Next Article in Journal
The Utility of Noninvasive Urinary Biomarkers for the Evaluation of Vesicoureteral Reflux in Children
Next Article in Special Issue
MLLT11 Regulates Endometrial Stroma Cell Adhesion, Proliferation and Survival in Ectopic Lesions of Women with Advanced Endometriosis
Previous Article in Journal
Green Biologics: Harnessing the Power of Plants to Produce Pharmaceuticals
Previous Article in Special Issue
Hyperandrogenism and Its Possible Effects on Endometrial Receptivity: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Copper in Gynecological Diseases

by
Rocío A. Conforti
1,
María B. Delsouc
1,
Edith Zorychta
2,
Carlos M. Telleria
2,3,*,† and
Marilina Casais
1,*,†
1
Facultad de Química, Bioquímica y Farmacia, Universidad Nacional de San Luis (UNSL), Instituto Multidisciplinario de Investigaciones Biológicas de San Luis (IMIBIO-SL-CONICET), San Luis CP D5700HHW, Argentina
2
Experimental Pathology Unit, Department of Pathology, Faculty of Medicine and Health Sciences, McGill University, 3775 University Street, Montreal, QC H3A 2B4, Canada
3
Cancer Research Program, Research Institute, McGill University Health Centre, Montreal, QC H4A 3J1, Canada
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(24), 17578; https://doi.org/10.3390/ijms242417578
Submission received: 21 November 2023 / Revised: 14 December 2023 / Accepted: 15 December 2023 / Published: 17 December 2023
(This article belongs to the Special Issue Molecular Research in Gynecological Diseases)

Abstract

:
Copper (Cu) is an essential micronutrient for the correct development of eukaryotic organisms. This metal plays a key role in many cellular and physiological activities, including enzymatic activity, oxygen transport, and cell signaling. Although the redox activity of Cu is crucial for enzymatic reactions, this property also makes it potentially toxic when found at high levels. Due to this dual action of Cu, highly regulated mechanisms are necessary to prevent both the deficiency and the accumulation of this metal since its dyshomeostasis may favor the development of multiple diseases, such as Menkes’ and Wilson’s diseases, neurodegenerative diseases, diabetes mellitus, and cancer. As the relationship between Cu and cancer has been the most studied, we analyze how this metal can affect three fundamental processes for tumor progression: cell proliferation, angiogenesis, and metastasis. Gynecological diseases are characterized by high prevalence, morbidity, and mortality, depending on the case, and mainly include benign and malignant tumors. The cellular processes that promote their progression are affected by Cu, and the mechanisms that occur may be similar. We analyze the crosstalk between Cu deregulation and gynecological diseases, focusing on therapeutic strategies derived from this metal.

1. Introduction

Copper (Cu) is an essential micronutrient for the proper development of eukaryotic organisms [1]. Because it cannot be created or destroyed through metabolic processes, this metal must be acquired from external sources, primarily food and water. According to recommendations, adults should consume approximately 0.9 mg of Cu/day, and in conditions such as pregnancy and lactation, around 1.3 mg/day [2]. In general, the average intake of most people meets or exceeds this requirement since it is estimated that Cu ingested through food, water, and supplements ranges between 1.1 and 1.7 mg/day in adults [3], of which only 15% is retained in tissues: the rest is excreted through the bile and, to a lesser extent, through urine. This micronutrient is present at high concentrations in foods such as liver, crustaceans, red meat, milk, chocolate, seeds, fish, mushrooms, and nuts [4]. Cu is mainly accumulated in the liver, kidneys, brain, heart, muscles, and skeleton [5]. The serum concentration of Cu in healthy adults ranges between 70 and 110 mg/dL, where 70% is bound to its principal transporter, ceruloplasmin (Cp) [6,7].
As a vital trace element, Cu plays a key role in many cellular and physiological processes, such as enzyme activities, oxygen transport, and cell signaling. Being a catalytic cofactor of redox proteins, it is clear that Cu plays a crucial role in carrying out biological functions necessary for growth and development [8]. These functions are due to its two oxidation states: the reduced form (Cu+) and the oxidized form (Cu2+), which give it the ability to act as an electron recipient or donor. The extracellular environment contains mainly Cu2+, while inside the cells, the reduced form of Cu is found [9,10]. Cu2+ may regulate various growth factors and membrane receptors, while Cu+ is involved in intracellular regulation by affecting the activation state of membrane receptors or binding to transcription factors to alter gene expression [11]. Analysis of the human proteome has so far identified more than fifty Cu-binding proteins, of which some examples include Cu/Zn superoxide dismutase (SOD1), cytochrome C oxidase (CCO), Cp, lysyl oxidase (LOX), tyrosinase, and dopamine-β-hydroxylase (DβH), among others (Table 1). The main functions of Cu involve oxidation-reduction reactions that ultimately produce free oxygen radicals. For this reason, free cellular Cu concentrations must be maintained at low levels [8]. Given its essential role in cellular physiology, it is important to understand the mechanisms related to Cu metabolism in biological systems.

1.1. Copper Metabolism

In mammals, Cu absorption, distribution, storage, and excretion take place at both systemic and cellular levels. During the last 20 years, the mechanisms related to these processes have been widely studied [12]. A schematic diagram of Cu metabolism in mammals is shown in Figure 1. Cu homeostasis depends mainly on the precise regulation of these processes by organ systems and individual cells. Studying the various alterations that may cause Cu dyshomeostasis is one of the most attractive focuses in Cu research at the present time.

1.1.1. Copper Uptake

Copper is acquired mainly from food and water and is absorbed through the intestinal epithelium to reach the liver through the portal vein [10]. In the digestive tract, Cu2+ can be incorporated by epithelial cells through the action of divalent metal transporter 1 (DMT1); however, specific deletion of Dmt1 in enterocytes does not prevent intestinal absorption of Cu [13], indicating that other mechanisms of transport are also in place. The reduction of Cu2+ by metalloreductases such as DCYTB and STEAP 2, 3, and 4 on the surface of mammalian cells [14,15] allows Cu+ ions to then be incorporated by CTR1 (copper transporter 1, encoded in humans by the Slc31a1 gene). CTR1 is a high-affinity Cu importer belonging to the SLC31 family, and it plays a fundamental role in Cu homeostasis, being the main pathway of Cu+ incorporation into cells [1]. It is located on the apical membrane of enterocytes [16], but it can also be found on the basolateral membrane and within intracellular organelles [17]. CTR1 is a homotrimeric protein that forms a pore in the membrane, where each monomer displays an extracellular N-terminal domain for Cu binding [18]. It also has three transmembrane domains (TMDs): TMD1 and TMD2 interact with Cu, and TMD3 is essential for CTR1 oligomerization. The cytoplasmic C-terminus allows intracellular delivery of Cu by undergoing conformational changes upon metal binding [19]. Enterocyte-specific Ctr1 knockout mice experience severe Cu deficiency in peripheral tissues, cardiac hypertrophy, liver iron overload, and severe growth and viability defects [20], while systemic inactivation of CTR1 leads to embryonic death [21], confirming the importance of CTR1 in Cu uptake and normal cellular function.

1.1.2. Copper Distribution

The Cu-transporting ATPase α (ATP7A) in intestinal epithelial cells is the essential protein transporting Cu from the intestine to the rest of the body. ATP7A is expressed in many tissues except the liver, where it is replaced by its paralog, Cu-transporting ATPase β (ATP7B) [22]. ATP7B is mainly expressed in the liver, kidney, heart, brain, placenta, and lung [22]. In the placenta and blood–brain barrier, ATP7A ensures sufficient amounts of Cu for proper development of the fetus and brain [23]. After absorption into the enterocyte, ATP7A secretes Cu into the portal circulation, where it binds to soluble chaperones, including albumin, transcuprein, and macroglobulins [24,25,26]. Upon reaching the liver, Cu enters hepatocytes through CTR1, and the liver becomes the main depot of Cu in the body, distributing it to peripheral organs through the bloodstream or excreting it through the bile [27]. Within the cytoplasm, Cu trafficking is tightly coordinated by high-affinity Cu chaperones that deliver Cu to specific proteins and metallothioneins (MTs) that bind Cu for storage [22,24,28,29]. The major Cu chaperones include cytochrome C oxidase (CCO), Cu chaperone for SOD (CCS), and antioxidant chaperone 1 (ATOX1).
Cytochrome C oxidase utilizes Cu for mitochondrial function and oxidative phosphorylation (Figure 1A). CCO consists of two subunits, COX1 and COX2, which bind Cu at conserved sites [30]. The Cu chaperone COX17, located in the mitochondrial intermembrane space (IMS), transports Cu from the cytosol to the IMS to contribute to the correct assembly of CCO [28]. In the IMS, COX17 delivers Cu+ to SCO1 (synthesis of cytochrome C oxidase 1) for transfer to the COX2 subunit or COX11 for delivery to the COX1 subunit [31,32]. Other participants could be involved in Cu trafficking to mitochondria, such as COX19 and a non-protein, anionic copper ligand [32]. Mitochondria provide the main intracellular reservoir of Cu, which is essential for their energy production through oxidative phosphorylation [8,32]. Within enterocytes and other cells, CCS delivers Cu to the SOD1 enzyme to scavenge free radicals (Figure 1B). A recent study suggested that CCS first acquires Cu from CTR1 and then delivers it to SOD1 by forming a CTR1-CCS-SOD1 complex that can be dissociated upon SOD1 activation [33]. CCS expression is regulated by cellular Cu content because when Cu levels decrease, CCS increases, while when Cu content increases, this chaperone is degraded [34]. The SOD family of proteins is critical in the defense against oxidative stress because they catalyze the degradation of superoxide radicals into hydrogen peroxide and oxygen [35]. There are several isoforms of SOD, of which SOD1 (intracellular dimeric) and SOD3 (extracellular tetrameric) contain Cu, whereas SOD2 is a mitochondrial enzyme that contains Mn. In addition, ATOX1 is responsible for transferring Cu to ATP7A and ATP7B, which are membrane pumps characterized by eight TMDs, including multiple Cu binding sites located mainly on TMD6, TMD7, and TMD8 (Figure 1C) [36,37]. These ATPases are located in the trans-Golgi network (TGN), in endocytic vesicles, or in the plasma membrane, pumping Cu+ from ATOX1 to the other side of the membrane [38]. The central role of ATOX1 is reflected in the perinatal death of Atox1 knockout mice due to the altered Cu balance [39].
Since free Cu ions have the potential to generate reactive oxygen species (ROS) in cells, excess intracellular Cu+ must be sequestered by molecules such as MTs and glutathione (GSH) (Figure 1D). MTs are a family of low-molecular-weight proteins capable of binding excess Cu+ ions through thiol groups [29]. In humans, four distinct MTs are known: MT1, MT2, MT3, and MT4. MT1 and MT2 are widely expressed throughout the body, while MT3 and MT4 are principally expressed in the central nervous system [29]. Glutathione is a tripeptide containing glutamate, cysteine, and glycine residues that is also capable of buffering excess Cu. It is probably the first acceptor of Cu as soon as it enters the cell [40,41]. Millimolar cytoplasmic GSH concentrations are estimated to markedly exceed Cu levels [42]. This fact enables GSH to act as a cytosolic Cu buffer that prevents the rise of free Cu ions and drives CTR1-mediated Cu influx by maintaining a negative concentration gradient at the plasma membrane [40]. GSH and other molecules with thiol groups, together with the enzyme, glutaredoxin 1, may generate a reducing environment conducive to the redox regulation of ATP7A and ATP7B, modulating the binding of Cu to cysteine residues, being fundamental for the export of the metal [41].

1.1.3. Copper Excretion

After being stored, Cu can be released into the bloodstream for subsequent distribution to specific tissues and organs [12,24]. This occurs through several pathways, where ATP7A and ATP7B are the central players. These ATPases have a dual role in the cell; first, they have a biosynthetic function because they promote the synthesis of enzymes loaded with Cu (cuproenzymes) in the TGN, such as Cp, LOX, and tyrosinase, which are then secreted out of the cells (Figure 1C) [23]. Cp is the main transport medium for Cu in the circulatory system; therefore, the abundance of Cp in plasma may serve as a biological marker of systemic concentration of this metal [43,44]. In addition, ATP7A and ATP7B have a homeostatic function because when the cellular concentration of Cu increases, they move within endocytic vesicles toward the plasma membrane to transfer excess Cu out of the cell (Figure 1E) [45]. In hepatocytes, ATP7B ensures the movement of Cu through the canalicular membrane for its subsequent elimination through the bile so that any overload is excreted through the digestive tract [38]. Although biliary excretion is the main form of endogenous Cu excretion, there are other routes for Cu elimination, such as urine, sweat, and menstruation [24].

1.2. Copper Homeostasis

Although the redox activity of Cu is essential for enzymatic reactions, this property also makes it potentially toxic at high levels [12]. During the change between Cu+ and Cu2+ states, electron transfer results in the generation of ROS, including superoxide anion (O2−), nitric oxide (NO), hydroxyl radical (OH), and hydrogen peroxide (H2O2), via the Fenton reaction [46]. ROS can attack bio-membranes, destabilizing their structure and affecting their cellular functions, and can also oxidize proteins and denature DNA and RNA, altering the repair mechanisms of these nucleic acids [47]. All of these changes may contribute to the development of cancer, neurodegenerative diseases, and cellular aging [48]. In contrast, a deficiency in Cu can lead to alterations in energy levels, glucose and cholesterol metabolism, and immune cell function, increasing the risks of infections and cardiovascular disorders [44,49,50]. The activities of SOD1, Cp, catalase, and glutathione peroxidase, as well as MT and GSH, are also compromised by an imbalance in the levels of Cu [44]. The dual roles of Cu as an essential and toxic element require specific regulatory mechanisms to prevent both deficiency and accumulation since dyshomeostasis can promote the development of multiple diseases, affecting liver function, lipid metabolism, the central nervous system, and resistance to chemotherapy, among others [51].
Copper homeostasis is highly regulated by transcriptional control and selective transport mechanisms [47]. High levels of cellular Cu negatively regulate the concentration of mammalian CTR1 at the plasma membrane, which trigger CTR1 removal via endocytosis-dependent internalization or degradation (Figure 1F) [52]. In contrast, when Cu concentration is reduced, internalized CTR1 returns to the plasma membrane [53]. In vitro studies have shown that transcription of the Ctr1 gene is regulated by the transcription factor Sp1 (Specificity protein 1) in a Cu-dependent manner, where overload produces a negative regulation of Ctr1 [54]. In vivo, mice fed a Cu-deficient diet had increased CTR1 expression in the intestine [55]. CTR1 function can also be regulated via the generation of a truncated protein (tCTR1) through the removal of its high-affinity Cu-binding domain [56]. tCTR1 is produced within endosomal compartments, has lower uptake activity than CTR1, and requires interactions with CTR2 (copper transporter 2), which is the only other SLC31 family protein in mammals. Initially, CTR2 was proposed as a low-affinity Cu transporter; however, it is currently believed that CTR2 has lost the ability to transport Cu and that its primary role is to produce tCTR1 [57].
Another protein involved in Cu homeostasis is the ATOX1 chaperone, which can act as a transcription factor stimulated by Cu, translocating to the nucleus to bind to promoters of genes that encode cyclin D1, the organizer of the nicotinamide adenine dinucleotide phosphate (NADPH) oxidase p47phox, and SOD3 (Figure 1G) [58,59,60]. It has also been reported that high concentrations of cellular Cu can improve MT gene transcription, mediated by metal-regulatory transcription factor 1 (MTF1) and nuclear factor erythroid 2-related factor 2 (Nrf2) [61,62].
Regulating the localization and function of ATP7A and ATP7B is essential in controlling Cu export from the cell [23]. At physiological levels of Cu, these transporters pump Cu from the cytosol into the lumen of the TGN to load Cu+ into cuproenzymes, which mediate the transport of Cu through the circulatory system [7]. When intracellular Cu increases, ATPases move to the post-Golgi vesicular compartments, which are loaded with Cu, and release this metal into the extracellular medium after fusion with the plasma membrane [38,45]. After Cu levels are restored to physiological levels, ATP7A and ATP7B are transported back to the TGN through the action of several protein complexes, such as AP-1, Arp2/3, WASH, and COMMD/CCDC22/CCDC93 [63].

1.3. Copper and Pathogenesis

The participation of Cu in both the development and progression of diseases has been documented in numerous reports that show an alteration of Cu homeostasis with aberrant levels of this metal. Mutations in the genes encoding ATP7A and ATP7B cause inherited disorders of Cu metabolism, known as Menkes’ disease and Wilson’s disease, respectively [64,65]. Menkes’ disease is an X-linked recessive disorder, fatal to male infants, in which the dysfunction of ATP7A leads to reduced Cu availability in tissues, causing growth retardation, hypotonia, kinky, brittle hair (pili torti), deterioration of the nervous system, and severe intellectual disability [64]. Wilson’s disease is an autosomal recessive disorder characterized by a profound accumulation of Cu, primarily in the liver, brain, and kidneys, due to mutations in the ATP7B gene that impair the ability to excrete Cu into the bile. This triggers hepatic and neuropsychiatric symptoms in these patients [65,66].
In addition to the genetic disorders described above, Cu dyshomeostasis has been associated with a large number of diseases, namely neurodegenerative disorders, such as Alzheimer’s, Parkinson’s, and Huntington’s diseases, and amyotrophic lateral sclerosis [67,68], as well as atherosclerosis [69], diabetes mellitus [70], and cancer [10,71,72]. Recent studies have demonstrated a strong correlation between Cu and three fundamental processes for tumor progression: cell proliferation, angiogenesis, and metastasis [71,72]. Cu also has a role in oxidative stress and chronic inflammation, which promote cell transformation [35,73]. Furthermore, gene expression analysis has revealed multiple alterations in Cu-sensitive or Cu-binding proteins [74], which indicate a relationship between Cu dyshomeostasis and cancer pathogenesis. Therefore, it has been proposed that an important risk factor for carcinogenesis could be elevated levels of Cu in tissues or serum [47,75]. Preclinical studies demonstrated that daily administration of CuSO4 through drinking water significantly increased tumor growth in a murine model of breast cancer [76]. In conjunction with these results, elevated Cu levels in serum and malignant tissues have been documented in different human cancers, including breast, gastrointestinal, and gynecological malignancies [77,78,79,80]. While Cu elevation in cancer cells may be involved in carcinogenesis, it could also be a feature of the cancer phenotype for two main reasons. Tumors, especially fast-growing ones, have greater metabolic demands than healthy tissues that do not divide [47]. As Cu is a cofactor for multiple enzymes in cellular energy metabolism, such as CCO, and in antioxidant defenses, such as SOD [47,77], the demand for Cu could increase in cancer cells. Second, in tissues undergoing hypoxia, upregulation of CTR1 has been observed [81]. Hypoxia-inducible factor 1-alpha (HIF-1α) may activate the transcription of genes related to Cu metabolism (e.g., those that control CTR1), contributing to higher Cu levels in hypoxic tumor cells [82].
Due to the popularity of copper intrauterine devices (Cu IUDs) as a contraceptive method interfering with fertilization and/or implantation [83], it would be interesting to evaluate whether their use could modify serum Cu levels. In a recent review [84], eight of twelve studies analyzed found that Cu IUDs would not change the serum concentrations of this metal. Although the in situ release of Cu ions is very low, research is limited, and there is no clear evidence. Since Cu IUD users generally experience abnormal uterine bleeding or abdominal pain and have not shown clinical signs of toxicity, it is believed that Cu homeostasis mechanisms may be sufficient to prevent the accumulation of this metal. In an animal study, Wistar rats were implanted with Cu IUDs of different doses, with no toxic effects evident [85]. Because they were exposed to much higher Cu levels than those used in humans, the results are reassuring about the Cu IUDs’ safety; however, more research is needed.
Although Cu is involved in a spectrum of diseases, its role in cancer has been the most studied, permitting an analysis of how this metal can affect different cellular processes related to tumor progression. This will be described in the following subsections, and subsequently, our focus will be to evaluate the crosstalk between Cu deregulation and gynecological diseases, which mainly include benign and malignant tumors. The mechanisms that occur in both types of tumors may be similar, where the starting point is abnormal cell proliferation [86]. Finally, we will focus on new Cu-based therapeutic strategies, especially for those gynecological diseases with high prevalence, morbidity, and mortality that do not respond adequately to other treatments.

1.3.1. Copper and Cell Proliferation

Cuproplasia is defined as Cu-dependent cell growth and proliferation that can lead to neoplasia and hyperplasia [10]. This process is related to mitochondrial respiration, redox signaling, autophagy, antioxidant defense, and kinase signaling and may involve enzymatic and non-enzymatic Cu activities [10]. It has been observed that CCS can promote carcinogenesis. For example, in patients with breast cancer, the levels of CCS were increased along with the ability of CCS to promote proliferation through the MAPK/ERK pathway [87]. It was also shown that a specific inhibitor of CCS and ATOX1 reduced cancer cell proliferation and tumor growth [88]. Another emerging concept, metalloallostery, has expanded knowledge about the contributions of Cu to cellular signaling events since it proposes a new paradigm in which the dynamic binding of Cu occurs at sites other than the active sites of proteins to regulate them [89]. In the context of positive metalloallostery, Cu directly binds to MEK1 and MEK2 kinases and enhances their ability to phosphorylate ERK1 and ERK2 in a dose-dependent manner, stimulating the RAF–MEK–ERK signaling cascade, and ultimately, further promoting tumor proliferation [90]; this makes Cu an attractive target as this signaling cascade is one of the best-defined axes that promote cell proliferation and it is abnormal in most human cancers.
Autophagy is a cellular degradation process that plays an essential role in the development and differentiation of cells, constituting a means to cope with intracellular and environmental stress and potentially promoting tumor progression [91]. Recent studies have shown that increased intracellular Cu promotes the growth and survival of cancer cells by activating autophagy, stimulating the autophagic kinases ULK1 and ULK2 [91,92].

1.3.2. Copper and Angiogenesis

One of the main processes involved in tumor growth is angiogenesis, where vascular endothelial cells migrate, proliferate, and differentiate to create a network of new blood vessels extending from surrounding vessels into the expanding tumor [93]. Angiogenesis is regulated by angiogenic-stimulating factors (angiogenin, vascular endothelial growth factor [VEGF], fibroblast growth factor [FGF], transforming growth factor beta [TGF-β]), and interleukins (IL-1, IL-6, IL-8), as well as through inhibitors (angiostatin and endostatin). The role of Cu as a pro-angiogenic metal was first proposed in 1980, with the discovery that Cu salts induce endothelial cell migration, an early step in angiogenesis [94]. Cu may be involved in the entire angiogenic signaling cascade, promoting the growth and mobility of vascular endothelial cells, regulating the synthesis and secretion of the main pro-angiogenic mediators (VEGF and FGF), and directly binding to angiogenin to modulate its affinity for endothelial cells [95]. Cu-dependent activation of HIF-1α transcriptional activity requires interaction with CCS, inducing the expression of pro-angiogenic genes [96]. Cu chelation has been shown to block HIF-1α-mediated VEGF expression [96,97] and to suppress the transcriptional activity of nuclear factor kappa B (NF-κB), thereby inhibiting the expression of FGF, VEGF, IL-1, IL-8, and IL-6 [98,99]. Overexpression of Cu-dependent SOD1 markedly increases VEGF production, while a reduction in SOD1 activity induces vascular abnormalities and impairs angiogenesis [100]. Additionally, Cu ions increase NO production, an inducer of vascular dilation, by activating endothelial nitric oxide synthase [95].
Copper transporters and chaperones also participate in angiogenesis. Upon stimulation by VEGF, cysteine 189 of the cytoplasmic C-terminal domain of CTR1 is sulfenylated, leading to the formation of a disulfide bond between CTR1-VEGFR2 and its co-internalization into early endosomes, promoting angiogenesis via VEGFR2 [101]. Indeed, silencing CTR1 expression in Cu-treated endothelial cells inhibits tube formation and reduces VEGF expression [102]. Regarding ATOX1, since it can act as a transcription factor for NADPH oxidase, it causes inflammatory neovascularization [58]. ATOX1 may also stimulate cyclin D1 in a Cu-dependent manner [59], potentially contributing to cancer cell proliferation and angiogenesis. In this regard, depletion of ATOX1 inhibits vascular smooth muscle cell migration stimulated by platelet-derived growth factor (PDGF), supporting a role for ATOX1 in vascular remodeling and tumor angiogenesis [103]. There may also be a role for ATP7A, as it can limit the degradation of VEGFR2, thereby promoting angiogenesis [104]. In summary, blocking Cu-dependent angiogenesis is an interesting strategy that can be further explored to inhibit tumor growth.

1.3.3. Copper and Metastasis

Processes such as the development of pre-metastatic niches, escape from immune defenses, and angiogenesis will advance and sustain cancer progression. Cu and its binding proteins are involved in the metastatic spread of tumors [105], playing a critical role in the metastatic cascade, both within cells and in the tumor microenvironment. Cu participates in the epithelial–mesenchymal transition (EMT), an early step of metastasis, conferring migratory and invasive capabilities to the cancer cells [71,106,107]. In EMT, molecular reprogramming occurs, deactivating the expression of genes that encode epithelial markers, such as E-cadherin and occludin, and activating mesenchymal genes, such as N-cadherin and vimentin, which are targets of several transcription factors (Snail, Twist, Slug) [106]. The participation of Cu in the remodeling of the extracellular matrix (ECM) and the establishment of a pre-metastatic niche occurs mainly through the activity of LOX and Cu-dependent LOX-like (LOXL) proteins [108,109], which catalyze the cross-linking of collagen and elastin in the ECM. When LOX is active, it stimulates transcription via Twist to promote EMT in the tumor environment [110], and increased expression of LOXL2 correlates with metastasis and poor survival in breast cancer patients [111]. Both angiogenesis and metastasis were suppressed with LOX inhibitors during carcinogenesis examined in vivo, and the decrease in LOX expression inhibited cell migration and neovascular formation in tumor endothelial cells [112].
Adaptation to microenvironmental stressors such as hypoxia is an early characteristic of growing tumors, where HIF-1α plays a key role [113]. Cu and CCS activate HIF-1α by regulating binding to hypoxia-response elements (HREs), promoting the transcription of the target genes involved in EMT [96]. Indeed, Cu depletion in a tumor cell line inhibited the cellular characteristics of hypoxia-induced EMT by downregulating the expression of vimentin and fibronectin genes, which are under the control of the HIF1-α/Snail/Twist signaling pathway, and Cu depletion also inhibited angiogenesis in a mouse model [107]. HIF-1α also induces the expression of LOX, which promotes the synthesis of the HIF-1α protein upon activation of the PI3K/Akt pathway. Therefore, the synergistic action and regulation of both proteins results in the promotion of tumor progression [113,114]. The mediator of the cell motility 1 (MEMO1) protein was identified as a pro-metastatic mediator in breast cancer, where it acts as a Cu-dependent redox protein that promotes a more oxidized intracellular environment through the production of ROS [115]. MEMO1 is thought to have a metal-binding pocket similar to that of metal-dependent redox enzymes, where Cu can be coordinated to favor ROS production [115].

2. Copper in Gynecological Diseases

Over the years, investigating the role of Cu has gained increasing importance, and researchers have joined forces in trying to understand its action. As we previously described, Cu is a crucial element involved in each step of cancer development, from tumorigenesis to metastasis, and there is a large amount of research on the role of Cu in various types of cancer. However, to date, there are very few studies on the specific role of Cu in gynecological diseases [73,77,116]. These diseases mainly include benign and malignant tumors and endocrine diseases [117]. We will evaluate the impact of Cu dyshomeostasis in these diseases, focusing on therapeutic strategies based on altering the role of Cu.

2.1. Ovarian Diseases

2.1.1. Ovarian Cancer

Gynecological cancers include all cancers that affect the female reproductive organs, including endometrial cancer, cervical cancer, ovarian cancer (OC), fallopian tube cancer, vaginal cancer, and vulvar cancer. Among the different gynecological cancers, OC is the most lethal worldwide [118,119]. More than 20 microscopically distinct types of OC can be identified, which are mainly classified into three groups: (1) epithelial cancers, (2) germ cell tumors, and (3) specialized stromal cell cancers [117]. Although significant progress has been made in early detection and treatment, OC is usually detected at a late stage and has a poor prognosis [120]. The overall 5-year survival rate for epithelial OC (EOC), which comprises about 90% of ovarian malignancies, is approximately 30% [119]. In addition to genetic and reproductive risk factors, it has been postulated that chronic inflammation, oxidative stress, and damage caused by free radicals to epithelial cells play a fundamental role in ovarian carcinogenesis [121]. EOC cells form spheroids to avoid immune detection and to resist cell death, where communication between these cells and the peritoneal ecosystem plays a crucial role in the progression and dissemination of the disease [122].
Elevated Cu levels have been reported in the serum of patients with OC [123,124,125], and it is also elevated in OC tumors [80], possibly due to alterations in trace elements with a reduced catabolism or an increase in the neoplastic synthesis of Cp, since elevated levels of both Cu and Cp have been found in patients with OC [123]. A meta-analysis demonstrated not only an increase in circulating Cu concentration but also a decrease in Zn levels in patients diagnosed with OC [125]. In another study, Cu levels were found to be elevated in patients with OC or endometrioma compared to the control groups [124]. However, a meta-analysis showed that using any type of IUD, including Cu IUDs, was associated with a lower incidence of OC [126]. A recent bioinformatics study demonstrated that analyzing the prognostic signature of Cu metabolism-related genes (CMRGs) could provide a useful predictive biomarker and a potential therapeutic target for patients with OC [127]. Additionally, the study showed that CMRGs help define the immune environment, which could serve to identify specific patient subgroups to receive specialized treatment.
The first-line treatments for OC are cytoreductive surgery and platinum-based chemotherapy [128]. Although the response rate is high, most patients typically experience relapses within 2 to 3 years [128]. At first relapse, 25% of patients have platinum resistance or refractory disease with a poor prognosis [129,130]. Numerous studies have identified the transport mechanisms of platinum-containing drugs [131], where it has been observed that many of the proteins involved also participate in Cu homeostasis. Reduced CTR1 expression has been related to cisplatin (CDDP) resistance in patients with OC [132], and higher CTR1 expression has been associated with a better response to CDDP treatment and favorable overall survival [133]. However, ATP7A and ATP7B are necessary to confer resistance to CDDP, carboplatin, and oxaliplatin in OC cell lines [131,134]. ATP7A- and ATP7B-dependent chemoresistance is linked to the impaired accumulation of CDDP in the nucleus and, consequently, the decreased formation of platinum-DNA adducts [131]. Other studies have proposed that CDDP binds to the Cu binding site of ATOX1 and is then transferred to ATP7B, promoting CDDP resistance [135,136]. However, the knockout of ATOX1 did not affect the acquisition of resistance to CDDP, indicating that other mechanisms are involved [137].

2.1.2. Polycystic Ovary Syndrome

Polycystic ovary syndrome (PCOS) is an endocrine and metabolic disorder that occurs in approximately 6% to 20% of women of reproductive age and is a leading cause of infertility [138]. According to a large community-based cohort study, 72% of PCOS patients were infertile compared to 16% of the control group [139]. This disease is characterized by menstrual disorders, polycystic ovaries, and phenotypes related to hyperandrogenism, such as acne, alopecia, and hirsutism [140,141], in addition to a higher risk of spontaneous abortion and pregnancy-related complications [142]. PCOS is associated with obesity, dyslipidemia, insulin resistance (IR), type 2 diabetes mellitus, cardiovascular diseases, and endometrial cancer [143,144,145,146,147]. Regarding the etiology of PCOS, increasing evidence suggests that it could be a multifactorial and polygenic disorder with considerable epigenetic and environmental implications, including dietary and lifestyle factors [148,149,150].
The role of Cu in PCOS is complex and may vary with the phenotype. Several studies have found elevated Cu levels in patients with PCOS [151,152,153,154,155,156,157,158,159,160], while others have found no differences from the control group [161,162,163]. Considering metabolic factors, a significant increase in serum Cu levels was found in both obese and non-obese patients with PCOS compared to healthy subjects [153], and this increase was linked to IR [152,164]. Consequently, controlling Cu in these patients has been recommended as a potential strategy to lower oxidative stress and IR that could be caused by this metal and to minimize long-term metabolic complications [154]. Another study confirmed that patients with PCOS and IR had higher Cu levels than those without IR; however, Cu levels were lower in patients with PCOS than in the control group [165], similar to another work [166]. When Cu levels were measured in the follicular fluid, concentrations were higher in patients with PCOS than in controls [155], and this increase could negatively affect the development of follicles and be related to anomalies in steroidogenesis. Consistently, other investigators found that dietary intake of Cu was positively correlated with the risk of PCOS, and that this metal altered ovarian steroidogenesis, affecting ovarian follicle development [167], promoting premature follicular atresia, and inhibiting follicular maturation and the formation of multiple follicles.

2.2. Uterine Diseases

2.2.1. Uterine Cervix Cancer

Uterine cervix cancer or cervical cancer (CC) is the fourth most common cancer in women worldwide, particularly in developing countries, making it a significant health problem [118]. High-risk human papillomavirus (HPV) infection is considered responsible for more than 90% of CC cases [168], so the prevalence varies depending on the prevalence of HPV infection [169]. Immunization against this virus can help prevent CC, and HPV testing is essential for early CC detection [170,171]. The overall 5-year survival rate is close to 66%; however, as treatment options are limited, patients with metastatic or recurrent disease have a lower survival rate [120]. Aside from HPV infection, some of the most relevant factors for the pathogenesis of CC are inflammation of the epithelium, elevated levels of lipid peroxides, reduced levels of non-enzymatic antioxidants, and altered activities of antioxidant enzymes [172]. One study found that Cu IUD use is not associated with cervical neoplasia [173], and patients have a lower risk of high-grade cervical lesions than oral contraceptive users [174].
High levels of Cu have been found in most studies in patients with CC. An early investigation observed a higher tissue concentration of Cu and a higher Cu/Zn ratio in patients with CC, along with a decrease in Zn levels compared to the control group [175], these results being confirmed in a later study [176]. A meta-analysis recently documented the association between increased serum Cu levels and CC risk [177], and subsequently, this association was confirmed for cervical and endometrial cancers as well as OC. CC patients had the highest Cu concentrations [178], and this increase was positively correlated with the stages of the disease, while Cu decreased after different treatments (surgery, chemotherapy, radiotherapy, or a combination of both) [179]. This result differs from a study where the authors observed that increased serum Cu levels were not modified after chemoradiotherapy in patients with CC [180]. In summary, Cu is indicated as a possible risk factor associated with CC that could be useful to monitor this type of cancer and potentially to control the progress of the disease [177,178,179].

2.2.2. Endometrial Cancer

Uterine cancer or endometrial cancer (EC) is the fifteenth most common cancer in general and the sixth most common cancer in women [118]. Risk factors for the development of EC are obesity, high levels of estrogen, low levels of progesterone, PCOS, IR, diabetes, and estrogen-secreting ovarian tumors [181]. Most patients with early-stage disease have a good prognosis; however, the 5-year overall survival rate for advanced EC is 47% to 69% in stage III and 15% to 17% in stage IV [182]. Specific serum markers have not been established for clinical use in patients with EC. Regarding Cu, there is limited research on EC. A recent analysis evaluated the serum concentrations of Cu and Zn in patients with EC, finding lower levels of these metals compared to the control group [183]. In turn, patients with a greater degree of myometrial invasion had lower Cu levels than those with less myometrial invasion. In contrast, one study found elevated Cu levels without alteration in Zn levels [184], and others reported no changes in tissue Cu in patients with EC [80] or serum Cu in patients with OC and EC [185]. Other investigators found higher mean Cu levels in the serum of EC patients, but the results were not statistically significant. However, the authors observed that the menopausal status and body mass index of the patients were risk factors for EC, which may be affected by Cu concentrations [186]. Furthermore, in a review, the authors found that Cu IUD use could reduce the risk of EC, but the mechanism of action is unclear [187]. It is evident that the results obtained over the years have contradicted each other; therefore, more studies evaluating Cu levels are required to determine the possible clinical relevance in patients with EC.

2.2.3. ‘Benign’ Diseases

Benign neoplasms have received less attention than malignant tumors, probably due to the biased view that ‘malignant’ is life-threatening and ‘benign’ has little effect. While this may be true, benign tumors can put pressure on vital organs, disrupt hormonal balance, and become malignant over time. Although both types of tumors have marked differences (for example, the ability to metastasize), they can be very similar at a mechanistic level, starting from abnormal cell proliferation [86].
Benign uterine diseases are common gynecological disorders in women of reproductive age, and this category includes endometrial polyps [188], uterine leiomyomas [189], and endometriosis [190], among others. Symptoms range from dysmenorrhea and irregular uterine bleeding to the risk of infertility. Associated factors are age, diet, lifestyle, pregnancy, abortion, and hormone use [188,189,190,191]. Endometrial polyps are common endothelial tumors that cause abnormal uterine bleeding and comprise endometrial glands, stroma, blood vessels, and fibrous tissue [188,192]. Although most are benign, malignant transformation has been observed in some cases [192], and estrogen and progesterone play an important role in their pathogenesis, controlling their growth and development [193,194]. Uterine leiomyomas (also called fibroids or myomas) are benign monoclonal neoplasms of the myometrium and represent the most common pelvic tumors in women, affecting more than 70% worldwide [189]. There are three cell populations in fibroids: well-differentiated cells, intermediately differentiated cells, and stem cells, which are believed to be the origin of fibroids [195]. Fibroid-initiating stem cells are more prevalent in women of Afro-American descent and lower in Caucasian women [196]. Searching the relationship between Cu and these uterine diseases revealed few articles on the subject. One study of patients diagnosed with polyps, fibroids, or other benign uterine diseases reported a significant increase in serum Cu levels compared to healthy women [197]. Another showed that Cu levels were higher in patients with uterine fibroids compared to the control group and significantly higher in patients with CC compared to those with fibroids [175]. In patients with endometrial polyps, no differences were found in serum Cu levels compared to the control group [198], but the Cu/Zn ratio was statistically higher, so the authors suggested that oxidative stress would play a role in the pathogenesis of endometrial polyps. By comparing different gynecological diseases, the lowest serum Cu values were found in patients with endometrial polyps and highest in patients with EC, along with elevated Zn levels in uterine fibroids [186].
Endometriosis (EDT) is an estrogen-dependent disease characterized by endometrial-like tissue growing outside the uterine cavity. EDT is considered a chronic and systemic disease, affecting 5–10% of reproductive-age patients in the world [190]. Although EDT is not cancer in itself, it presents similar characteristics: progressive and invasive growth, recurrence, ability to develop its own blood supply, and tendency to metastasize [199]; therefore, it is interesting to know whether Cu has some relationship with this pathology. The first studies reported elevated Cu levels in serum and urine samples from patients with EDT [200,201], which were associated with oxidative stress [200,202]. In patients with advanced-stage EDT [200], a positive correlation was found between Cu and the total oxidant status and between Cu and the oxidative stress index. In another work carried out in animals with induced EDT, the authors demonstrated that elevated Cu levels were positively correlated with the volume of endometriotic-like lesions, high nitrite levels in peritoneal fluid, and increased catalase and glutathione peroxidase activity [202]. In endometriotic lesions, SOD1 has also been found to have increased activity compared to controls [203], which is important for tumor formation [204]. Cu could also stimulate the main signaling pathways of cell proliferation in EDT [205], contributing to malignant transformation within this pathology. Considering that EDT is an estrogen-dependent disease, it is interesting to highlight that Cu is capable of modulating steroidogenesis. It has been observed that, at low levels, this metal can decrease the concentration of estradiol precursor hormones [206], while at high levels, it promotes the expression of enzymes related to the synthesis of this estrogen [207]. We found that the surgical establishment of EDT in mice increased the concentrations of Cu and estradiol, and the administration of a Cu chelating drug decreased both concentrations to values similar to the group with placebo surgery [208]. We also found similar results in another study [209], in which elevated Cu and estradiol levels were efficiently reduced by a Cu chelator in a murine model deficient in tumor necrosis factor (TNF-α) receptor 1 (TNFR1), which presents an aggravated state of EDT [202,210,211].

3. Therapeutic Strategies

The recognition that Cu can have a crucial role in disease pathogenesis has led to the development of therapeutic strategies designed to modulate Cu transport and concentrations; to date, the main focus has been on strategies to treat different types of cancer [72,212,213]. Currently, there is significant interest in strategies to mitigate Cu dyshomeostasis, and hence, the attractiveness of applying these therapeutic strategies to the gynecological diseases analyzed in this review. We will focus on the two main strategies, namely Cu chelators, which decrease the bioavailability, and Cu ionophores, which increase intracellular Cu levels. We will subsequently analyze new strategies related to nanotechnology and plant-derived natural compounds, which have been gaining ground as potential treatments for gynecological diseases.

3.1. Copper Chelators

A chelator is a chemical compound capable of selectively binding to a particular metal atom or ion through a coordination bond, forming a stable structure [214]. The mechanism of action of Cu chelators involves binding to the metal with subsequent excretion of Cu to inhibit cuproplasia. Historically, Cu chelators were developed to treat Wilson’s disease [66], and the most representative examples are D-penicillamine, trientine, and tetrathiomolybdate (Table 2). While the first two have been used clinically for Wilson’s disease for many years, tetrathiomolybdate is a more recent addition—it has been approved in Europe but is still undergoing clinical trials in the US [65]. Considering the critical role of Cu in cancer progression, different authors investigated whether Cu chelators could serve as an antitumor strategy in animal models and clinical trials [77], promoting the emergence of many reports with interesting results. Cu chelation therapy is promising, not only because of its effectiveness but because these agents have the ability to act selectively on malignant tumors, exerting little toxicity on normal cells [47,79].

3.1.1. D-penicillamine

D-penicillamine is a byproduct derived from penicillin that, in addition to binding Cu with great strength, has the ability to chelate other divalent cations such as Ni, Zn, and Pb [77]. The mechanism of action is based on the chelation of Cu2+ ions with the subsequent formation of a stable complex that is excreted through the urine [215]. D-penicillamine is commonly used to treat Wilson’s disease; however, it has been associated with severe toxicity due to numerous adverse effects, such as dystonia, hypersensitivity, pancytopenia, fever, renal failure, congestive heart failure, and tremor, among others [216]. This drug can also be used for cystinuria, rheumatoid arthritis, and heavy metal poisoning [216]. It was shown that Cu chelation by D-penicillamine inhibited neovascularization and human endothelial cell proliferation, affecting angiogenesis [217] and decreasing tumor growth [218]. It also inhibited LOX activity and reduced VEGF expression, causing deficient collagen cross-link formation and delaying tumor progression [219]. In a recent study, the authors observed that treatment with D-penicillamine (but not trientine) caused the inhibition of cell proliferation and EMT by affecting TGF-β/Smad signaling in glioblastoma cells [220]. In oxaliplatin-resistant cervical cancer cells, the combination of D-penicillamine with oxaliplatin or CDDP had a synergistic lethal effect, promoting a greater formation of platinum-DNA adducts, with an increase in the expression of CTR1 and a decrease in ATP7A through the transcription factor Sp1 [221]. Clinical trials with D-penicillamine have been developed for Wilson’s disease, rheumatoid arthritis, cystinuria, and brain tumors.

3.1.2. Trientine

Due to the severe side effects induced by D-penicillamine treatment, triethylenetetramine or trientine was introduced. This drug has a lower Cu-chelating capacity and better tolerability than D-penicillamine and is indicated in those patients with Wilson’s disease who do not tolerate D-penicillamine [222]. Trientine has a polyamine structure that chelates Cu through a stable ring, promoting cupriuresis [222]. The risk of neurological deterioration with trientine is similar to that of D-penicillamine, which usually resolves by reducing the dose [223]. Other adverse events are headache, anemia, arthralgia, rash, and gastrointestinal upset. Trientine has been investigated as a potential anticancer agent. It suppressed tumor development in mice [224] and in hepatocellular carcinoma cell lines [225] and reduced tumor growth in a murine fibrosarcoma model [226]. It also inhibited tumor angiogenesis by decreasing endothelial cell proliferation and expression of CD31 [224] and IL-8 [225]. In addition, trientine is an inhibitor of telomerase [227], an essential factor for cell immortalization that is expressed in most human cancers [228]. Because Cu chelation has been shown to enhance platinum uptake by tumor cells, a small clinical trial combined trientine with carboplatin and pegylated liposomal doxorubicin for the treatment of OC, fallopian tube cancer, and recurrent peritoneal cancer refractory to platinum therapy (ClinicalTrials.gov ID: NCT03480750, Table 3) [229]. The results showed that the combination was safe, but antitumor activity was modest, with no correlation between the clinical response and Cu or Cp levels. This finding was inconclusive, possibly due to the small sample size or the potential influence of ethnic distribution [229].

3.1.3. Tetrathiomolybdate

Another highly specific and widely studied Cu chelator is tetrathiomolybdate, particularly ammonium tetrathiomolybdate (TM), which is rapidly absorbed and has a good safety profile. The first indication of its Cu-binding capacity was the recognition that ruminant animals fed Mo-rich grasses developed a Cu deficiency syndrome (tear disease) [230]. The initial report suggested the administration of molybdates to treat Wilson’s disease; however, a subsequent study showed no clinical benefit in patients [231]. Unlike ruminants, in which rumen cellulose disulfide reacts with Mo, the human gastric mucosa cannot reduce molybdate to the form that can bind Cu [232]. Eventually TM, a reduced form of molybdate, was introduced to diminish Cu levels in humans. If Cu levels are normal, TM is converted to molybdate, incapable of binding Cu, and is excreted via the urine. In the presence of excessive levels of Cu, TM interferes with Cu absorption at the intestinal level when taken with meals and, between meals, forms a stable tripartite complex with serum albumin and circulating Cu to promote biliary excretion, reducing excessive levels of the metal [100,222]. The side effects can be anemia, leukopenia, and increased transaminases, which are easily reversed by reducing the daily dose of TM [233]. Despite the low toxicity of TM, its clinical use is somewhat limited due to the instability of ammonium with oxygen, so a more stable and pharmacologically equivalent TM derivative, bis-choline tetrathiomolybdate (ALXN1840), is also available and is being investigated for the therapy of Wilson’s disease.
Ammonium tetrathiomolybdate has also been shown to reduce tumor growth and function as an effective antiangiogenic agent in both preclinical studies and clinical trials in cancer [98,234,235,236,237,238]. TM can (a) suppress the transcriptional activity of NF-κB, which in turn decreases the expression of angiogenic factors, such as VEGF, FGF, IL-1α, IL-8 [99], (b) induce the degradation of HIF-1α and therefore, reduce the expression of pro-angiogenic factors [97], and (c) suppress Cu chaperone proteins, inhibiting the delivery of Cu to cuproenzymes such as LOX [100,239]. Among these enzymes, inhibition of SOD1 is one of the main therapeutic targets of TM, producing antiangiogenic and antiproliferative effects [240]. Recently, other studies have suggested that TM-induced Cu depletion inhibits MEK1/2 kinase activity, suppressing BRAFV600E-driven tumorigenesis [90,241,242]. In research carried out with OC and EC cell lines, it was found that treatment with TM decreased the protein levels of HIF-1α by mediating its degradation independently of Akt signaling, affecting VEGF levels [97]. It was also observed that trientine or D-penicillamine does not decrease HIF-1α, even at a concentration three times higher than that used with TM [97]. If high Cu levels are reduced, it is possible to sensitize cells to chemotherapy and radiotherapy; therefore, combining these treatments with TM is of interest [237,243,244,245,246]. Research confirmed that Cu depletion sensitized OC cells to therapy with mitomycin C, fenretinide, and 5-fluorouracil by increasing ROS production and inducing DNA damage [237]. TM treatment also improved the efficacy of CDDP in EC and OC cells [245], exerting an antiproliferative effect. TM also enhanced the cytotoxic effects of doxorubicin in EC and OC cells by increasing ROS levels and inducing apoptosis [237,246]. A recent study evaluated the combined effect of TM and lenvatinib (a VEGFR inhibitor) in a model of hepatocellular carcinoma [247]. Tumor burden was positively correlated with Cu concentration, and TM in combination with lenvatinib suppressed tumor growth and angiogenesis to a greater extent than either drug alone, indicating the potential value of this combination as an anticancer treatment.
There is currently no cure for endometriosis, so it is necessary to investigate new treatments that allow the control of EDT progression [248]. Due to the demonstrated implication of Cu in the progression of EDT, our research group first investigated TM as a potential therapy [208,209] and found that it was highly effective in a model with induced EDT. TM decreased the size of the lesions and reduced the elevated levels of Cu and estradiol to physiological levels, along with antiproliferative and antiangiogenic effects [208]. Observing these promising results, we investigated the therapeutic potential of TM in a TNFR1-deficient murine model with induced EDT, which presents an aggravation of the pathology [209]. TM inhibited the EDT progression in the deficient mice, notably affecting cell proliferation, angiogenesis, and oxidative stress while restoring the levels of Cu and estradiol, which are higher in this aggravated version of EDT [209]. TNF-α secretion can be regulated by Cu [212], and several studies have reported the crucial role played by TNF-α, TNFR1, and TNF-α receptor 2 (TNFR2) in EDT [202,210,249,250]. Cell survival, cell proliferation, and cell death occur as a balance between the TNFR1 and TNFR2 signaling pathways, demonstrating the significant crosstalk between them [251,252]. Without TNFR1 expression, TNFR2-dependent pathways that promote tumor progression become relevant [251]. We found that TM also decreased the expression of Tnfr2 [209], and this is important since blocking TNFR2 has been shown to reduce tumor growth [253] and EDT development [254]. Table 3 shows some of the clinical trials where the effectiveness of Cu chelators in gynecological diseases has been evaluated. As can be observed, despite the promising preclinical results, TM has not yet been investigated as a single or combination therapy in gynecological diseases. The last active clinical trial is a Phase 2 study in breast cancer (ClinicalTrials. gov ID: NCT00195091, Table 3), where it has so far shown that patients with triple-negative breast cancer were more responsive to TM treatment than patients with other breast cancer subtypes [238]. ALXN1840 has only been tested for the therapy of Wilson’s disease (Table 3).
Table 3. Examples of clinical trials on drugs related to Cu. Information was obtained from the public database (http://www.clinicaltrials.gov/), accessed on 11 November 2023.
Table 3. Examples of clinical trials on drugs related to Cu. Information was obtained from the public database (http://www.clinicaltrials.gov/), accessed on 11 November 2023.
DiseaseTrial PhaseInterventionTrial IDStatusStudy
Completion
Breast CancerPhase 2TMNCT00195091Active, not
recruiting
06/2025
Wilson’s DiseasePhase 2ALXN1840NCT04422431Completed05/2023
EOC, TC, PPCPhase 1–2Trientine 2HC
+ PLD + carboplatin
NCT03480750Completed12/2019
Advanced cancersPhase 1Trientine 4HC
+ carboplatin
NCT01178112 Completed08/2014
EOC, TC, PPCPhase 2Elesclomol + paclitaxelNCT00888615Completed08/2016
CCPhase 264CuII(atsm)NCT00794339Terminated12/2011
CINPhase 2CurcuminNCT04266275Not yet
recruiting
03/2025
CCPhase 1–2Curcumin
+ radiotherapy
NCT05947513Not yet
recruiting
11/2024
CCPhase 2CurcuminNCT04294836Withdrawn12/2023
EDTPhase 2CurcuminNCT04493476Unknown
Status
12/2022
CC, ECPhase 2Pembrolizumab
+ radiation + curcumin + immune modulatory cocktail
NCT03192059Completed06/2021
ECPhase 2CurcuminNCT02017353Completed10/2016

3.2. Copper Ionophores

Unlike the sequestering nature of Cu chelators, Cu ionophores transport this metal into cells, forcing an increase in the intracellular Cu concentration and exerting cytotoxic effects through different pathways [10,255]. Examples of Cu ionophores are disulfiram, clioquinol, elesclomol, and bis(thiosemicarbazone) analogs [mainly CuII(atsm) and CuII(gtsm)]. Several years ago, it was determined that tumor cells were more sensitive to elevated levels of ROS than were normal cells [256]. Despite the promoting effects of Cu on tumor progression, inducing Cu accumulation within cancer cells could promote ROS elevation to take advantage of ROS toxicity as a potential antitumor therapy [257]. Cuproptosis is a specific type of cell death recently postulated by Tsvetkov et al. [258], which is triggered by the accumulation of intracellular Cu. The authors showed the ability of Cu to bind to lipoylated proteins of the tricarboxylic acid (TCA) cycle, promoting increased mitochondrial energy metabolism and toxicity stress, which ultimately causes cell death. The mode of action of Cu ionophores is believed to be interaction with DNA, inhibition of the proteasome, and the ability to displace other metals from the binding sites on critical proteins [259,260].

3.2.1. Disulfiram and Dithiocarbamates

The best-known dithiocarbamates are pyrrolidine dithiocarbamate and diethyldithiocarbamate, the active form of disulfiram (DSF). DSF has been used for many years to treat alcohol dependence since it inhibits the enzyme, aldehyde dehydrogenase (ALDH) [261], and the first evidence of its effectiveness in cancer was in 1977 when it was used in an alcoholic patient with metastatic breast cancer who received DSF and went into spontaneous remission [262]. Since then, its possible use as an anticancer agent has gained interest [263,264,265,266]. DSF has been shown to inhibit cell proliferation, migration, and invasion by altering the nuclear translocation of NF-κB and the expression of Smad4 [267]. This downregulates proteins such as Snail and Slug, inhibiting EMT and hindering tumor metastasis. In OC cells, DSF also inhibits ALDH [268,269], which has been related to poor prognosis because it promotes resistance to therapy, the maintenance of cancer stem cells, and the mitigation of oxidative stress [270,271]. DSF also prevents the growth of endometriotic lesions by reducing angiogenesis, cell proliferation, and NF-κB expression. In an animal model of endometriosis, DSF increased the serum concentration of malondialdehyde (a marker of lipid peroxidation) and lowered the total antioxidants, TNF-α, and IL-1β compared to the control group [272]. DSF also enhances the anticancer activity of chemotherapeutic drugs, such as CDDP and temozolomide [266,273], which is why DSF is often used in combination therapy. Beneficial effects have been observed in OC, where the combination of DSF with docosahexaenoic acid (DHA) [274] and PARP inhibitors [275] suppressed tumor growth, improving drug sensitivity. In these studies, DSF ameliorated DHA-induced oxidative stress by upregulating Nrf2-mediated HO-1 (heme oxygenase 1) gene transcription [274] and inhibiting the expression of genes associated with DNA damage repair [275]. In turn, it was demonstrated in chemoresistant OC cells that DSF combined with CDDP synergistically inhibited tumor growth, possibly promoting the downregulation of Smad3 [276]. By adding Cu, it is possible to enhance the DSF activity in some cases by the DSF/Cu complex formation [266,277,278,279]. Evidence has shown that the main targets of DSF/Cu may be the levels of ROS, the ubiquitin–proteasome system, and NF-κB [263,265,273,280]. DSF/Cu preferentially targets cancer cells and cancer stem cells rather than normal cells [281,282,283,284]. An example of this was observed by Xu et al., where DSF/Cu was cytotoxic in a dose-dependent manner for leukemia stem cells without affecting normal hematopoietic progenitor cells [282]. In another study, DSF increased Cu absorption in cancer cells, with an increase in Cu redox reactions, promoting oxidative stress [285]. In human osteosarcoma cells, DSF/Cu reduced cell growth by autophagy and apoptosis in a ROS-dependent manner with the implication of the ROS/JNK pathway [286], similar to the effects observed in CC cell lines [287]. Although several dithiocarbamates and their derivatives have demonstrated Cu-dependent anticancer activity [288], and promising preclinical results have been observed with DSF, clinical studies in cancer patients have not been successful. When DSF/Cu was administered as monotherapy, it did not produce significant benefits in patients with solid tumors, probably due to insufficient bioavailability of DSF and its metabolite in blood [273].

3.2.2. Clioquinol

The best-known derivative of the 8-hydroxyquinoline class of drugs is clioquinol, which was initially synthesized as an antimicrobial agent for shigellosis and intestinal amebiasis [289]. It has subsequently been studied in different diseases ranging from neurodegenerative disorders to cancer [280]. The first study to evaluate clioquinol as an antitumor agent showed that it decreased viability by inducing apoptosis in eight different cancer cell lines and prevented the growth of OC xenografts in mice [290], with similar results in prostate cancer cells and xenografts [291]. Another study reported that a different OC cell line was sensitive to the combination of clioquinol and DHA, with toxicity mediated by the action of PPARα [292]. Similar to DSF, the anticancer activity of clioquinol is enhanced by Cu and has been linked to proteasome inhibition and oxidative stress [290,293,294,295]. One of the targets of clioquinol is the X-linked inhibitor of an apoptosis protein (XIAP), which modulates caspase activity, allowing selective action with apoptosis being only triggered in cancer cells [293] and an insignificant effect in normal cells. Clioquinol increases the tissue content of Cu2+, indicating that the clioquinol–Cu2+ complex could be the metabolite that triggers the death of cancer cells, and it could be formed intracellularly or extracellularly and transported into the cells [296]. Clioquinol can trigger autophagy by inducing LC3 lipidation and autophagosome formation in myeloma and leukemia cells [295]. It can also exacerbate the anticancer activity of macrophages toward CC cells, promoting the secretion of interleukins and cytokines, such as TNF-α [297]. Although clioquinol has shown selective promise in cancer chemotherapy, it has also caused serious neurotoxicity that led to its clinical prohibition [298]. Various routes of administration or combination with other drugs for safer application are still being investigated [255,299]. Further derivatives of 8-hydroxyquinoline, such as PBT2 and nitroxoline, might have greater effectiveness as anticancer agents by inhibiting the proliferation of cancer cells with fewer side effects [300,301]; nevertheless, they have not yet been tested in gynecological diseases.

3.2.3. Elesclomol and Derivatives

Elesclomol is a carbohydrazide, bis(thio-hydrazide amide), developed from a parent molecule, which had anticancer activity but was chemically and metabolically unstable [302]. Elesclomol is stable and causes a 10-fold increase in cancer cell cytotoxicity compared to the parent molecule. It induces oxidative stress in cancer cells [303,304,305,306] and alters mitochondrial metabolism, particularly the TCA cycle, promoting cuproptosis [258,307]. The anticancer activity is due to the formation of an elesclomol–Cu2+ complex [308] that facilitates transport into the mitochondria, where reduction to Cu+ leads to oxidative stress and subsequent cell death [305]. Using CRISPR-Cas9 deletion, the mitochondrial protein ferredoxin-1 (FDX1) was shown to bind to the elesclomol–Cu2+ complex, reducing Cu2+ to Cu+ and promoting the anticancer activity of this ionophore [307]. In a mouse model, treatment with elesclomol–Cu2+ inhibited the development of endometriosis through FDX1-mediated cuproptosis [309].
Inactivating mutations in the AT-rich interactive domain-containing protein 1A (ARID1A) are found more frequently in gynecological cancers [310], and in 14 gynecological cancer cell lines, loss of ARID1A caused increased levels of ROS. Elesclomol inhibited tumor growth and induced apoptosis in these ARID1A mutant cells [303]. In another in vitro study, elesclomol with anisomycin inhibited the proliferation of OC stem cells, while elesclomol alone was ineffective [311]. In an OC relapse model, both disulfiram and elesclomol promoted cell death following treatment with carboplatin compared to carboplatin alone [312]. Although these laboratory studies have been promising, when elesclomol was administered in clinical trials as monotherapy or in combination with other chemotherapeutics for different types of tumors [313,314,315], the benefit has been small or negligible. In a phase II clinical study, elesclomol with paclitaxel was used as a treatment for cisplatin-resistant OC, fallopian tube cancer, and peritoneal cancer [313]. Although this combination showed a good safety profile, it did not produce the expected response, possibly because elesclomol is not effective at elevated levels of the enzyme, lactate dehydrogenase (LDH) [313,316], suggesting that elesclomol may be less effective in situations with a high rate of glycolysis. Hypoxia has been associated with more aggressive tumors that have elevated LDH levels. Elesclomol is more effective in non-hypoxic conditions because it interferes with metabolic processes in oxygenated tumor cells [305]. For more information, in a recent review, special attention is paid to elesclomol as an anticancer therapy [317].

3.2.4. Bis(thiosemicarbazones)

Thiosemicarbazones and bis(thiosemicarbazones) are capable of binding to metals, forming stable, lipophilic, and often neutral complexes [318]. Diacetyl-bis-(N4-methylthiosemicarbazonato)-copper(II) [CuII(atsm)] and glyoxal-bis-(N4-methylthiosemicarbazonato)-copper(II) [CuII(gtsm)] have a similar structure, but differences in their redox behavior [319]. Due to the elevated Cu levels in cancer, these ionophoric Cu compounds have been investigated to determine if they could selectively treat tumor cells without altering normal cells. A study on a TRAMP (transgenic adenocarcinoma of the mouse prostate) model documented CuII(gtsm) selectivity for cancer cells with high Cu levels [260]. CuII(gtsm) increased ROS in TRAMP cells along with decreased GSH but did not do so in normal mouse prostate epithelial cells. In another study investigating CuII(atsm) and CuII(gtsm) as anticancer agents [320], CuII(gtsm) was cytotoxic against prostate cancer cells and significantly reduced the tumor burden, while the CuII(atsm) action was insignificant. It is important to note that CuII(gtsm) dissociates upon entering the cell, increasing the intracellular bioavailability of Cu and causing toxicity, while the ligand (H2gtsm) is recycled out of the cell and re-enters with more re-coordinated Cu [320]. This property explains how increasing extracellular Cu improves the anticancer activity of CuII(gtsm), which could be applicable in patients with elevated serum Cu levels. In contrast, CuII(atsm) retains Cu due to its lower reduction potential in intracellular reducing environments [318,321]. CuII(atsm) is selective toward cells with low oxygen levels since a more forced-reducing environment (such as hypoxia) leads to the reduction of CuII(atsm) and its dissociation [319,322], as demonstrated in hypoxic neuroblastoma cells where CuII(atsm) caused higher intracellular Cu levels compared to control cells [321]. As a result of this characteristic, radiolabeled Cu complexes have been synthesized that are theranostic, i.e., they allow simultaneous imaging diagnosis and therapy [323], especially with the 64Cu isotope [324]. In one of the first studies, 64CuII(atsm) demonstrated anticancer activity as a radiotherapy agent in a hamster colon cancer model, increasing survival time without toxic effects [325]. Several studies in cancer patients have been performed to evaluate survival concerning the uptake of these isotopes. When 60CuII(atsm) was used as a marker of hypoxia, higher uptake predicted a worse prognosis in patients with CC [326]. In another study of CC, 60CuII(atsm) promoted the overexpression of VEGF, cyclooxygenase-2, epidermal growth factor receptor (EGFR), carbonic anhydrase 9 (CA-9), along with an increase in cell death [327]. In a comparative study, 64CuII(atsm) was shown to be more effective than 60CuII(atsm) in obtaining better-quality images for patients with CC [328]. Cu ionophores may offer great selectivity toward cancer cells with antitumor activity against different cancer types but, to date, most preclinical results have not been replicated in patient trials, reflecting the need to better understand the action mechanism and pharmacokinetics of these compounds [47].

3.3. New Therapeutic Strategies

Recognizing that alterations in Cu homeostasis are involved in the pathogenesis of various diseases and the potential value of Cu-based therapies has prompted the development of new compounds based on Cu [329]. Nanotechnological strategies and natural plant-derived compounds have been gaining ground as potential treatments for gynecological diseases.

3.3.1. Cu-Based Nanoparticles

Nano-oncology involves the use of nanotechnological strategies for cancer treatment. In this sense, nanoparticles (NPs) can function directly as an antitumor treatment or as a vehicle to mediate the controlled administration of drugs to increase their effectiveness and decrease their side effects [330]. Cu-based NPs (CuNPs) form a stable structure with a diameter of 10–50 nanometers, and they are used in a variety of industrial processes that release them into the environment. CuNPs pass through wastewater treatment plants into water systems and enter vegetation through the agricultural use of fertilizers and pesticides [331], but current levels of environmental exposure have not been linked to disease pathogenesis. Due to their high surface-to-volume ratio, CuNPs can interact efficiently with tissues, an attractive characteristic for use in oncology. The Cu-induced toxicity of CuNPs is related to oxidative damage through increasing ROS, the formation of peroxy radicals, lipid peroxidation, and reduction in CCO activity [331]. The production of NPs can be accomplished via ‘green synthesis’ using plants, algae, and microorganisms, which is presumably an environmentally friendly process.
Copper-based nanoparticles have been investigated as antitumor agents in several types of cancer [332,333,334,335]. In OC cell lines, CuNPs synthesized from a Camellia sinensis leaf extract were effective in causing tumor cell death [332]. In CC lines, CuNPs synthesized from a pumpkin seed extract caused a decrease in cell viability, increased production of ROS, apoptosis induction, and the suppression of cell migration with the antitumor effect linked to inhibition of the PI3K/Akt signaling pathway [335]. Similarly, other CuNPs synthesized using an extract of Houttuynia cordata were effective as antitumor agents in CC cells [336]. The development of CuNP–transferrin loaded with doxorubicin also successfully inhibited tumor growth in mice [337], and these NPs were able to specifically enter CC and breast cancer cell lines that overexpressed the transferrin receptor. CuNPs synthesized from the red alga Pterocladia capillacea and loaded with nedaplatin improved the antitumor activity in OC compared to treatment with nedaplatin alone [338]. There have recently been reports on an innovative and promising strategy to improve the precision of cancer treatment by using NPs in photothermal therapy. Copper sulfide (CuS) NPs target tumor cells and enter the nucleus; subsequent near-infrared laser irradiation activates the NPs to increase the temperature within the nucleus, leading to apoptosis of the tumor cell. The main goal is to target both the primary tumor and malignant cells that have escaped, thereby minimizing metastasis. Initial studies in mice have demonstrated that photothermal therapy was effective and safe in eliminating residual CC cells and preventing tumor recurrence [339]. Another investigation showed that CuS NPs, together with laser irradiation, effectively killed tumor cells in mouse models of OC with a minimal effect on surrounding healthy tissue [340].

3.3.2. Natural Compounds Derived from Plants

In recent years, attempts have been made to identify natural molecules that can be used in oncology. Although these compounds are considered to act as antioxidants, the objective is to have them work as pro-oxidants in the presence of Cu, catalyzing ROS formation and DNA degradation. In this regard, several plant-derived Cu-binding molecules have been reported to exert anticancer effects and increase the antitumor activity of other known chemotherapeutics with low side effects.

Curcumin

Curcumin, a bioactive turmeric polyphenol derived from the rhizomes of Curcuma longa, chelates Cu with a wide range of biological effects, including antioxidant, anti-inflammatory, and antimicrobial properties when examined in a variety of laboratory models [341]. It may also have protective effects against different types of cancer, including lung cancer, breast cancer, colon cancer, and gynecological cancers [342,343,344], but when given orally, it is poorly absorbed and rapidly inactivated, limiting the potential for clinical use. Strategies to improve the pharmacokinetics have included the creation of curcumin-metal NP, and a curcumin–Cu complex was shown to have higher anticancer activity compared to curcumin alone [345,346]. In an in vitro study with EC cells, curcumin treatment suppressed tumor growth, inhibited cell proliferation, and promoted ROS-induced apoptosis [347]. It also attenuated cell migration by increasing the expression of the Slit2 protein, causing the downregulation of SDF-1 (stromal cell-derived factor 1) and CXCR4 (C-X-C motif chemokine receptor 4) and, therefore, of MMP (matrix metalloproteinase) 2 and 9. The decrease in the expression of MMPs, with implications for invasion, migration [348], and cell proliferation [349], has been documented by other investigators. In CC cells, curcumin suppresses proliferation and invasion by affecting the Wnt/β-catenin and NF-κB pathways [350] and elevates intracellular ROS levels but not in healthy epithelial cells, leading to cell-specific apoptosis [351]. Regarding OC, curcumin has shown great anticancer potential because it suppresses cell cycle progression, promotes apoptosis and autophagy, and inhibits tumor metastasis, so current efforts are focused on finding suitable derivatives to overcome the pharmacokinetic limitations (reviewed in detail by Liu et al. [352]).
Curcumin has also been studied in animal models of PCOS, where it (a) reduces testosterone levels and increases estrogen levels [353], (b) promotes an anti-inflammatory mechanism by reducing proteins involved [354], (c) improves ovarian function [354,355], (d) improves the levels of total cholesterol, HDL, LDL, and triglycerides [353,355], and (e) decreases malondialdehyde levels and increases the activities of SOD, catalase, and GSH [353,355], among other effects. However, the results in clinical trials are discrepant, probably due to the inclusion and exclusion criteria used and the number of participants. For example, in a clinical trial in patients with PCOS, curcumin reduced serum insulin, fasting glucose, and the index of insulin resistance (HOMA-IR) [356], while no differences were found in another clinical trial [357]. Regarding EDT, in an in vitro study, curcumin induced a lower expression of ICAM-1, VCAM-1, IL-6, IL-8, and MCP-1 by inhibiting the activation of NF-κB induced by TNF-α without affecting the viability of endometriotic stromal cells [358]. However, another study showed that the number of endometriotic lesions, their volume, and the degree of adhesions, along with the levels of IL-1β, IL-6, HIF-1α, and VEGF, were reduced in mice treated with curcumin compared with the control group [359]. Reduced secretion of pro-angiogenic chemokines and pro-inflammatory cytokines, upregulation of IL-10 and IL-12, and abrogation of IKKα/β, NF-κB, STAT3, and JNK signaling pathways have been demonstrated in eutopic endometrial stromal cells from patients with EDT treated with curcumin [360]. Curcumin reduces cell survival, VEGF expression, and cell proliferation in endometriotic cells [361] and lowers estradiol levels that are elevated in EDT [362]. In addition, curcumin may decrease EDT by promoting apoptosis through p53-dependent and -independent mitochondrial pathways [363]. Patients with EDT receiving a combination of quercetin, turmeric, and N-acetylcysteine reported a reduction in pain and lower use of non-steroidal anti-inflammatory drugs (NSAIDs) [364]. The role of curcumin and other plant-derived compounds as potential treatments for EDT has been reviewed in detail by Meresman et al. [365]. Table 3 shows some clinical trials that use or have used curcumin alone or with other treatments.

Coumarins

Coumarins are found in plants, such as Rutaceae and Umbelliferae, and belong to the benzo-α-pyrone family. These compounds have anti-inflammatory, antioxidant, and antitumor activities [366]. A coumarin–Cu complex [367,368] and a coumarin–Cu–thiosemicarbazone hybrid have been effective antiproliferative agents in cell lines of different types of cancer [369], and a coumarin–amide–Cu complex was shown to have greater antitumor capacity than CDDP in a breast cancer cell line [370]. Two studies in OC cells have demonstrated the antitumor effect of two natural derivatives of coumarin, 4-methylumbelliferone, and Osthole. These compounds reduce cell proliferation by affecting the PI3K/Akt and MAPK pathways [371] or induce several cell death mechanisms [372]. Osthole has also been tested in CC cells, where it reduces cell viability, proliferation, migration, and invasion, along with inducing apoptosis [373]. The combination of Osthole with CDDP reduced cell proliferation and enhanced apoptosis in CC cells to a greater extent than CDDP alone, notably downregulating the PI3K/Akt pathway [374]. Other coumarin derivatives have similar effects on this pathway in CC cells [375,376]. Imperatorin, a furanocoumarin derivative, was effective in an animal model of EDT [377], significantly inhibiting the growth of ectopic endometrium, improving the histopathological characteristics, and inhibiting the PI3K/Akt/NF-κB pathway. Auraptene, a coumarin derivative found in citrus fruits, decreased the inflammation and elevated the fertilization rate in isolated oocytes in a mouse model of PCOS [378]. The drug lowered ROS levels and elevated intracellular GSH levels, indicating that auraptene could be a potential candidate to improve oocyte maturation and fertilization capacity in patients with PCOS [378]. This has subsequently been confirmed in a mouse model of in vitro fertilization and early embryo development [379].

4. Concluding Remarks

Gynecological diseases are characterized by high prevalence, morbidity, and mortality and it is essential to investigate the pathogenesis and possible diagnostic and therapeutic strategies for these disorders. Over the years, the importance of Cu in health and disease has been increasingly recognized, and research on Cu has gained prominence, with extensive efforts to document and understand its complex roles and diverse mechanisms of action. Although Cu is crucial for many physiological functions, it is also potentially toxic at altered levels, and specific regulatory mechanisms normally prevent Cu dyshomeostasis. Documenting these mechanisms and the alterations that can occur has revealed that Cu has a critical role in the pathogenesis of several diseases, particularly cancer, where Cu is involved in every step from tumorigenesis to metastasis. With time, different therapeutic options based on Cu have emerged for a variety of disorders with promising results, both in animal models and in clinical trials. Some strategies are based on reducing high levels of Cu with chelators to slow the progression of specific gynecological diseases. Other drugs, such as Cu ionophores, can force Cu into cells to take full advantage of its toxic role and induce tumor cell death. Notably, these two categories of drugs mediate opposite actions: Cu chelators inhibit cuproplasia, while Cu ionophores induce cuproptosis. Given the attractiveness of altering Cu levels as a therapeutic strategy, the need to continue investigating these types of drugs is evident, and this will also require further understanding of the pathogenesis of each disorder and the potential role of Cu dyshomeostasis. Ongoing research is an essential stage in the discovery of more effective treatments to target specific genes and influence distinct signaling pathways. The development of new Cu-based compounds holds great promise to revolutionize diagnostic and therapeutic strategies, especially for those gynecologic diseases with high mortality.

Author Contributions

R.A.C. drafted the manuscript, compiled the figure and tables, and participated in the final editing of the manuscript. M.B.D., E.Z., C.M.T., and M.C. contributed with a critical revision, writing, and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from Universidad Nacional de San Luis (UNSL), Argentina [PROICO 02-0720], from the Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET), Argentina [PIP 2021-2023/00969], and the Agencia Nacional de Promoción Científica y Tecnológica (ANPCyT), Argentina [PICT-2020-SERIEA-03793] (to M.C.). This research was also funded by Health Canada to Ovarian Cancer Canada in support of the OvCAN research initiative (to C.M.T.) and a Merk grant (to E.Z.). R.A.C. was the recipient of a scholarship from the Emerging Leaders in the Americas Program (ELAP), provided with the support of the Government of Canada. This work is part of the Doctoral thesis of R.A.C.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AktProtein kinase B
ALDHAldehyde dehydrogenase
ALXN1840Bis-choline tetrathiomolybdate
ARID1AAT-rich interactive domain-containing protein 1A
ArpActin-related proteins
ATOX1Antioxidant chaperone 1
ATPAdenosine triphosphate
ATP7ACopper-transporting ATPase alpha
ATP7BCopper-transporting ATPase beta
BRAFSerine/threonine-protein kinase B-raf
CA-9Carbonic anhydrase 9
Cas9CRISPR-associated protein 9
CCCervical cancer
CCDCCoiled-coil domain containing protein
CCOCytochrome C oxidase
CCSCopper chaperone for superoxide dismutase
CD31Cluster of differentiation 31
CDDPCisplatin
CINCervical intraepithelial neoplasia
CMRGsCopper-metabolism related genes
COMMDCopper metabolism MURR1 domain-containing protein
COX11CCO copper chaperone 11
COX17CCO copper chaperone 17
COX19CCO assembly factor 19
CpCeruloplasmin
CRISPRClustered regularly interspaced short palindromic repeats
CTR1Copper transporter 1
CTR2Copper transporter 2
CuCopper
Cu IUDsCopper intrauterine devices
CuS NPsCopper sulfide nanoparticles
CXCRC-X-C motif chemokine receptor
DβHDopamine-β-hydroxylase
DCYTBDuodenal cytochrome B
DHADocosahexaenoic acid
DMT1Divalent metal transporter 1
DSFDisulfiram
ECEndometrial cancer
ECMExtracellular matrix
EDTEndometriosis
EGFREpidermal growth factor receptor
EMTEpithelial-mesenchymal transition
EOC Epithelial ovarian cancer
ERKExtracellular signal-regulated kinase
FDX1Ferredoxin-1
FGFFibroblast growth factor
GSHGlutathione
2HC Dihydrochloride
4HC Tetrahydrochloride
HIF-1αHypoxia-inducible factor 1-alpha
HO-1Heme oxygenase 1
HOMA-IRHomeostatic model assessment for insulin resistance
HPVHuman papillomavirus
HREsHypoxia response elements
ICAMIntercellular adhesion molecule
IKKsInhibitory kappa B kinases
ILInterleukin
IMSMitochondrial intermembrane space
IRInsulin resistance
JNKc-Jun N-terminal kinase
LC3Microtubule-associated protein light chain 3
LDHLactate dehydrogenase
LOXLysyl oxidase
LOXLLOX-like proteins
MAPKMitogen-activated protein kinase
MCP-1Monocyte chemoattractant protein-1
MEKMitogen-activated protein kinase kinase
MEMO1Mediator of cell motility 1
MMPMatrix metalloproteinase
MTMetallothionein
MTF1Metal-regulatory transcription factor 1
NADPHNicotinamide adenine dinucleotide phosphate
NF-κB Nuclear factor kappa B
NONitric oxide
NPsNanoparticles
Nrf2Nuclear factor erythroid 2-related factor 2
NSAIDNon-steroidal anti-inflammatory drug
OCOvarian cancer
p53Tumor protein p53
PARPPoly (ADP-ribose) polymerase
PBT25,7-dichloro-2-[(dimethylamino)methyl]-8-hydroxyquinoline
PCOSPolycystic ovary syndrome
PDGFPlatelet-derived growth factor
PI3KPhosphoinositide 3-kinase
PLDPegylated liposomal doxorubicin
PPARPeroxisome proliferator-activated receptor
PPC Primary peritoneal cancer
RAFRapidly accelerated fibrosarcoma
ROSReactive oxygen species
SCO1Synthesis of cytochrome C oxidase 1
SDF-1Stromal cell-derived factor 1
SODSuperoxide dismutase
STATSignal transducer and activator of transcription
STEAPSix-transmembrane epithelial antigen of the prostate
TCFallopian tube cancer
TCATricarboxylic acid
TGF-βTransforming growth factor beta
TGNTrans-Golgi network
TMAmmonium tetrathiomolybdate
TMDTransmembrane domain
TNFTumor necrosis factor
TNFRTNF receptor
TRAMPTransgenic adenocarcinoma of the mouse prostate
ULKUnc-51 like autophagy activating kinase
VCAMVascular cell adhesion protein
VEGFVascular endothelial growth factor
VEGFRVEGF receptor
WASHWiskott–Aldrich syndrome protein and SCAR homolog
XIAPX-linked inhibitor of apoptosis protein

References

  1. Nevitt, T.; Öhrvik, H.; Thiele, D.J. Charting the Travels of Copper in Eukaryotes from Yeast to Mammals. Biochim. Biophys. Acta (BBA)-Mol. Cell Res. 2012, 1823, 1580–1593. [Google Scholar] [CrossRef]
  2. Trumbo, P.; Yates, A.A.; Schlicker, S.; Poos, M. Dietary Reference Intakes. J. Am. Diet. Assoc. 2001, 101, 294–301. [Google Scholar] [CrossRef]
  3. Moshfegh, A.J.; Goldman, J.D.; Rhodes, D.G.; Friday, J.E. Usual Nutrient Intake from Food and Beverages, by Gender and Age, What We Eat In America, NHANES 2017-March 2020 Prepandemic. 2023. Available online: www.ars.usda.gov/nea/bhnrc/fsrg (accessed on 11 November 2023).
  4. Myint, Z.W.; Oo, T.H.; Thein, K.Z.; Tun, A.M.; Saeed, H. Copper Deficiency Anemia. Ann. Hematol. 2018, 97, 1527–1534. [Google Scholar] [CrossRef]
  5. Grochowski, C.; Blicharska, E.; Baj, J.; Mierzwińska, A.; Brzozowska, K.; Forma, A.; Maciejewski, R. Serum Iron, Magnesium, Copper, and Manganese Levels in Alcoholism: A Systematic Review. Molecules 2019, 24, 1361. [Google Scholar] [CrossRef]
  6. Feng, Y.; Zeng, J.-W.; Ma, Q.; Zhang, S.; Tang, J.; Feng, J.-F. Serum Copper and Zinc Levels in Breast Cancer: A Meta-Analysis. J. Trace Elem. Med. Biol. 2020, 62, 126629. [Google Scholar] [CrossRef]
  7. Linder, M.C. Ceruloplasmin and Other Copper Binding Components of Blood Plasma and Their Functions: An Update. Metallomics 2016, 8, 887–905. [Google Scholar] [CrossRef]
  8. Tsang, T.; Davis, C.I.; Brady, D.C. Copper Biology. Curr. Biol. 2021, 31, R421–R427. [Google Scholar] [CrossRef]
  9. Shi, H.; Jiang, Y.; Yang, Y.; Peng, Y.; Li, C. Copper Metabolism in Saccharomyces Cerevisiae: An Update. Biometals 2021, 34, 3–14. [Google Scholar] [CrossRef]
  10. Ge, E.J.; Bush, A.I.; Casini, A.; Cobine, P.A.; Cross, J.R.; DeNicola, G.M.; Dou, Q.P.; Franz, K.J.; Gohil, V.M.; Gupta, S.; et al. Connecting Copper and Cancer: From Transition Metal Signalling to Metalloplasia. Nat. Rev. Cancer 2022, 22, 102–113. [Google Scholar] [CrossRef]
  11. Grubman, A.; White, A.R. Copper as a Key Regulator of Cell Signalling Pathways. Expert Rev. Mol. Med. 2014, 16, e11. [Google Scholar] [CrossRef]
  12. Shanbhag, V.C.; Gudekar, N.; Jasmer, K.; Papageorgiou, C.; Singh, K.; Petris, M.J. Copper Metabolism as a Unique Vulnerability in Cancer. Biochim. Biophys. Acta (BBA)-Mol. Cell Res. 2021, 1868, 118893. [Google Scholar] [CrossRef]
  13. Shawki, A.; Anthony, S.R.; Nose, Y.; Engevik, M.A.; Niespodzany, E.J.; Barrientos, T.; Öhrvik, H.; Worrell, R.T.; Thiele, D.J.; Mackenzie, B. Intestinal DMT1 Is Critical for Iron Absorption in the Mouse but Is Not Required for the Absorption of Copper or Manganese. Am. J. Physiol. Liver Physiol. 2015, 309, G635–G647. [Google Scholar] [CrossRef]
  14. Wyman, S.; Simpson, R.J.; McKie, A.T.; Sharp, P.A. Dcytb (Cybrd1) Functions as Both a Ferric and a Cupric Reductase in Vitro. FEBS Lett. 2008, 582, 1901–1906. [Google Scholar] [CrossRef]
  15. Ozumi, K.; Sudhahar, V.; Kim, H.W.; Chen, G.-F.; Kohno, T.; Finney, L.; Vogt, S.; McKinney, R.D.; Ushio-Fukai, M.; Fukai, T. Role of Copper Transport Protein Antioxidant 1 in Angiotensin II–Induced Hypertension: A Key Regulator of Extracellular Superoxide Dismutase. Hypertension 2012, 60, 476–486. [Google Scholar] [CrossRef]
  16. Nose, Y.; Wood, L.K.; Kim, B.-E.; Prohaska, J.R.; Fry, R.S.; Spears, J.W.; Thiele, D.J. Ctr1 Is an Apical Copper Transporter in Mammalian Intestinal Epithelial Cells in Vivo That Is Controlled at the Level of Protein Stability. J. Biol. Chem. 2010, 285, 32385–32392. [Google Scholar] [CrossRef]
  17. Zimnicka, A.M.; Maryon, E.B.; Kaplan, J.H. Human Copper Transporter HCTR1 Mediates Basolateral Uptake of Copper into Enterocytes: Implications for Copper Homeostasis. J. Biol. Chem. 2007, 282, 26471–26480. [Google Scholar] [CrossRef]
  18. Galler, T.; Lebrun, V.; Raibaut, L.; Faller, P.; Wezynfeld, N.E. How Trimerization of CTR1 N-Terminal Model Peptides Tunes Cu-Binding and Redox-Chemistry. Chem. Commun. 2020, 56, 12194–12197. [Google Scholar] [CrossRef]
  19. Schushan, M.; Barkan, Y.; Haliloglu, T.; Ben-Tal, N. Cα-Trace Model of the Transmembrane Domain of Human Copper Transporter 1, Motion and Functional Implications. Proc. Natl. Acad. Sci. USA 2010, 107, 10908–10913. [Google Scholar] [CrossRef]
  20. Nose, Y.; Kim, B.-E.; Thiele, D.J. Ctr1 Drives Intestinal Copper Absorption and Is Essential for Growth, Iron Metabolism, and Neonatal Cardiac Function. Cell Metab. 2006, 4, 235–244. [Google Scholar] [CrossRef]
  21. Kuo, Y.-M.; Zhou, B.; Cosco, D.; Gitschier, J. The Copper Transporter CTR1 Provides an Essential Function in Mammalian Embryonic Development. Proc. Natl. Acad. Sci. USA 2001, 98, 6836–6841. [Google Scholar] [CrossRef]
  22. Lelièvre, P.; Sancey, L.; Coll, J.-L.; Deniaud, A.; Busser, B. The Multifaceted Roles of Copper in Cancer: A Trace Metal Element with Dysregulated Metabolism, but Also a Target or a Bullet for Therapy. Cancers 2020, 12, 3594. [Google Scholar] [CrossRef]
  23. Lutsenko, S.; Barnes, N.L.; Bartee, M.Y.; Dmitriev, O.Y. Function and Regulation of Human Copper-Transporting ATPases. Physiol. Rev. 2007, 87, 1011–1046. [Google Scholar] [CrossRef]
  24. Chen, L.; Min, J.; Wang, F. Copper Homeostasis and Cuproptosis in Health and Disease. Signal Transduct. Target. Ther. 2022, 7, 378. [Google Scholar] [CrossRef]
  25. Ramos, D.; Mar, D.; Ishida, M.; Vargas, R.; Gaite, M.; Montgomery, A.; Linder, M.C. Mechanism of Copper Uptake from Blood Plasma Ceruloplasmin by Mammalian Cells. PLoS ONE 2016, 11, e0149516. [Google Scholar] [CrossRef]
  26. Moriya, M.; Ho, Y.-H.; Grana, A.; Nguyen, L.; Alvarez, A.; Jamil, R.; Ackland, M.L.; Michalczyk, A.; Hamer, P.; Ramos, D.; et al. Copper Is Taken up Efficiently from Albumin and A2-Macroglobulin by Cultured Human Cells by More than One Mechanism. Am. J. Physiol. Physiol. 2008, 295, C708–C721. [Google Scholar] [CrossRef]
  27. Pierson, H.; Yang, H.; Lutsenko, S. Copper Transport and Disease: What Can We Learn from Organoids? Annu. Rev. Nutr. 2019, 39, 75–94. [Google Scholar] [CrossRef]
  28. Heaton, D.N.; George, G.N.; Garrison, G.; Winge, D.R. The Mitochondrial Copper Metallochaperone Cox17 Exists as an Oligomeric, Polycopper Complex. Biochemistry 2001, 40, 743–751. [Google Scholar] [CrossRef] [PubMed]
  29. Calvo, J.; Jung, H.; Meloni, G. Copper Metallothioneins. IUBMB Life 2017, 69, 236–245. [Google Scholar] [CrossRef] [PubMed]
  30. Nývltová, E.; Dietz, J.V.; Seravalli, J.; Khalimonchuk, O.; Barrientos, A. Coordination of Metal Center Biogenesis in Human Cytochrome c Oxidase. Nat. Commun. 2022, 13, 3615. [Google Scholar] [CrossRef] [PubMed]
  31. Horng, Y.-C.; Cobine, P.A.; Maxfield, A.B.; Carr, H.S.; Winge, D.R. Specific Copper Transfer from the Cox17 Metallochaperone to Both Sco1 and Cox11 in the Assembly of Yeast Cytochrome C Oxidase. J. Biol. Chem. 2004, 279, 35334–35340. [Google Scholar] [CrossRef]
  32. Zischka, H.; Einer, C. Mitochondrial Copper Homeostasis and Its Derailment in Wilson Disease. Int. J. Biochem. Cell Biol. 2018, 102, 71–75. [Google Scholar] [CrossRef] [PubMed]
  33. Skopp, A.; Boyd, S.D.; Ullrich, M.S.; Liu, L.; Winkler, D.D. Copper-Zinc Superoxide Dismutase (Sod1) Activation Terminates Interaction between Its Copper Chaperone (Ccs) and the Cytosolic Metal-Binding Domain of the Copper Importer Ctr1. Biometals 2019, 32, 695–705. [Google Scholar] [CrossRef] [PubMed]
  34. Bertinato, J.; L’Abbé, M.R. Copper Modulates the Degradation of Copper Chaperone for Cu, Zn Superoxide Dismutase by the 26 S Proteosome. J. Biol. Chem. 2003, 278, 35071–35078. [Google Scholar] [CrossRef] [PubMed]
  35. Reuter, S.; Gupta, S.C.; Chaturvedi, M.M.; Aggarwal, B.B. Oxidative Stress, Inflammation, and Cancer: How Are They Linked? Free Radic. Biol. Med. 2010, 49, 1603–1616. [Google Scholar] [CrossRef]
  36. Inesi, G.; Pilankatta, R.; Tadini-Buoninsegni, F. Biochemical Characterization of P-Type Copper ATPases. Biochem. J. 2014, 463, 167–176. [Google Scholar] [CrossRef] [PubMed]
  37. Jayakanthan, S.; Braiterman, L.T.; Hasan, N.M.; Unger, V.M.; Lutsenko, S. Human Copper Transporter ATP7B (Wilson Disease Protein) Forms Stable Dimers in Vitro and in Cells. J. Biol. Chem. 2017, 292, 18760–18774. [Google Scholar] [CrossRef]
  38. Polishchuk, E.V.; Concilli, M.; Iacobacci, S.; Chesi, G.; Pastore, N.; Piccolo, P.; Paladino, S.; Baldantoni, D.; van IJzendoorn, S.C.D.; Chan, J.; et al. Wilson Disease Protein ATP7B Utilizes Lysosomal Exocytosis to Maintain Copper Homeostasis. Dev. Cell 2014, 29, 686–700. [Google Scholar] [CrossRef]
  39. Hamza, I.; Prohaska, J.; Gitlin, J.D. Essential Role for Atox1 in the Copper-Mediated Intracellular Trafficking of the Menkes ATPase. Proc. Natl. Acad. Sci. USA 2003, 100, 1215–1220. [Google Scholar] [CrossRef]
  40. Maryon, E.B.; Molloy, S.A.; Kaplan, J.H. Cellular Glutathione Plays a Key Role in Copper Uptake Mediated by Human Copper Transporter 1. Am. J. Physiol. Physiol. 2013, 304, C768–C779. [Google Scholar] [CrossRef]
  41. Singleton, W.C.J.; McInnes, K.T.; Cater, M.A.; Winnall, W.R.; McKirdy, R.; Yu, Y.; Taylor, P.E.; Ke, B.-X.; Richardson, D.R.; Mercer, J.F.B.; et al. Role of Glutaredoxin1 and Glutathione in Regulating the Activity of the Copper-Transporting P-Type ATPases, ATP7A and ATP7B. J. Biol. Chem. 2010, 285, 27111–27121. [Google Scholar] [CrossRef]
  42. Jiang, X.; Chen, J.; Bajić, A.; Zhang, C.; Song, X.; Carroll, S.L.; Cai, Z.-L.; Tang, M.; Xue, M.; Cheng, N.; et al. Quantitative Real-Time Imaging of Glutathione. Nat. Commun. 2017, 8, 16087. [Google Scholar] [CrossRef]
  43. Harvey, L.J.; Ashton, K.; Hooper, L.; Casgrain, A.; Fairweather-Tait, S.J. Methods of Assessment of Copper Status in Humans: A Systematic Review. Am. J. Clin. Nutr. 2009, 89, 2009S–2024S. [Google Scholar] [CrossRef]
  44. Hordyjewska, A.; Popiołek, Ł.; Kocot, J. The Many “Faces” of Copper in Medicine and Treatment. Biometals 2014, 27, 611–621. [Google Scholar] [CrossRef]
  45. Mercer, J.F.B.; Barnes, N.; Stevenson, J.; Strausak, D.; Llanos, R.M. Copper-Induced Trafficking of the Cu-ATPases: A Key Mechanism for Copper Homeostasis. Biometals 2003, 16, 175–184. [Google Scholar] [CrossRef]
  46. Valko, M.; Morris, H.; Cronin, M.T.D. Metals, Toxicity and Oxidative Stress. Curr. Med. Chem. 2005, 12, 1161–1208. [Google Scholar] [CrossRef]
  47. Denoyer, D.; Masaldan, S.; La Fontaine, S.; Cater, M.A. Targeting Copper in Cancer Therapy: “Copper That Cancer”. Metallomics 2015, 7, 1459–1476. [Google Scholar] [CrossRef]
  48. Pizzino, G.; Irrera, N.; Cucinotta, M.; Pallio, G.; Mannino, F.; Arcoraci, V.; Squadrito, F.; Altavilla, D.; Bitto, A. Oxidative Stress: Harms and Benefits for Human Health. Oxid. Med. Cell. Longev. 2017, 2017, 8416763. [Google Scholar] [CrossRef]
  49. Tosco, A.; Fontanella, B.; Danise, R.; Cicatiello, L.; Grober, O.; Ravo, M.; Weisz, A.; Marzullo, L. Molecular Bases of Copper and Iron Deficiency-Associated Dyslipidemia: A Microarray Analysis of the Rat Intestinal Transcriptome. Genes Nutr. 2010, 5, 1–8. [Google Scholar] [CrossRef]
  50. Bonham, M.; O’Connor, J.M.; Hannigan, B.M.; Strain, J.J. The Immune System as a Physiological Indicator of Marginal Copper Status? Br. J. Nutr. 2002, 87, 393–403. [Google Scholar] [CrossRef]
  51. Gaetke, L.M.; Chow-Johnson, H.S.; Chow, C.K. Copper: Toxicological Relevance and Mechanisms. Arch. Toxicol. 2014, 88, 1929–1938. [Google Scholar] [CrossRef] [PubMed]
  52. Molloy, S.A.; Kaplan, J.H. Copper-Dependent Recycling of HCTR1, the Human High Affinity Copper Transporter. J. Biol. Chem. 2009, 284, 29704–29713. [Google Scholar] [CrossRef]
  53. Clifford, R.J.; Maryon, E.B.; Kaplan, J.H. Dynamic Internalization and Recycling of a Metal Ion Transporter: Cu Homeostasis and CTR1, the Human Cu+ Uptake System. J. Cell Sci. 2016, 129, 1711–1721. [Google Scholar] [CrossRef]
  54. Liang, Z.D.; Tsai, W.-B.; Lee, M.-Y.; Savaraj, N.; Kuo, M.T. Specificity Protein 1 (Sp1) Oscillation Is Involved in Copper Homeostasis Maintenance by Regulating Human High-Affinity Copper Transporter 1 Expression. Mol. Pharmacol. 2012, 81, 455–464. [Google Scholar] [CrossRef]
  55. Kuo, Y.-M.; Gybina, A.A.; Pyatskowit, J.W.; Gitschier, J.; Prohaska, J.R. Copper Transport Protein (Ctr1) Levels in Mice Are Tissue Specific and Dependent on Copper Status. J. Nutr. 2006, 136, 21–26. [Google Scholar] [CrossRef]
  56. Öhrvik, H.; Logeman, B.; Turk, B.; Reinheckel, T.; Thiele, D.J. Cathepsin Protease Controls Copper and Cisplatin Accumulation via Cleavage of the Ctr1 Metal-Binding Ectodomain. J. Biol. Chem. 2016, 291, 13905–13916. [Google Scholar] [CrossRef]
  57. Logeman, B.L.; Wood, L.K.; Lee, J.; Thiele, D.J. Gene Duplication and Neo-Functionalization in the Evolutionary and Functional Divergence of the Metazoan Copper Transporters Ctr1 and Ctr2. J. Biol. Chem. 2017, 292, 11531–11546. [Google Scholar] [CrossRef]
  58. Chen, G.-F.; Sudhahar, V.; Youn, S.-W.; Das, A.; Cho, J.; Kamiya, T.; Urao, N.; McKinney, R.D.; Surenkhuu, B.; Hamakubo, T.; et al. Copper Transport Protein Antioxidant-1 Promotes Inflammatory Neovascularization via Chaperone and Transcription Factor Function. Sci. Rep. 2015, 5, 14780. [Google Scholar] [CrossRef]
  59. Itoh, S.; Kim, H.W.; Nakagawa, O.; Ozumi, K.; Lessner, S.M.; Aoki, H.; Akram, K.; McKinney, R.D.; Ushio-Fukai, M.; Fukai, T. Novel Role of Antioxidant-1 (Atox1) as a Copper-Dependent Transcription Factor Involved in Cell Proliferation. J. Biol. Chem. 2008, 283, 9157–9167. [Google Scholar] [CrossRef]
  60. Kamiya, T.; Takeuchi, K.; Fukudome, S.; Hara, H.; Adachi, T. Copper Chaperone Antioxidant-1, Atox-1, Is Involved in the Induction of SOD3 in THP-1 Cells. Biometals 2018, 31, 61–68. [Google Scholar] [CrossRef]
  61. Palmiter, R.D. Regulation of Metallothionein Genes by Heavy Metals Appears to Be Mediated by a Zinc-Sensitive Inhibitor That Interacts with a Constitutively Active Transcription Factor, MTF-1. Proc. Natl. Acad. Sci. USA 1994, 91, 1219–1223. [Google Scholar] [CrossRef]
  62. Song, M.O.; Mattie, M.D.; Lee, C.-H.; Freedman, J.H. The Role of Nrf1 and Nrf2 in the Regulation of Copper-Responsive Transcription. Exp. Cell Res. 2014, 322, 39–50. [Google Scholar] [CrossRef]
  63. Hartwig, C.; Zlatic, S.A.; Wallin, M.; Vrailas-Mortimer, A.; Fahrni, C.J.; Faundez, V. Trafficking Mechanisms of P-Type ATPase Copper Transporters. Curr. Opin. Cell Biol. 2019, 59, 24–33. [Google Scholar] [CrossRef]
  64. Ojha, R.; Prasad, A.N. Menkes Disease: What a Multidisciplinary Approach Can Do. J. Multidiscip. Healthc. 2016, 9, 371–385. [Google Scholar] [CrossRef]
  65. Dev, S.; Kruse, R.L.; Hamilton, J.P.; Lutsenko, S. Wilson Disease: Update on Pathophysiology and Treatment. Front. Cell Dev. Biol. 2022, 10, 871877. [Google Scholar] [CrossRef]
  66. Członkowska, A.; Litwin, T.; Dusek, P.; Ferenci, P.; Lutsenko, S.; Medici, V.; Rybakowski, J.K.; Weiss, K.H.; Schilsky, M.L. Wilson Disease. Nat. Rev. Dis. Prim. 2018, 4, 21. [Google Scholar] [CrossRef]
  67. Gromadzka, G.; Tarnacka, B.; Flaga, A.; Adamczyk, A. Copper Dyshomeostasis in Neurodegenerative Diseases—Therapeutic Implications. Int. J. Mol. Sci. 2020, 21, 9259. [Google Scholar] [CrossRef]
  68. Gil-Bea, F.J.; Aldanondo, G.; Lasa-Fernández, H.; de Munain, A.L.; Vallejo-Illarramendi, A. Insights into the Mechanisms of Copper Dyshomeostasis in Amyotrophic Lateral Sclerosis. Expert Rev. Mol. Med. 2017, 19, e7. [Google Scholar] [CrossRef]
  69. Chen, X.; Cai, Q.; Liang, R.; Zhang, D.; Liu, X.; Zhang, M.; Xiong, Y.; Xu, M.; Liu, Q.; Li, P.; et al. Copper Homeostasis and Copper-Induced Cell Death in the Pathogenesis of Cardiovascular Disease and Therapeutic Strategies. Cell Death Dis. 2023, 14, 105. [Google Scholar] [CrossRef]
  70. Pal, I.; Dey, S.G. The Role of Heme and Copper in Alzheimer’s Disease and Type 2 Diabetes Mellitus. JACS Au 2023, 3, 657–681. [Google Scholar] [CrossRef]
  71. Tang, X.; Yan, Z.; Miao, Y.; Ha, W.; Li, Z.; Yang, L.; Mi, D. Copper in Cancer: From Limiting Nutrient to Therapeutic Target. Front. Oncol. 2023, 13, 1209156. [Google Scholar] [CrossRef] [PubMed]
  72. Kong, R.; Sun, G. Targeting Copper Metabolism: A Promising Strategy for Cancer Treatment. Front. Pharmacol. 2023, 14, 1203447. [Google Scholar] [CrossRef]
  73. Michalczyk, K.; Cymbaluk-Płoska, A. The Role of Zinc and Copper in Gynecological Malignancies. Nutrients 2020, 12, 3732. [Google Scholar] [CrossRef]
  74. Barresi, V.; Trovato-Salinaro, A.; Spampinato, G.; Musso, N.; Castorina, S.; Rizzarelli, E.; Condorelli, D.F. Transcriptome Analysis of Copper Homeostasis Genes Reveals Coordinated Upregulation of SLC 31A1, SCO 1, and COX 11 in Colorectal Cancer. FEBS Open Bio. 2016, 6, 794–806. [Google Scholar] [CrossRef]
  75. Mulware, S.J. Comparative Trace Elemental Analysis in Cancerous and Noncancerous Human Tissues Using PIXE. J. Biophys. 2013, 2013, 192026. [Google Scholar] [CrossRef]
  76. Ishida, S.; Andreux, P.; Poitry-Yamate, C.; Auwerx, J.; Hanahan, D. Bioavailable Copper Modulates Oxidative Phosphorylation and Growth of Tumors. Proc. Natl. Acad. Sci. USA 2013, 110, 19507–19512. [Google Scholar] [CrossRef]
  77. Lopez, J.; Ramchandani, D.; Vahdat, L. Copper Depletion as a Therapeutic Strategy in Cancer. Met. Ions Life Sci. 2019, 19, 303–330. [Google Scholar] [CrossRef]
  78. Zowczak, M.; Iskra, M.; Torliński, L.; Cofta, S. Analysis of Serum Copper and Zinc Concentrations in Cancer Patients. Biol. Trace Elem. Res. 2001, 82, 1–8. [Google Scholar] [CrossRef]
  79. Gupte, A.; Mumper, R.J. Elevated Copper and Oxidative Stress in Cancer Cells as a Target for Cancer Treatment. Cancer Treat. Rev. 2009, 35, 32–46. [Google Scholar] [CrossRef] [PubMed]
  80. Yaman, M.; Kaya, G.; Simsek, M. Comparison of Trace Element Concentrations in Cancerous and Noncancerous Human Endometrial and Ovary Tissues. Int. J. Gynecol. Cancer 2007, 17, 220–228. [Google Scholar] [CrossRef] [PubMed]
  81. Zimnicka, A.M.; Tang, H.; Guo, Q.; Kuhr, F.K.; Oh, M.-J.; Wan, J.; Chen, J.; Smith, K.A.; Fraidenburg, D.R.; Choudhury, M.S.R.; et al. Upregulated Copper Transporters in Hypoxia-Induced Pulmonary Hypertension. PLoS ONE 2014, 9, e90544. [Google Scholar] [CrossRef] [PubMed]
  82. Su, Y.; Zhang, X.; Li, S.; Xie, W.; Guo, J. Emerging Roles of the Copper-CTR1 Axis in Tumorigenesis. Mol. Cancer Res. 2022, 20, 1339–1353. [Google Scholar] [CrossRef]
  83. United Nations; Department of Economic and Social Affairs. Population Division. In Contraceptive Use by Method 2019: Data Booklet; United Nations: New York, NY, USA, 2019; ISBN 978-92-1-148329-1. [Google Scholar]
  84. Crandell, L.; Mohler, N. A Literature Review of the Effects of Copper Intrauterine Devices on Blood Copper Levels in Humans. Nurs. Womens Health 2021, 25, 71–81. [Google Scholar] [CrossRef]
  85. Zhao, X.; Liu, Q.; Sun, H.; Hu, Y.; Wang, Z. Chronic Systemic Toxicity Study of Copper Intrauterine Devices in Female Wistar Rats. Med. Sci. Monit. 2017, 23, 3961–3970. [Google Scholar] [CrossRef]
  86. Boutry, J.; Tissot, S.; Ujvari, B.; Capp, J.-P.; Giraudeau, M.; Nedelcu, A.M.; Thomas, F. The Evolution and Ecology of Benign Tumors. Biochim. Biophys. Acta (BBA)-Rev. Cancer 2022, 1877, 188643. [Google Scholar] [CrossRef]
  87. Li, Y.; Liang, R.; Zhang, X.; Wang, J.; Shan, C.; Liu, S.; Li, L.; Zhang, S. Copper Chaperone for Superoxide Dismutase Promotes Breast Cancer Cell Proliferation and Migration via ROS-Mediated MAPK/ERK Signaling. Front. Pharmacol. 2019, 10, 356. [Google Scholar] [CrossRef]
  88. Wang, J.; Luo, C.; Shan, C.; You, Q.; Lu, J.; Elf, S.; Zhou, Y.; Wen, Y.; Vinkenborg, J.L.; Fan, J.; et al. Inhibition of Human Copper Trafficking by a Small Molecule Significantly Attenuates Cancer Cell Proliferation. Nat. Chem. 2015, 7, 968–979. [Google Scholar] [CrossRef]
  89. Pham, V.N.; Chang, C.J. Metalloallostery and Transition Metal Signaling: Bioinorganic Copper Chemistry Beyond Active Sites. Angew. Chemie. 2023, 62, e202213644. [Google Scholar] [CrossRef]
  90. Brady, D.C.; Crowe, M.S.; Greenberg, D.N.; Counter, C.M. Copper Chelation Inhibits BRAFV600E-Driven Melanomagenesis and Counters Resistance to BRAFV600E and MEK1/2 Inhibitors. Cancer Res. 2017, 77, 6240–6252. [Google Scholar] [CrossRef] [PubMed]
  91. Polishchuk, E.V.; Merolla, A.; Lichtmannegger, J.; Romano, A.; Indrieri, A.; Ilyechova, E.Y.; Concilli, M.; De Cegli, R.; Crispino, R.; Mariniello, M.; et al. Activation of Autophagy, Observed in Liver Tissues from Patients with Wilson Disease and from ATP7B-Deficient Animals, Protects Hepatocytes from Copper-Induced Apoptosis. Gastroenterology 2019, 156, 1173–1189. [Google Scholar] [CrossRef] [PubMed]
  92. Tsang, T.; Posimo, J.M.; Gudiel, A.A.; Cicchini, M.; Feldser, D.M.; Brady, D.C. Copper Is an Essential Regulator of the Autophagic Kinases ULK1/2 to Drive Lung Adenocarcinoma. Nat. Cell Biol. 2020, 22, 412–424. [Google Scholar] [CrossRef] [PubMed]
  93. Zhao, Y.; Adjei, A.A. Targeting Angiogenesis in Cancer Therapy: Moving beyond Vascular Endothelial Growth Factor. Oncologist 2015, 20, 660–673. [Google Scholar] [CrossRef] [PubMed]
  94. McAuslan, B.R.; Reilly, W. Endothelial Cell Phagokinesis in Response to Specific Metal Ions. Exp. Cell Res. 1980, 130, 147–157. [Google Scholar] [CrossRef] [PubMed]
  95. Urso, E.; Maffia, M. Behind the Link between Copper and Angiogenesis: Established Mechanisms and an Overview on the Role of Vascular Copper Transport Systems. J. Vasc. Res. 2015, 52, 172–196. [Google Scholar] [CrossRef] [PubMed]
  96. Feng, W.; Ye, F.; Xue, W.; Zhou, Z.; Kang, Y.J. Copper Regulation of Hypoxia-Inducible Factor-1 Activity. Mol. Pharmacol. 2009, 75, 174–182. [Google Scholar] [CrossRef] [PubMed]
  97. Kim, K.K.; Abelman, S.; Yano, N.; Ribeiro, J.R.; Singh, R.K.; Tipping, M.; Moore, R.G. Tetrathiomolybdate Inhibits Mitochondrial Complex IV and Mediates Degradation of Hypoxia-Inducible Factor-1α in Cancer Cells. Sci. Rep. 2015, 5, 14296. [Google Scholar] [CrossRef] [PubMed]
  98. Pan, Q.; Kleer, C.G.; Van Golen, K.L.; Irani, J.; Bottema, K.M.; Bias, C.; De Carvalho, M.; Mesri, E.A.; Robins, D.M.; Dick, R.D.; et al. Copper Deficiency Induced by Tetrathiomolybdate Suppresses Tumor Growth and Angiogenesis. Cancer Res. 2002, 62, 4854–4859. [Google Scholar] [PubMed]
  99. Pan, Q.; Bao, L.W.; Merajver, S.D. Tetrathiomolybdate Inhibits Angiogenesis and Metastasis through Suppression of the NFκB Signaling Cascade. Mol. Cancer Res. 2003, 1, 701–706. [Google Scholar] [PubMed]
  100. Denoyer, D.; Clatworthy, S.A.S.; Cater, M.A. Copper Complexes in Cancer Therapy. In Metal Ions in Life Sciences; Europe PMC: London, UK, 2018; Volume 18, pp. 469–506. ISBN 9783110470734. [Google Scholar]
  101. Das, A.; Ash, D.; Fouda, A.Y.; Sudhahar, V.; Kim, Y.-M.; Hou, Y.; Hudson, F.Z.; Stansfield, B.K.; Caldwell, R.B.; McMenamin, M.; et al. Cysteine Oxidation of Copper Transporter CTR1 Drives VEGFR2 Signalling and Angiogenesis. Nat. Cell Biol. 2022, 24, 35–50. [Google Scholar] [CrossRef]
  102. Narayanan, G.; Vuyyuru, H.; Muthuvel, B.; Konerirajapuram Natrajan, S. CTR1 Silencing Inhibits Angiogenesis by Limiting Copper Entry into Endothelial Cells. PLoS ONE 2013, 8, e71982. [Google Scholar] [CrossRef]
  103. Kohno, T.; Urao, N.; Ashino, T.; Sudhahar, V.; McKinney, R.D.; Hamakubo, T.; Iwanari, H.; Ushio-Fukai, M.; Fukai, T. Novel Role of Copper Transport Protein Antioxidant-1 in Neointimal Formation after Vascular Injury. Arterioscler. Thromb. Vasc. Biol. 2013, 33, 805–813. [Google Scholar] [CrossRef]
  104. Ash, D.; Sudhahar, V.; Youn, S.-W.; Okur, M.N.; Das, A.; O’Bryan, J.P.; McMenamin, M.; Hou, Y.; Kaplan, J.H.; Fukai, T.; et al. The P-Type ATPase Transporter ATP7A Promotes Angiogenesis by Limiting Autophagic Degradation of VEGFR2. Nat. Commun. 2021, 12, 3091. [Google Scholar] [CrossRef] [PubMed]
  105. Fukai, T.; Ushio-Fukai, M.; Kaplan, J.H. Copper Transporters and Copper Chaperones: Roles in Cardiovascular Physiology and Disease. Am. J. Physiol. Physiol. 2018, 315, C186–C201. [Google Scholar] [CrossRef] [PubMed]
  106. Dongre, A.; Weinberg, R.A. New Insights into the Mechanisms of Epithelial–Mesenchymal Transition and Implications for Cancer. Nat. Rev. Mol. Cell Biol. 2019, 20, 69–84. [Google Scholar] [CrossRef] [PubMed]
  107. Li, S.; Zhang, J.; Yang, H.; Wu, C.; Dang, X.; Liu, Y. Copper Depletion Inhibits CoCl2-Induced Aggressive Phenotype of MCF-7 Cells via Downregulation of HIF-1 and Inhibition of Snail/Twist-Mediated Epithelial-Mesenchymal Transition. Sci. Rep. 2015, 5, 12410. [Google Scholar] [CrossRef] [PubMed]
  108. Xiao, Q.; Ge, G. Lysyl Oxidase, Extracellular Matrix Remodeling and Cancer Metastasis. Cancer Microenviron. 2012, 5, 261–273. [Google Scholar] [CrossRef] [PubMed]
  109. Yang, N.; Cao, D.-F.; Yin, X.-X.; Zhou, H.-H.; Mao, X.-Y. Lysyl Oxidases: Emerging Biomarkers and Therapeutic Targets for Various Diseases. Biomed. Pharmacother. 2020, 131, 110791. [Google Scholar] [CrossRef]
  110. El-Haibi, C.P.; Bell, G.W.; Zhang, J.; Collmann, A.Y.; Wood, D.; Scherber, C.M.; Csizmadia, E.; Mariani, O.; Zhu, C.; Campagne, A.; et al. Critical Role for Lysyl Oxidase in Mesenchymal Stem Cell-Driven Breast Cancer Malignancy. Proc. Natl. Acad. Sci. USA 2012, 109, 17460–17465. [Google Scholar] [CrossRef]
  111. Barker, H.E.; Chang, J.; Cox, T.R.; Lang, G.; Bird, D.; Nicolau, M.; Evans, H.R.; Gartland, A.; Erler, J.T. LOXL2-Mediated Matrix Remodeling in Metastasis and Mammary Gland Involution. Cancer Res. 2011, 71, 1561–1572. [Google Scholar] [CrossRef]
  112. Osawa, T.; Ohga, N.; Akiyama, K.; Hida, Y.; Kitayama, K.; Kawamoto, T.; Yamamoto, K.; Maishi, N.; Kondoh, M.; Onodera, Y.; et al. Lysyl Oxidase Secreted by Tumour Endothelial Cells Promotes Angiogenesis and Metastasis. Br. J. Cancer 2013, 109, 2237–2247. [Google Scholar] [CrossRef]
  113. Semenza, G.L. Molecular Mechanisms Mediating Metastasis of Hypoxic Breast Cancer Cells. Trends Mol. Med. 2012, 18, 534–543. [Google Scholar] [CrossRef]
  114. Pez, F.; Dayan, F.; Durivault, J.; Kaniewski, B.; Aimond, G.; Le Provost, G.S.; Deux, B.; Clézardin, P.; Sommer, P.; Pouysségur, J.; et al. The HIF-1-Inducible Lysyl Oxidase Activates HIF-1 via the Akt Pathway in a Positive Regulation Loop and Synergizes with HIF-1 in Promoting Tumor Cell Growth. Cancer Res. 2011, 71, 1647–1657. [Google Scholar] [CrossRef] [PubMed]
  115. MacDonald, G.; Nalvarte, I.; Smirnova, T.; Vecchi, M.; Aceto, N.; Doelemeyer, A.; Frei, A.; Lienhard, S.; Wyckoff, J.; Hess, D.; et al. Memo Is a Copper-Dependent Redox Protein with an Essential Role in Migration and Metastasis. Sci. Signal. 2014, 7, ra56. [Google Scholar] [CrossRef] [PubMed]
  116. Lukanović, D.; Herzog, M.; Kobal, B.; Černe, K. The Contribution of Copper Efflux Transporters ATP7A and ATP7B to Chemoresistance and Personalized Medicine in Ovarian Cancer. Biomed. Pharmacother. 2020, 129, 110401. [Google Scholar] [CrossRef] [PubMed]
  117. Mok, S.C.; Wong, K.K.; Lu, K.H.; Munger, K.; Nagymanyoki, Z. Molecular Basis of Gynecologic Diseases. In Essential Concepts in Molecular Pathology; Elsevier: Amsterdam, The Netherlands, 2020; pp. 409–424. ISBN 9780128132579. [Google Scholar]
  118. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA. Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  119. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer Statistics, 2022. CA. Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef] [PubMed]
  120. Wang, Q.; Peng, H.; Qi, X.; Wu, M.; Zhao, X. Targeted Therapies in Gynecological Cancers: A Comprehensive Review of Clinical Evidence. Signal Transduct. Target. Ther. 2020, 5, 137. [Google Scholar] [CrossRef]
  121. Savant, S.S.; Sriramkumar, S.; O’Hagan, H.M. The Role of Inflammation and Inflammatory Mediators in the Development, Progression, Metastasis, and Chemoresistance of Epithelial Ovarian Cancer. Cancers 2018, 10, 251. [Google Scholar] [CrossRef]
  122. Ritch, S.J.; Telleria, C.M. The Transcoelomic Ecosystem and Epithelial Ovarian Cancer Dissemination. Front. Endocrinol. 2022, 13, 886533. [Google Scholar] [CrossRef]
  123. Nayak, S.B.; Bhat, V.R.; Mayya, S.S. Serum Copper, Ceruloplasmin and Thiobarbituric Acid Reactive Substance Status in Patients with Ovarian Cancer. Indian J. Physiol. Pharmacol. 2004, 48, 486–488. [Google Scholar]
  124. Korun, Z.E.U.; Erdem, M.; Erdem, A.; Onan, A.; Bozkurt, N.; Öktem, M.; Biberoğlu, K. Use of Serum Copper and Zinc Levels in the Diagnostic Evaluation of Endometrioma and Epithelial Ovarian Carcinoma. Česká Gynekol. 2023, 88, 279–286. [Google Scholar] [CrossRef]
  125. Lin, S.; Yang, H. Ovarian Cancer Risk According to Circulating Zinc and Copper Concentrations: A Meta-Analysis and Mendelian Randomization Study. Clin. Nutr. 2021, 40, 2464–2468. [Google Scholar] [CrossRef] [PubMed]
  126. Wheeler, L.J.; Desanto, K.; Teal, S.B.; Sheeder, J.; Guntupalli, S.R. Intrauterine Device Use and Ovarian Cancer Risk: A Systematic Review and Meta-Analysis. Obstet. Gynecol. 2019, 134, 791–800. [Google Scholar] [CrossRef] [PubMed]
  127. Zhao, S.; Zhang, X.; Gao, F.; Chi, H.; Zhang, J.; Xia, Z.; Cheng, C.; Liu, J. Identification of Copper Metabolism-Related Subtypes and Establishment of the Prognostic Model in Ovarian Cancer. Front. Endocrinol. 2023, 14, 1145797. [Google Scholar] [CrossRef] [PubMed]
  128. González-Martín, A.; Harter, P.; Leary, A.; Lorusso, D.; Miller, R.E.; Pothuri, B.; Ray-Coquard, I.; Tan, D.S.P.; Bellet, E.; Oaknin, A.; et al. Newly Diagnosed and Relapsed Epithelial Ovarian Cancer: ESMO Clinical Practice Guideline for Diagnosis, Treatment and Follow-Up. Ann. Oncol. 2023, 34, 833–848. [Google Scholar] [CrossRef] [PubMed]
  129. Christie, E.L.; Bowtell, D.D.L. Acquired Chemotherapy Resistance in Ovarian Cancer. Ann. Oncol. 2017, 28, viii13–viii15. [Google Scholar] [CrossRef] [PubMed]
  130. Pignata, S.; Pisano, C.; Di Napoli, M.; Cecere, S.C.; Tambaro, R.; Attademo, L. Treatment of Recurrent Epithelial Ovarian Cancer. Cancer 2019, 125, 4609–4615. [Google Scholar] [CrossRef] [PubMed]
  131. Katano, K.; Kondo, A.; Safaei, R.; Holzer, A.; Samimi, G.; Mishima, M.; Kuo, Y.-M.; Rochdi, M.; Howell, S.B. Acquisition of Resistance to Cisplatin Is Accompanied by Changes in the Cellular Pharmacology of Copper. Cancer Res. 2002, 62, 6559–6565. [Google Scholar]
  132. Ishida, S.; McCormick, F.; Smith-McCune, K.; Hanahan, D. Enhancing Tumor-Specific Uptake of the Anticancer Drug Cisplatin with a Copper Chelator. Cancer Cell 2010, 17, 574–583. [Google Scholar] [CrossRef]
  133. Lee, Y.-Y.; Choi, C.H.; Do, I.-G.; Song, S.Y.; Lee, W.; Park, H.S.; Song, T.J.; Kim, M.K.; Kim, T.-J.; Lee, J.-W.; et al. Prognostic Value of the Copper Transporters, CTR1 and CTR2, in Patients with Ovarian Carcinoma Receiving Platinum-Based Chemotherapy. Gynecol. Oncol. 2011, 122, 361–365. [Google Scholar] [CrossRef]
  134. Samimi, G.; Safaei, R.; Katano, K.; Holzer, A.K.; Rochdi, M.; Tomioka, M.; Goodman, M.; Howell, S.B. Increased Expression of the Copper Efflux Transporter ATP7A Mediates Resistance to Cisplatin, Carboplatin, and Oxaliplatin in Ovarian Cancer Cells. Clin. Cancer Res. 2004, 10, 4661–4669. [Google Scholar] [CrossRef]
  135. Dolgova, N.V.; Nokhrin, S.; Yu, C.H.; George, G.N.; Dmitriev, O.Y. Copper Chaperone Atox1 Interacts with the Metal-Binding Domain of Wilson’s Disease Protein in Cisplatin Detoxification. Biochem. J. 2013, 454, 147–156. [Google Scholar] [CrossRef] [PubMed]
  136. E Palm-Espling, M.; Lundin, C.; Bjorn, E.; Naredi, P.; Wittung-Stafshede, P. Interaction between the Anticancer Drug Cisplatin and the Copper Chaperone Atox1 in Human Melanoma Cells. Protein Pept. Lett. 2014, 21, 63–68. [Google Scholar] [CrossRef] [PubMed]
  137. Bompiani, K.M.; Tsai, C.-Y.; Achatz, F.P.; Liebig, J.K.; Howell, S.B. Copper Transporters and Chaperones CTR1, CTR2, ATOX1, and CCS as Determinants of Cisplatin Sensitivity. Metallomics 2016, 8, 951–962. [Google Scholar] [CrossRef] [PubMed]
  138. Siddiqui, S.; Mateen, S.; Ahmad, R.; Moin, S. A Brief Insight into the Etiology, Genetics, and Immunology of Polycystic Ovarian Syndrome (PCOS). J. Assist. Reprod. Genet. 2022, 39, 2439–2473. [Google Scholar] [CrossRef] [PubMed]
  139. Kiel, I.A.; Lionett, S.; Parr, E.B.; Jones, H.; Røset, M.A.H.; Salvesen, Ø.; Vanky, E.; Moholdt, T. Improving Reproductive Function in Women with Polycystic Ovary Syndrome with High-Intensity Interval Training (IMPROV-IT): Study Protocol for a Two-Centre, Three-Armed Randomised Controlled Trial. BMJ Open 2020, 10, e034733. [Google Scholar] [CrossRef]
  140. Wang, Z.; Zhai, D.; Zhang, D.; Bai, L.; Yao, R.; Yu, J.; Cheng, W.; Yu, C. Quercetin Decreases Insulin Resistance in a Polycystic Ovary Syndrome Rat Model by Improving Inflammatory Microenvironment. Reprod. Sci. 2017, 24, 682–690. [Google Scholar] [CrossRef]
  141. Torshizi, F.F.; Chamani, M.; Khodaei, H.R.; Sadeghi, A.A.; Hejazi, S.H.; Heravi, R.M. Therapeutic Effects of Organic Zinc on Reproductive Hormones, Insulin Resistance and MTOR Expression, as a Novel Component, in a Rat Model of Polycystic Ovary Syndrome. Iran. J. Basic Med. Sci. 2020, 23, 36. [Google Scholar] [CrossRef]
  142. Palomba, S.; De Wilde, M.A.; Falbo, A.; Koster, M.P.H.; La Sala, G.B.; Fauser, B.C.J.M. Pregnancy Complications in Women with Polycystic Ovary Syndrome. Hum. Reprod. Update 2015, 21, 575–592. [Google Scholar] [CrossRef]
  143. Naderpoor, N.; Shorakae, S.; Joham, A.; Boyle, J.; De Courten, B.; Teede, H.J. Obesity and Polycystic Ovary Syndrome. Minerva Endocrinol. 2014, 40, 37–51. [Google Scholar]
  144. Chen, C.; Jing, G.; Li, Z.; Juan, S.; Bin, C.; Jie, H. Insulin Resistance and Polycystic Ovary Syndrome in a Chinese Population. Endocr. Pract. Off. J. Am. Coll. Endocrinol. Am. Assoc. Clin. Endocrinol. 2017. [Google Scholar] [CrossRef]
  145. Ollila, M.-M.; West, S.; Keinänen-Kiukaanniemi, S.; Jokelainen, J.; Auvinen, J.; Puukka, K.; Ruokonen, A.; Järvelin, M.-R.; Tapanainen, J.S.; Franks, S.; et al. Overweight and Obese but Not Normal Weight Women with PCOS Are at Increased Risk of Type 2 Diabetes Mellitus—A Prospective, Population-Based Cohort Study. Hum. Reprod. 2017, 32, 423–431. [Google Scholar] [CrossRef] [PubMed]
  146. Berni, T.R.; Morgan, C.L.; Rees, D.A. Women with Polycystic Ovary Syndrome Have an Increased Risk of Major Cardiovascular Events: A Population Study. J. Clin. Endocrinol. Metab. 2021, 106, e3369–e3380. [Google Scholar] [CrossRef] [PubMed]
  147. Barry, J.A.; Azizia, M.M.; Hardiman, P.J. Risk of Endometrial, Ovarian and Breast Cancer in Women with Polycystic Ovary Syndrome: A Systematic Review and Meta-Analysis. Hum. Reprod. Update 2014, 20, 748–758. [Google Scholar] [CrossRef] [PubMed]
  148. Joham, A.E.; Norman, R.J.; Stener-Victorin, E.; Legro, R.S.; Franks, S.; Moran, L.J.; Boyle, J.; Teede, H.J. Polycystic Ovary Syndrome. Lancet Diabetes Endocrinol. 2022, 10, 668–680. [Google Scholar] [CrossRef] [PubMed]
  149. Günalan, E.; Yaba, A.; Yılmaz, B. The Effect of Nutrient Supplementation in the Management of Polycystic Ovary Syndrome-Associated Metabolic Dysfunctions: A Critical Review. J. Turkish Ger. Gynecol. Assoc. 2018, 19, 220–232. [Google Scholar] [CrossRef] [PubMed]
  150. Dapas, M.; Lin, F.T.J.; Nadkarni, G.N.; Sisk, R.; Legro, R.S.; Urbanek, M.; Hayes, M.G.; Dunaif, A. Distinct Subtypes of Polycystic Ovary Syndrome with Novel Genetic Associations: An Unsupervised, Phenotypic Clustering Analysis. PLoS Med. 2020, 17, e1003132. [Google Scholar] [CrossRef] [PubMed]
  151. Jiang, Q.; Zhang, F.; Han, L.; Zhu, B.; Liu, X. Serum Copper Level and Polycystic Ovarian Syndrome: A Meta-Analysis. Gynecol. Obstet. Investig. 2021, 86, 239–246. [Google Scholar] [CrossRef]
  152. Yin, J.; Hong, X.; Ma, J.; Bu, Y.; Liu, R. Serum Trace Elements in Patients with Polycystic Ovary Syndrome: A Systematic Review and Meta-Analysis. Front. Endocrinol. 2020, 11, 572384. [Google Scholar] [CrossRef]
  153. Mohmmed, A.H.; Awad, N.A.; AL-Fartosy, A.J.M. Study of Trace Elements Selenium, Copper, Zinc and Manganese Level in Polycystic Ovary Syndrome (PCOS). Int. J. Res. Appl. Sci. Biotechnol. 2019, 6, 16–22. [Google Scholar] [CrossRef]
  154. Kanafchian, M.; Esmaeilzadeh, S.; Mahjoub, S.; Rahsepar, M.; Ghasemi, M. Status of Serum Copper, Magnesium, and Total Antioxidant Capacity in Patients with Polycystic Ovary Syndrome. Biol. Trace Elem. Res. 2020, 193, 111–117. [Google Scholar] [CrossRef]
  155. Sun, Y.; Wang, W.; Guo, Y.; Zheng, B.; Li, H.; Chen, J.; Zhang, W. High Copper Levels in Follicular Fluid Affect Follicle Development in Polycystic Ovary Syndrome Patients: Population-Based and in Vitro Studies. Toxicol. Appl. Pharmacol. 2019, 365, 101–111. [Google Scholar] [CrossRef] [PubMed]
  156. Li, M.; Tang, Y.; Lin, C.; Huang, Q.; Lei, D.; Hu, Y. Serum Macroelement and Microelement Concentrations in Patients with Polycystic Ovary Syndrome: A Cross-Sectional Study. Biol. Trace Elem. Res. 2017, 176, 73–80. [Google Scholar] [CrossRef] [PubMed]
  157. Spritzer, P.M.; Lecke, S.B.; Fabris, V.C.; Ziegelmann, P.K.; Amaral, L. Blood Trace Element Concentrations in Polycystic Ovary Syndrome: Systematic Review and Meta-Analysis. Biol. Trace Elem. Res. 2017, 175, 254–262. [Google Scholar] [CrossRef] [PubMed]
  158. Zheng, G.; Wang, L.; Guo, Z.; Sun, L.; Wang, L.; Wang, C.; Zuo, Z.; Qiu, H. Association of Serum Heavy Metals and Trace Element Concentrations with Reproductive Hormone Levels and Polycystic Ovary Syndrome in a Chinese Population. Biol. Trace Elem. Res. 2015, 167, 1–10. [Google Scholar] [CrossRef] [PubMed]
  159. Celik, C.; Bastu, E.; Abali, R.; Alpsoy, S.; Guzel, E.C.; Aydemir, B.; Yeh, J. The Relationship between Copper, Homocysteine and Early Vascular Disease in Lean Women with Polycystic Ovary Syndrome. Gynecol. Endocrinol. 2013, 29, 488–491. [Google Scholar] [CrossRef] [PubMed]
  160. Mehde, A.A.; Resan, A.K. Study of Several Biochemical Features in Sera of Patients with Polycystic Ovaries and Compared with the Control Group. Aust. J. Basic Appl. Sci. 2014, 8, 620–627. [Google Scholar]
  161. Sharif, M.E.; Adam, I.; Ahmed, M.A.; Rayis, D.A.; Hamdan, H.Z. Serum Level of Zinc and Copper in Sudanese Women with Polycystic Ovarian Syndrome. Biol. Trace Elem. Res. 2017, 180, 23–27. [Google Scholar] [CrossRef]
  162. Khalaf, B.H.; Ouda, M.H.; Alghurabi, H.S.; Shubbar, A.S. Zinc and Copper Levels and Their Correlation with Polycystic Ovary Syndrome Biochemical Changes. Int. J. Pharm. Sci. Res 2018, 9, 3036–3041. [Google Scholar] [CrossRef]
  163. Schmalbrock, L.J.; Weiss, G.; Rijntjes, E.; Reinschissler, N.; Sun, Q.; Schenk, M.; Schomburg, L. Pronounced Trace Element Variation in Follicular Fluids of Subfertile Women Undergoing Assisted Reproduction. Nutrients 2021, 13, 4134. [Google Scholar] [CrossRef]
  164. Chakraborty, P.; Ghosh, S.; Goswami, S.K.; Kabir, S.N.; Chakravarty, B.; Jana, K. Altered Trace Mineral Milieu Might Play an Aetiological Role in the Pathogenesis of Polycystic Ovary Syndrome. Biol. Trace Elem. Res. 2013, 152, 9–15. [Google Scholar] [CrossRef]
  165. Bizoń, A.; Tchórz, A.; Madej Pawełand Leśniewski, M.; Wójtowicz, M.; Piwowar, A.; Franik, G. The Activity of Superoxide Dismutase, Its Relationship with the Concentration of Zinc and Copper and the Prevalence of Rs2070424 Superoxide Dismutase Gene in Women with Polycystic Ovary Syndrome—Preliminary Study. J. Clin. Med. 2022, 11, 2548. [Google Scholar] [CrossRef] [PubMed]
  166. Kirmizi, D.A.; Baser, E.; Turksoy, V.A.; Kara, M.; Yalvac, E.S.; Gocmen, A.Y. Are Heavy Metal Exposure and Trace Element Levels Related to Metabolic and Endocrine Problems in Polycystic Ovary Syndrome? Biol. Trace Elem. Res. 2020, 198, 77–86. [Google Scholar] [CrossRef] [PubMed]
  167. Wang, Q.; Sun, Y.; Zhao, A.; Cai, X.; Yu, A.; Xu, Q.; Liu, W.; Zhang, N.; Wu, S.; Chen, Y.; et al. High Dietary Copper Intake Induces Perturbations in the Gut Microbiota and Affects Host Ovarian Follicle Development. Ecotoxicol. Environ. Saf. 2023, 255, 114810. [Google Scholar] [CrossRef] [PubMed]
  168. Ojha, P.S.; Maste, M.M.; Tubachi, S.; Patil, V.S. Human Papillomavirus and Cervical Cancer: An Insight Highlighting Pathogenesis and Targeting Strategies. Virus Dis. 2022, 33, 132–154. [Google Scholar] [CrossRef] [PubMed]
  169. Bruni, L.; Albero, G.; Serrano, B.; Mena, M.; Collado, J.; Gómez, D.; Muñoz, J.; Bosch, F.; de Sanjosé, S. ICO/IARC Information Centre on HPV and Cancer (HPV Information Centre). Human Papillomavirus and Related Diseases in the World. Summ. Rep. 10 March 2023. Available online: https://hpvcentre.net/statistics/reports/XWX.pdf (accessed on 11 November 2023).
  170. Kamolratanakul, S.; Pitisuttithum, P. Human Papillomavirus Vaccine Efficacy and Effectiveness against Cancer. Vaccines 2021, 9, 1413. [Google Scholar] [CrossRef] [PubMed]
  171. Wentzensen, N.; Schiffman, M.; Palmer, T.; Arbyn, M. Triage of HPV Positive Women in Cervical Cancer Screening. J. Clin. Virol. 2016, 76, S49–S55. [Google Scholar] [CrossRef] [PubMed]
  172. Preci, D.P.; Almeida, A.; Weiler, A.L.; Franciosi, M.L.M.; Cardoso, A.M. Oxidative Damage and Antioxidants in Cervical Cancer. Int. J. Gynecol. Cancer 2021, 31, 265–271. [Google Scholar] [CrossRef]
  173. Averbach, S.; Silverberg, M.J.; Leyden, W.; Smith-McCune, K.; Raine-Bennett, T.; Sawaya, G.F. Recent Intrauterine Device Use and the Risk of Precancerous Cervical Lesions and Cervical Cancer. Contraception 2018, 98, 130–134. [Google Scholar] [CrossRef]
  174. Skorstengaard, M.; Lynge, E.; Napolitano, G.; Blaakær, J.; Bor, P. Risk of Precancerous Cervical Lesions in Women Using a Hormone-Containing Intrauterine Device and Other Contraceptives: A Register-Based Cohort Study from Denmark. Hum. Reprod. 2021, 36, 1796–1807. [Google Scholar] [CrossRef]
  175. Cunzhi, H.; Jiexian, J.; Xianwen, Z.; Jingang, G.; Shumin, Z.; Lili, D. Serum and Tissue Levels of Six Trace Elements and Copper/Zinc Ratio in Patients with Cervical Cancer and Uterine Myoma. Biol. Trace Elem. Res. 2003, 94, 113–122. [Google Scholar] [CrossRef]
  176. Naidu, M.S.K.; Suryakar, A.N.; Swami, S.C.; Katkam, R.V.; Kumbar, K.M. Oxidative Stress and Antioxidant Status in Cervical Cancer Patients. Indian J. Clin. Biochem. 2007, 22, 140–144. [Google Scholar] [CrossRef] [PubMed]
  177. Zhang, M.; Shi, M.; Zhao, Y. Association between Serum Copper Levels and Cervical Cancer Risk: A Meta-Analysis. Biosci. Rep. 2018, 38, BSR20180161. [Google Scholar] [CrossRef]
  178. Okonkwo, C.A.; Amegor, F.O.; Gbolade, J.O. Relationship between Trace Elements and Major Gynaecological Malignancies. Asian J. Med. Sci. 2013, 5, 124–127. [Google Scholar] [CrossRef]
  179. Hijam, D.; Dubey, A.; Laishram, V.; Jaichand, L.; Devi, T.I. Serum Copper Levels in Different Stages of Cervical Cancer in Manipur. Int. J. Med. Res. Prof. 2016, 24, 194–197. [Google Scholar] [CrossRef]
  180. Shah, S.; Kalal, B.S. Oxidative Stress in Cervical Cancer and Its Response to Chemoradiation. Turkish J. Obstet. Gynecol. 2019, 16, 124–128. [Google Scholar] [CrossRef] [PubMed]
  181. Brooks, R.A.; Fleming, G.F.; Lastra, R.R.; Lee, N.K.; Moroney, J.W.; Son, C.H.; Tatebe, K.; Veneris, J.L. Current Recommendations and Recent Progress in Endometrial Cancer. CA. Cancer J. Clin. 2019, 69, 258–279. [Google Scholar] [CrossRef]
  182. Lee, Y.C.; Lheureux, S.; Oza, A.M. Treatment Strategies for Endometrial Cancer: Current Practice and Perspective. Curr. Opin. Obstet. Gynecol. 2017, 29, 47–58. [Google Scholar] [CrossRef]
  183. Atakul, T.; Altinkaya, S.O.; Abas, B.I.; Yenisey, C. Serum Copper and Zinc Levels in Patients with Endometrial Cancer. Biol. Trace Elem. Res. 2020, 195, 46–54. [Google Scholar] [CrossRef]
  184. Rzymski, P.; Niedzielski, P.; Rzymski Pawełand Tomczyk, K.; Kozak, L.; Poniedziałek, B. Metal Accumulation in the Human Uterus Varies by Pathology and Smoking Status. Fertil. Steril. 2016, 105, 1511–1518. [Google Scholar] [CrossRef]
  185. Wieder-Huszla, S.; Chudecka-Głaz, A.; Cymbaluk-Płoska, A.; Karakiewicz, B.; Bosiacki, M.; Chlubek, D.; Jurczak, A. Evaluation of the Concentration of Selected Elements in Patients with Cancer of the Reproductive Organs with Respect to Treatment Stage—Preliminary Study. Nutrients 2022, 14, 2368. [Google Scholar] [CrossRef]
  186. Michalczyk, K.; Kapczuk, P.; Kupnicka, P.; Witczak, G.; Michalczyk, B.; Bosiacki, M.; Chlubek, D.; Cymbaluk-Płoska, A. Assessment of Serum Zn, Cu, Mn, and Fe Concentration in Women with Endometrial Cancer and Different Endometrial Pathologies. Nutrients 2023, 15, 3605. [Google Scholar] [CrossRef] [PubMed]
  187. Bahamondes, L.; Bahamondes, M.V.; Shulman, L.P. Non-Contraceptive Benefits of Hormonal and Intrauterine Reversible Contraceptive Methods. Hum. Reprod. Update 2015, 21, 640–651. [Google Scholar] [CrossRef] [PubMed]
  188. Raz, N.; Feinmesser, L.; Moore, O.; Haimovich, S. Endometrial Polyps: Diagnosis and Treatment Options—A Review of Literature. Minim. Invasive Ther. Allied Technol. 2021, 30, 278–287. [Google Scholar] [CrossRef] [PubMed]
  189. Yang, Q.; Ciebiera, M.; Bariani, M.V.; Ali, M.; Elkafas, H.; Boyer, T.G.; Al-Hendy, A. Comprehensive Review of Uterine Fibroids: Developmental Origin, Pathogenesis, and Treatment. Endocr. Rev. 2022, 43, 678–719. [Google Scholar] [CrossRef] [PubMed]
  190. Taylor, H.S.; Kotlyar, A.M.; Flores, V.A. Endometriosis Is a Chronic Systemic Disease: Clinical Challenges and Novel Innovations. Lancet 2021, 397, 839–852. [Google Scholar] [CrossRef] [PubMed]
  191. Hart, R.J. Physiological Aspects of Female Fertility: Role of the Environment, Modern Lifestyle, and Genetics. Physiol. Rev. 2016, 96, 873–909. [Google Scholar] [CrossRef] [PubMed]
  192. Lee, S.C.; Kaunitz, A.M.; Sanchez-Ramos, L.; Rhatigan, R.M. The Oncogenic Potential of Endometrial Polyps: A Systematic Review and Meta-Analysis. Obstet. Gynecol. 2010, 116, 1197–1205. [Google Scholar] [CrossRef]
  193. Peng, X.; Li, T.; Xia, E.; Xia, C.; Liu, Y.; Yu, D. A Comparison of Oestrogen Receptor and Progesterone Receptor Expression in Endometrial Polyps and Endometrium of Premenopausal Women. J. Obstet. Gynaecol. 2009, 29, 340–346. [Google Scholar] [CrossRef]
  194. Liu, Z.; Kuokkanen, S.; Pal, L. Steroid Hormone Receptor Profile of Premenopausal Endometrial Polyps. Reprod. Sci. 2010, 17, 377–383. [Google Scholar] [CrossRef]
  195. Yin, P.; Ono, M.; Moravek, M.B.; Coon, J.S.; Navarro, A.; Monsivais, D.; Dyson, M.T.; Druschitz, S.A.; Malpani, S.S.; Serna, V.A.; et al. Human Uterine Leiomyoma Stem/Progenitor Cells Expressing CD34 and CD49b Initiate Tumors in Vivo. J. Clin. Endocrinol. Metab. 2015, 100, E601–E606. [Google Scholar] [CrossRef]
  196. Mas, A.; Stone, L.; O’Connor, P.M.; Yang, Q.; Kleven, D.; Simon, C.; Walker, C.L.; Al-Hendy, A. Developmental Exposure to Endocrine Disruptors Expands Murine Myometrial Stem Cell Compartment as a Prerequisite to Leiomyoma Tumorigenesis. Stem Cells 2017, 35, 666–678. [Google Scholar] [CrossRef] [PubMed]
  197. Li, D.; Jiang, T.; Wang, X.; Yin, T.; Shen, L.; Zhang, Z.; Zou, W.; Liu, Y.; Zong, K.; Liang, D.; et al. Serum Essential Trace Element Status in Women and the Risk of Endometrial Diseases: A Case-Control Study. Biol. Trace Elem. Res. 2023, 201, 2151–2161. [Google Scholar] [CrossRef] [PubMed]
  198. Yılmaz, B.K.; Evliyaoğlu, Ö.; Yorgancı, A.; Özyer, Ş.; Üstün, Y.E. Serum Concentrations of Heavy Metals in Women with Endometrial Polyps. J. Obstet. Gynaecol. 2020, 40, 541–545. [Google Scholar] [CrossRef] [PubMed]
  199. Flores, I.; Rivera, E.; Ruiz, L.A.; Santiago, O.I.; Vernon, M.W.; Appleyard, C.B. Molecular Profiling of Experimental Endometriosis Identified Gene Expression Patterns in Common with Human Disease. Fertil. Steril. 2007, 87, 1180–1199. [Google Scholar] [CrossRef] [PubMed]
  200. Turgut, A.I.; Ozler, A.; Goruk, N.Y.; Tunc, S.Y.; Evliyaoglu, O.; Gul, T. Copper, Ceruloplasmin and Oxidative Stress in Patients with Advanced-Stage Endometriosis. Eur. Rev. Med. Pharmacol. Sci. 2013, 17, 1472–1478. [Google Scholar] [PubMed]
  201. Pollack, A.Z.; Louis, G.M.B.; Chen, Z.; Peterson, C.M.; Sundaram, R.; Croughan, M.S.; Sun, L.; Hediger, M.L.; Stanford, J.B.; Varner, M.W.; et al. Trace Elements and Endometriosis: The ENDO Study. Reprod. Toxicol. 2013, 42, 41–48. [Google Scholar] [CrossRef] [PubMed]
  202. Delsouc, M.B.; Ghersa, F.; Ramírez, D.; Della Vedova, M.C.; Gil, R.A.; Vallcaneras, S.S.; Casais, M. Endometriosis Progression in Tumor Necrosis Factor Receptor P55-Deficient Mice: Impact on Oxidative/Nitrosative Stress and Metallomic Profile. J. Trace Elem. Med. Biol. 2019, 52, 157–165. [Google Scholar] [CrossRef]
  203. Ngô, C.; Chéreau, C.; Nicco, C.; Weill, B.; Chapron, C.; Batteux, F. Reactive Oxygen Species Controls Endometriosis Progression. Am. J. Pathol. 2009, 175, 225–234. [Google Scholar] [CrossRef]
  204. Tsang, C.K.; Chen, M.; Cheng, X.; Qi, Y.; Chen, Y.; Das, I.; Li, X.; Vallat, B.; Fu, L.-W.; Qian, C.-N.; et al. SOD1 Phosphorylation by MTORC1 Couples Nutrient Sensing and Redox Regulation. Mol. Cell 2018, 70, 502–515. [Google Scholar] [CrossRef]
  205. McKinnon, B.D.; Kocbek, V.; Nirgianakis, K.; Bersinger, N.A.; Mueller, M.D. Kinase Signalling Pathways in Endometriosis: Potential Targets for Non-Hormonal Therapeutics. Hum. Reprod. Update 2016, 22, 382–403. [Google Scholar] [CrossRef]
  206. Klevay, L.M.; Christopherson, D.M. Copper Deficiency Halves Serum Dehydroepiandrosterone in Rats. J. Trace Elem. Med. Biol. 2000, 14, 143–145. [Google Scholar] [CrossRef] [PubMed]
  207. Soni, R.K.; Gupta, P.S.P.; Nandi, S.; Mondal, S.; Ippala, J.R.; Mor, A.; Mishra, A.; Tripathi, S.K. Effect of in Vitro Copper Supplementation on Granulosa Cell Estradiol Synthesis and Associated Genes. Indian J. Anim. Res. 2018, 52, 652–657. [Google Scholar] [CrossRef]
  208. Delsouc, M.B.; Conforti, R.A.; Vitale, D.L.; Alaniz, L.; Pacheco, P.; Andujar, S.; Vallcaneras, S.S.; Casais, M. Antiproliferative and Antiangiogenic Effects of Ammonium Tetrathiomolybdate in a Model of Endometriosis. Life Sci. 2021, 287, 120099. [Google Scholar] [CrossRef] [PubMed]
  209. Conforti, R.A.; Delsouc, M.B.; Zabala, A.S.; Vallcaneras, S.S.; Casais, M. The Copper Chelator Ammonium Tetrathiomolybdate Inhibits the Progression of Experimental Endometriosis in TNFR1-Deficient Mice. Sci. Rep. 2023, 13, 10354. [Google Scholar] [CrossRef] [PubMed]
  210. Vallcaneras, S.; Ghersa, F.; Bastón, J.; Delsouc, M.B.; Meresman, G.; Casais, M. TNFRp55 Deficiency Promotes the Development of Ectopic Endometriotic-like Lesions in Mice. J. Endocrinol. 2017, 234, 269–278. [Google Scholar] [CrossRef] [PubMed]
  211. Ghersa, F.; Delsouc, M.B.; Goyeneche, A.A.; Vallcaneras, S.S.; Meresman, G.; Telleria, C.M.; Casais, M. Reduced Inflammatory State Promotes Reinnervation of Endometriotic-like Lesions in TNFRp55 Deficient Mice. Mol. Hum. Reprod. 2019, 25, 385–396. [Google Scholar] [CrossRef]
  212. Li, Y. Copper Homeostasis: Emerging Target for Cancer Treatment. IUBMB Life 2020, 72, 1900–1908. [Google Scholar] [CrossRef]
  213. Babak, M.V.; Ahn, D. Modulation of Intracellular Copper Levels as the Mechanism of Action of Anticancer Copper Complexes: Clinical Relevance. Biomedicines 2021, 9, 852. [Google Scholar] [CrossRef]
  214. Kim, J.-J.; Kim, Y.-S.; Kumar, V. Heavy Metal Toxicity: An Update of Chelating Therapeutic Strategies. J. Trace Elem. Med. Biol. 2019, 54, 226–231. [Google Scholar] [CrossRef]
  215. Peisach, J.; Blumberg, W.E. A Mechanism for the Action of Penicillamine in the Treatment of Wilson’s Disease. Mol. Pharmacol. 1969, 5, 200–209. [Google Scholar]
  216. Kumar, V.; Singh, A.P.; Wheeler, N.; Galindo, C.L.; Kim, J.-J. Safety Profile of D-Penicillamine: A Comprehensive Pharmacovigilance Analysis by FDA Adverse Event Reporting System. Expert Opin. Drug Saf. 2021, 20, 1443–1450. [Google Scholar] [CrossRef] [PubMed]
  217. Matsubara, T.; Saura, R.; Hirohata, K.; Ziff, M. Inhibition of Human Endothelial Cell Proliferation in Vitro and Neovascularization in Vivo by D-Penicillamine. J. Clin. Investig. 1989, 83, 158–167. [Google Scholar] [CrossRef] [PubMed]
  218. Crowe, A.; Jackaman, C.; Beddoes, K.M.; Ricciardo, B.; Nelson, D.J. Rapid Copper Acquisition by Developing Murine Mesothelioma: Decreasing Bioavailable Copper Slows Tumor Growth, Normalizes Vessels and Promotes T Cell Infiltration. PLoS ONE 2013, 8, e73684. [Google Scholar] [CrossRef] [PubMed]
  219. Mammoto, T.; Jiang, A.; Jiang, E.; Panigrahy, D.; Kieran, M.W.; Mammoto, A. Role of Collagen Matrix in Tumor Angiogenesis and Glioblastoma Multiforme Progression. Am. J. Pathol. 2013, 183, 1293–1305. [Google Scholar] [CrossRef] [PubMed]
  220. Kim, H.; Jo, S.; Kim, I.-G.; Kim, R.-K.; Kahm, Y.-J.; Jung, S.-H.; Lee, J.H. Effect of Copper Chelators via the TGF-β Signaling Pathway on Glioblastoma Cell Invasion. Molecules 2022, 27, 8851. [Google Scholar] [CrossRef]
  221. Chen, S.-J.; Kuo, C.-C.; Pan, H.-Y.; Tsou, T.-C.; Yeh, S.-C.; Chang, J.-Y. Mechanistic Basis of a Combination D-Penicillamine and Platinum Drugs Synergistically Inhibits Tumor Growth in Oxaliplatin-Resistant Human Cervical Cancer Cells In Vitro and In Vivo. Biochem. Pharmacol. 2015, 95, 28–37. [Google Scholar] [CrossRef]
  222. Horn, N.; Møller, L.B.; Nurchi, V.M.; Aaseth, J. Chelating Principles in Menkes and Wilson Diseases: Choosing the Right Compounds in the Right Combinations at the Right Time. J. Inorg. Biochem. 2019, 190, 98–112. [Google Scholar] [CrossRef]
  223. Weiss, K.H.; Thurik, F.; Gotthardt, D.N.; Schäfer, M.; Teufel, U.; Wiegand, F.; Merle, U.; Ferenci-Foerster, D.; Maieron, A.; Stauber, R.; et al. Efficacy and Safety of Oral Chelators in Treatment of Patients with Wilson Disease. Clin. Gastroenterol. Hepatol. 2013, 11, 1028–1035. [Google Scholar] [CrossRef]
  224. Yoshii, J.; Yoshiji, H.; Kuriyama, S.; Ikenaka, Y.; Noguchi, R.; Okuda, H.; Tsujinoue, H.; Nakatani, T.; Kishida, H.; Nakae, D.; et al. The Copper-Chelating Agent, Trientine, Suppresses Tumor Development and Angiogenesis in the Murine Hepatocellular Carcinoma Cells. Int. J. Cancer 2001, 94, 768–773. [Google Scholar] [CrossRef]
  225. Moriguchi, M.; Nakajima, T.; Kimura, H.; Watanabe, T.; Takashima, H.; Mitsumoto, Y.; Katagishi, T.; Okanoue, T.; Kagawa, K. The Copper Chelator Trientine Has an Antiangiogenic Effect against Hepatocellular Carcinoma, Possibly through Inhibition of Interleukin-8 Production. Int. J. Cancer 2002, 102, 445–452. [Google Scholar] [CrossRef]
  226. Hayashi, M.; Nishiya, H.; Chiba, T.; Endoh, D.; Kon, Y.; Okui, T. Trientine, a Copper-Chelating Agent, Induced Apoptosis in Murine Fibrosarcoma Cells in Vivo and in Vitro. J. Vet. Med. Sci. 2007, 69, 137–142. [Google Scholar] [CrossRef]
  227. Liu, J.; Guo, L.; Yin, F.; Zheng, X.; Chen, G.; Wang, Y. Characterization and Antitumor Activity of Triethylene Tetramine, a Novel Telomerase Inhibitor. Biomed. Pharmacother. 2008, 62, 480–485. [Google Scholar] [CrossRef] [PubMed]
  228. Guterres, A.N.; Villanueva, J. Targeting Telomerase for Cancer Therapy. Oncogene 2020, 39, 5811–5824. [Google Scholar] [CrossRef] [PubMed]
  229. Huang, Y.-F.; Kuo, M.T.; Liu, Y.-S.; Cheng, Y.-M.; Wu, P.-Y.; Chou, C.-Y. A Dose Escalation Study of Trientine plus Carboplatin and Pegylated Liposomal Doxorubicin in Women with a First Relapse of Epithelial Ovarian, Tubal, and Peritoneal Cancer within 12 Months after Platinum-Based Chemotherapy. Front. Oncol. 2019, 9, 437. [Google Scholar] [CrossRef] [PubMed]
  230. Ferguson, W.S.; Lewis, A.H.; Watson, S.J. The Teart Pastures of Somerset: I. The Cause and Cure of Teartness. J. Agric. Sci. 1943, 33, 44–51. [Google Scholar] [CrossRef]
  231. Bickel, H.; Neale, F.C.; Hall, G. A Clinical and Biochemical Study of Hepatolenticular Degeneration (Wilson’s Disease). QJM An Int. J. Med. 1957, 26, 527–558. [Google Scholar]
  232. Dick, A.T.; Dewey, D.W.; Gawthorne, J.M. Thiomolybdates and the Copper–Molybdenum–Sulphur Interaction in Ruminant Nutrition. J. Agric. Sci. 1975, 85, 567–568. [Google Scholar] [CrossRef]
  233. Brewer, G.J.; Askari, F.; Lorincz, M.T.; Carlson, M.; Schilsky, M.; Kluin, K.J.; Hedera, P.; Moretti, P.; Fink, J.K.; Tankanow, R.; et al. Treatment of Wilson Disease with Ammonium Tetrathiomolybdate: IV. Comparison of Tetrathiomolybdate and Trientine in a Double-Blind Study of Treatment of the Neurologic Presentation of Wilson Disease. Arch. Neurol. 2006, 63, 521. [Google Scholar] [CrossRef]
  234. Cox, C.; Teknos, T.N.; Barrios, M.; Brewer, G.J.; Dick, R.D.; Merajver, S.D. The Role of Copper Suppression as an Antiangiogenic Strategy in Head and Neck Squamous Cell Carcinoma. Laryngoscope 2001, 111, 696–701. [Google Scholar] [CrossRef]
  235. Khan, M.K.; Miller, M.W.; Taylor, J.; Gill, N.K.; Dick, R.D.; Van Goled, K.; Brewert, G.J.; Merajver, S.D. Radiotherapy and Antiangiogenic TM in Lung Cancer. Neoplasia 2002, 4, 164–170. [Google Scholar] [CrossRef]
  236. Van Golen, K.L.; Bao, L.; Brewert, G.J.; Pienta, K.J.; Kamradt, J.M.; Livant, D.L.; Merajver, S.D. Suppression of Tumor Recurrence and Metastasis by a Combination of the PHSCN Sequence and the Antiangiogenic Compound Tetrathiomolybdate in Prostate Carcinoma. Neoplasia 2002, 4, 373–379. [Google Scholar] [CrossRef] [PubMed]
  237. Kim, K.K.; Lange, T.S.; Singh, R.K.; Brard, L.; Moore, R.G. Tetrathiomolybdate Sensitizes Ovarian Cancer Cells to Anticancer Drugs Doxorubicin, Fenretinide, 5-Fluorouracil and Mitomycin C. BMC Cancer 2012, 12, 147. [Google Scholar] [CrossRef] [PubMed]
  238. Chan, N.; Willis, A.; Kornhauser, N.; Ward, M.M.; Lee, S.B.; Nackos, E.; Seo, B.R.; Chuang, E.; Cigler, T.; Moore, A.; et al. Influencing the Tumor Microenvironment: A Phase II Study of Copper Depletion Using Tetrathiomolybdate in Patients with Breast Cancer at High Risk for Recurrence and in Preclinical Models of Lung Metastases. Clin. Cancer Res. 2017, 23, 666–676. [Google Scholar] [CrossRef] [PubMed]
  239. Alvarez, H.M.; Xue, Y.; Robinson, C.D.; Canalizo-Hernández, M.A.; Marvin, R.G.; Kelly, R.A.; Mondragón, A.; Penner-Hahn, J.E.; O’Halloran, T.V. Tetrathiomolybdate Inhibits Copper Trafficking Proteins through Metal Cluster Formation. Science 2010, 327, 331–334. [Google Scholar] [CrossRef] [PubMed]
  240. Juarez, J.C.; Betancourt, O.; Pirie-Shepherd, S.R.; Guan, X.; Price, M.L.; Shaw, D.E.; Mazar, A.P.; Doñate, F. Copper Binding by Tetrathiomolybdate Attenuates Angiogenesis and Tumor Cell Proliferation through the Inhibition of Superoxide Dismutase 1. Clin. Cancer Res. 2006, 12, 4974–4982. [Google Scholar] [CrossRef]
  241. Baldari, S.; Di Rocco, G.; Heffern, M.C.; Su, T.A.; Chang, C.J.; Toietta, G. Effects of Copper Chelation on BRAFV600E Positive Colon Carcinoma Cells. Cancers 2019, 11, 659. [Google Scholar] [CrossRef] [PubMed]
  242. Kim, Y.-J.; Tsang, T.; Anderson, G.R.; Posimo, J.M.; Brady, D.C. Inhibition of BCL2 Family Members Increases the Efficacy of Copper Chelation in BRAFV600E-Driven Melanoma. Cancer Res. 2020, 80, 1387–1400. [Google Scholar] [CrossRef]
  243. Ryumon, S.; Okui, T.; Kunisada, Y.; Kishimoto, K.; Shimo, T.; Hasegawa, K.; Ibaragi, S.; Akiyama, K.; Thu Ha, N.T.; Monsur Hassan, N.M.; et al. Ammonium Tetrathiomolybdate Enhances the Antitumor Effect of Cisplatin via the Suppression of ATPase Copper Transporting Beta in Head and Neck Squamous Cell Carcinoma. Oncol. Rep. 2019, 42, 2611–2621. [Google Scholar] [CrossRef]
  244. Schneider, B.J.; Lee, J.S.-J.; Hayman, J.A.; Chang, A.C.; Orringer, M.B.; Pickens, A.; Pan, C.C.; Merajver, S.D.; Urba, S.G. Pre-Operative Chemoradiation Followed by Post-Operative Adjuvant Therapy with Tetrathiomolybdate, a Novel Copper Chelator, for Patients with Resectable Esophageal Cancer. Investig. New Drugs 2013, 31, 435–442. [Google Scholar] [CrossRef]
  245. Kim, K.K.; Han, A.; Yano, N.; Ribeiro, J.R.; Lokich, E.; Singh, R.K.; Moore, R.G. Tetrathiomolybdate Mediates Cisplatin-Induced P38 Signaling and EGFR Degradation and Enhances Response to Cisplatin Therapy in Gynecologic Cancers. Sci. Rep. 2015, 5, 1–11. [Google Scholar] [CrossRef]
  246. Kim, K.K.; Kawar, N.M.; Singh, R.K.; Lange, T.S.; Brard, L.; Moore, R.G. Tetrathiomolybdate Induces Doxorubicin Sensitivity in Resistant Tumor Cell Lines. Gynecol. Oncol. 2011, 122, 183–189. [Google Scholar] [CrossRef] [PubMed]
  247. Nan, L.; Yuan, W.; Guodong, C.; Yonghui, H. Multitargeting Strategy Using Tetrathiomolybdate and Lenvatinib: Maximizing Antiangiogenesis Activity in a Preclinical Liver Cancer Model. Anti-Cancer Agents Med. Chem. (Former. Curr. Med. Chem. Agents) 2023, 23, 786–793. [Google Scholar] [CrossRef] [PubMed]
  248. Rogers, P.A.W.; Adamson, G.D.; Al-Jefout, M.; Becker, C.M.; D’Hooghe, T.M.; Dunselman, G.A.J.; Fazleabas, A.; Giudice, L.C.; Horne, A.W.; Hull, M.L.; et al. Research Priorities for Endometriosis. Reprod. Sci. 2017, 24, 202–226. [Google Scholar] [CrossRef] [PubMed]
  249. Richter, O.N.; Dorn, C.; Rösing, B.; Flaskamp, C.; Ulrich, U. Tumor Necrosis Factor Alpha Secretion by Peritoneal Macrophages in Patients with Endometriosis. Arch. Gynecol. Obstet. 2005, 271, 143–147. [Google Scholar] [CrossRef] [PubMed]
  250. Braun, D.P.; Ding, J.; Dmowski, W.P. Peritoneal Fluid-Mediated Enhancement of Eutopic and Ectopic Endometrial Cell Proliferation Is Dependent on Tumor Necrosis Factor-α in Women with Endometriosis. Fertil. Steril. 2002, 78, 727–732. [Google Scholar] [CrossRef] [PubMed]
  251. Sheng, Y.; Li, F.; Qin, Z. TNF Receptor 2 Makes Tumor Necrosis Factor a Friend of Tumors. Front. Immunol. 2018, 9, 1170. [Google Scholar] [CrossRef] [PubMed]
  252. Gough, P.; Myles, I.A. Tumor Necrosis Factor Receptors: Pleiotropic Signaling Complexes and Their Differential Effects. Front. Immunol. 2020, 11, 585880. [Google Scholar] [CrossRef]
  253. Rivas, M.A.; Carnevale, R.P.; Proietti, C.J.; Rosemblit, C.; Beguelin, W.; Salatino, M.; Charreau, E.H.; Frahm, I.; Sapia, S.; Brouckaert, P.; et al. TNFα Acting on TNFR1 Promotes Breast Cancer Growth via P42/P44 MAPK, JNK, Akt and NF-ΚB-Dependent Pathways. Exp. Cell Res. 2008, 314, 509–529. [Google Scholar] [CrossRef]
  254. Islimye, M.; Kilic, S.; Zulfikaroglu, E.; Topcu, O.; Zergeroglu, S.; Batioglu, S. Regression of Endometrial Autografts in a Rat Model of Endometriosis Treated with Etanercept. Eur. J. Obstet. Gynecol. Reprod. Biol. 2011, 159, 184–189. [Google Scholar] [CrossRef]
  255. Oliveri, V. Selective Targeting of Cancer Cells by Copper Ionophores: An Overview. Front. Mol. Biosci. 2022, 9, 841814. [Google Scholar] [CrossRef]
  256. Trachootham, D.; Alexandre, J.; Huang, P. Targeting Cancer Cells by ROS-Mediated Mechanisms: A Radical Therapeutic Approach? Nat. Rev. Drug Discov. 2009, 8, 579–591. [Google Scholar] [CrossRef] [PubMed]
  257. Shimada, K.; Reznik, E.; Stokes, M.E.; Krishnamoorthy, L.; Bos, P.H.; Song, Y.; Quartararo, C.E.; Pagano, N.C.; Carpizo, D.R.; DeCarvalho, A.C.; et al. Copper-Binding Small Molecule Induces Oxidative Stress and Cell-Cycle Arrest in Glioblastoma-Patient-Derived Cells. Cell Chem. Biol. 2018, 25, 585–594. [Google Scholar] [CrossRef] [PubMed]
  258. Tsvetkov, P.; Coy, S.; Petrova, B.; Dreishpoon, M.; Verma, A.; Abdusamad, M.; Rossen, J.; Joesch-Cohen, L.; Humeidi, R.; Spangler, R.D.; et al. Copper Induces Cell Death by Targeting Lipoylated TCA Cycle Proteins. Science 2022, 375, 1254–1261. [Google Scholar] [CrossRef] [PubMed]
  259. Xiao, Y.A.N.; Chen, D.I.; Zhang, X.I.A.; Cui, Q.; Fan, Y.; Bi, C.; Dou, Q.P. Molecular Study on Copper-Mediated Tumor Proteasome Inhibition and Cell Death. Int. J. Oncol. 2010, 37, 81–87. [Google Scholar] [CrossRef] [PubMed]
  260. Denoyer, D.; Pearson, H.B.; Clatworthy, S.A.S.; Smith, Z.M.; Francis, P.S.; Llanos, R.M.; Volitakis, I.; Phillips, W.A.; Meggyesy, P.M.; Masaldan, S.; et al. Copper as a Target for Prostate Cancer Therapeutics: Copper-Ionophore Pharmacology and Altering Systemic Copper Distribution. Oncotarget 2016, 7, 37064–37080. [Google Scholar] [CrossRef] [PubMed]
  261. Lu, C.; Li, X.; Ren, Y.; Zhang, X. Disulfiram: A Novel Repurposed Drug for Cancer Therapy. Cancer Chemother. Pharmacol. 2021, 87, 159–172. [Google Scholar] [CrossRef] [PubMed]
  262. Lewison, E.F. Spontaneous Regression of Breast Cancer. Prog. Clin. Biol. Res. 1977, 12, 47–53. [Google Scholar]
  263. Ekinci, E.; Rohondia, S.; Khan, R.; Dou, Q.P. Repurposing Disulfiram as an Anti-Cancer Agent: Updated Review on Literature and Patents. Recent Pat. Anticancer. Drug Discov. 2019, 14, 113–132. [Google Scholar] [CrossRef]
  264. Jia, Y.; Huang, T. Overview of Antabuse®(Disulfiram) in Radiation and Cancer Biology. Cancer Manag. Res. 2021, 13, 4095–4101. [Google Scholar] [CrossRef]
  265. Kannappan, V.; Ali, M.; Small, B.; Rajendran, G.; Elzhenni, S.; Taj, H.; Wang, W.; Dou, Q.P. Recent Advances in Repurposing Disulfiram and Disulfiram Derivatives as Copper-Dependent Anticancer Agents. Front. Mol. Biosci. 2021, 8, 741316. [Google Scholar] [CrossRef]
  266. Li, H.; Wang, J.; Wu, C.; Wang, L.; Chen, Z.-S.; Cui, W. The Combination of Disulfiram and Copper for Cancer Treatment. Drug Discov. Today 2020, 25, 1099–1108. [Google Scholar] [CrossRef] [PubMed]
  267. Li, Y.; Wang, L.-H.; Zhang, H.-T.; Wang, Y.-T.; Liu, S.; Zhou, W.-L.; Yuan, X.-Z.; Li, T.-Y.; Wu, C.-F.; Yang, J.-Y. Disulfiram Combined with Copper Inhibits Metastasis and Epithelial-Mesenchymal Transition in Hepatocellular Carcinoma through the NF-ΚB and TGF-β Pathways. J. Cell. Mol. Med. 2018, 22, 439–451. [Google Scholar] [CrossRef] [PubMed]
  268. Caminear, M.W.; Harrington, B.S.; Kamdar, R.D.; Kruhlak, M.J.; Annunziata, C.M. Disulfiram Transcends ALDH Inhibitory Activity When Targeting Ovarian Cancer Tumor-Initiating Cells. Front. Oncol. 2022, 12, 762820. [Google Scholar] [CrossRef] [PubMed]
  269. Guo, F.; Yang, Z.; Sehouli, J.; Kaufmann, A.M. Blockade of ALDH in Cisplatin-Resistant Ovarian Cancer Stem Cells in Vitro Synergistically Enhances Chemotherapy-Induced Cell Death. Curr. Oncol. 2022, 29, 2808–2822. [Google Scholar] [CrossRef] [PubMed]
  270. Dinavahi, S.S.; Bazewicz, C.G.; Gowda, R. Aldehyde Dehydrogenase Inhibitors for Cancer Therapeutics. Trends Pharmacol. Sci. 2019, 40, 774–789. [Google Scholar] [CrossRef] [PubMed]
  271. Silva, I.A.; Bai, S.; McLean, K.; Yang, K.; Griffith, K.; Thomas, D.; Ginestier, C.; Johnston, C.; Kueck, A.; Reynolds, R.K.; et al. Aldehyde Dehydrogenase in Combination with CD133 Defines Angiogenic Ovarian Cancer Stem Cells That Portend Poor Patient Survival. Cancer Res. 2011, 71, 3991–4001. [Google Scholar] [CrossRef]
  272. Çelik, Ö.; Erşahin, A.; Acet, M.; Çelik, N.; Baykuş, Y.; Deniz, R.; Özerol, E.; Özerol, İ. Disulfiram, as a Candidate NF-Kappa B and Proteasome Inhibitor, Prevents Endometriotic Implant Growing in a Rat Model of Endometriosis. Eur. Rev. Med. Pharmacol. Sci. 2016, 20, 4380–4389. [Google Scholar]
  273. Meraz-Torres, F.; Plöger, S.; Garbe, C.; Niessner, H.; Sinnberg, T. Disulfiram as a Therapeutic Agent for Metastatic Malignant Melanoma—Old Myth or New Logos? Cancers 2020, 12, 3538. [Google Scholar] [CrossRef]
  274. Jiao, Y.; Hannafon, B.N.; Zhang, R.R.; Fung, K.-M.; Ding, W.-Q. Docosahexaenoic Acid and Disulfiram Act in Concert to Kill Cancer Cells: A Mutual Enhancement of Their Anticancer Actions. Oncotarget 2017, 8, 17908–17920. [Google Scholar] [CrossRef]
  275. Tang, B.; Wu, M.; Zhang, L.; Jian, S.; Lv, S.; Lin, T.; Zhu, S.; Liu, L.; Wang, Y.; Yi, Z.; et al. Combined Treatment of Disulfiram with PARP Inhibitors Suppresses Ovarian Cancer. Front. Oncol. 2023, 13, 1154073. [Google Scholar] [CrossRef]
  276. Du, R.; Sun, F.; Li, K.; Qi, J.; Zhong, W.; Wang, W.; Sun, Q.; Deng, Q.; Wang, H.; Nie, J.; et al. Proteomics Analysis Revealed Smad3 as A Potential Target of the Synergistic Anti-Tumor Activity of Disulfiram and Cisplatin in Ovarian Cancer. Anticancer. Agents Med. Chem. 2023, 23, 1754–1764. [Google Scholar] [CrossRef] [PubMed]
  277. Liu, Y.; Guan, X.; Wang, M.; Wang, N.; Chen, Y.; Li, B.; Xu, Z.; Fu, F.; Du, C.; Zheng, Z. Disulfiram/Copper Induces Antitumor Activity against Gastric Cancer via the ROS/MAPK and NPL4 Pathways. Bioengineered 2022, 13, 6579–6589. [Google Scholar] [CrossRef] [PubMed]
  278. Safi, R.; Nelson, E.R.; Chitneni, S.K.; Franz, K.J.; George, D.J.; Zalutsky, M.R.; McDonnell, D.P. Copper Signaling Axis as a Target for Prostate Cancer Therapeutics. Cancer Res. 2014, 74, 5819–5831. [Google Scholar] [CrossRef] [PubMed]
  279. Lun, X.; Wells, J.C.; Grinshtein, N.; King, J.C.; Hao, X.; Dang, N.-H.; Wang, X.; Aman, A.; Uehling, D.; Datti, A.; et al. Disulfiram When Combined with Copper Enhances the Therapeutic Effects of Temozolomide for the Treatment of Glioblastoma. Clin. Cancer Res. 2016, 22, 3860–3875. [Google Scholar] [CrossRef] [PubMed]
  280. Oliveri, V. Biomedical Applications of Copper Ionophores. Coord. Chem. Rev. 2020, 422, 213474. [Google Scholar] [CrossRef]
  281. Liu, X.; Wang, L.; Cui, W.; Yuan, X.; Lin, L.; Cao, Q.; Wang, N.; Li, Y.; Guo, W.; Zhang, X.; et al. Targeting ALDH1A1 by Disulfiram/Copper Complex Inhibits Non-Small Cell Lung Cancer Recurrence Driven by ALDH-Positive Cancer Stem Cells. Oncotarget 2016, 7, 58516–58530. [Google Scholar] [CrossRef] [PubMed]
  282. Xu, B.; Wang, S.; Li, R.; Chen, K.; He, L.; Deng, M.; Kannappan, V.; Zha, J.; Dong, H.; Wang, W. Disulfiram/Copper Selectively Eradicates AML Leukemia Stem Cells in Vitro and in Vivo by Simultaneous Induction of ROS-JNK and Inhibition of NF-ΚB and Nrf2. Cell Death Dis. 2017, 8, e2797. [Google Scholar] [CrossRef]
  283. Serra, R.; Zhao, T.; Huq, S.; Gorelick, N.L.; Casaos, J.; Cecia, A.; Mangraviti, A.; Eberhart, C.; Bai, R.; Olivi, A.; et al. Disulfiram and Copper Combination Therapy Targets NPL4, Cancer Stem Cells and Extends Survival in a Medulloblastoma Model. PLoS ONE 2021, 16, e0251957. [Google Scholar] [CrossRef]
  284. Sun, T.; Yang, W.; Toprani, S.M.; Guo, W.; He, L.; DeLeo, A.B.; Ferrone, S.; Zhang, G.; Wang, E.; Lin, Z.; et al. Induction of Immunogenic Cell Death in Radiation-Resistant Breast Cancer Stem Cells by Repurposing Anti-Alcoholism Drug Disulfiram. Cell Commun. Signal. 2020, 18, 36. [Google Scholar] [CrossRef]
  285. Falls-Hubert, K.C.; Butler, A.L.; Gui, K.; Anderson, M.; Li, M.; Stolwijk, J.M.; Rodman III, S.N.; Solst, S.R.; Tomanek-Chalkley, A.; Searby, C.C.; et al. Disulfiram Causes Selective Hypoxic Cancer Cell Toxicity and Radio-Chemo-Sensitization via Redox Cycling of Copper. Free Radic. Biol. Med. 2020, 150, 1–11. [Google Scholar] [CrossRef]
  286. Guo, W.; Zhang, X.; Lin, L.; Wang, H.; He, E.; Wang, G.; Zhao, Q. The Disulfiram/Copper Complex Induces Apoptosis and Inhibits Tumour Growth in Human Osteosarcoma by Activating the ROS/JNK Signalling Pathway. J. Biochem. 2021, 170, 275–287. [Google Scholar] [CrossRef] [PubMed]
  287. Zhang, W.; Zhai, Q.; Li, M.; Huang, S.; Sun, Z.; Yan, Z.; Li, J.; Li, L.; Li, Y. Anti-Cancer Effects of Disulfiram in Cervical Cancer Cell Lines Are Mediated by Both Autophagy and Apoptosis. Bull. Exp. Biol. Med. 2022, 172, 642–648. [Google Scholar] [CrossRef] [PubMed]
  288. Shinde, S.D.; Sakla, A.P.; Shankaraiah, N. An Insight into Medicinal Attributes of Dithiocarbamates: Bird’s Eye View. Bioorg. Chem. 2020, 105, 104346. [Google Scholar] [CrossRef] [PubMed]
  289. Wykowski, R.; Fuentefria, A.M.; de Andrade, S.F. Antimicrobial Activity of Clioquinol and Nitroxoline: A Scoping Review. Arch. Microbiol. 2022, 204, 535. [Google Scholar] [CrossRef] [PubMed]
  290. Ding, W.-Q.; Liu, B.; Vaught, J.L.; Yamauchi, H.; Lind, S.E. Anticancer Activity of the Antibiotic Clioquinol. Cancer Res. 2005, 65, 3389–3395. [Google Scholar] [CrossRef] [PubMed]
  291. Chen, D.; Cui, Q.C.; Yang, H.; Barrea, R.A.; Sarkar, F.H.; Sheng, S.; Yan, B.; Reddy, G.P.V.; Dou, Q.P. Clioquinol, a Therapeutic Agent for Alzheimer’s Disease, Has Proteasome-Inhibitory, Androgen Receptor-Suppressing, Apoptosis-Inducing, and Antitumor Activities in Human Prostate Cancer Cells and Xenografts. Cancer Res. 2007, 67, 1636–1644. [Google Scholar] [CrossRef]
  292. Tuller, E.R.; Brock, A.L.; Yu, H.; Lou, J.R.; Benbrook, D.M.; Ding, W.-Q. PPARα Signaling Mediates the Synergistic Cytotoxicity of Clioquinol and Docosahexaenoic Acid in Human Cancer Cells. Biochem. Pharmacol. 2009, 77, 1480–1486. [Google Scholar] [CrossRef]
  293. Cater, M.A.; Haupt, Y. Clioquinol Induces Cytoplasmic Clearance of the X-Linked Inhibitor of Apoptosis Protein (XIAP): Therapeutic Indication for Prostate Cancer. Biochem. J. 2011, 436, 481–491. [Google Scholar] [CrossRef]
  294. Mao, X.; Li, X.; Sprangers, R.; Wang, X.; Venugopal, A.; Wood, T.; Zhang, Y.; Kuntz, D.A.; Coe, E.; Trudel, S.; et al. Clioquinol Inhibits the Proteasome and Displays Preclinical Activity in Leukemia and Myeloma. Leukemia 2009, 23, 585–590. [Google Scholar] [CrossRef]
  295. Cao, B.; Li, J.; Zhou, X.; Juan, J.; Han, K.; Zhang, Z.; Kong, Y.; Wang, J.; Mao, X. Clioquinol Induces Pro-Death Autophagy in Leukemia and Myeloma Cells by Disrupting the MTOR Signaling Pathway. Sci. Rep. 2014, 4, 5749. [Google Scholar] [CrossRef]
  296. Barrea, R.A.; Chen, D.; Irving, T.C.; Dou, Q.P. Synchrotron X-Ray Imaging Reveals a Correlation of Tumor Copper Speciation with Clioquinol’s Anticancer Activity. J. Cell. Biochem. 2009, 108, 96–105. [Google Scholar] [CrossRef] [PubMed]
  297. Du, T.; Filiz, G.; Caragounis, A.; Crouch, P.J.; White, A.R. Clioquinol Promotes Cancer Cell Toxicity through Tumor Necrosis Factor α Release from Macrophages. J. Pharmacol. Exp. Ther. 2008, 324, 360–367. [Google Scholar] [CrossRef] [PubMed]
  298. Bareggi, S.R.; Cornelli, U. Clioquinol: Review of Its Mechanisms of Action and Clinical Uses in Neurodegenerative Disorders. CNS Neurosci. Ther. 2012, 18, 41–46. [Google Scholar] [CrossRef] [PubMed]
  299. Khan, R.; Khan, H.; Abdullah, Y.; Dou, Q.P. Feasibility of Repurposing Clioquinol for Cancer Therapy. Recent Pat. Anticancer. Drug Discov. 2020, 15, 14–31. [Google Scholar] [CrossRef] [PubMed]
  300. Jiang, H.; Taggart, J.E.; Zhang, X.; Benbrook, D.M.; Lind, S.E.; Ding, W.-Q. Nitroxoline (8-Hydroxy-5-Nitroquinoline) Is More a Potent Anti-Cancer Agent than Clioquinol (5-Chloro-7-Iodo-8-Quinoline). Cancer Lett. 2011, 312, 11–17. [Google Scholar] [CrossRef] [PubMed]
  301. Summers, K.L.; Dolgova, N.V.; Gagnon, K.B.; Sopasis, G.J.; James, A.K.; Lai, B.; Sylvain, N.J.; Harris, H.H.; Nichol, H.K.; George, G.N.; et al. PBT2 Acts through a Different Mechanism of Action than Other 8-Hydroxyquinolines: An X-ray Fluorescence Imaging Study. Metallomics 2020, 12, 1979–1994. [Google Scholar] [CrossRef] [PubMed]
  302. Chen, S.; Sun, L.; Koya, K.; Tatsuta, N.; Xia, Z.; Korbut, T.; Du, Z.; Wu, J.; Liang, G.; Jiang, J.; et al. Syntheses and Antitumor Activities of N′ 1, N′ 3-Dialkyl-N′ 1, N′ 3-Di-(Alkylcarbonothioyl) Malonohydrazide: The Discovery of Elesclomol. Bioorganic Med. Chem. Lett. 2013, 23, 5070–5076. [Google Scholar] [CrossRef] [PubMed]
  303. Kwan, S.-Y.; Cheng, X.; Tsang, Y.T.M.; Choi, J.-S.; Kwan, S.-Y.; Izaguirre, D.I.; Kwan, H.-S.; Gershenson, D.M.; Wong, K.-K. Loss of ARID1A Expression Leads to Sensitivity to ROS-Inducing Agent Elesclomol in Gynecologic Cancer Cells. Oncotarget 2016, 7, 56933–56943. [Google Scholar] [CrossRef]
  304. Buccarelli, M.; D’Alessandris, Q.G.; Matarrese, P.; Mollinari, C.; Signore, M.; Cappannini, A.; Martini, M.; D’Aliberti, P.; De Luca, G.; Pedini, F.; et al. Elesclomol-Induced Increase of Mitochondrial Reactive Oxygen Species Impairs Glioblastoma Stem-like Cell Survival and Tumor Growth. J. Exp. Clin. Cancer Res. 2021, 40, 228. [Google Scholar] [CrossRef]
  305. Nagai, M.; Vo, N.H.; Ogawa, L.S.; Chimmanamada, D.; Inoue, T.; Chu, J.; Beaudette-Zlatanova, B.C.; Lu, R.; Blackman, R.K.; Barsoum, J.; et al. The Oncology Drug Elesclomol Selectively Transports Copper to the Mitochondria to Induce Oxidative Stress in Cancer Cells. Free Radic. Biol. Med. 2012, 52, 2142–2150. [Google Scholar] [CrossRef]
  306. Hasinoff, B.B.; Yadav, A.A.; Patel, D.; Wu, X. The Cytotoxicity of the Anticancer Drug Elesclomol Is Due to Oxidative Stress Indirectly Mediated through Its Complex with Cu (II). J. Inorg. Biochem. 2014, 137, 22–30. [Google Scholar] [CrossRef] [PubMed]
  307. Tsvetkov, P.; Detappe, A.; Cai, K.; Keys, H.R.; Brune, Z.; Ying, W.; Thiru, P.; Reidy, M.; Kugener, G.; Rossen, J.; et al. Mitochondrial Metabolism Promotes Adaptation to Proteotoxic Stress. Nat. Chem. Biol. 2019, 15, 681–689. [Google Scholar] [CrossRef] [PubMed]
  308. Yadav, A.A.; Patel, D.; Wu, X.; Hasinoff, B.B. Molecular Mechanisms of the Biological Activity of the Anticancer Drug Elesclomol and Its Complexes with Cu (II), Ni (II) and Pt (II). J. Inorg. Biochem. 2013, 126, 1–6. [Google Scholar] [CrossRef] [PubMed]
  309. Lu, J.; Ling, X.; Sun, Y.; Liu, L.; Liu, L.; Wang, X.; Lu, C.; Ren, C.; Han, X.; Yu, Z. FDX1 Enhances Endometriosis Cell Cuproptosis via G6PD-Mediated Redox Homeostasis. Apoptosis 2023, 28, 1128–1140. [Google Scholar] [CrossRef] [PubMed]
  310. Takeda, T.; Banno, K.; Okawa, R.; Yanokura, M.; Iijima, M.; Irie-Kunitomi, H.; Nakamura, K.; Iida, M.; Adachi, M.; Umene, K.; et al. ARID1A Gene Mutation in Ovarian and Endometrial Cancers. Oncol. Rep. 2016, 35, 607–613. [Google Scholar] [CrossRef]
  311. Nie, X.; Chen, H.; Xiong, Y.; Chen, J.; Liu, T. Anisomycin Has a Potential Toxicity of Promoting Cuproptosis in Human Ovarian Cancer Stem Cells by Attenuating YY1/Lipoic Acid Pathway Activation. J. Cancer 2022, 13, 3503–3514. [Google Scholar] [CrossRef]
  312. Harrington, B.S.; Ozaki, M.K.; Caminear, M.W.; Hernandez, L.F.; Jordan, E.; Kalinowski, N.J.; Goldlust, I.S.; Guha, R.; Ferrer, M.; Thomas, C.; et al. Drugs Targeting Tumor-Initiating Cells Prolong Survival in a Post-Surgery, Post-Chemotherapy Ovarian Cancer Relapse Model. Cancers 2020, 12, 1645. [Google Scholar] [CrossRef]
  313. Monk, B.J.; Kauderer, J.T.; Moxley, K.M.; Bonebrake, A.J.; Dewdney, S.B.; Secord, A.A.; Ueland, F.R.; Johnston, C.M.; Aghajanian, C. A Phase II Evaluation of Elesclomol Sodium and Weekly Paclitaxel in the Treatment of Recurrent or Persistent Platinum-Resistant Ovarian, Fallopian Tube or Primary Peritoneal Cancer: An NRG Oncology/Gynecologic Oncology Group Study. Gynecol. Oncol. 2018, 151, 422–427. [Google Scholar] [CrossRef]
  314. Hedley, D.; Shamas-Din, A.; Chow, S.; Sanfelice, D.; Schuh, A.C.; Brandwein, J.M.; Seftel, M.D.; Gupta, V.; Yee, K.W.L.; Schimmer, A.D. A Phase I Study of Elesclomol Sodium in Patients with Acute Myeloid Leukemia. Leuk. Lymphoma 2016, 57, 2437–2440. [Google Scholar] [CrossRef]
  315. O’Day, S.; Gonzalez, R.; Lawson, D.; Weber, R.; Hutchins, L.; Anderson, C.; Haddad, J.; Kong, S.; Williams, A.; Jacobson, E. Phase II, Randomized, Controlled, Double-Blinded Trial of Weekly Elesclomol plus Paclitaxel versus Paclitaxel Alone for Stage IV Metastatic Melanoma. J. Clin. Oncol. 2009, 27, 5452–5458. [Google Scholar] [CrossRef]
  316. O’Day, S.J.; Eggermont, A.M.M.; Chiarion-Sileni, V.; Kefford, R.; Grob, J.J.; Mortier, L.; Robert, C.; Schachter, J.; Testori, A.; Mackiewicz, J.; et al. Final Results of Phase III SYMMETRY Study: Randomized, Double-Blind Trial of Elesclomol plus Paclitaxel versus Paclitaxel Alone as Treatment for Chemotherapy-Naive Patients with Advanced Melanoma. J. Clin. Oncol. 2013, 31, 1211–1218. [Google Scholar] [CrossRef] [PubMed]
  317. Zheng, P.; Zhou, C.; Lu, L.; Liu, B.; Ding, Y. Elesclomol: A Copper Ionophore Targeting Mitochondrial Metabolism for Cancer Therapy. J. Exp. Clin. Cancer Res. 2022, 41, 271. [Google Scholar] [CrossRef] [PubMed]
  318. Helsel, M.E.; Franz, K.J. Pharmacological Activity of Metal Binding Agents That Alter Copper Bioavailability. Dalt. Trans. 2015, 44, 8760–8770. [Google Scholar] [CrossRef] [PubMed]
  319. Xiao, Z.; Donnelly, P.S.; Zimmermann, M.; Wedd, A.G. Transfer of Copper between Bis (Thiosemicarbazone) Ligands and Intracellular Copper-Binding Proteins. Insights into Mechanisms of Copper Uptake and Hypoxia Selectivity. Inorg. Chem. 2008, 47, 4338–4347. [Google Scholar] [CrossRef]
  320. Cater, M.A.; Pearson, H.B.; Wolyniec, K.; Klaver, P.; Bilandzic, M.; Paterson, B.M.; Bush, A.I.; Humbert, P.O.; La Fontaine, S.; Donnelly, P.S.; et al. Increasing Intracellular Bioavailable Copper Selectively Targets Prostate Cancer Cells. ACS Chem. Biol. 2013, 8, 1621–1631. [Google Scholar] [CrossRef]
  321. Donnelly, P.S.; Liddell, J.R.; Lim, S.; Paterson, B.M.; Cater, M.A.; Savva, M.S.; Mot, A.I.; James, J.L.; Trounce, I.A.; White, A.R.; et al. An Impaired Mitochondrial Electron Transport Chain Increases Retention of the Hypoxia Imaging Agent Diacetylbis (4-Methylthiosemicarbazonato) CopperII. Proc. Natl. Acad. Sci. USA 2012, 109, 47–52. [Google Scholar] [CrossRef]
  322. Holland, J.P.; Barnard, P.J.; Collison, D.; Dilworth, J.R.; Edge, R.; Green, J.C.; McInnes, E.J.L. Spectroelectrochemical and Computational Studies on the Mechanism of Hypoxia Selectivity of Copper Radiopharmaceuticals. Chem. Eur. J. 2008, 14, 5890–5907. [Google Scholar] [CrossRef]
  323. Paterson, B.M.; Donnelly, P.S. Copper Complexes of Bis (Thiosemicarbazones): From Chemotherapeutics to Diagnostic and Therapeutic Radiopharmaceuticals. Chem. Soc. Rev. 2011, 40, 3005–3018. [Google Scholar] [CrossRef]
  324. Boschi, A.; Martini, P.; Janevik-Ivanovska, E.; Duatti, A. The Emerging Role of Copper-64 Radiopharmaceuticals as Cancer Theranostics. Drug Discov. Today 2018, 23, 1489–1501. [Google Scholar] [CrossRef]
  325. Lewis, J.S.; Laforest, R.; Buettner, T.L.; Song, S.-K.; Fujibayashi, Y.; Connett, J.M.; Welch, M.J. Copper-64-Diacetyl-Bis (N 4-Methylthiosemicarbazone): An Agent for Radiotherapy. Proc. Natl. Acad. Sci. USA 2001, 98, 1206–1211. [Google Scholar] [CrossRef]
  326. Dehdashti, F.; Grigsby, P.W.; Lewis, J.S.; Laforest, R.; Siegel, B.A.; Welch, M.J. Assessing Tumor Hypoxia in Cervical Cancer by PET with 60Cu-Labeled Diacetyl-Bis (N4-Methylthiosemicarbazone). J. Nucl. Med. 2008, 49, 201–205. [Google Scholar] [CrossRef] [PubMed]
  327. Grigsby, P.W.; Malyapa, R.S.; Higashikubo, R.; Schwarz, J.K.; Welch, M.J.; Huettner, P.C.; Dehdashti, F. Comparison of Molecular Markers of Hypoxia and Imaging with 60 Cu-ATSM in Cancer of the Uterine Cervix. Mol. Imaging Biol. 2007, 9, 278–283. [Google Scholar] [CrossRef] [PubMed]
  328. Lewis, J.S.; Laforest, R.; Dehdashti, F.; Grigsby, P.W.; Welch, M.J.; Siegel, B.A. An Imaging Comparison of 64Cu-ATSM and 60Cu-ATSM in Cancer of the Uterine Cervix. J. Nucl. Med. 2008, 49, 1177–1182. [Google Scholar] [CrossRef]
  329. Santini, C.; Pellei, M.; Gandin, V.; Porchia, M.; Tisato, F.; Marzano, C. Advances in Copper Complexes as Anticancer Agents. Chem. Rev. 2014, 114, 815–862. [Google Scholar] [CrossRef] [PubMed]
  330. Chaturvedi, V.K.; Singh, A.; Singh, V.K.; Singh, M.P. Cancer Nanotechnology: A New Revolution for Cancer Diagnosis and Therapy. Curr. Drug Metab. 2019, 20, 416–429. [Google Scholar] [CrossRef]
  331. Ameh, T.; Sayes, C.M. The Potential Exposure and Hazards of Copper Nanoparticles: A Review. Environ. Toxicol. Pharmacol. 2019, 71, 103220. [Google Scholar] [CrossRef]
  332. Dou, L.; Zhang, X.; Zangeneh, M.M.; Zhang, Y. Efficient Biogenesis of Cu2O Nanoparticles Using Extract of Camellia Sinensis Leaf: Evaluation of Catalytic, Cytotoxicity, Antioxidant, and Anti-Human Ovarian Cancer Properties. Bioorg. Chem. 2021, 106, 104468. [Google Scholar] [CrossRef]
  333. Tabrez, S.; Khan, A.U.; Mirza, A.A.; Suhail, M.; Jabir, N.R.; Zughaibi, T.A.; Alam, M. Biosynthesis of Copper Oxide Nanoparticles and Its Therapeutic Efficacy against Colon Cancer. Nanotechnol. Rev. 2022, 11, 1322–1331. [Google Scholar] [CrossRef]
  334. Zhao, H.; Maruthupandy, M.; Al-mekhlafi, F.A.; Chackaravarthi, G.; Ramachandran, G.; Chelliah, C.K. Biological Synthesis of Copper Oxide Nanoparticles Using Marine Endophytic Actinomycetes and Evaluation of Biofilm Producing Bacteria and A549 Lung Cancer Cells. J. King Saud Univ. 2022, 34, 101866. [Google Scholar] [CrossRef]
  335. Zughaibi, T.A.; Jabir, N.R.; Khan, A.U.; Khan, M.S.; Tabrez, S. Screening of Cu4O3 NPs Efficacy and Its Anticancer Potential against Cervical Cancer. Cell Biochem. Funct. 2023, 41, 1174–1187. [Google Scholar] [CrossRef]
  336. Chen, H.; Feng, X.; Gao, L.; Mickymaray, S.; Paramasivam, A.; Abdulaziz Alfaiz, F.; Almasmoum, H.A.; Ghaith, M.M.; Almaimani, R.A.; Aziz Ibrahim, I.A. Inhibiting the PI3K/AKT/MTOR Signalling Pathway with Copper Oxide Nanoparticles from Houttuynia Cordata Plant: Attenuating the Proliferation of Cervical Cancer Cells. Artif. Cells Nanomed. Biotechnol. 2021, 49, 240–249. [Google Scholar] [CrossRef] [PubMed]
  337. Goswami, U.; Dutta, A.; Raza, A.; Kandimalla, R.; Kalita, S.; Ghosh, S.S.; Chattopadhyay, A. Transferrin–Copper Nanocluster–Doxorubicin Nanoparticles as Targeted Theranostic Cancer Nanodrug. ACS Appl. Mater. Interfaces 2018, 10, 3282–3294. [Google Scholar] [CrossRef] [PubMed]
  338. Aboeita, N.M.; Fahmy, S.A.; El-Sayed, M.M.H.; Azzazy, H.M.E.-S.; Shoeib, T. Enhanced Anticancer Activity of Nedaplatin Loaded onto Copper Nanoparticles Synthesized Using Red Algae. Pharmaceutics 2022, 14, 418. [Google Scholar] [CrossRef] [PubMed]
  339. Li, N.; Sun, Q.; Yu, Z.; Gao, X.; Pan, W.; Wan, X.; Tang, B. Nuclear-Targeted Photothermal Therapy Prevents Cancer Recurrence with near-Infrared Triggered Copper Sulfide Nanoparticles. ACS Nano 2018, 12, 5197–5206. [Google Scholar] [CrossRef] [PubMed]
  340. Zhou, M.; Melancon, M.; Stafford, R.J.; Li, J.; Nick, A.M.; Tian, M.; Sood, A.K.; Li, C. Precision Nanomedicine Using Dual PET and MR Temperature Imaging–Guided Photothermal Therapy. J. Nucl. Med. 2016, 57, 1778–1783. [Google Scholar] [CrossRef] [PubMed]
  341. Shah, M.; Murad, W.; Mubin, S.; Ullah, O.; Rehman, N.U.; Rahman, M.H. Multiple Health Benefits of Curcumin and Its Therapeutic Potential. Environ. Sci. Pollut. Res. 2022, 29, 43732–43744. [Google Scholar] [CrossRef] [PubMed]
  342. Yang, Y.; Liang, S.; Geng, H.; Xiong, M.; Li, M.; Su, Q.; Jia, F.; Zhao, Y.; Wang, K.; Jiang, J.; et al. Proteomics Revealed the Crosstalk between Copper Stress and Cuproptosis, and Explored the Feasibility of Curcumin as Anticancer Copper Ionophore. Free Radic. Biol. Med. 2022, 193, 638–647. [Google Scholar] [CrossRef]
  343. Giordano, A.; Tommonaro, G. Curcumin and Cancer. Nutrients 2019, 11, 2376. [Google Scholar] [CrossRef]
  344. Pourhanifeh, M.H.; Darvish, M.; Tabatabaeian, J.; Fard, M.R.; Mottaghi, R.; Azadchehr, M.J.; Jahanshahi, M.; Sahebkar, A.; Mirzaei, H. Therapeutic Role of Curcumin and Its Novel Formulations in Gynecological Cancers. J. Ovarian Res. 2020, 13, 130. [Google Scholar] [CrossRef]
  345. Greish, K.; Pittalà, V.; Taurin, S.; Taha, S.; Bahman, F.; Mathur, A.; Jasim, A.; Mohammed, F.; El-Deeb, I.M.; Fredericks, S.; et al. Curcumin–Copper Complex Nanoparticles for the Management of Triple-Negative Breast Cancer. Nanomaterials 2018, 8, 884. [Google Scholar] [CrossRef]
  346. Luo, C.-Q.; Xing, L.; Cui, P.-F.; Qiao, J.-B.; He, Y.-J.; Chen, B.-A.; Jin, L.; Jiang, H.-L. Curcumin-Coordinated Nanoparticles with Improved Stability for Reactive Oxygen Species-Responsive Drug Delivery in Lung Cancer Therapy. Int. J. Nanomed. 2017, 12, 855–869. [Google Scholar] [CrossRef]
  347. Sirohi, V.K.; Popli, P.; Sankhwar, P.; Kaushal, J.B.; Gupta, K.; Manohar, M.; Dwivedi, A. Curcumin Exhibits Anti-Tumor Effect and Attenuates Cellular Migration via Slit-2 Mediated down-Regulation of SDF-1 and CXCR4 in Endometrial Adenocarcinoma Cells. J. Nutr. Biochem. 2017, 44, 60–70. [Google Scholar] [CrossRef] [PubMed]
  348. Chen, Q.; Gao, Q.; Chen, K.; Wang, Y.; Chen, L.; Li, X.U. Curcumin Suppresses Migration and Invasion of Human Endometrial Carcinoma Cells. Oncol. Lett. 2015, 10, 1297–1302. [Google Scholar] [CrossRef]
  349. Sun, M.-X.; Yu, F.; Gong, M.-L.; Fan, G.-L.; Liu, C.-X. Effects of Curcumin on the Role of MMP-2 in Endometrial Cancer Cell Proliferation and Invasion. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 5033–5041. [Google Scholar] [CrossRef] [PubMed]
  350. Ghasemi, F.; Shafiee, M.; Banikazemi, Z.; Pourhanifeh, M.H.; Khanbabaei, H.; Shamshirian, A.; Moghadam, S.A.; ArefNezhad, R.; Sahebkar, A.; Avan, A.; et al. Curcumin Inhibits NF-KB and Wnt/β-Catenin Pathways in Cervical Cancer Cells. Pathol. Pract. 2019, 215, 152556. [Google Scholar] [CrossRef] [PubMed]
  351. Kim, B.; Kim, H.S.; Jung, E.-J.; Lee, J.Y.; Tsang, B.K.; Lim, J.M.; Song, Y.S. Curcumin Induces ER Stress-Mediated Apoptosis through Selective Generation of Reactive Oxygen Species in Cervical Cancer Cells. Mol. Carcinog. 2016, 55, 918–928. [Google Scholar] [CrossRef] [PubMed]
  352. Liu, X.; Qi, M.; Li, X.; Wang, J.; Wang, M. Curcumin: A Natural Organic Component That Plays a Multi-Faceted Role in Ovarian Cancer. J. Ovarian Res. 2023, 16, 47. [Google Scholar] [CrossRef]
  353. Reddy, P.S.; Begum, N.; Mutha, S.; Bakshi, V. Beneficial Effect of Curcumin in Letrozole Induced Polycystic Ovary Syndrome. Asian Pac. J. Reprod. 2016, 5, 116–122. [Google Scholar] [CrossRef]
  354. Mohammadi, S.; Kayedpoor, P.; Karimzadeh-Bardei, L.; Nabiuni, M. The Effect of Curcumin on TNF-α, IL-6 and CRP Expression in a Model of Polycystic Ovary Syndrome as an Inflammation State. J. Reprod. Infertil. 2017, 18, 352–360. [Google Scholar]
  355. Abuelezz, N.Z.; Shabana, M.E.; Abdel-Mageed, H.M.; Rashed, L.; Morcos, G.N.B. Nanocurcumin Alleviates Insulin Resistance and Pancreatic Deficits in Polycystic Ovary Syndrome Rats: Insights on PI3K/AkT/MTOR and TNF-α Modulations. Life Sci. 2020, 256, 118003. [Google Scholar] [CrossRef]
  356. Jamilian, M.; Foroozanfard, F.; Kavossian, E.; Aghadavod, E.; Shafabakhsh, R.; Hoseini, A.; Asemi, Z. Effects of Curcumin on Body Weight, Glycemic Control and Serum Lipids in Women with Polycystic Ovary Syndrome: A Randomized, Double-Blind, Placebo-Controlled Trial. Clin. Nutr. ESPEN 2020, 36, 128–133. [Google Scholar] [CrossRef] [PubMed]
  357. Sohaei, S.; Amani, R.; Tarrahi, M.J.; Ghasemi-Tehrani, H. The Effects of Curcumin Supplementation on Glycemic Status, Lipid Profile and Hs-CRP Levels in Overweight/Obese Women with Polycystic Ovary Syndrome: A Randomized, Double-Blind, Placebo-Controlled Clinical Trial. Complement. Ther. Med. 2019, 47, 102201. [Google Scholar] [CrossRef] [PubMed]
  358. Kim, K.-H.; Lee, E.N.; Park, J.K.; Lee, J.-R.; Kim, J.-H.; Choi, H.-J.; Kim, B.-S.; Lee, H.-W.; Lee, K.-S.; Yoon, S. Curcumin Attenuates TNF-α-Induced Expression of Intercellular Adhesion Molecule-1, Vascular Cell Adhesion Molecule-1 and Proinflammatory Cytokines in Human Endometriotic Stromal Cells. Phyther. Res. 2012, 26, 1037–1047. [Google Scholar] [CrossRef] [PubMed]
  359. Ding, J.; Mei, S.; Cheng, W.; Ni, Z.; Yu, C. Curcumin Treats Endometriosis in Mice by the HIF Signaling Pathway. Am. J. Transl. Res. 2022, 14, 2184–2198. [Google Scholar] [PubMed]
  360. Chowdhury, I.; Banerjee, S.; Driss, A.; Xu, W.; Mehrabi, S.; Nezhat, C.; Sidell, N.; Taylor, R.N.; Thompson, W.E. Curcumin Attenuates Proangiogenic and Proinflammatory Factors in Human Eutopic Endometrial Stromal Cells through the NF-ΚB Signaling Pathway. J. Cell. Physiol. 2019, 234, 6298–6312. [Google Scholar] [CrossRef] [PubMed]
  361. Cao, H.; Wei, Y.-X.; Zhou, Q.; Zhang, Y.; Guo, X.-P.; Zhang, J. Inhibitory Effect of Curcumin in Human Endometriosis Endometrial Cells via Downregulation of Vascular Endothelial Growth Factor. Mol. Med. Rep. 2017, 16, 5611–5617. [Google Scholar] [CrossRef] [PubMed]
  362. Zhang, Y.; Cao, H.; Yu, Z.; Peng, H.-Y.; Zhang, C. Curcumin Inhibits Endometriosis Endometrial Cells by Reducing Estradiol Production. Iran. J. Reprod. Med. 2013, 11, 415–422. [Google Scholar] [PubMed]
  363. Jana, S.; Paul, S.; Swarnakar, S. Curcumin as Anti-Endometriotic Agent: Implication of MMP-3 and Intrinsic Apoptotic Pathway. Biochem. Pharmacol. 2012, 83, 797–804. [Google Scholar] [CrossRef]
  364. Fadin, M.; Nicoletti, M.C.; Pellizzato, M.; Accardi, M.; Baietti, M.G.; Fratter, A. Effectiveness of the Integration of Quercetin, Turmeric, and N-Acetylcysteine in Reducing Inflammation and Pain Associated with Endometriosis. In-Vitro and in-Vivo Studies. Minerva Ginecol. 2020, 72, 285–291. [Google Scholar] [CrossRef]
  365. Meresman, G.F.; Götte, M.; Laschke, M.W. Plants as Source of New Therapies for Endometriosis: A Review of Preclinical and Clinical Studies. Hum. Reprod. Update 2021, 27, 367–392. [Google Scholar] [CrossRef]
  366. Zhu, J.-J.; Jiang, J.-G. Pharmacological and Nutritional Effects of Natural Coumarins and Their Structure–Activity Relationships. Mol. Nutr. Food Res. 2018, 62, e1701073. [Google Scholar] [CrossRef] [PubMed]
  367. Pivetta, T.; Valletta, E.; Ferino, G.; Isaia, F.; Pani, A.; Vascellari, S.; Castellano, C.; Demartin, F.; Cabiddu, M.G.; Cadoni, E. Novel Coumarins and Related Copper Complexes with Biological Activity: DNA Binding, Molecular Docking and in Vitro Antiproliferative Activity. J. Inorg. Biochem. 2017, 177, 101–109. [Google Scholar] [CrossRef] [PubMed]
  368. Khan, S.; Zafar, A.; Naseem, I. Redox Cycling of Copper by Coumarin-Di (2-Picolyl) Amine Hybrid Molecule Leads to ROS-Mediated Modulation of Redox Scavengers, DNA Damage and Cell Death in Diethylnitrosamine Induced Hepatocellular Carcinoma. Bioorg. Chem. 2020, 99, 103818. [Google Scholar] [CrossRef] [PubMed]
  369. Stepanenko, I.; Babak, M.V.; Spengler, G.; Hammerstad, M.; Popovic-Bijelic, A.; Shova, S.; Büchel, G.E.; Darvasiova, D.; Rapta, P.; Arion, V.B. Coumarin-Based Triapine Derivatives and Their Copper (II) Complexes: Synthesis, Cytotoxicity and MR2 RNR Inhibition Activity. Biomolecules 2021, 11, 862. [Google Scholar] [CrossRef] [PubMed]
  370. Lu, W.; Tang, J.; Gu, Z.; Sun, L.; Wei, H.; Wang, Y.; Yang, S.; Chi, X.; Xu, L. Crystal Structure, in Vitro Cytotoxicity, DNA Binding and DFT Calculations of New Copper (II) Complexes with Coumarin-Amide Ligand. J. Inorg. Biochem. 2023, 238, 112030. [Google Scholar] [CrossRef]
  371. An, G.; Park, S.; Lee, M.; Lim, W.; Song, G. Antiproliferative Effect of 4-Methylumbelliferone in Epithelial Ovarian Cancer Cells Is Mediated by Disruption of Intracellular Homeostasis and Regulation of PI3K/AKT and MAPK Signaling. Pharmaceutics 2020, 12, 640. [Google Scholar] [CrossRef]
  372. Liang, J.; Zhou, J.; Xu, Y.; Huang, X.; Wang, X.; Huang, W.; Li, H. Osthole Inhibits Ovarian Carcinoma Cells through LC3-Mediated Autophagy and GSDME-Dependent Pyroptosis except for Apoptosis. Eur. J. Pharmacol. 2020, 874, 172990. [Google Scholar] [CrossRef]
  373. Che, Y.; Li, J.; Li, Z.; Li, J.; Wang, S.; Yan, Y.; Zou, K.; Zou, L. Osthole Enhances Antitumor Activity and Irradiation Sensitivity of Cervical Cancer Cells by Suppressing ATM/NF-ΚB Signaling. Oncol. Rep. 2018, 40, 737–747. [Google Scholar] [CrossRef]
  374. Su, J.; Zhang, F.; Li, X.; Liu, Z. Osthole Promotes the Suppressive Effects of Cisplatin on NRF2 Expression to Prevent Drug-Resistant Cervical Cancer Progression. Biochem. Biophys. Res. Commun. 2019, 514, 510–517. [Google Scholar] [CrossRef]
  375. Zhu, M.-L.; Li, J.-C.; Wang, L.; Zhong, X.; Zhang, Y.-W.; Tan, R.-Z.; Wang, H.-L.; Fan, J.-M.; Wang, L. Decursin Inhibits the Growth of HeLa Cervical Cancer Cells through PI3K/Akt Signaling. J. Asian Nat. Prod. Res. 2021, 23, 584–595. [Google Scholar] [CrossRef]
  376. Tian, Q.; Wang, L.; Sun, X.; Zeng, F.; Pan, Q.; Xue, M. Scopoletin Exerts Anticancer Effects on Human Cervical Cancer Cell Lines by Triggering Apoptosis, Cell Cycle Arrest, Inhibition of Cell Invasion and PI3K/AKT Signalling Pathway. J BUON 2019, 24, 997–1002. [Google Scholar] [PubMed]
  377. Ma, T.; Liu, P.; Wei, J.; Zhao, M.; Yao, X.; Luo, X.; Xu, S. Imperatorin Alleviated Endometriosis by Inhibiting the Activation of PI3K/Akt/NF-ΚB Pathway in Rats. Life Sci. 2021, 274, 119291. [Google Scholar] [CrossRef] [PubMed]
  378. Abizadeh, M.; Novin, M.G.; Amidi, F.; Ziaei, S.A.; Abdollahifar, M.A.; Nazarian, H. Potential of Auraptene in Improvement of Oocyte Maturation, Fertilization Rate, and Inflammation in Polycystic Ovary Syndrome Mouse Model. Reprod. Sci. 2020, 27, 1742–1751. [Google Scholar] [CrossRef]
  379. Kim, Y.H.; Lee, S.Y.; Kim, E.Y.; Kim, K.H.; Koong, M.K.; Lee, K.A. The Antioxidant Auraptene Improves Aged Oocyte Quality and Embryo Development in Mice. Antioxidants 2023, 12, 87. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram of copper metabolism in mammals. After intestinal absorption, Cu travels through the portal vein bound to soluble proteins, such as albumin and transcuprein. On the surface of mammalian cells, metalloreductases such as STEAP 2, 3, and 4 reduce Cu2+ ions to Cu+ so that cells can absorb Cu through CTR1. (A) In the mitochondrial intermembrane space, COX17 is responsible for delivering Cu+ to either SCO1 or COX11 to contribute to the correct assembly of CCO, which utilizes Cu for energy production through oxidative phosphorylation. (B) CCS chaperone transfers Cu+ to SOD1, which is critical in the defense against oxidative stress because it catalyzes the degradation of superoxide radicals. (C) ATOX1 is responsible for providing Cu to the ATPases (ATP7A and ATP7B) that are principally located in the trans-Golgi network (TGN). ATPases pump Cu+ from the cytosol into the lumen of the TGN to promote the synthesis of cuproenzymes, such as Cp, LOX, and SOD3, which are secreted out of the cells to mediate the Cu transport through the circulatory system. (D) Since free Cu ions have the potential to generate reactive oxygen species, excess intracellular Cu+ is sequestered mainly by glutathione (GSH) and metallothioneins (MTs) that uptake Cu for storage. GSH can also deliver Cu to MTs. (E) When the cytoplasmic Cu concentration increases, ATP7A and ATP7B move within endocytic vesicles toward the plasma membrane to transfer excess Cu into the bloodstream. ATP7A is expressed in many tissues except in the liver, where it is replaced by ATP7B. In hepatocytes, ATP7B ensures the movement of Cu through the canalicular membrane for its subsequent elimination through the bile. (F) The concentration of mammalian CTR1 at the plasma membrane is negatively regulated in response to elevated Cu levels (red dotted arrow), with CTR1 being removed from the cell surface. (G) ATOX1 can carry Cu into the cell nucleus and act as a transcription factor for the expression of genes encoding cyclin D1 and SOD3 (green dotted arrow). High concentrations of cellular Cu may also stimulate the transcription of MT genes. Created with BioRender.com (accessed on 11 November 2023).
Figure 1. Schematic diagram of copper metabolism in mammals. After intestinal absorption, Cu travels through the portal vein bound to soluble proteins, such as albumin and transcuprein. On the surface of mammalian cells, metalloreductases such as STEAP 2, 3, and 4 reduce Cu2+ ions to Cu+ so that cells can absorb Cu through CTR1. (A) In the mitochondrial intermembrane space, COX17 is responsible for delivering Cu+ to either SCO1 or COX11 to contribute to the correct assembly of CCO, which utilizes Cu for energy production through oxidative phosphorylation. (B) CCS chaperone transfers Cu+ to SOD1, which is critical in the defense against oxidative stress because it catalyzes the degradation of superoxide radicals. (C) ATOX1 is responsible for providing Cu to the ATPases (ATP7A and ATP7B) that are principally located in the trans-Golgi network (TGN). ATPases pump Cu+ from the cytosol into the lumen of the TGN to promote the synthesis of cuproenzymes, such as Cp, LOX, and SOD3, which are secreted out of the cells to mediate the Cu transport through the circulatory system. (D) Since free Cu ions have the potential to generate reactive oxygen species, excess intracellular Cu+ is sequestered mainly by glutathione (GSH) and metallothioneins (MTs) that uptake Cu for storage. GSH can also deliver Cu to MTs. (E) When the cytoplasmic Cu concentration increases, ATP7A and ATP7B move within endocytic vesicles toward the plasma membrane to transfer excess Cu into the bloodstream. ATP7A is expressed in many tissues except in the liver, where it is replaced by ATP7B. In hepatocytes, ATP7B ensures the movement of Cu through the canalicular membrane for its subsequent elimination through the bile. (F) The concentration of mammalian CTR1 at the plasma membrane is negatively regulated in response to elevated Cu levels (red dotted arrow), with CTR1 being removed from the cell surface. (G) ATOX1 can carry Cu into the cell nucleus and act as a transcription factor for the expression of genes encoding cyclin D1 and SOD3 (green dotted arrow). High concentrations of cellular Cu may also stimulate the transcription of MT genes. Created with BioRender.com (accessed on 11 November 2023).
Ijms 24 17578 g001
Table 1. Functions of the main cuproenzymes.
Table 1. Functions of the main cuproenzymes.
CuproenzymeFunction
LOXRequired for the formation of the extracellular matrix.
SODCatalyzes the conversion of superoxide radicals to molecular oxygen and hydrogen peroxide.
CpMulticopper ferroxidase; principal Cu carrier in serum.
HephaestinMulticopper ferroxidase. It supports the transportation of Fe released from intestinal enterocytes.
CCOElectron transfer protein. It catalyzes ATP production.
TyrosinaseCatalyzes phenol oxidation; it is required for melanin synthesis, a fundamental pigment for hair, skin, and eyes.
DβHOxidoreductase. It catalyzes the conversion of dopamine to epinephrine.
MEKKinases that belong to the mitogen-activated protein kinase cascade and that mainly promote cell proliferation and survival.
ULK1/2Autophagy-initiating kinases.
MEMO1Regulation of cell motility and ROS production.
Table 2. Examples of Cu chelators.
Table 2. Examples of Cu chelators.
Compound NameChemical FormulaStructural Formula 1
D-penicillamineC5H11NO2SIjms 24 17578 i001
TrientineC6H18N4Ijms 24 17578 i002
TetrathiomolybdateMoS4Ijms 24 17578 i003
1 Structural formulas were obtained from the DrugBank public database (http://www.drugbank.com/), accessed on 11 November 2023.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Conforti, R.A.; Delsouc, M.B.; Zorychta, E.; Telleria, C.M.; Casais, M. Copper in Gynecological Diseases. Int. J. Mol. Sci. 2023, 24, 17578. https://doi.org/10.3390/ijms242417578

AMA Style

Conforti RA, Delsouc MB, Zorychta E, Telleria CM, Casais M. Copper in Gynecological Diseases. International Journal of Molecular Sciences. 2023; 24(24):17578. https://doi.org/10.3390/ijms242417578

Chicago/Turabian Style

Conforti, Rocío A., María B. Delsouc, Edith Zorychta, Carlos M. Telleria, and Marilina Casais. 2023. "Copper in Gynecological Diseases" International Journal of Molecular Sciences 24, no. 24: 17578. https://doi.org/10.3390/ijms242417578

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