6,11,12,13,48,49]. In EOC, FRα is highly expressed and increases with disease stage [12,14,36,56]. FRα in tumors is generally accessible to the circulation. However, in normal (polarized) tissues, with the exception of the placenta, FRα localizes to luminal membranes without exposure to systemic circulation [6,11,13]. These features provide a compelling rationale for developing selective FR-targeted therapeutics for solid tumors. The FOLR1 gene encoding FRα includes two TATA-less promoters (P1, P4) [57]. In malignant cells, the FOLR1 P4 promoter appears to predominate. Tissue-specific expression of FRα involves both transcriptional and post-transcriptional mechanisms, including alternate FOLR1 promoters and mRNA splicing [57], as well as differential translation of mRNAs [58]. In contrast to FRα, FRβ is transcribed from a single FOLR2 promoter [59]. The FOLR1 P4 promoter is repressed by estrogen receptor [60] and activated by dexamethasone [61]. In ovarian cancer, FRα may [36] or may not [62] be associated with FOLR1 gene amplification (chromosome 11q13.3).

2.2. C1 Metabolism

C1 metabolism and folates are compartmentalized in the cytosol and the mitochondria of mammalian cells (Figure 1) [21,22,25]. Following internalization, cytosolic folates are transported into the mitochondria by a mitochondrial folate transporter (SLC25A32) [63,64]. Both cytosolic and mitochondrial folates are metabolized to polyglutamate conjugates, catalyzed by alternate isoforms of folylpoly-γ-glutamate synthetase (FPGS) encoded by a single FPGS gene [65]. FPGS catalyzes the addition of up to eight additional glutamate residues to the γ-carboxyl of the terminal glutamate of the parent folate molecule [66]. Once polyglutamylated, cytosolic folates are preferentially retained in cells [66] and do not exchange with the mitochondrial folate polyglutamyl forms [65]. For most C1 transfer reactions, polyglutamyl folate forms are the preferred substrates over the non-polyglutamyl or monoglutamate folate forms [66].
Serine is synthesized from glucose via the glycolytic intermediate 3-phosphoglycerate. Phosphoglycerate dehydrogenase reduces 3-phosphoglycerate to 3-phosphohydroxypyruvate, which is then converted to serine via transamination and phosphoserine hydrolysis reactions. Serine serves as the principal source of C1 units for cellular biosynthesis. Following its transport into the mitochondria [67], serine is metabolized sequentially by serine hydroxymethyl transferase (SHMT) 2 to glycine and 5,10-methylene THF (Figure 1). The 5,10-methylene THF is oxidized in an NAD(P)+ dependent reaction by 5,10-methylene THF dehydrogenase (MTHFD) −2 or −2L to 10-formyl THF. The 10-formyl THF is hydrolyzed by MTHFDL1 to THF and formate, which passes to the cytosol for utilization in C1 biosynthetic reactions (Figure 1) [21,22,68,69,70]. Serine catabolism in the mitochondria serves as the principal source of C1 units and glycine for cellular biosynthesis, including the de novo synthesis of purine nucleotides and thymidylate in the cytosol [21,22,68,69,70]. Further, mitochondrial C1 metabolism is an important source of NAD(P)H and glycine for the synthesis of glutathione as well as ATP (Figure 1) [21,22,68,69,70]. Mitochondrial NADPH is also generated by the catabolism of 10-formyl-THF to CO2 and THF through aldehyde dehydrogenase 1 family member L2 (ALDH1L2) [71].
In the cytosol, 10-formyl THF is resynthesized from THF and formate by the 10-formyl THF synthetase activity of the trifunctional enzyme MTHFD1 (Figure 1). The MTHFD1 cyclohydrolase and dehydrogenase activities further convert the 10-formyl THF to 5,10-methylene THF and the oxidation of NADPH to NADP+. For de novo purine biosynthesis, 10-formyl THF is required, which involves 10 reactions from phosphoribosyl pyrophosphate (PRPP) to IMP [72] (Figure 1). Glycine is incorporated into glycinamide ribonucleotide (GAR) by GAR synthetase. C1 units from 10-formyl THF are used in reactions catalyzed by glycinamide ribonucleotide formyl transferase (GARFTase) and 5-aminoimidazole-4-carboxamide ribonucleotide (AICAR) formyl transferase (ATIC). Thus, both the glycine and formate resulting from mitochondrial C1 metabolism are incorporated into the purine ring. To facilitate an efficient flux of pathway intermediates, de novo purine biosynthetic enzymes are assembled into a “purinosome” [73], which mediates passage of the key intermediates, including the THF cofactors between the sequential steps from PRPP to AMP and GMP. Purinosomes are reported to co-localize with mitochondria to further maximize the metabolic efficiency of purine biosynthesis [73].
SHMT1 is the cytosolic homolog of SHMT2 in mitochondria and catalyzes the reverse reaction, whereby glycine is converted to serine with a C1 unit from 5,10-methylene THF, forming THF (Figure 1). Thymidylate synthase converts dUMP to thymidylate, with 5,10-methylene THF providing a C1 unit and reducing equivalent, generating dihydrofolate. Dihydrofolate is reduced back to active THF by dihydrofolate reductase. 5,10-Methylene THF can also be reduced to 5-methyl THF in a NAD(P)H-dependent reaction catalyzed by 5,10-methylene THF reductase (MTHFR) (Figure 1). 5-Methyl THF provides C1 units for the vitamin B12-dependent conversion of homocysteine to methionine catalyzed by methionine synthase (MTR). Methionine is further metabolized to SAM by methionine adenosyltransferase (MAT), required for the methylation of DNA, phospholipids, and proteins [74] (Figure 1). DNA methylation mediates the silencing of transcription and aberrant gene promoter methylation. The role of DNA methylation in the progression and chemoresistance in high-grade serous ovarian cancer at both the genome-wide and individual gene level has been recently reviewed [75].
Mitochondrial SHMT2 and serine provide C1 units for cellular anabolism in the cytosol. The opposing directionality of the C1 flux in each compartment (i.e., serine-to-formate in the mitochondria; formate-to-serine in the cytosol) reflects differences in the electrochemical potential between the mitochondrial NAD(P)H [high NAD(P)+/NAD(P)H] and cytosolic NADPH [low NAD(P)+/NAD(P)H] [25]. Without this, the cytosolic NAD+ pools would soon be depleted, resulting in decreased glycolysis accompanying serine oxidation and the generation of NADH [22]. Interestingly, mitochondrial C1 also stabilizes the THF from degradation in the cytosol by providing formate for the synthesis of 10-formyl-THF by MTHFD1 [76].
Notably, serine catabolism is elevated in cancer, and the genes encoding MTHFD2 and SHMT2 are among the top five differentially expressed genes in tumors and normal tissues for a wide range of cancers, including EOC [77]. Their high levels of expression in cancer cells may reflect their regulation by MYC as MYC binds to the SHMT2 and MTHFD2 gene promoters, along with that for MTHFD1L [78,79]. Whereas MTHFD2 is considered a tumor-selective target [80,81], a recent study reported that normal and cancer cells express both MTHFD2 and MTHFD2L [81].

3. Therapeutic Challenges of Treating Ovarian Cancer

For 2021, it is estimated that there will be 21,410 new cases of ovarian cancer and that an estimated 13,770 women will die of this disease [82]. Ovarian cancer accounts for 5% of cancer deaths among women, far greater than any other gynecologic cancer [1]. Ovarian cancer is a dismal disease that is frequently diagnosed at a late stage. The 5-year overall survival rate is less than 40% [83]. Women present at an advanced stage due to a lack of early detection strategies and symptoms frequently being confused with benign conditions [84,85]. The World Health Organization classification includes six subtypes of epithelial ovarian tumors including serous, mucinous, endometroid, clear cell, transitional cells, and squamous carcinoma [86]. Approximately 10% of all ovarian cancers are non-epithelial (i.e., germ cell tumors, sex cord origin, small cell carcinomas, and sarcomas) [87]. Of the EOC subtypes, 70% are classified as high-grade serous adenocarcinomas [88].
Although ovarian cancer is a disease of aging with >80% of the cases diagnosed in postmenopausal women over age 50, family history and genetics nonetheless play a role in the risk of disease [86]. Women who have a first-degree relative with ovarian cancer have nearly a two-fold increased risk of disease. However, even when genetic testing is performed after diagnosis, only a small number of cases have identifiable and actionable mutations. In terms of mutation status, 5–15% of ovarian cancers show BRCA mutations [89,90]. Germline BRCA1 and BRCA2 mutations are present in 14–15% of EOCs [91,92] and up to 22.6% of high-grade serous ovarian cancers [91,92]. Somatic BRCA1 and BRCA2 mutations have been reported in 6–7% of high-grade serous EOCs [93]. BRCA1 promoter hypermethylation occurs in 10–20% of high-grade serous ovarian cancers [94,95]. Other reported mutations include Fanconi anemia genes (PALB2, FANCA, FANCI, FANCL, and FANCC), RAD genes (RAD50, RAD51, RAD51C, and RAD54L) and DNA damage response genes (ATM, ATR, CHEK1, and CHEK2) [96]. More than 80% of ovarian cancers have p53 gene mutations [97,98].
Treatment of ovarian cancer has largely remained unchanged over several decades and generally consists of debulking surgery, followed by six-to-eight cycles of chemotherapy. The principal goal of surgery is to achieve complete cytoreduction of all macroscopic disease, as this is associated with a substantial increase in long-term survival and progression-free survival (PFS) [99,100,101]. The standard therapeutic regimen for ovarian cancer includes a combination of paclitaxel and carboplatin administered intravenously once every three weeks [102]. Three or more cycles of neoadjuvant chemotherapy prior to the debulking surgery and adjuvant chemotherapy is an alternative for some patients [103] and provides an opportunity to test for chemosensitivity a priori to identify patients at higher risk of relapse. Most women initially respond to platinum-based therapies; however, subsequent rounds of therapy generally lead to less efficacy and eventually to chemotherapy resistance [104]. Patients are initially characterized as platinum sensitive if their progression-free interval is more than 6 months, and patients who have disease progression in less than 6 months are considered platinum resistant [104].
In recent years, treatment has been increasingly concentrated in specialized centers with an emphasis on centralizing ovarian cancer care and through the use of multidisciplinary teams [105,106]. Targeted and combination therapies have assumed a greater role in the treatment of ovarian cancer. Combination therapies with a platinum-based agent and gemcitabine and/or bevacizumab are extensively used and show significantly better PFS than platinum monotherapy [107]. An exciting advancement has been the development of PARP inhibitors (e.g., olaparib, rucaparib, and niraparib) which are less toxic than the standard platinum agents for both treatment and maintenance therapies [4,108]. PARP inhibitors interfere with the ability to repair DNA damage and exploit BRCA1 and −2 germline mutations and deficiencies in the DNA damage response which collectively are believed to occur in up to 50% of high-grade EOCs [4].

4. The Role of FRα in the Treatment of EOC

FRα is expressed in ~85% of EOCs [6,11,12,13]. Differences in FRα expression have been reported in all of the different histologic subtypes of ovarian cancer [109]. FRα expression increases with disease stage and can span a wide range within a particular subtype [12,14,56]. These findings have prompted interest in using FRα for therapeutic targeting of EOC, particularly for women whose tumors express substantial levels of FRα. Further rationale includes the narrow expression of FRs in non-malignant tissues and the comparatively inefficient uptake of folate substrates compared to facilitative folate transporters, such as the RFC or PCFT [6,9,29]. Different approaches for targeting FRα in EOC patients include monoclonal antibodies, antibody-drug conjugates, folate drug conjugates, and targeted antifolates [110].
Farletuzumab is a humanized IgG1 antibody that targets FRα. Farletuzumab promotes cell death via antibody-dependent cytotoxicity and complement-dependent cytotoxicity, autophagy, and inhibition of Lyn kinase signaling [111,112,113,114]. Farletuzumab was evaluated in a phase I clinical trial with manageable toxicities [115]. A phase II clinical trial with ovarian cancer patients who received farletuzumab, carboplatin, and a taxane, followed by maintenance therapy with farletuzumab, showed a promising overall response rate [116]. A phase III randomized, placebo-controlled trial assessed the efficacy and safety of farletuzumab combined with weekly paclitaxel in platinum-resistant recurrent or refractory EOC (MORAb-003-003; NCT00738699) [117] (Table 1). This study was terminated early after recruiting 417 patients because it did not meet the pre-specified criteria for continuation. A phase III randomized, double-blind, placebo-controlled trial (MORAb-003-004; NCT00849667) compared the efficacy and safety of carboplatin and a taxane with and without weekly farletuzumab in patients with platinum-sensitive ovarian cancer in their first relapse [116,118] (Table 1). Altogether, 1100 women were randomized to three arms (1.25 mg/kg farletuzumab plus combination therapy, 2.5 mg/kg farletuzumab plus combination therapy, or placebo plus combination therapy). No significant differences in PFS among the treatment arms were observed. While the primary PFS endpoint was not achieved, there was a trend toward improved PFS in some patient subsets.
Mirvetuximab soravtansine (IMGN853) is an antibody-drug conjugate to FRα, developed by Immunogen, which was “fast-tracked” toward clinical trials with the objective of determining the clinical efficacy against platinum-resistant ovarian cancers. Mirvetiximab soravtansine consists of a humanized anti-FRα antibody conjugated to a cytotoxic payload of maytansinoid DM4 [119]. An initial study showed that mirvetuximab soravtansine was well tolerated as a single agent, and an expansion phase study in patients with platinum-resistant FRα-overexpressing ovarian cancer afforded promising results (overall response rate of 26%) [120]. In a randomized, open-label, phase III study (FORWARD I), which compared mirvetuximab soravtansine and chemotherapy in 366 patients with platinum-resistant EOC, mirvetuximab soravtansine did not result in a significant improvement in PFS compared with chemotherapy (Table 1). Secondary endpoints consistently favored mirvetuximab soravtansine, particularly in patients with high FRα expression. Mirvetuximab soravtansine showed a more manageable safety profile than chemotherapy [121]. Another Phase III clinical trial (IMGN853-0416; NCT04209855) is recruiting patients and is designed to compare the efficacy and safety of mirvetuximab soravtansine versus investigator’s choice chemotherapy in patients with platinum-resistant high-grade EOC, primary peritoneal cancer, or fallopian tube cancer, whose tumors express a high level of FRα.
Another approach for targeting FRα was pioneered by Endocyte in collaboration with Purdue University. This concept involved a conjugate delivery system, including a folic acid-targeting moiety, which binds to exposed FRs on the surface of FR-expressing tumors, permitting internalization by endocytosis [5]. The targeting entity is conjugated to a cytotoxic “warhead”, separated by a hydrophilic cleavable linker, permitting the release of the cytotoxic moiety. Perhaps the most successful conjugate of this series was EC145 (Vintafolide), which included a folic acid-targeting moiety conjugated to desacetylvinblastine monohydrazide (DAVLB), a derivative of the microtubule inhibitor vinblastine [5]. This agent is specific for FR-expressing tumors such that, following its internalization and DAVLB release from the vesicle, EC145 inhibits dividing cancer cells by disrupting mitotic spindle formation. The preclinical data with EC145 were promising, leading to near cures in a high FRα-expressing KB tumor model in immune-compromised mice with limited toxicity. Based on the preclinical data, EC145 advanced to clinical trials [122]. A phase I clinical trial showed an acceptable toxicity profile [123]. A randomized phase II study in women with platinum-resistant ovarian cancer treated with EC145 and pegylated liposomal doxorubicin versus pegylated liposomal doxorubicin alone showed a benefit in PFS for the combined treatment [124]. However, EC145 did not progress past phase III as there was a lack of improvement in PFS for women with predominantly FRα-positive platinum-resistant EOC (Table 1) [125].
Table 1. Phase III clinical trial involving FRα-targeted therapies.
Table 1. Phase III clinical trial involving FRα-targeted therapies.
Farlatuzumab
TitlePatientsResultsReference
Randomized, double-blind, placebo-controlled, phase III study to assess the efficacy and safety of weekly MORAb-003 in combination with carboplatin and taxane in subjects with platinum-sensitive ovarian cancer in first relapse (MORAb-003-004; NCT00849667)Platinum-sensitive EOC (1100 patients) in first relapseNo significant differences in PFS among the treatment arms were observed. The primary end point of PFS was not met.[116,118]
Phase III randomized clinical trial of weekly paclitaxel with or without farletuzumab (MORAb-003-003; NCT00738699)Platinum-resistant ovarian cancer (417 patients)Study was terminated due to failure to meet pre-specified criteria.[117]
Vintafolide (EC145)
Study for women with platinum resistant ovarian cancer evaluating EC145 in combination with Doxil® (PROCEED) (EC-FV-06; NCT01170650)FRα-positive platinum-resistant ovarian cancer (640 patients)Trial was terminated owing to failure to meet pre-specified PFS criteria.[125]
Mirvetuximab soravtansine
Phase III RCT (FORWARD I) evaluating chemotherapy (paclitaxel, pegylated liposomal doxorubicin, or topotecan) vs. mirvetuximab soravtansine (IMGN853-0403; NCT02631876)FRα-positive platinum-resistant ovarian cancer (366 patients)Mirvetuximab soravtansine did not result in a significant improvement in PFS compared with chemotherapy.[121]
CT900 (BGC945, ONX0801) is a cyclopenta[g]quinazoline-based antifolate that targets TS (Figure 2, Table 2). CT900 is transported solely by FR rather than the RFC, thus protecting bone marrow and intestinal cells [26]. Preclinical in vivo studies using KB tumor xenografts showed that the terminal half-life of CT900 was longer in tumors than in normal tissues, such as liver and kidney [26]. In 2017, CT900 completed a successful phase I clinical trial with high-grade serous ovarian cancers overexpressing FRα [126]. A retrospective data analysis of an expansion cohort from the trial suggested that 50% of patients with FRα-overexpressing high-grade serous ovarian cancer had a partial response [126].
Cancer imaging through the FRα is complementary to FRα-targeted therapies for EOC. Approaches include FRα-targeted contrast-enhanced magnetic resonance imaging (MRI) [127,128] and radiolabeled folate derivatives [124,129,130]. The latter have been tested in clinical trials, where 111In-diethylenetriaminepentaacetic acid-folate was used for whole-body single-photon emission computed tomography in women with endometrial or ovarian cancer [129]. A folate conjugate (etarfolatide) conjugated with 99mTc was investigated as a predictive biomarker in clinical trials [124,130], including a phase II clinical trial with vintafolide in women with recurrent platinum-resistant ovarian cancer [124]. The antibody-based agent 89Zr-DFO-M9346A was used to measure FRα in murine ovarian cancer xenografts treated with mirvetuximab soravtansine [131].
Fluorescent folate probes have been developed for intraoperative visualization and improved resections of FRα-expressing ovarian, breast, and lung cancers [132]. Folate conjugated to fluorescein isothiocyanate (EC17) was tested intraoperatively in women with ovarian cancer [133,134]. In November 2021, the U.S. Food and Drug Administration approved Cytalux (pafolacianine) [135] for intraoperative visualization of FRα-expressing ovarian cancer lesions.

5. Targeting EOC via Targeted Antifolates: C1 Metabolism as a Unique Vulnerability for EOC

Pemetrexed (Alimta) has been used both singly and in combination therapy with recurrent platinum-sensitive and resistant EOC and showed a favorable toxicity profile and response rate comparable to other agents used in first-line combination therapy [136]. Pemetrexed inhibits thymidylate synthase with secondary inhibitions of de novo purine biosynthesis at GARFTase and ATIC [137]. Further, as noted above, the thymidylate synthase inhibitor CT900 (ONX0801) is transported into EOCs by FRα [26] and has shown promising clinical activity toward FRα-overexpressing high-grade serous ovarian cancer [126].
Of particular interest are novel investigational drugs that target the potential metabolic vulnerabilities in EOC, such as de novo purine nucleotide biosynthesis. The expression of GARFTase and ATIC in high-grade serous ovarian cancer from patients (n = 39) is dramatically increased (5.2- and 4.4-, respectively) over the levels in normal ovaries (n = 8) (A. Wallace-Povirk, L.H. Matherly, unpublished), suggesting the importance of this key anabolic pathway to the malignant phenotype. By extension, anti-purine inhibitors are particularly suited for EOC as they are cytotoxic independent of p53 and BRCA mutant status [138,139]. Further, they are tumor-selective, resulting from the loss of purine salvage [140,141], and effect a suppression of mTOR signaling [142,143].
Particularly notable are 6-subsituted pyrrolo[2,3-d]pyrimidine antifolates related to pemetrexed (a 5-substituted pyrrolo[2,3-d]pyrimidine benzoyl antifolate with a 2-carbon bridge) (Figure 2B; Table 2). Whereas pemetrexed is a good substrate for both the RFC and PCFT, it is poorly transported by FRα [19]. However, the related 6-substituted pyrrolo[2,3-d]pyrimidine benzoyl compounds AGF17 and AGF23 (Figure 2B) are transported by both FRα and PCFT with negligible transport by the RFC [144,145]. A 3-carbon bridge length appears to be optimal for PCFT transport [27,144,146,147]. Moreover, PCFT transport is further enhanced by a replacement of the side-chain phenyl ring by a 2′,3′- and 2′,4′-thiophene, as in AGF94 and AGF154, respectively (Figure 2B) [41,147,148]. Side-chain 3′-fluorinated thienoyl analogs of this series (AGF278, AGF283; Figure 2B) were also synthesized and show high levels of FRα- and PCFT-targeted activity [149].
Figure 2. Structures of inhibitors of C1 metabolism. Panel (A): structures are shown for the dual SHMT1 and SHMT2 inhibitors SHIN1 and SHIN2 [150,151], the MTHFD2 inhibitor DS18561882 [152], as well as pemetrexed [137] and CT900 (BGC945, ONX0801) [26] (both principally thymidylate synthase inhibitors). Panel (B): structures are shown for pyrrolo[2,3-d]pyrimidine GARFTase inhibitors, AGF17 [145], AGF23 [145], AGF94 [147], AGF154 [148], AGF278 [149], and AGF283 [149], as well as the pyrrolo[3,2-d]pyrimidine antifolate AGF347 [153], which acts as a multitargeted inhibitor of SHMT2 in the mitochondria and of SHMT1, GARFTase, and ATIC in the cytosol.
Figure 2. Structures of inhibitors of C1 metabolism. Panel (A): structures are shown for the dual SHMT1 and SHMT2 inhibitors SHIN1 and SHIN2 [150,151], the MTHFD2 inhibitor DS18561882 [152], as well as pemetrexed [137] and CT900 (BGC945, ONX0801) [26] (both principally thymidylate synthase inhibitors). Panel (B): structures are shown for pyrrolo[2,3-d]pyrimidine GARFTase inhibitors, AGF17 [145], AGF23 [145], AGF94 [147], AGF154 [148], AGF278 [149], and AGF283 [149], as well as the pyrrolo[3,2-d]pyrimidine antifolate AGF347 [153], which acts as a multitargeted inhibitor of SHMT2 in the mitochondria and of SHMT1, GARFTase, and ATIC in the cytosol.
Cancers 14 00191 g002
AGF94, AGF154, AGF278, and AGF283 are potent (nanomolar) inhibitors of EOC cell lines (IGROV1, SKOV3, A2780) with a wide (~3000-fold) range of FRα levels and relatively constant PCFT, although the uptake by FRα predominated when both systems were present [14,149]. In IGROV1 cells, FRα knockdown (~95%) had a modest impact on in vitro drug efficacy, establishing transport redundancy for pyrrolo[2,3-d]pyrimidine antifolates such as AGF94 by the PCFT and FRα [14,149]. For AGF94, this finding was extended in vivo [14].
As EOCs express FRα over an extraordinarily wide range [14], the notion of targeting the PCFT in EOC is especially appealing. The PCFT is constitutively expressed in EOC [14] and has a modest expression in most normal tissues [8,29,39]. Further, PCFT is active under the acidic pH conditions associated with the tumor microenvironment, although significant PCFT transport is detectable up to pH 7 [154]. Above pH 7, PCFT transport activity is low [30]. Notably, the recent failures of FRα-targeted therapies in clinical trials with ovarian cancer patients (Table 1) reflect, at least in part, the challenge in identifying FRα-positivity, hence the patients most likely to respond to FR-targeted treatments. This suggests opportunities for the PCFT-targeted therapeutics for EOC independent of FRα expression.
Table 2. FR and PCFT-targeted antifolates.
Table 2. FR and PCFT-targeted antifolates.
InhibitorTransporterIntracellular TargetReferences
PemetrexedPCFT, RFCThymidylate synthase, GARFTase, ATIC[137]
AGF17FRα, FRβ, PCFTGARFTase[144,145]
AGF23FRα, FRβ, PCFTGARFTase[145]
AGF94FRα, FRβ, PCFTGARFTase[14,147]
AGF154FRα, FRβ, PCFTGARFTase[148]
AGF278FRα, FRβ, PCFTGARFTase[149]
AGF283FRα, FRβ, PCFTGARFTase[149]
AGF347FRα, PCFT, RFCSHMT1, SHMT2, GARFTase, ATIC[153,155,156]
CT900FRαThymidylate synthase [26]
(±) SHIN1Not determinedSHMT1, SHMT2[150]
(+) SHIN2Not determinedSHMT1, SHMT2[151]
DS18561882Not determinedMTHFD1, MTHFD2[152]
With the recognition of the essential role of mitochondrial C1 metabolism (Figure 1) in malignancies including EOC [77], attention has turned toward identifying the inhibitors of this pathway [21]. SHMT2 and MTHFD2 are highly expressed in EOC [77,157], suggesting the therapeutic potential for targeting this pathway in this disease. While MTHFD2 inhibitors have been reported [152,158,159,160], to date the most promising compound is DS18561882 (Figure 2A; Table 2). DS18451882 is a potent (nanomolar) inhibitor of MTHFD2 with a 90-fold selectivity for MTHFD2 over MTHFD1 [152]. Further, DS18561882 showed in vivo efficacy against MDA-MB 231 triple-negative breast cancer xenografts in immune-compromised mice with minimal toxicity [152]. The roles of FRα, RFC, and PCFT in cellular uptake and anti-tumor efficacy of DS18561882 are unclear.
Rabinowitz and his colleagues reported pyrazolopyran inhibitors of SHMT1 and −2 (SHIN1, SHIN2) (Figure 2A; Table 2) that inhibited the proliferation of human tumor cell lines at sub-micromolar concentrations [150,151]. Notably, SHIN2 circumvented the apparent pharmacologic shortcomings of SHIN1 and showed in vivo efficacy toward a PDX T-cell leukemia mouse model [151]. Of particular interest is the 5-substituted pyrrolo[3,2-d]pyrimidine antifolate AGF347 [153,155] (Figure 2B; Table 2). AGF347 is a broad-spectrum anti-tumor agent with demonstrated in vitro activity toward colon cancer, lung adenocarcinoma, and pancreatic adenocarcinoma cell lines [153,155], as well as impressive in vivo efficacy toward MIA PaCa-2 pancreatic cancer xenografts [153]. AGF347 is transported by both FRα and PCFT, as well as RFC, and is a potent inhibitor of SHMT2 in the mitochondria and of SHMT1, GARFTase, and ATIC in the cytosol [153,155,156]. AGF347 inhibited the in vitro proliferation of EOC cell lines including cisplatin-resistant SKOV3 EOC, TOV112D, and A2780 EOC cells; the in vivo anti-tumor efficacy was demonstrated with SKOV3 EOC xenografts in SCID mice [156]. For both SHIN1/2 and AGF347, inhibition of both SHMT1 and SHMT2 was essential to their anti-tumor effects, as this prevents metabolic compensation for the loss of SHMT2 activity by reversal of the SHMT1 reaction and the synthesis of glycine and 5,10-methylene THF [150,153,161].

6. Targeting the Tumor Microenvironment in Epithelial Ovarian Cancer

6.1. The Role of the Dynamic TME in EOC Progression

Ovarian cancer involves a complex intraperitoneal milieu, which includes not only cells within the bulk tumor, but also fibroblasts, endothelial cells, and assorted immune cells [83,162]. Tumor and non-tumor cells secrete a host of bioactive constituents, including growth factors, hormones, phospholipids, and cytokines, that contribute to the TME [83,162]. Thus, the TME in ovarian cancer involves a dynamic interplay between the malignant ascites and the surrounding tissues via receptor-mediated (e.g., autocrine and paracrine) or contact-dependent signaling and epigenetic regulation that enables tumor progression and immune evasion [83].
Fibroblasts synthesize matrix-metalloproteases and other proteins comprising the extracellular matrix, including collagens, fibronectin, and laminin [163]. Cancer-associated fibroblasts (CAFs) promote tumor cell proliferation, invasion, and migration and also promote immune suppression and angiogenesis [164,165,166,167]. Endothelial cells line blood vessels and are associated with angiogenesis in response to angiogenesis activators (e.g., VEGF and PDGF) and inhibitors (e.g., angiopoetin) [168,169]. Immune cells include macrophages, dendritic cells, myeloid-derived suppressor cells (MDSCs), and lymphocytes that impact both tumor progression and suppression, as well as participate in tumorigenesis, metastasis, and angiogenesis [170,171].
The following section describes strategies for leveraging advances in the biology of FRs and C1 metabolism that target the TME in ovarian cancer with a particular focus on T lymphocytes and TAMs.

6.2. Targeting the TME in EOC

The ascites contains a host of infiltrating immune cells, including T lymphocytes and TAMs, although TAMs are considered the principal immune cellular component that results in an immunosuppressive environment in EOC [172,173]. Further, TAMs contribute to metastasis and angiogenesis by releasing pro-angiogenic factors, such as vascular endothelial growth factor and matrix metalloproteinase [162,174].
Initial optimism for the use of immune checkpoint inhibitors with EOC was tempered by disappointing results with these agents in clinical trials as single agents or combined with standard chemotherapy [175]. Nonetheless, the lymphoid compartment remains an intriguing target for novel TME therapies in EOC [83]. Patients with increased levels of CD3+ tumor-infiltrating lymphocytes (TILs) showed delayed recurrence of disease; when greater TIL infiltration was accompanied by increased levels of interferon γ, the EOC patients experienced longer overall survivals [83].
The high levels of FRα in ovarian cancer suggested opportunities for a targeted immunotherapy of EOC employing CAR T. Engineered CAR T cells containing an FRα-specific epitope coupled to the T cell receptor chain CD3ζ induced regression in preclinical models [176]. Based on these results, a phase I clinical trial (NCT03585764) is recruiting patients with ovarian or fallopian tube cancers or primary peritoneal cancers. A recent study described a novel tandem CAR encoding an anti-FRα scFv, an anti-mesothelin scFv, and two peptide sequences of IL-12 to improve the efficacy, infiltration, persistence, and proliferation of CAR T cells in ovarian cancer [177].
There has been increasing interest in targeting TAMs as a cancer therapy, as inhibiting TAMs could, in principle, suppress tumor progression [162]. A screen to discover differentially expressed genes in the presence of macrophages skewed toward an M1-like or M2-like state, revealed that FOLR2, which encodes FRβ was differentially expressed in M2-like human macrophages. FRβ was expressed on IL-10-producing M2-like macrophages (CD163+, CD68+, CD14+ IL-10), corresponding to the anti-inflammatory/pro-tumor TAM subtype [16].
As FRβ expression on TAMs was induced by malignant ascites and conditioned media from fibroblasts, it was suggested that the presence of FRβ on the surface of macrophages could offer an opportunity for depleting TAMs with cytotoxic folate-conjugates or antifolates as a component of therapy [16]. In both lung adenocarcinoma and squamous cell carcinoma, FRβ showed an increased expression that was predictive of a worse outcome and prognosis [18]. A BIM (BCL-2-interacting mediator of cell death) plasmid encapsulated in a folate “lipoplex” was used to target the TME of lung cancer. This therapeutic was selectively delivered to FRβ-expressing TAMs, resulting in their substantial depletion in the TME without significant off-target toxicities. Importantly, there were significantly fewer tumor nodules in the lungs of the treated mice compared to the control mice [18]. In an experimental C6 rat glioma model, targeting FRβ-expressing TAMs with an anti-mouse FRβ monoclonal antibody conjugated to Pseudomonas exotoxin A significantly depleted TAMs and reduced tumor growth [17]. A folate-conjugated TLR7 agonist against FRβ-expressing macrophages not only showed in vivo activity in a range of tumor models, including metastatic lesions, but also reversed the expression of a high M2-like to M1-like macrophage ratio and increased the infiltration of cytotoxic CD8 T cells [15].
Our recent studies in a syngeneic mouse model of high-grade serous ovarian cancer (BR-Luc [178,179]) treated with the FR-targeted pyrrolo[2,3-d]pyrimidine inhibitor AGF94 (Figure 2B, Table 2) [147] demonstrated anti-tumor efficacy, accompanied by a direct impact on the TME, including significantly decreased FRβ-expressing TAMs and no evidence of CD3+ T cell depletion or impact on the relative proportions of CD4+ and CD8+ T cells after drug treatment (A. Wallace-Povirk, L. Rubinsak, A. Malysa, S.H. Dzinic, M. Ravindra, C. O’Connor, Z. Hou, S. Kim, J. Back, L. Polin, et al., manuscript submitted). As AGF94 is transported by both FRs and the PCFT [147], this could represent a new approach for therapy of high-grade serous ovarian cancer through its ability to directly target the tumor via uptake by FRα and/or the PCFT and its effects on the TME, particularly FRβ-expressing TAMs.
AGF94 is an inhibitor of de novo purine biosynthesis at GARFTase [147]; this raises the intriguing question of whether C1 metabolism could represent in a unique vulnerability for TAMs, in addition to the primary tumor. Serine metabolism was associated with an inflammatory response in macrophages via its impact on glutathione synthesis [180], and depletion of S-adenosyl methionine was reported to inhibit inflammatory macrophages [181], suggesting an important role for C1 metabolism in TAM biology. Future studies will be necessary to further explore the potential of targeting C1 metabolism in inflammatory macrophages with a new generation of C1 inhibitors as a novel therapeutic strategy in EOC.

7. Conclusions

The review summarizes the advances in the biology and treatment of EOC with FRα- and PCFT-targeted therapeutics. Although studies involving FRα-targeted therapies with antibodies (such as farletuzumab) [111,112,113,114], antibody-drug conjugates (such as mirvetuximab soravtansine) [119], or drug conjugates (such as EC145) [5] combined with high frequency FRα expression have established FRα as a viable and tumor-specific approach for treating EOC, clinical results have been mixed [110,115,116,118,120,124,125]. This, in part, reflects the wide range of FRα expression and the challenges in predicting which patients are most likely to respond to FRα-targeted therapies.
Of particular interest are therapies targeting C1 metabolism in relation to EOC, such as CT900, which is transported by FRα and inhibits thymidylate synthase [26], as CT900 has shown promising single-agent activity in early phase clinical trials [126]. Of further interest is the potential of treating EOC with FRα and the PCFT dual-targeted antifolates, such as AGF94, that derive their tumor selectivity from their FRα and PCFT specificities over the ubiquitously expressed RFC [14,41,147,148,149]. The near constitutive expression of the PCFT in EOC [14], combined with its tumor specificity and the limited RFC transport, would result in fewer toxicities than chemotherapy, making this approach highly appealing. As an anti-purine biosynthesis inhibitor for EOC, AGF94 would inhibit cells independent of wild-type/mutant p53 and BRCA status [138,139], show selectivity based on impaired purine salvage [140,141] and suppress mTOR signaling [142,143]. With the recognition of the importance of mitochondrial C1 catabolism from serine to the malignant phenotype, including ovarian cancer, an exciting new generation of investigational agents which target this pathway (e.g., SHIN2 and AGF347) is being developed [150,151,153,155], some of which are transported by FRα and the PCFT and potently inhibit EOC cells [156].
While EOC is considered an immunologically “cold” tumor, there is substantial interest in targeting the TME given its role in disease progression and therapy resistance in high-grade serous EOC [175]. Particularly exciting is the use of engineered CAR T cells containing a FRα-specific epitope linked to T cell receptor [176], and the development of novel FR-targeted cytotoxic conjugates or pyrrolopyrimidine antifolates such as AGF94 [15,16,17,18,147]. The latter could offer an entirely new approach for EOC therapy through its multi-targeting of the tumor via FRα and the PCFT and TAMs via FRβ. Of particular interest will be better understanding the metabolic basis for the relationship of C1 metabolism to inflammatory responses mediated by TAMs.
FR- and C1-targeted therapeutics, as described in this review, have extraordinary potential to improve the treatment of patients with EOC. The excitement is palpable as work progresses to demonstrate whether this potential for combatting this devastating disease can be realized.

Author Contributions

Conceptualization, A.W.-P. and L.H.M.; writing—original draft preparation, A.W.-P. and L.H.M.; writing, review and editing, A.W.-P., Z.H., M.J.N., A.G. and L.H.M.; project administration, L.H.M., Z.H. and A.G.; funding acquisition, L.H.M., A.G., Z.H. and A.W.-P. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by grants R01 CA53535 (L.H. Matherly, Z. Hou) and R01 CA250469 (L.H. Matherly, A. Gangjee) from the National Cancer Institute, the Eunice and Milton Ring Endowed Chair for Cancer Research (L.H. Matherly), the Duquesne University Adrian Van Kaam Chair in Scholarly Excellence (Aleem Gangjee), and the Michigan Ovarian Cancer Alliance foundation (Z. Hou, L.H. Matherly). A. Wallace-Povirk was supported by F31CA243215 from the National Cancer Institute.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Torre, L.A.; Trabert, B.; DeSantis, C.E.; Miller, K.D.; Samimi, G.; Runowicz, C.D.; Gaudet, M.M.; Jemal, A.; Siegel, R.L. Ovarian cancer statistics, 2018. CA Cancer J. Clin. 2018, 68, 284–296. [Google Scholar] [CrossRef] [PubMed]
  2. Lheureux, S.; Gourley, C.; Vergote, I.; Oza, A.M. Epithelial ovarian cancer. Lancet 2019, 393, 1240–1253. [Google Scholar] [CrossRef] [Green Version]
  3. Nero, C.; Ciccarone, F.; Pietragalla, A.; Duranti, S.; Daniele, G.; Salutari, V.; Carbone, M.V.; Scambia, G.; Lorusso, D. Ovarian Cancer Treatments Strategy: Focus on PARP Inhibitors and Immune Check Point Inhibitors. Cancers 2021, 13, 1298. [Google Scholar] [CrossRef] [PubMed]
  4. Kurnit, K.C.; Fleming, G.F.; Lengyel, E. Updates and New Options in Advanced Epithelial Ovarian Cancer Treatment. Obstet. Gynecol. 2021, 137, 108–121. [Google Scholar] [CrossRef] [PubMed]
  5. Assaraf, Y.G.; Leamon, C.P.; Reddy, J.A. The folate receptor as a rational therapeutic target for personalized cancer treatment. Drug Resist. Updates 2014, 17, 89–95. [Google Scholar] [CrossRef] [PubMed]
  6. Elnakat, H.; Ratnam, M. Distribution, functionality and gene regulation of folate receptor isoforms: Implications in targeted therapy. Adv. Drug Deliv. Rev. 2004, 56, 1067–1084. [Google Scholar] [CrossRef]
  7. Vergote, I.B.; Marth, C.; Coleman, R.L. Role of the folate receptor in ovarian cancer treatment: Evidence, mechanism, and clinical implications. Cancer Metastasis Rev. 2015, 34, 41–52. [Google Scholar] [CrossRef]
  8. Desmoulin, S.K.; Hou, Z.; Gangjee, A.; Matherly, L.H. The human proton-coupled folate transporter: Biology and therapeutic applications to cancer. Cancer Biol. Ther. 2012, 13, 1355–1373. [Google Scholar] [CrossRef] [Green Version]
  9. Matherly, L.H.; Hou, Z.; Deng, Y. Human reduced folate carrier: Translation of basic biology to cancer etiology and therapy. Cancer Metastasis Rev. 2007, 26, 111–128. [Google Scholar] [CrossRef] [PubMed]
  10. Matherly, L.H.; Hou, Z.; Gangjee, A. The promise and challenges of exploiting the proton-coupled folate transporter for selective therapeutic targeting of cancer. Cancer Chemother. Pharmacol. 2018, 81, 1–15. [Google Scholar] [CrossRef]
  11. Parker, N.; Turk, M.J.; Westrick, E.; Lewis, J.D.; Low, P.S.; Leamon, C.P. Folate receptor expression in carcinomas and normal tissues determined by a quantitative radioligand binding assay. Anal. Biochem. 2005, 338, 284–293. [Google Scholar] [CrossRef]
  12. Toffoli, G.; Cernigoi, C.; Russo, A.; Gallo, A.; Bagnoli, M.; Boiocchi, M. Overexpression of folate binding protein in ovarian cancers. Int. J. Cancer 1997, 74, 193–198. [Google Scholar] [CrossRef]
  13. Weitman, S.D.; Lark, R.H.; Coney, L.R.; Fort, D.W.; Frasca, V.; Zurawski, V.R., Jr.; Kamen, B.A. Distribution of the folate receptor GP38 in normal and malignant cell lines and tissues. Cancer Res. 1992, 52, 3396–3401. [Google Scholar] [PubMed]
  14. Hou, Z.; Gattoc, L.; O’Connor, C.; Yang, S.; Wallace-Povirk, A.; George, C.; Orr, S.; Polin, L.; White, K.; Kushner, J.; et al. Dual Targeting of Epithelial Ovarian Cancer Via Folate Receptor alpha and the Proton-Coupled Folate Transporter with 6-Substituted Pyrrolo[2,3-d]pyrimidine Antifolates. Mol. Cancer Ther. 2017, 16, 819–830. [Google Scholar] [CrossRef] [Green Version]
  15. Cresswell, G.M.; Wang, B.; Kischuk, E.M.; Broman, M.M.; Alfar, R.A.; Vickman, R.E.; Dimitrov, D.S.; Kularatne, S.A.; Sundaram, C.P.; Singhal, S.; et al. Folate Receptor Beta Designates Immunosuppressive Tumor-Associated Myeloid Cells That Can Be Reprogrammed with Folate-Targeted Drugs. Cancer Res. 2021, 81, 671–684. [Google Scholar] [CrossRef] [PubMed]
  16. Puig-Kroger, A.; Sierra-Filardi, E.; Dominguez-Soto, A.; Samaniego, R.; Corcuera, M.T.; Gomez-Aguado, F.; Ratnam, M.; Sanchez-Mateos, P.; Corbi, A.L. Folate receptor beta is expressed by tumor-associated macrophages and constitutes a marker for M2 anti-inflammatory/regulatory macrophages. Cancer Res. 2009, 69, 9395–9403. [Google Scholar] [CrossRef] [Green Version]
  17. Nagai, T.; Tanaka, M.; Tsuneyoshi, Y.; Xu, B.; Michie, S.A.; Hasui, K.; Hirano, H.; Arita, K.; Matsuyama, T. Targeting tumor-associated macrophages in an experimental glioma model with a recombinant immunotoxin to folate receptor beta. Cancer Immunol. Immunother. 2009, 58, 1577–1586. [Google Scholar] [CrossRef]
  18. Tie, Y.; Zheng, H.; He, Z.; Yang, J.; Shao, B.; Liu, L.; Luo, M.; Yuan, X.; Liu, Y.; Zhang, X.; et al. Targeting folate receptor beta positive tumor-associated macrophages in lung cancer with a folate-modified liposomal complex. Signal Transduct. Target. Ther. 2020, 5, 6. [Google Scholar] [CrossRef]
  19. Matherly, L.H.; Wilson, M.R.; Hou, Z. The major facilitative folate transporters solute carrier 19A1 and solute carrier 46A1: Biology and role in antifolate chemotherapy of cancer. Drug Metab. Dispos. 2014, 42, 632–649. [Google Scholar] [CrossRef] [Green Version]
  20. Visentin, M.; Zhao, R.; Goldman, I.D. The antifolates. Hematol. Oncol. Clin. N. Am. 2012, 26, 629–648. [Google Scholar] [CrossRef] [Green Version]
  21. Dekhne, A.S.; Hou, Z.; Gangjee, A.; Matherly, L.H. Therapeutic Targeting of Mitochondrial One-Carbon Metabolism in Cancer. Mol. Cancer Ther. 2020, 19, 2245–2255. [Google Scholar] [CrossRef]
  22. Ducker, G.S.; Rabinowitz, J.D. One-Carbon Metabolism in Health and Disease. Cell Metab. 2017, 25, 27–42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Lucock, M. Folic acid: Nutritional biochemistry, molecular biology, and role in disease processes. Mol. Genet. Metab. 2000, 71, 121–138. [Google Scholar] [CrossRef] [PubMed]
  24. Stover, P.J. Physiology of folate and vitamin B12 in health and disease. Nutr. Rev. 2004, 62, S3–S12. [Google Scholar] [CrossRef] [PubMed]
  25. Tibbetts, A.S.; Appling, D.R. Compartmentalization of Mammalian folate-mediated one-carbon metabolism. Ann. Rev. Nutr. 2010, 30, 57–81. [Google Scholar] [CrossRef] [PubMed]
  26. Gibbs, D.D.; Theti, D.S.; Wood, N.; Green, M.; Raynaud, F.; Valenti, M.; Forster, M.D.; Mitchell, F.; Bavetsias, V.; Henderson, E.; et al. BGC 945, a novel tumor-selective thymidylate synthase inhibitor targeted to alpha-folate receptor-overexpressing tumors. Cancer Res. 2005, 65, 11721–11728. [Google Scholar] [CrossRef] [Green Version]
  27. Parker, J.L.; Deme, J.C.; Kuteyi, G.; Wu, Z.; Huo, J.; Goldman, I.D.; Owens, R.J.; Biggin, P.C.; Lea, S.M.; Newstead, S. Structural basis of antifolate recognition and transport by PCFT. Nature 2021, 595, 130–134. [Google Scholar] [CrossRef]
  28. Matherly, L.H.; Goldman, D.I. Membrane transport of folates. Vitam. Horm. 2003, 66, 403–456. [Google Scholar] [PubMed]
  29. Zhao, R.; Diop-Bove, N.; Visentin, M.; Goldman, I.D. Mechanisms of membrane transport of folates into cells and across epithelia. Annu. Rev. Nutr. 2011, 31, 177–201. [Google Scholar] [CrossRef] [Green Version]
  30. Zhao, R.; Goldman, I.D. The molecular identity and characterization of a Proton-coupled Folate Transporter--PCFT; biological ramifications and impact on the activity of pemetrexed. Cancer Metastasis Rev. 2007, 26, 129–139. [Google Scholar] [CrossRef]
  31. Saier, M.H., Jr. Transporter Classification Database. 2012. Available online: http://www.tcdb.org/ (accessed on 1 December 2021).
  32. Whetstine, J.R.; Flatley, R.M.; Matherly, L.H. The human reduced folate carrier gene is ubiquitously and differentially expressed in normal human tissues: Identification of seven non-coding exons and characterization of a novel promoter. Biochem. J. 2002, 367, 629–640. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, Y.; Zhao, R.; Russell, R.G.; Goldman, I.D. Localization of the murine reduced folate carrier as assessed by immunohistochemical analysis. Biochim. Biophys. Acta Biomembr. 2001, 1513, 49–54. [Google Scholar] [CrossRef] [Green Version]
  34. Waes, J.G.-V.; Heller, S.; Bauer, L.K.; Wilberding, J.; Maddox, J.R.; Aleman, F.; Rosenquist, T.H.; Finnell, R.H. Embryonic development in the reduced folate carrier knockout mouse is modulated by maternal folate supplementation. Birth Defects Res. Part A Clin. Mol. Teratol. 2008, 82, 494–507. [Google Scholar] [CrossRef]
  35. Zhao, R.; Russell, R.G.; Wang, Y.; Liu, L.; Gao, F.; Kneitz, B.; Edelmann, W.; Goldman, I.D. Rescue of embryonic lethality in reduced folate carrier-deficient mice by maternal folic acid supplementation reveals early neonatal failure of hematopoietic organs. J. Biol. Chem. 2001, 276, 10224–10228. [Google Scholar] [CrossRef] [Green Version]
  36. Siu, M.K.Y.; Kong, D.S.H.; Chan, H.Y.; Wong, E.S.Y.; Philip, P.C.I.; Jiang, L.; Hextan, Y.S.N.; Le, X.F.; Cheung, A.N.Y. Paradoxical Impact of Two Folate Receptors, FRa and RFC, in Ovarian Cancer: Effect on Cell Proliferation, Invasion and Clinical Outcome. PLoS ONE 2012, 7, e47201. [Google Scholar] [CrossRef] [Green Version]
  37. Qiu, A.; Jansen, M.; Sakaris, A.; Min, S.H.; Chattopadhyay, S.; Tsai, E.; Sandoval, C.; Zhao, R.; Akabas, M.H.; Goldman, I.D. Identification of an intestinal folate transporter and the molecular basis for hereditary folate malabsorption. Cell 2006, 127, 917–928. [Google Scholar] [CrossRef] [Green Version]
  38. Zhao, R.; Aluri, S.; Goldman, I.D. The proton-coupled folate transporter (PCFT-SLC46A1) and the syndrome of systemic and cerebral folate deficiency of infancy: Hereditary folate malabsorption. Mol. Asp. Med. 2017, 53, 57–72. [Google Scholar] [CrossRef] [Green Version]
  39. Qiu, A.; Min, S.H.; Jansen, M.; Malhotra, U.; Tsai, E.; Cabelof, D.C.; Matherly, L.H.; Zhao, R.; Akabas, M.H.; Goldman, I.D. Rodent intestinal folate transporters (SLC46A1): Secondary structure, functional properties, and response to dietary folate restriction. Am. J. Physiol. Cell Physiol. 2007, 293, C1669–C1678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Zhao, R.; Visentin, M.; Suadicani, S.O.; Goldman, I.D. Inhibition of the proton-coupled folate transporter (PCFT-SLC46A1) by bicarbonate and other anions. Mol. Pharmacol. 2013, 84, 95–103. [Google Scholar] [CrossRef] [Green Version]
  41. Desmoulin, S.K.; Wang, L.; Hales, E.; Polin, L.; White, K.; Kushner, J.; Stout, M.; Hou, Z.; Cherian, C.; Gangjee, A.; et al. Therapeutic targeting of a novel 6-substituted pyrrolo [2,3-d]pyrimidine thienoyl antifolate to human solid tumors based on selective uptake by the proton-coupled folate transporter. Mol. Pharmacol. 2011, 80, 1096–1107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Giovannetti, E.; Zucali, P.A.; Assaraf, Y.G.; Funel, N.; Gemelli, M.; Stark, M.; Thunnissen, E.; Hou, Z.; Muller, I.B.; Struys, E.A.; et al. Role of proton-coupled folate transporter in pemetrexed resistance of mesothelioma: Clinical evidence and new pharmacological tools. Ann. Oncol 2017, 28, 2725–2732. [Google Scholar] [CrossRef]
  43. Wilson, M.R.; Hou, Z.; Yang, S.; Polin, L.; Kushner, J.; White, K.; Huang, J.; Ratnam, M.; Gangjee, A.; Matherly, L.H. Targeting Nonsquamous Nonsmall Cell Lung Cancer via the Proton-Coupled Folate Transporter with 6-Substituted Pyrrolo[2,3-d]Pyrimidine Thienoyl Antifolates. Mol. Pharmacol. 2016, 89, 425–434. [Google Scholar] [CrossRef] [Green Version]
  44. Diop-Bove, N.K.; Wu, J.; Zhao, R.; Locker, J.; Goldman, I.D. Hypermethylation of the human proton-coupled folate transporter (SLC46A1) minimal transcriptional regulatory region in an antifolate-resistant HeLa cell line. Mol. Cancer Ther. 2009, 8, 2424–2431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Gonen, N.; Bram, E.E.; Assaraf, Y.G. PCFT/SLC46A1 promoter methylation and restoration of gene expression in human leukemia cells. Biochem. Biophys. Res. Commun. 2008, 376, 787–792. [Google Scholar] [CrossRef]
  46. Webb, B.A.; Chimenti, M.; Jacobson, M.P.; Barber, D.L. Dysregulated pH: A perfect storm for cancer progression. Nat. Rev. Cancer 2011, 11, 671–677. [Google Scholar] [CrossRef] [PubMed]
  47. Han, L.; Nishimura, K.; Al Hosseini, H.S.; Bianchi, E.; Wright, G.J.; Jovine, L. Divergent evolution of vitamin B9 binding underlies Juno-mediated adhesion of mammalian gametes. Curr. Biol. 2016, 26, R100–R101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Christoph, D.C.; Asuncion, B.R.; Hassan, B.; Tran, C.; Maltzman, J.D.; O’Shannessy, D.J.; Wynes, M.W.; Gauler, T.C.; Wohlschlaeger, J.; Hoiczyk, M.; et al. Significance of folate receptor alpha and thymidylate synthase protein expression in patients with non-small-cell lung cancer treated with pemetrexed. J. Thorac. Oncol. 2013, 8, 19–30. [Google Scholar] [CrossRef] [Green Version]
  49. Nunez, M.I.; Behrens, C.; Woods, D.M.; Lin, H.; Suraokar, M.; Kadara, H.; Hofstetter, W.; Kalhor, N.; Lee, J.J.; Franklin, W.; et al. High expression of folate receptor alpha in lung cancer correlates with adenocarcinoma histology and EGFR [corrected] mutation. J. Thorac. Oncol. 2012, 7, 833–840. [Google Scholar] [CrossRef] [Green Version]
  50. Czeizel, A.E.; Dudas, I. Prevention of the first occurrence of neural-tube defects by periconceptional vitamin supplementation. N. Engl. J. Med. 1992, 327, 1832–1835. [Google Scholar] [CrossRef] [PubMed]
  51. Nakashima-Matsushita, N.; Homma, T.; Yu, S.; Matsuda, T.; Sunahara, N.; Nakamura, T.; Tsukano, M.; Ratnam, M.; Matsuyama, T. Selective expression of folate receptor beta and its possible role in methotrexate transport in synovial macrophages from patients with rheumatoid arthritis. Arthritis Rheum. 1999, 42, 1609–1616. [Google Scholar] [CrossRef]
  52. Ratnam, M.; Marquardt, H.; Duhring, J.L.; Freisheim, J.H. Homologous membrane folate binding proteins in human placenta: Cloning and sequence of a cDNA. Biochemistry 1989, 28, 8249–8254. [Google Scholar] [CrossRef] [PubMed]
  53. Ross, J.F.; Wang, H.; Behm, F.G.; Mathew, P.; Wu, M.; Booth, R.; Ratnam, M. Folate receptor type beta is a neutrophilic lineage marker and is differentially expressed in myeloid leukemia. Cancer 1999, 85, 348–357. [Google Scholar] [CrossRef]
  54. Shen, F.; Wu, M.; Ross, J.F.; Miller, D.; Ratnam, M. Folate receptor type gamma is primarily a secretory protein due to lack of an efficient signal for glycosylphosphatidylinositol modification: Protein characterization and cell type specificity. Biochemistry 1995, 34, 5660–5665. [Google Scholar] [CrossRef] [PubMed]
  55. Kamen, B.A.; Wang, M.T.; Streckfuss, A.J.; Peryea, X.; Anderson, R.G. Delivery of folates to the cytoplasm of MA104 cells is mediated by a surface membrane receptor that recycles. J. Biol. Chem. 1988, 263, 13602–13609. [Google Scholar] [CrossRef]
  56. Chen, Y.L.; Chang, M.C.; Huang, C.Y.; Chiang, Y.C.; Lin, H.W.; Chen, C.A.; Hsieh, C.Y.; Cheng, W.F. Serous ovarian carcinoma patients with high alpha-folate receptor had reducing survival and cytotoxic chemo-response. Mol. Oncol. 2012, 6, 360–369. [Google Scholar] [CrossRef] [Green Version]
  57. Elwood, P.C.; Nachmanoff, K.; Saikawa, Y.; Page, S.T.; Pacheco, P.; Roberts, S.; Chung, K.N. The divergent 5’ termini of the alpha human folate receptor (hFR) mRNAs originate from two tissue-specific promoters and alternative splicing: Characterization of the alpha hFR gene structure. Biochemistry 1997, 36, 1467–1478. [Google Scholar] [CrossRef] [PubMed]
  58. Roberts, S.J.; Chung, K.N.; Nachmanoff, K.; Elwood, P.C. Tissue-specific promoters of the alpha human folate receptor gene yield transcripts with divergent 5’ leader sequences and different translational efficiencies. Biochem. J. 1997, 326, 439–447. [Google Scholar] [CrossRef] [PubMed]
  59. Page, S.T.; Owen, W.C.; Price, K.; Elwood, P.C. Expression of the human placental folate receptor transcript is regulated in human tissues. Organization and full nucleotide sequence of the gene. J. Mol. Biol. 1993, 229, 1175–1183. [Google Scholar] [CrossRef] [Green Version]
  60. Kelley, K.M.; Rowan, B.G.; Ratnam, M. Modulation of the folate receptor alpha gene by the estrogen receptor: Mechanism and implications in tumor targeting. Cancer Res. 2003, 63, 2820–2828. [Google Scholar]
  61. Tran, T.; Shatnawi, A.; Zheng, X.; Kelley, K.M.; Ratnam, M. Enhancement of folate receptor alpha expression in tumor cells through the glucocorticoid receptor: A promising means to improved tumor detection and targeting. Cancer Res. 2005, 65, 4431–4441. [Google Scholar] [CrossRef] [Green Version]
  62. Campbell, I.G.; Jones, T.A.; Foulkes, W.D.; Trowsdale, J. Folate-binding protein is a marker for ovarian cancer. Cancer Res. 1991, 51, 5329–5338. [Google Scholar]
  63. Lawrence, S.A.; Hackett, J.C.; Moran, R.G. Tetrahydrofolate Recognition by the Mitochondrial Folate Transporter. J. Biol. Chem. 2011, 286, 31480–31489. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. McCarthy, E.A.; Titus, S.A.; Taylor, S.M.; Jackson-Cook, C.; Moran, R.G. A mutation inactivating the mitochondrial inner membrane folate transporter creates a glycine requirement for survival of chinese hamster cells. J. Biol. Chem. 2004, 279, 33829–33836. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Lawrence, S.A.; Titus, S.A.; Ferguson, J.; Heineman, A.L.; Taylor, S.M.; Moran, R.G. Mammalian mitochondrial and cytosolic folylpolyglutamate synthetase maintain the subcellular compartmentalization of folates. J. Biol. Chem. 2014, 289, 29386–29396. [Google Scholar] [CrossRef] [Green Version]
  66. Shane, B. Folylpolyglutamate synthesis and role in the regulation of one-carbon metabolism. Vitam. Horm. 1989, 45, 263–335. [Google Scholar] [CrossRef] [PubMed]
  67. Kory, N.; Wyant, G.A.; Prakash, G.; Uit de Bos, J.; Bottanelli, F.; Pacold, M.E.; Chan, S.H.; Lewis, C.A.; Wang, T.; Keys, H.R.; et al. SFXN1 is a mitochondrial serine transporter required for one-carbon metabolism. Science 2018, 362, eaat9528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Locasale, J.W. Serine, glycine and the one-carbon cycle: Cancer metabolism in full circle. Nat. Rev. Cancer 2013, 13, 572–583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Yang, M.; Vousden, K.H. Serine and one-carbon metabolism in cancer. Nat. Rev. Cancer 2016, 16, 650–662. [Google Scholar] [CrossRef]
  70. Ye, J.; Fan, J.; Venneti, S.; Wan, Y.W.; Pawel, B.R.; Zhang, J.; Finley, L.W.; Lu, C.; Lindsten, T.; Cross, J.R.; et al. Serine catabolism regulates mitochondrial redox control during hypoxia. Cancer Discov. 2014, 4, 1406–1417. [Google Scholar] [CrossRef] [Green Version]
  71. Krupenko, S.A.; Krupenko, N.I. ALDH1L1 and ALDH1L2 Folate Regulatory Enzymes in Cancer. Adv. Exp. Med. Biol. 2018, 1032, 127–143. [Google Scholar] [CrossRef]
  72. Pedley, A.M.; Benkovic, S.J. A New View into the Regulation of Purine Metabolism: The Purinosome. Trends Biochem. Sci. 2017, 42, 141–154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Pareek, V.; Pedley, A.M.; Benkovic, S.J. Human de novo purine biosynthesis. Crit. Rev. Biochem. Mol. Biol. 2021, 56, 1–16. [Google Scholar] [CrossRef]
  74. Lu, S.C. S-Adenosylmethionine. Int. J. Biochem. Cell Biol. 2000, 32, 391–395. [Google Scholar] [CrossRef]
  75. Matthews, B.G.; Bowden, N.A.; Wong-Brown, M.W. Epigenetic Mechanisms and Therapeutic Targets in Chemoresistant High-Grade Serous Ovarian Cancer. Cancers 2021, 13, 5993. [Google Scholar] [CrossRef] [PubMed]
  76. Zheng, Y.; Lin, T.Y.; Lee, G.; Paddock, M.N.; Momb, J.; Cheng, Z.; Li, Q.; Fei, D.L.; Stein, B.D.; Ramsamooj, S.; et al. Mitochondrial One-Carbon Pathway Supports Cytosolic Folate Integrity in Cancer Cells. Cell 2018, 175, 1546–1560. [Google Scholar] [CrossRef] [Green Version]
  77. Nilsson, R.; Jain, M.; Madhusudhan, N.; Sheppard, N.G.; Strittmatter, L.; Kampf, C.; Huang, J.; Asplund, A.; Mootha, V.K. Metabolic enzyme expression highlights a key role for MTHFD2 and the mitochondrial folate pathway in cancer. Nat. Commun. 2014, 5, 4128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Badar, T.; Patel, K.P.; Thompson, P.A.; DiNardo, C.; Takahashi, K.; Cabrero, M.; Borthakur, G.; Cortes, J.; Konopleva, M.; Kadia, T.; et al. Detectable FLT3-ITD or RAS mutation at the time of transformation from MDS to AML predicts for very poor outcomes. Leuk. Res. 2015, 39, 1367–1374. [Google Scholar] [CrossRef] [Green Version]
  79. Nikiforov, M.A.; Chandriani, S.; O’Connell, B.; Petrenko, O.; Kotenko, I.; Beavis, A.; Sedivy, J.M.; Cole, M.D. A functional screen for Myc-responsive genes reveals serine hydroxymethyltransferase, a major source of the one-carbon unit for cell metabolism. Mol. Cell. Biol. 2002, 22, 5793–5800. [Google Scholar] [CrossRef] [Green Version]
  80. Shin, M.; Momb, J.; Appling, D.R. Human mitochondrial MTHFD2 is a dual redox cofactor-specific methylenetetrahydrofolate dehydrogenase/methenyltetrahydrofolate cyclohydrolase. Cancer Metab. 2017, 5, 11. [Google Scholar] [CrossRef]
  81. Nilsson, R.; Nicolaidou, V.; Koufaris, C. Mitochondrial MTHFD isozymes display distinct expression, regulation, and association with cancer. Gene 2019, 716, 144032. [Google Scholar] [CrossRef] [PubMed]
  82. Available online: https://seer.cancer.gov/statfacts/html/ovary.html (accessed on 1 December 2021).
  83. Worzfeld, T.; Strandmann, E.P.V.; Huber, M.; Adhikary, T.; Wagner, U.; Reinartz, S.; Muller, R. The Unique Molecular and Cellular Microenvironment of Ovarian Cancer. Front. Oncol. 2017, 7, 24. [Google Scholar] [CrossRef] [Green Version]
  84. Lokadasan, R.; James, F.V.; Narayanan, G.; Prabhakaran, P.K. Targeted agents in epithelial ovarian cancer: Review on emerging therapies and future developments. Ecancermedicalscience 2016, 10, 626. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Kim, A.; Ueda, Y.; Naka, T.; Enomoto, T. Therapeutic strategies in epithelial ovarian cancer. J. Exp. Clin. Cancer Res. 2012, 31, 14. [Google Scholar] [CrossRef] [Green Version]
  86. Ledermann, J.A.; Raja, F.A.; Fotopoulou, C.; Gonzalez-Martin, A.; Colombo, N.; Sessa, C. Newly diagnosed and relapsed epithelial ovarian carcinoma: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann. Oncol. 2013, 24 (Suppl. S6), vi24–vi32. [Google Scholar] [CrossRef]
  87. Boussios, S.; Zarkavelis, G.; Seraj, E.; Zerdes, I.; Tatsi, K.; Pentheroudakis, G. Non-epithelial Ovarian Cancer: Elucidating Uncommon Gynaecological Malignancies. Anticancer Res. 2016, 36, 5031–5042. [Google Scholar] [CrossRef] [Green Version]
  88. Devouassoux-Shisheboran, M.; Genestie, C. Pathobiology of ovarian carcinomas. Chin. J. Cancer 2015, 34, 50–55. [Google Scholar] [CrossRef] [Green Version]
  89. Neff, R.T.; Senter, L.; Salani, R. BRCA mutation in ovarian cancer: Testing, implications and treatment considerations. Ther. Adv. Med. Oncol. 2017, 9, 519–531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Ramus, S.J.; Gayther, S.A. The contribution of BRCA1 and BRCA2 to ovarian cancer. Mol. Oncol. 2009, 3, 138–150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Alsop, K.; Fereday, S.; Meldrum, C.; deFazio, A.; Emmanuel, C.; George, J.; Dobrovic, A.; Birrer, M.J.; Webb, P.M.; Stewart, C.; et al. BRCA mutation frequency and patterns of treatment response in BRCA mutation-positive women with ovarian cancer: A report from the Australian Ovarian Cancer Study Group. J. Clin. Oncol. 2012, 30, 2654–2663. [Google Scholar] [CrossRef] [Green Version]
  92. Pal, T.; Permuth-Wey, J.; Betts, J.A.; Krischer, J.P.; Fiorica, J.; Arango, H.; LaPolla, J.; Hoffman, M.; Martino, M.A.; Wakeley, K.; et al. BRCA1 and BRCA2 mutations account for a large proportion of ovarian carcinoma cases. Cancer 2005, 104, 2807–2816. [Google Scholar] [CrossRef]
  93. Hennessy, B.T.; Timms, K.M.; Carey, M.S.; Gutin, A.; Meyer, L.A.; Flake, D.D., 2nd; Abkevich, V.; Potter, J.; Pruss, D.; Glenn, P.; et al. Somatic mutations in BRCA1 and BRCA2 could expand the number of patients that benefit from poly (ADP ribose) polymerase inhibitors in ovarian cancer. J. Clin. Oncol. 2010, 28, 3570–3576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Baldwin, R.L.; Nemeth, E.; Tran, H.; Shvartsman, H.; Cass, I.; Narod, S.; Karlan, B.Y. BRCA1 promoter region hypermethylation in ovarian carcinoma: A population-based study. Cancer Res. 2000, 60, 5329–5333. [Google Scholar] [PubMed]
  95. Esteller, M.; Silva, J.M.; Dominguez, G.; Bonilla, F.; Matias-Guiu, X.; Lerma, E.; Bussaglia, E.; Prat, J.; Harkes, I.C.; Repasky, E.A.; et al. Promoter hypermethylation and BRCA1 inactivation in sporadic breast and ovarian tumors. J. Natl. Cancer Inst. 2000, 92, 564–569. [Google Scholar] [CrossRef]
  96. Konstantinopoulos, P.A.; Ceccaldi, R.; Shapiro, G.I.; D’Andrea, A.D. Homologous Recombination Deficiency: Exploiting the Fundamental Vulnerability of Ovarian Cancer. Cancer Discov. 2015, 5, 1137–1154. [Google Scholar] [CrossRef] [Green Version]
  97. Corney, D.C.; Flesken-Nikitin, A.; Choi, J.; Nikitin, A.Y. Role of p53 and Rb in ovarian cancer. Adv. Exp. Med. Biol. 2008, 622, 99–117. [Google Scholar] [CrossRef] [Green Version]
  98. Cancer Genome Atlas Research, N. Integrated genomic analyses of ovarian carcinoma. Nature 2011, 474, 609–615. [Google Scholar] [CrossRef] [PubMed]
  99. du Bois, A.; Reuss, A.; Pujade-Lauraine, E.; Harter, P.; Ray-Coquard, I.; Pfisterer, J. Role of surgical outcome as prognostic factor in advanced epithelial ovarian cancer: A combined exploratory analysis of 3 prospectively randomized phase 3 multicenter trials: By the Arbeitsgemeinschaft Gynaekologische Onkologie Studiengruppe Ovarialkarzinom (AGO-OVAR) and the Groupe d’Investigateurs Nationaux Pour les Etudes des Cancers de l’Ovaire (GINECO). Cancer 2009, 115, 1234–1244. [Google Scholar] [CrossRef]
  100. van der Burg, M.E.; van Lent, M.; Buyse, M.; Kobierska, A.; Colombo, N.; Favalli, G.; Lacave, A.J.; Nardi, M.; Renard, J.; Pecorelli, S. The effect of debulking surgery after induction chemotherapy on the prognosis in advanced epithelial ovarian cancer. Gynecological Cancer Cooperative Group of the European Organization for Research and Treatment of Cancer. N. Engl. J. Med. 1995, 332, 629–634. [Google Scholar] [CrossRef]
  101. Vergote, I.B.; De Wever, I.; Decloedt, J.; Tjalma, W.; Van Gramberen, M.; van Dam, P. Neoadjuvant chemotherapy versus primary debulking surgery in advanced ovarian cancer. Semin. Oncol. 2000, 27, 31–36. [Google Scholar] [CrossRef]
  102. Ozols, R.F.; Bundy, B.N.; Greer, B.E.; Fowler, J.M.; Clarke-Pearson, D.; Burger, R.A.; Mannel, R.S.; DeGeest, K.; Hartenbach, E.M.; Baergen, R. Phase III trial of carboplatin and paclitaxel compared with cisplatin and paclitaxel in patients with optimally resected stage III ovarian cancer: A Gynecologic Oncology Group study. J. Clin. Oncol. 2003, 21, 3194–3200. [Google Scholar] [CrossRef]
  103. Sato, S.; Itamochi, H. Neoadjuvant chemotherapy in advanced ovarian cancer: Latest results and place in therapy. Ther. Adv. Med. Oncol. 2014, 6, 293–304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Gupta, S.; Nag, S.; Aggarwal, S.; Rauthan, A.; Warrier, N. Maintenance therapy for recurrent epithelial ovarian cancer: Current therapies and future perspectives—A review. J. Ovarian Res. 2019, 12, 103. [Google Scholar] [CrossRef] [PubMed]
  105. Falzone, L.; Scandurra, G.; Lombardo, V.; Gattuso, G.; Lavoro, A.; Distefano, A.B.; Scibilia, G.; Scollo, P. A multidisciplinary approach remains the best strategy to improve and strengthen the management of ovarian cancer (Review). Int. J. Oncol. 2021, 59, 1–14. [Google Scholar] [CrossRef] [PubMed]
  106. Vernooij, F.; Heintz, A.P.; Coebergh, J.W.; Massuger, L.F.; Witteveen, P.O.; van der Graaf, Y. Specialized and high-volume care leads to better outcomes of ovarian cancer treatment in the Netherlands. Gynecol. Oncol. 2009, 112, 455–461. [Google Scholar] [CrossRef] [PubMed]
  107. Cortez, A.J.; Tudrej, P.; Kujawa, K.A.; Lisowska, K.M. Advances in ovarian cancer therapy. Cancer Chemother. Pharmacol. 2018, 81, 17–38. [Google Scholar] [CrossRef] [Green Version]
  108. Evans, T.; Matulonis, U. PARP inhibitors in ovarian cancer: Evidence, experience and clinical potential. Ther. Adv. Med. Oncol. 2017, 9, 253–267. [Google Scholar] [CrossRef]
  109. Kobel, M.; Madore, J.; Ramus, S.J.; Clarke, B.A.; Pharoah, P.D.; Deen, S.; Bowtell, D.D.; Odunsi, K.; Menon, U.; Morrison, C.; et al. Evidence for a time-dependent association between FOLR1 expression and survival from ovarian carcinoma: Implications for clinical testing. An Ovarian Tumour Tissue Analysis consortium study. Br. J. Cancer 2014, 111, 2297–2307. [Google Scholar] [CrossRef] [Green Version]
  110. Scaranti, M.; Cojocaru, E.; Banerjee, S.; Banerji, U. Exploiting the folate receptor alpha in oncology. Nat. Rev. Clin. Oncol. 2020, 17, 349–359. [Google Scholar] [CrossRef]
  111. Ebel, W.; Routhier, E.L.; Foley, B.; Jacob, S.; McDonough, J.M.; Patel, R.K.; Turchin, H.A.; Chao, Q.; Kline, J.B.; Old, L.J.; et al. Preclinical evaluation of MORAb-003, a humanized monoclonal antibody antagonizing folate receptor-alpha. Cancer Immun. 2007, 7, 6. [Google Scholar]
  112. Ledermann, J.A.; Canevari, S.; Thigpen, T. Targeting the folate receptor: Diagnostic and therapeutic approaches to personalize cancer treatments. Ann. Oncol. 2015, 26, 2034–2043. [Google Scholar] [CrossRef] [PubMed]
  113. Lin, J.; Spidel, J.L.; Maddage, C.J.; Rybinski, K.A.; Kennedy, R.P.; Krauthauser, C.L.; Park, Y.C.; Albone, E.F.; Jacob, S.; Goserud, M.T.; et al. The antitumor activity of the human FOLR1-specific monoclonal antibody, farletuzumab, in an ovarian cancer mouse model is mediated by antibody-dependent cellular cytotoxicity. Cancer Biol. Ther. 2013, 14, 1032–1038. [Google Scholar] [CrossRef] [Green Version]
  114. Wen, Y.; Graybill, W.S.; Previs, R.A.; Hu, W.; Ivan, C.; Mangala, L.S.; Zand, B.; Nick, A.M.; Jennings, N.B.; Dalton, H.J.; et al. Immunotherapy targeting folate receptor induces cell death associated with autophagy in ovarian cancer. Clin. Cancer Res. 2015, 21, 448–459. [Google Scholar] [CrossRef] [Green Version]
  115. Konner, J.A.; Bell-McGuinn, K.M.; Sabbatini, P.; Hensley, M.L.; Tew, W.P.; Pandit-Taskar, N.; Vander Els, N.; Phillips, M.D.; Schweizer, C.; Weil, S.C.; et al. Farletuzumab, a humanized monoclonal antibody against folate receptor alpha, in epithelial ovarian cancer: A phase I study. Clin. Cancer Res. 2010, 16, 5288–5295. [Google Scholar] [CrossRef] [Green Version]
  116. Armstrong, D.K.; White, A.J.; Weil, S.C.; Phillips, M.; Coleman, R.L. Farletuzumab (a monoclonal antibody against folate receptor alpha) in relapsed platinum-sensitive ovarian cancer. Gynecol. Oncol. 2013, 129, 452–458. [Google Scholar] [CrossRef]
  117. Sato, S.; Itamochi, H. Profile of farletuzumab and its potential in the treatment of solid tumors. OncoTargets Ther. 2016, 9, 1181–1188. [Google Scholar] [CrossRef] [Green Version]
  118. Vergote, I.; Armstrong, D.; Scambia, G.; Teneriello, M.; Sehouli, J.; Schweizer, C.; Weil, S.C.; Bamias, A.; Fujiwara, K.; Ochiai, K.; et al. A Randomized, Double-Blind, Placebo-Controlled, Phase III Study to Assess Efficacy and Safety of Weekly Farletuzumab in Combination with Carboplatin and Taxane in Patients with Ovarian Cancer in First Platinum-Sensitive Relapse. J. Clin. Oncol. 2016, 34, 2271–2278. [Google Scholar] [CrossRef]
  119. Ab, O.; Whiteman, K.R.; Bartle, L.M.; Sun, X.; Singh, R.; Tavares, D.; LaBelle, A.; Payne, G.; Lutz, R.J.; Pinkas, J.; et al. IMGN853, a Folate Receptor-alpha (FRalpha)-Targeting Antibody-Drug Conjugate, Exhibits Potent Targeted Antitumor Activity against FRalpha-Expressing Tumors. Mol. Cancer Ther. 2015, 14, 1605–1613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Moore, K.N.; Borghaei, H.; O’Malley, D.M.; Jeong, W.; Seward, S.M.; Bauer, T.M.; Perez, R.P.; Matulonis, U.A.; Running, K.L.; Zhang, X.; et al. Phase 1 dose-escalation study of mirvetuximab soravtansine (IMGN853), a folate receptor alpha-targeting antibody-drug conjugate, in patients with solid tumors. Cancer 2017, 123, 3080–3087. [Google Scholar] [CrossRef]
  121. Moore, K.N.; Oza, A.M.; Colombo, N.; Oaknin, A.; Scambia, G.; Lorusso, D.; Konecny, G.E.; Banerjee, S.; Murphy, C.G.; Tanyi, J.L.; et al. Phase III, randomized trial of mirvetuximab soravtansine versus chemotherapy in patients with platinum-resistant ovarian cancer: Primary analysis of FORWARD I. Ann. Oncol. 2021, 32, 757–765. [Google Scholar] [CrossRef] [PubMed]
  122. Reddy, J.A.; Dorton, R.; Westrick, E.; Dawson, A.; Smith, T.; Xu, L.C.; Vetzel, M.; Kleindl, P.; Vlahov, I.R.; Leamon, C.P. Preclinical evaluation of EC145, a folate-vinca alkaloid conjugate. Cancer Res. 2007, 67, 4434–4442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Lorusso, P.M.; Edelman, M.J.; Bever, S.L.; Forman, K.M.; Pilat, M.; Quinn, M.F.; Li, J.; Heath, E.I.; Malburg, L.M.; Klein, P.J.; et al. Phase I study of folate conjugate EC145 (Vintafolide) in patients with refractory solid tumors. J. Clin. Oncol. 2012, 30, 4011–4016. [Google Scholar] [CrossRef]
  124. Naumann, R.W.; Coleman, R.L.; Burger, R.A.; Sausville, E.A.; Kutarska, E.; Ghamande, S.A.; Gabrail, N.Y.; Depasquale, S.E.; Nowara, E.; Gilbert, L.; et al. PRECEDENT: A randomized phase II trial comparing vintafolide (EC145) and pegylated liposomal doxorubicin (PLD) in combination versus PLD alone in patients with platinum-resistant ovarian cancer. J. Clin. Oncol. 2013, 31, 4400–4406. [Google Scholar] [CrossRef] [PubMed]
  125. Vergote, I.; Leamon, C.P. Vintafolide: A novel targeted therapy for the treatment of folate receptor expressing tumors. Ther. Adv. Med. Oncol. 2015, 7, 206–218. [Google Scholar] [CrossRef] [Green Version]
  126. Banerji, U.; Garces, A.H.I.; Michalarea, V.; Ruddle, R.; Raynaud, F.I.; Riisnaes, R.; Rodrigues, D.N.; Tunariu, N.; Porter, J.C.; Ward, S.E.; et al. An investigator-initiated phase I study of ONX-0801 a first-in-class alpha folate receptor targeted, small molecule thymidylate synthase inhibitor in solid tumors. J. Clin. Oncol. 2017, 35, 2503. [Google Scholar] [CrossRef]
  127. Konda, S.D.; Aref, M.; Brechbiel, M.; Wiener, E.C. Development of a tumor-targeting MR contrast agent using the high-affinity folate receptor: Work in progress. Investig. Radiol. 2000, 35, 50–57. [Google Scholar] [CrossRef] [PubMed]
  128. Meier, R.; Henning, T.D.; Boddington, S.; Tavri, S.; Arora, S.; Piontek, G.; Rudelius, M.; Corot, C.; Daldrup-Link, H.E. Breast cancers: MR imaging of folate-receptor expression with the folate-specific nanoparticle P1133. Radiology 2010, 255, 527–535. [Google Scholar] [CrossRef] [Green Version]
  129. Siegel, B.A.; Dehdashti, F.; Mutch, D.G.; Podoloff, D.A.; Wendt, R.; Sutton, G.P.; Burt, R.W.; Ellis, P.R.; Mathias, C.J.; Green, M.A.; et al. Evaluation of 111In-DTPA-folate as a receptor-targeted diagnostic agent for ovarian cancer: Initial clinical results. J. Nucl. Med. 2003, 44, 700–707. [Google Scholar]
  130. Fisher, R.E.; Siegel, B.A.; Edell, S.L.; Oyesiku, N.M.; Morgenstern, D.E.; Messmann, R.A.; Amato, R.J. Exploratory study of 99mTc-EC20 imaging for identifying patients with folate receptor-positive solid tumors. J. Nucl. Med. 2008, 49, 899–906. [Google Scholar] [CrossRef] [Green Version]
  131. Brand, C.; Sadique, A.; Houghton, J.L.; Gangangari, K.; Ponte, J.F.; Lewis, J.S.; Pillarsetty, N.V.K.; Konner, J.A.; Reiner, T. Leveraging PET to image folate receptor alpha therapy of an antibody-drug conjugate. EJNMMI Res. 2018, 8, 87. [Google Scholar] [CrossRef] [PubMed]
  132. Kennedy, M.D.; Jallad, K.N.; Thompson, D.H.; Ben-Amotz, D.; Low, P.S. Optical imaging of metastatic tumors using a folate-targeted fluorescent probe. J. Biomed. Opt. 2003, 8, 636–641. [Google Scholar] [CrossRef]
  133. Tummers, Q.R.; Hoogstins, C.E.; Gaarenstroom, K.N.; de Kroon, C.D.; van Poelgeest, M.I.; Vuyk, J.; Bosse, T.; Smit, V.T.; van de Velde, C.J.; Cohen, A.F.; et al. Intraoperative imaging of folate receptor alpha positive ovarian and breast cancer using the tumor specific agent EC17. Oncotarget 2016, 7, 32144–32155. [Google Scholar] [CrossRef] [Green Version]
  134. van Dam, G.M.; Themelis, G.; Crane, L.M.; Harlaar, N.J.; Pleijhuis, R.G.; Kelder, W.; Sarantopoulos, A.; de Jong, J.S.; Arts, H.J.; van der Zee, A.G.; et al. Intraoperative tumor-specific fluorescence imaging in ovarian cancer by folate receptor-alpha targeting: First in-human results. Nat. Med. 2011, 17, 1315–1319. [Google Scholar] [CrossRef] [PubMed]
  135. Available online: https://www.fda.gov/news-events/press-announcements/fda-approves-new-imaging-drug-help-identify-ovarian-cancer-lesions (accessed on 1 December 2021).
  136. Roche, M.; Parisi, L.; Li, L.; Knehans, A.; Phaeton, R.; Kesterson, J.P. The role of pemetrexed in recurrent epithelial ovarian cancer: A scoping review. Oncol. Rev. 2018, 12, 346. [Google Scholar] [CrossRef] [Green Version]
  137. Chattopadhyay, S.; Moran, R.G.; Goldman, I.D. Pemetrexed: Biochemical and cellular pharmacology, mechanisms, and clinical applications. Mol. Cancer Ther. 2007, 6, 404–417. [Google Scholar] [CrossRef] [Green Version]
  138. Issaeva, N.; Thomas, H.D.; Djureinovic, T.; Jaspers, J.E.; Stoimenov, I.; Kyle, S.; Pedley, N.; Gottipati, P.; Zur, R.; Sleeth, K.; et al. 6-thioguanine selectively kills BRCA2-defective tumors and overcomes PARP inhibitor resistance. Cancer Res. 2010, 70, 6268–6276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Bronder, J.L.; Moran, R.G. A defect in the p53 response pathway induced by de novo purine synthesis inhibition. J. Biol. Chem. 2003, 278, 48861–48871. [Google Scholar] [CrossRef] [Green Version]
  140. Bertino, J.R.; Waud, W.R.; Parker, W.B.; Lubin, M. Targeting tumors that lack methylthioadenosine phosphorylase (MTAP) activity: Current strategies. Cancer Biol. Ther. 2011, 11, 627–632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Hori, H.; Tran, P.; Carrera, C.J.; Hori, Y.; Rosenbach, M.D.; Carson, D.A.; Nobori, T. Methylthioadenosine phosphorylase cDNA transfection alters sensitivity to depletion of purine and methionine in A549 lung cancer cells. Cancer Res. 1996, 56, 5653–5658. [Google Scholar]
  142. Hoxhaj, G.; Hughes-Hallett, J.; Timson, R.C.; Ilagan, E.; Yuan, M.; Asara, J.M.; Ben-Sahra, I.; Manning, B.D. The mTORC1 Signaling Network Senses Changes in Cellular Purine Nucleotide Levels. Cell Rep. 2017, 21, 1331–1346. [Google Scholar] [CrossRef] [Green Version]
  143. Rothbart, S.B.; Racanelli, A.C.; Moran, R.G. Pemetrexed indirectly activates the metabolic kinase AMPK in human carcinomas. Cancer Res. 2010, 70, 10299–10309. [Google Scholar] [CrossRef] [Green Version]
  144. Desmoulin, S.K.; Wang, Y.; Wu, J.; Stout, M.; Hou, Z.; Fulterer, A.; Chang, M.H.; Romero, M.F.; Cherian, C.; Gangjee, A.; et al. Targeting the proton-coupled folate transporter for selective delivery of 6-substituted pyrrolo[2,3-d]pyrimidine antifolate inhibitors of de novo purine biosynthesis in the chemotherapy of solid tumors. Mol. Pharmacol. 2010, 78, 577–587. [Google Scholar] [CrossRef] [Green Version]
  145. Deng, Y.; Wang, Y.; Cherian, C.; Hou, Z.; Buck, S.A.; Matherly, L.H.; Gangjee, A. Synthesis and discovery of high affinity folate receptor-specific glycinamide ribonucleotide formyltransferase inhibitors with antitumor activity. J. Med. Chem. 2008, 51, 5052–5063. [Google Scholar] [CrossRef] [Green Version]
  146. Wang, L.; Cherian, C.; Desmoulin, S.K.; Polin, L.; Deng, Y.; Wu, J.; Hou, Z.; White, K.; Kushner, J.; Matherly, L.H.; et al. Synthesis and antitumor activity of a novel series of 6-substituted pyrrolo[2,3-d]pyrimidine thienoyl antifolate inhibitors of purine biosynthesis with selectivity for high affinity folate receptors and the proton-coupled folate transporter over the reduced folate carrier for cellular entry. J. Med. Chem. 2010, 53, 1306–1318. [Google Scholar] [CrossRef] [Green Version]
  147. Wang, L.; Desmoulin, S.K.; Cherian, C.; Polin, L.; White, K.; Kushner, J.; Fulterer, A.; Chang, M.H.; Mitchell-Ryan, S.; Stout, M.; et al. Synthesis, biological, and antitumor activity of a highly potent 6-substituted pyrrolo[2,3-d]pyrimidine thienoyl antifolate inhibitor with proton-coupled folate transporter and folate receptor selectivity over the reduced folate carrier that inhibits beta-glycinamide ribonucleotide formyltransferase. J. Med. Chem. 2011, 54, 7150–7164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Wang, L.; Wallace, A.; Raghavan, S.; Deis, S.M.; Wilson, M.R.; Yang, S.; Polin, L.; White, K.; Kushner, J.; Orr, S.; et al. 6-Substituted Pyrrolo[2,3-d]pyrimidine Thienoyl Regioisomers as Targeted Antifolates for Folate Receptor alpha and the Proton-Coupled Folate Transporter in Human Tumors. J. Med. Chem. 2015, 58, 6938–6959. [Google Scholar] [CrossRef] [Green Version]
  149. Ravindra, M.; Wilson, M.R.; Tong, N.; O’Connor, C.; Karim, M.; Polin, L.; Wallace-Povirk, A.; White, K.; Kushner, J.; Hou, Z.; et al. Fluorine-Substituted Pyrrolo[2,3-d]Pyrimidine Analogues with Tumor Targeting via Cellular Uptake by Folate Receptor alpha and the Proton-Coupled Folate Transporter and Inhibition of de Novo Purine Nucleotide Biosynthesis. J. Med. Chem. 2018, 61, 4228–4248. [Google Scholar] [CrossRef] [PubMed]
  150. Ducker, G.S.; Ghergurovich, J.M.; Mainolfi, N.; Suri, V.; Jeong, S.K.; Hsin-Jung Li, S.; Friedman, A.; Manfredi, M.G.; Gitai, Z.; Kim, H.; et al. Human SHMT inhibitors reveal defective glycine import as a targetable metabolic vulnerability of diffuse large B-cell lymphoma. Proc. Natl. Acad. Sci. USA 2017, 114, 11404–11409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Garcia-Canaveras, J.C.; Lancho, O.; Ducker, G.S.; Ghergurovich, J.M.; Xu, X.; da Silva-Diz, V.; Minuzzo, S.; Indraccolo, S.; Kim, H.; Herranz, D.; et al. SHMT inhibition is effective and synergizes with methotrexate in T-cell acute lymphoblastic leukemia. Leukemia 2021, 35, 377–388. [Google Scholar] [CrossRef]
  152. Kawai, J.; Toki, T.; Ota, M.; Inoue, H.; Takata, Y.; Asahi, T.; Suzuki, M.; Shimada, T.; Ono, K.; Suzuki, K.; et al. Discovery of a Potent, Selective, and Orally Available MTHFD2 Inhibitor (DS18561882) with in Vivo Antitumor Activity. J. Med. Chem. 2019, 62, 10204–10220. [Google Scholar] [CrossRef]
  153. Dekhne, A.S.; Shah, K.; Ducker, G.S.; Katinas, J.M.; Wong-Roushar, J.; Nayeen, M.J.; Doshi, A.; Ning, C.; Bao, X.; Fruhauf, J.; et al. Novel pyrrolo[3,2-d]pyrimidine compounds target mitochondrial and cytosolic one-carbon metabolism with broad-spectrum antitumor efficacy. Mol. Cancer Ther. 2019, 18, 1787–1799. [Google Scholar] [CrossRef] [Green Version]
  154. O’Connor, C.; Wallace-Povirk, A.; Ning, C.; Fruhauf, J.; Tong, N.; Gangjee, A.; Matherly, L.H.; Hou, Z. Folate transporter dynamics and therapy with classic and tumor-targeted antifolates. Sci Rep. 2021, 11, 6389. [Google Scholar] [CrossRef] [PubMed]
  155. Dekhne, A.S.; Ning, C.; Nayeen, M.J.; Shah, K.; Kalpage, H.; Frühauf, J.; Wallace-Povirk, A.; O’Connor, C.; Hou, Z.; Kim, S.; et al. Cellular Pharmacodynamics of a Novel Pyrrolo[3,2-d]pyrimidine Inhibitor Targeting Mitochondrial and Cytosolic One-Carbon Metabolism. Mol. Pharmacol. 2020, 97, 9–22. [Google Scholar] [CrossRef] [PubMed]
  156. Wallace-Povirk, A.; O’Connor, C.; Bao, X.; Katinas, J.; Wong-Roushar, J.; Dekhne, A.; Hou, Z.; Nayeen, M.J.; Shah, K.; Nunez, J.; et al. Abstract 2348: Targeting mitochondrial and cytosolic one-carbon metabolism in epithelial ovarian cancer via folate receptor alpha. In Proceedings of the AACR Annual Meeting, Philadelphia, PA, USA, 10–15 April 2021; p. 2348. [Google Scholar]
  157. Li, Q.; Yang, F.; Shi, X.; Bian, S.; Shen, F.; Wu, Y.; Zhu, C.; Fu, F.; Wang, J.; Zhou, J.; et al. MTHFD2 promotes ovarian cancer growth and metastasis via activation of the STAT3 signaling pathway. FEBS Open Bio 2021, 11, 2845–2857. [Google Scholar] [CrossRef] [PubMed]
  158. Gustafsson, R.; Jemth, A.S.; Gustafsson, N.M.; Farnegardh, K.; Loseva, O.; Wiita, E.; Bonagas, N.; Dahllund, L.; Llona-Minguez, S.; Haggblad, M.; et al. Crystal Structure of the Emerging Cancer Target MTHFD2 in Complex with a Substrate-Based Inhibitor. Cancer Res. 2017, 77, 937–948. [Google Scholar] [CrossRef] [Green Version]
  159. Ju, H.Q.; Lu, Y.X.; Chen, D.L.; Zuo, Z.X.; Liu, Z.X.; Wu, Q.N.; Mo, H.Y.; Wang, Z.X.; Wang, D.S.; Pu, H.Y.; et al. Modulation of Redox Homeostasis by Inhibition of MTHFD2 in Colorectal Cancer: Mechanisms and Therapeutic Implications. J. Natl. Cancer Inst. 2019, 111, 584–596. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Kawai, J.; Ota, M.; Ohki, H.; Toki, T.; Suzuki, M.; Shimada, T.; Matsui, S.; Inoue, H.; Sugihara, C.; Matsuhashi, N.; et al. Structure-Based Design and Synthesis of an Isozyme-Selective MTHFD2 Inhibitor with a Tricyclic Coumarin Scaffold. ACS Med. Chem. Lett. 2019, 10, 893–898. [Google Scholar] [CrossRef]
  161. Ducker, G.S.; Chen, L.; Morscher, R.J.; Ghergurovich, J.M.; Esposito, M.; Teng, X.; Kang, Y.; Rabinowitz, J.D. Reversal of Cytosolic One-Carbon Flux Compensates for Loss of the Mitochondrial Folate Pathway. Cell Metab. 2016, 23, 1140–1153. [Google Scholar] [CrossRef] [Green Version]
  162. Bejarano, L.; Jordao, M.J.C.; Joyce, J.A. Therapeutic Targeting of the Tumor Microenvironment. Cancer Discov. 2021, 11, 933–959. [Google Scholar] [CrossRef]
  163. Ishii, G.; Ochiai, A.; Neri, S. Phenotypic and functional heterogeneity of cancer-associated fibroblast within the tumor microenvironment. Adv. Drug Deliv. Rev. 2016, 99, 186–196. [Google Scholar] [CrossRef]
  164. Gilead, A.; Meir, G.; Neeman, M. The role of angiogenesis, vascular maturation, regression and stroma infiltration in dormancy and growth of implanted MLS ovarian carcinoma spheroids. Int. J. Cancer 2004, 108, 524–531. [Google Scholar] [CrossRef]
  165. Schauer, I.G.; Sood, A.K.; Mok, S.; Liu, J. Cancer-associated fibroblasts and their putative role in potentiating the initiation and development of epithelial ovarian cancer. Neoplasia 2011, 13, 393–405. [Google Scholar] [CrossRef] [Green Version]
  166. Zhang, Y.; Tang, H.; Cai, J.; Zhang, T.; Guo, J.; Feng, D.; Wang, Z. Ovarian cancer-associated fibroblasts contribute to epithelial ovarian carcinoma metastasis by promoting angiogenesis, lymphangiogenesis and tumor cell invasion. Cancer Lett. 2011, 303, 47–55. [Google Scholar] [CrossRef]
  167. Givel, A.M.; Kieffer, Y.; Scholer-Dahirel, A.; Sirven, P.; Cardon, M.; Pelon, F.; Magagna, I.; Gentric, G.; Costa, A.; Bonneau, C.; et al. miR200-regulated CXCL12beta promotes fibroblast heterogeneity and immunosuppression in ovarian cancers. Nat. Commun. 2018, 9, 1056. [Google Scholar] [CrossRef]
  168. Ahmed, Z.; Bicknell, R. Angiogenic signalling pathways. Angiogenesis Protoc. 2009, 467, 3–24. [Google Scholar] [CrossRef]
  169. Carmeliet, P.; Jain, R.K. Angiogenesis in cancer and other diseases. Nature 2000, 407, 249–257. [Google Scholar] [CrossRef]
  170. Hanahan, D.; Coussens, L.M. Accessories to the crime: Functions of cells recruited to the tumor microenvironment. Cancer Cell 2012, 21, 309–322. [Google Scholar] [CrossRef] [Green Version]
  171. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Colvin, E.K. Tumor-associated macrophages contribute to tumor progression in ovarian cancer. Front. Oncol. 2014, 4, 137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Takaishi, K.; Komohara, Y.; Tashiro, H.; Ohtake, H.; Nakagawa, T.; Katabuchi, H.; Takeya, M. Involvement of M2-polarized macrophages in the ascites from advanced epithelial ovarian carcinoma in tumor progression via Stat3 activation. Cancer Sci. 2010, 101, 2128–2136. [Google Scholar] [CrossRef] [PubMed]
  174. Kowal, J.; Kornete, M.; Joyce, J.A. Re-education of macrophages as a therapeutic strategy in cancer. Immunotherapy 2019, 11, 677–689. [Google Scholar] [CrossRef] [PubMed]
  175. Wu, J.W.Y.; Dand, S.; Doig, L.; Papenfuss, A.T.; Scott, C.L.; Ho, G.; Ooi, J.D. T-Cell Receptor Therapy in the Treatment of Ovarian Cancer: A Mini Review. Front. Immunol. 2021, 12, 1141. [Google Scholar] [CrossRef] [PubMed]
  176. Kandalaft, L.E.; Powell, D.J., Jr.; Coukos, G. A phase I clinical trial of adoptive transfer of folate receptor-alpha redirected autologous T cells for recurrent ovarian cancer. J. Transl. Med. 2012, 10, 157. [Google Scholar] [CrossRef] [Green Version]
  177. Liang, Z.; Dong, J.; Yang, N.; Li, S.D.; Yang, Z.Y.; Huang, R.; Li, F.J.; Wang, W.T.; Ren, J.K.; Lei, J.; et al. Tandem CAR-T cells targeting FOLR1 and MSLN enhance the antitumor effects in ovarian cancer. Int. J. Biol. Sci. 2021, 17, 4365–4376. [Google Scholar] [CrossRef]
  178. Sale, S.; Orsulic, S. Models of ovarian cancer metastasis: Murine models. Drug Discov. Today Dis. Models 2006, 3, 149–154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Xing, D.; Orsulic, S. A mouse model for the molecular characterization of brca1-associated ovarian carcinoma. Cancer Res. 2006, 66, 8949–8953. [Google Scholar] [CrossRef] [Green Version]
  180. Rodriguez, A.E.; Ducker, G.S.; Billingham, L.K.; Martinez, C.A.; Mainolfi, N.; Suri, V.; Friedman, A.; Manfredi, M.G.; Weinberg, S.E.; Rabinowitz, J.D.; et al. Serine Metabolism Supports Macrophage IL-1beta Production. Cell Metab. 2019, 29, 1003–1011. [Google Scholar] [CrossRef] [Green Version]
  181. Yu, W.; Wang, Z.; Zhang, K.; Chi, Z.; Xu, T.; Jiang, D.; Chen, S.; Li, W.; Yang, X.; Zhang, X.; et al. One-Carbon Metabolism Supports S-Adenosylmethionine and Histone Methylation to Drive Inflammatory Macrophages. Mol. Cell 2019, 75, 1147–1160. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Folate transport and C1 metabolism. A schematic is shown depicting cellular uptake by facilitative transport via RFC or PCFT or by endocytosis via FRα. Intracellular folates are metabolized to polyglutamate conjugates. Folate “monoglutamates” are transported into the mitochondria by SLC25A32. In the mitochondria, serine is catabolized by sequential SHMT2, MTHFD2/L, and MTHFD1L through which the C1 moiety from serine C3 is incorporated into formate, thus providing C1 units for cellular biosynthesis in the cytosol. Abbreviations are as follows: 10-CHO-THF, 10-formyl tetrahydrofolate; 5,10-me-THF, 5,10-methylene tetrahydrofolate; AICAR, 5-aminoimidazole-4-carboxamide; ALDH1L2, aldehyde dehydrogenase 1 family member L2; ATIC, 5-aminoimidazole-4-carboxamide ribonucleotide formyltransferase; DHF, dihydrofolate; DHFR, dihydrofolate reductase; FAICAR, formyl 5-aminoimidazole-4-carboxamide ribonucleotide; fGAR, formyl glycinamide ribonucleotide; FPGS, folylpoly-γ-glutamate synthetase; GAR, glycinamide ribonucleotide; GARFTase, glycinamide ribonucleotide formyltransferase; GR, glutathione reductase; GS, glutathione synthetase; GSH, glutathione; MTFMT, methionyl tRNA formyltransferase; MTHFD1, methylenetetrahydrofolate dehydrogenase 1; MTHFD2(L), methylene tetrahydrofolate dehydrogenase 2(-like); MTHFR, methylenetetrahydrofolate reductase; MTR, methionine synthase; PCFT, proton-coupled folate transporter; PGs, polyglutamates; PRPP, phosphoribosyl pyrophosphate; RFC, reduced folate carrier; SAM, S-adenosylmethionine; SHMT1/2, serine hydroxymethyltransferase 1/2; THF, tetrahydrofolate; and TS, thymidylate synthase.
Figure 1. Folate transport and C1 metabolism. A schematic is shown depicting cellular uptake by facilitative transport via RFC or PCFT or by endocytosis via FRα. Intracellular folates are metabolized to polyglutamate conjugates. Folate “monoglutamates” are transported into the mitochondria by SLC25A32. In the mitochondria, serine is catabolized by sequential SHMT2, MTHFD2/L, and MTHFD1L through which the C1 moiety from serine C3 is incorporated into formate, thus providing C1 units for cellular biosynthesis in the cytosol. Abbreviations are as follows: 10-CHO-THF, 10-formyl tetrahydrofolate; 5,10-me-THF, 5,10-methylene tetrahydrofolate; AICAR, 5-aminoimidazole-4-carboxamide; ALDH1L2, aldehyde dehydrogenase 1 family member L2; ATIC, 5-aminoimidazole-4-carboxamide ribonucleotide formyltransferase; DHF, dihydrofolate; DHFR, dihydrofolate reductase; FAICAR, formyl 5-aminoimidazole-4-carboxamide ribonucleotide; fGAR, formyl glycinamide ribonucleotide; FPGS, folylpoly-γ-glutamate synthetase; GAR, glycinamide ribonucleotide; GARFTase, glycinamide ribonucleotide formyltransferase; GR, glutathione reductase; GS, glutathione synthetase; GSH, glutathione; MTFMT, methionyl tRNA formyltransferase; MTHFD1, methylenetetrahydrofolate dehydrogenase 1; MTHFD2(L), methylene tetrahydrofolate dehydrogenase 2(-like); MTHFR, methylenetetrahydrofolate reductase; MTR, methionine synthase; PCFT, proton-coupled folate transporter; PGs, polyglutamates; PRPP, phosphoribosyl pyrophosphate; RFC, reduced folate carrier; SAM, S-adenosylmethionine; SHMT1/2, serine hydroxymethyltransferase 1/2; THF, tetrahydrofolate; and TS, thymidylate synthase.
Cancers 14 00191 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wallace-Povirk, A.; Hou, Z.; Nayeen, M.J.; Gangjee, A.; Matherly, L.H. Folate Transport and One-Carbon Metabolism in Targeted Therapies of Epithelial Ovarian Cancer. Cancers 2022, 14, 191. https://doi.org/10.3390/cancers14010191

AMA Style

Wallace-Povirk A, Hou Z, Nayeen MJ, Gangjee A, Matherly LH. Folate Transport and One-Carbon Metabolism in Targeted Therapies of Epithelial Ovarian Cancer. Cancers. 2022; 14(1):191. https://doi.org/10.3390/cancers14010191

Chicago/Turabian Style

Wallace-Povirk, Adrianne, Zhanjun Hou, Md. Junayed Nayeen, Aleem Gangjee, and Larry H. Matherly. 2022. "Folate Transport and One-Carbon Metabolism in Targeted Therapies of Epithelial Ovarian Cancer" Cancers 14, no. 1: 191. https://doi.org/10.3390/cancers14010191

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