Abstract
Copper, a trace element and vital cofactor, plays a crucial role in the maintenance of biological functions. Recent evidence has established significant correlations between copper levels, cancer development and metastasis. The strong redox‐active properties of copper offer both benefits and disadvantages to cancer cells. The intestinal tract, which is primarily responsible for copper uptake and regulation, may suffer from an imbalance in copper homeostasis. Colorectal cancer (CRC) is the most prevalent primary cancer of the intestinal tract and is an aggressive malignant disease with limited therapeutic options. Current research is primarily focused on the relationship between copper and CRC. Innovative concepts, such as cuproplasia and cuproptosis, are being explored to understand copper‐related cellular proliferation and death. Cuproplasia is the regulation of cell proliferation that is mediated by both enzymatic and nonenzymatic copper‐modulated activities. Whereas, cuproptosis refers to cell death induced by excess copper via promoting the abnormal oligomerisation of lipoylated proteins within the tricarboxylic acid cycle, as well as by diminishing the levels of iron‐sulphur cluster proteins. A comprehensive understanding of copper‐related cellular proliferation and death mechanisms offers new avenues for CRC treatment. In this review, we summarise the evolving molecular mechanisms, ranging from abnormal intracellular copper concentrations to the copper‐related proteins that are being discovered, and discuss the role of copper in the pathogenesis, progression and potential therapies for CRC. Understanding the relationship between copper and CRC will help provide a comprehensive theoretical foundation for innovative treatment strategies in CRC management.
Keywords: cancer therapy, colorectal cancer, copper, cuproptosis
Physiological state copper ion transport process in vivo
The role of copper ions in the carcinogenesis and development of colorectal cancer
Copper ions mediate a variety of cell death modes
The role of targeted copper ions in the diagnosis and treatment of colorectal cancer

1. INTRODUCTION
Colorectal cancer (CRC) is the most common primary intestinal cancer and its incidence continues to increase. Currently, it is the third most frequent and second leading cause of cancer‐related mortality worldwide. 1 CRC is characterised by multistage progression, often associated with poor prognosis, and frequently diagnosed in advanced stages. 2 The proliferation and metastasis of CRC is multifactorial, including genetic predispositions, inflammatory bowel disease, dietary pattern with excessive meat intake and lifestyle factors. 3 Recent studies have shown that copper ions are absorbed through the gastrointestinal tract and are present at elevated levels in CRC, mediating its carcinogenesis and development. Furthermore, copper ions can also trigger several types of cell death. Therefore, understanding the mechanism of action of copper ions in CRC and targeting them are crucial for the diagnosis and treatment of CRC.
A considerable number of studies have shown that copper, as a trace element in the human body, plays a vital role in various biological signalling pathways, which is mainly attributed to its redox‐active properties. 4 The intricate copper homeostasis arises from the regulation of intracellular copper levels through a complex interplay of absorption, transport, storage and excretion mechanisms. Dysregulation of copper homeostasis has been associated with various diseases, particularly cancers. 5 , 6 , 7 , 8 The influence of copper on cancer cell characteristics is closely associated with cellular metabolism and antioxidant defence because many types of cancers require high levels of aerobic respiration and exhibit enhanced mitochondrial metabolism. 9 , 10
Copper can be absorbed from the daily diet, including copper‐rich foods such as offal and shellfish, 11 , 12 and is highly distributed in the liver, muscles, eyes and brain. Serum copper concentrations in healthy adults typically range from 70 mg/dL to 110 mg/dL. 13 To maintain cellular health and avoid toxicity, the concentration of free copper in the cytoplasm varies from 10−15 M to 10−21 M. 14 The intestinal tract is the centre of copper absorption and regulation, primarily mediated by Copper Transport Protein 1 (CTR1, Solute Carrier Family 31 member 1, SLC31A1, encoded by slc31a1) localised in the apical side of epithelial cells. After absorption through the gastrointestinal tract, copper ions are secreted into the bloodstream and bind to soluble chaperones such as ceruloplasmin, transcuprein, histidines, albumin and macroglobulins. 15 , 16 , 17 , 18 The gut is particularly vulnerable to copper imbalances because of its crucial role in copper absorption. Consequently, it is crucial to investigate the relationship between aberrant intestinal absorption of copper ions and the pathogenesis of CRC.
In this review article, we systematically summarise existing studies on the role of copper in CRC and describe the underlying mechanisms. First, the transport mechanisms of copper ions and their biological functions are reviewed in detail, highlighting diseases induced by aberrant copper ion levels and corresponding mechanisms. Furthermore, we describe the precise regulation of copper ion concentrations, which provides vital insights for clinical serological monitoring. Subsequently, this review elucidates in detail the role of copper ions in CRC proliferation and metastasis, and emphasises the importance of serum copper ions (mainly detected by plasma ceruloplasmin levels) in CRC surveillance, thus laying a solid foundation for basic cancer research. Considering that cuproptosis is mainly caused by copper overload, in this review, we also describe the mechanism of cuproptosis in the carcinogenesis and development of CRC, providing theoretical support for cuproptosis‐mediated malignant progression of CRC. Finally, we systematically explore various copper ion‐targeted treatment strategies for CRC, offering substantial theoretical support for therapeutic approaches. We hope that this review will deepen the understanding of this important metal ion, and thus facilitate the development of targeted therapies and biomarkers for clinical diagnostics.
2. THE PHYSIOLOGICAL HOMEOSTASIS AND METABOLISM OF COPPER
Copper, an essential transition metal, has an important bidirectional regulatory role with a U‐shaped dose response and is a dynamic cofactor for numerous enzymes, either donating or accepting electrons. 4 Copper ions deficiency can impair biological systems, and may lead to various human diseases, including cardiovascular disorders, 19 anaemia, 20 osteoporosis 21 and Menkes disease, a hereditary disorder caused by mutations in the ATPase copper transporting alpha (ATP7A) gene, which results in ATP7A deficiency. 22 , 23 Conversely, excessive copper accumulation can cause various metabolic disorders, ultimately leading to cell death. 24 Preclinical studies have suggested that elevated copper levels promote cancer cell proliferation and progression both in vivo and in vitro. 5 , 6 , 25 , 26 , 27 , 28 This phenomenon, termed ‘cuproplasia’, may be associated with an increase in reactive oxygen species (ROS), and moderately elevated levels of copper can destabilise the genome and interfere with signalling pathways associated with cancer. However, copper overload can lead to severe health risks, including Wilson's disease, 29 a rare genetic disorder caused primarily by mutations in the ATP7B gene, which is characterised by an excessive accumulation of copper in vital organs, leading to liver degeneration, organ damage and sclerosis. 30 Previous studies have shown that disturbances in copper metabolism can lead to cancer cell death via apoptosis, 31 , 32 paraptosis, 33 , 34 ferroptosis 35 and caspase independent cell death. 36 In terms of molecular mechanisms, excessive copper ions can trigger cell death through multiple signalling pathways. Due to their redox‐active properties, copper ions participate in the Fenton reaction to produce ROS. This process leads to the release of cytochrome c and mitochondrial apoptosis‐inducing factor 1 into the cytoplasm, initiating caspase activation and DNA fragmentation. 31 , 37 Copper‐induced ROS also increase lipid peroxidation, deplete glutathione (GSH) and enhance cellular susceptibility to oxidative damage. 38 Moreover, copper accumulation in the cell nucleus damages DNA in various biological systems and inactivates sulphur‐containing enzymes by binding to their sulphur groups. 39 Additionally, copper ions inhibit the chymotrypsin‐like activity of the 20S proteasome and 19S proteasomal deubiquitinases, further inducing apoptosis. 40 Research has also shown that excess copper induces paraptosis by inhibiting the ubiquitin‐proteasome system, triggering endoplasmic reticulum (ER) stress and caspase‐3 inhibition, leading to copper complex toxicity in HT‐1080 cells. 41 Intracellular copper overload induces a noncaspase‐dependent form of paraptosis via ROS and an unfolded protein response. 42 Concurrently, evidence indicates that copper plays dual role in promoting or inhibiting ferroptosis under specific conditions. Copper complexes, such as disulphiram/Cu2+ and elesclomol/Cu2+, disrupt mitochondrial homeostasis, leading to oxidative stress, thus inducing ferroptosis in hepatocellular carcinoma and CRC. 43 , 44 This is consistent with the observation that excessive copper accumulation produces a significant amount of ROS in cancer cells. Copper also promotes ferroptosis in pancreatic cancer by facilitating autophagic degradation of GPX4. 45 Conversely, copper ions may also inhibit the occurrence of ferroptosis by limiting GPX4 expression in bathocuproinedisulphonic acid treatment. 46 , 47 Copper nanoparticles provide a feasible method of copper delivery by inducing ferroptosis in CRC cells. 48 Therefore, ferroptosis is a metal‐dependent mode of cell death that involves both iron and copper. However, a recent study by Tsvetkov et al. demonstrated that copper‐dependent cell death is a unique form of cell death termed as cuproptosis. 49 Cuproptosis is a recently identified cell death pathway that is triggered by the influx of excessive copper ions into cells. This pathway involves aberrant oligomerisation of copper‐dependent lipoylated proteins in the tricarboxylic acid (TCA) cycle, and a concurrent decrease in iron‐sulphur (Fe‐S) cluster proteins. Central to this process is Ferredoxin1 (FDX1), which plays a pivotal role in reducing Cu2+ ions to Cu+, facilitating dehydrolipoyltransacetylase (DLAT)‐mediated lipid acylation, and ultimately triggering cell death by decreasing the levels of Fe‐S cluster proteins. Inhibition of FDX1 or the enzymatic processes associated with lipid acylation halts cuproptosis, highlighting the significance of biomarkers such as FDX1, LIAS, LIPT1, DLD, DLAT, PDHA1, PDHB, MTF1, GLS and CDKN2A in the study and understanding of cuproptosis. Among these, the first seven genes mediate resistance to cuproptosis, whereas the latter three genes promote cuproptosis. 50 These biomarkers are poised to become key indicators of this unique cell death mechanism. 49 , 51 Therefore, elucidating the mechanisms underlying copper‐induced cellular proliferation and death is essential for the advancement of anticancer therapies (Figure 1).
FIGURE 1.

Three roles of copper ion in the body. Deficient copper levels can compromise biological systems, leading to disorders such as cardiovascular complications, anaemia, osteoporosis and Menkes disease. Conversely, elevated copper levels may facilitate tumour growth through the induction of ROS production, exacerbation of genomic instability and alteration of tumour‐associated signal transduction pathways. Additionally, when copper concentrations surpass a specific threshold, they can paradoxically induce tumour cell death. ROS, reactive oxygen species.
Herein, we summarise the major factors and signalling pathways involved in copper homeostasis, including cuproenzymes, membrane transporters and copper chaperones, which collectively regulate intracellular copper homeostasis. These proteins synergistically coordinate copper input, intracellular utilisation and output to maintain intracellular copper levels within a specific range. Therefore, we categorise the major regulators of copper metabolism according to their roles and highlight their essential functions in the modulation of this process (Table 1).
TABLE 1.
Major regulators involved in copper homeostasis.
| Gene | Name | Subcellular location | Function | Refs | |
|---|---|---|---|---|---|
| Copper absorption | CTR1 | Copper transporter 1 | PM | High‐affinity copper importer | 17 |
| CTR2 | Copper transporter 2 | Organelle membrane | Regulation of CTR1 | 67, 68 | |
| DMT1 | Divalent metal transporter | PM | Copper uptake transporter, mainly complementary to CTR1 for copper uptake | 62 | |
| STEAP | Six‐transmembrane epithelial antigen of the prostate | PM | Metalloreductase, reduce Cu2+ to Cu+ | 56 | |
| DCYTB | Duodenal Cytochrome b | PM | Metalloreductase, reduce Cu2+ to Cu+ | 54 | |
| Copper excretion | ATP7A | ATPase copper transporter 7A | Organelle Membrane, Regulated by the level of copper ions | Transport copper across cell membranes | 6 |
| ATP7B | ATPase copper transporter 7B | Organelle membrane, regulated by the level of copper ions | Excrete copper into bile and plasma | 22 | |
| Copper utilisation | ATOX1 | Antioxidant protein | CP | Metallochaperone that delivers copper to ATP7A and ATP7B Cu+ transporters | 92 |
| CP | Ceruloplasmin | Secreted | Major exchangeable plasma Cu carrier | 16 | |
| CCS | Copper chaperone for superoxide dismutase | CP | Metallochaperone that delivers copper to SOD | 83 | |
| SOD1 | Superoxide dismutase 1 | CP, Mit | Superoxide scavenger | 84, 85, 86 | |
| MT1/2 | Metallothionein 1/2 | CP | Bind divalent heavy metal ions, altering the intracellular concentration of heavy metals | 82 | |
| SCO1 | Synthesis of cytochrome oxidase 1 | Mit | Metallochaperone, required for copper delivery to the CuA site | 127 | |
| SCO2 | Synthesis of cytochrome oxidase 2 | Mit | Copper‐binding thiol‐disulphide oxidoreductase of COX2, a component of the copper delivery pathway to the CuA site | 127 | |
| COX11 | Copper chaperone for cytochrome c oxidase 11 | Mit | Metallochaperone, a component of the copper delivery pathway to the CuB site | 104 | |
| COX17 | Copper chaperone for cytochrome c oxidase 17 | Mit and CP | Metallochaperone that transfers copper to SCO1 and COX11 for cytochrome oxidase copper loading in mitochondria | 87 | |
| LOX | Lysyl oxidase | Secreted | Copper‐dependent enzymes, oxidise lysyl and hydroxylysyl residues in collagen and elastin | 97 |
Abbreviations: CP, cytoplasm; Mit, mitochondria; NC, nucleus; PM, plasma membrane.
2.1. Copper absorption and excretion
Copper is absorbed via the intestinal tract, mainly in the small intestine, especially in the proximal part of the small intestine, including the duodenum and jejunum. 52 Copper ions exist in two valence states, Cu+ and Cu2+, with Cu+ being predominant in the reducing intracellular environment. Physiologically, CTR1 primarily facilitates the transport of monovalent copper ions into the intestinal epithelium. Consequently, metalloreductases such as the Six‐Transmembrane Epithelial Antigen of the Prostate (STEAP) and duodenal cytochrome b act as copper reductases, 53 , 54 , 55 , 56 converting divalent copper into monovalent copper, thereby enhancing the uptake of copper ions by CTR1. The STEAP family of proteins consists of four members, namely, STEAP 1−4, whose primary role is to promote metal homeostasis by reducing the oxidation states of iron and copper, thereby facilitating the absorption of these metals. 57 , 58 However, STEAP1 is an exception in that it does not exhibit metalloreductase activity, probably because of the lack of the F420H2:NADP+ oxidoreductase structural domain and Rossmann folded structure. 59 Furthermore, copper is also transported via divalent metal transporter 1 to a lesser extent. 60 , 61 , 62 , 63 Copper is then transported to the opposite side of the epithelium through the copper chaperone antioxidant 1 (ATOX1) and subsequently released into the bloodstream via ATP7A. 64
Accumulating evidence indicates that CTR1 is indispensable for copper transport to specific organs and that its regulation is copper‐dependent. 65 Studies by Nose et al. and Lee et al. established that CTR1 plays a crucial role in copper ion uptake in intestinal‐specific and systemic knockouts, respectively. Notably, mice with systemic Ctr1 knockout (Ctr1 − / −) experienced uterine death during the second trimester. 66 , 67 Additionally, CTR2 acts as a copper sensor within endosomal compartments. 68 , 69 Studies including those by Helena et al. have shown that CTR2 knockout in mice leads to progressive age‐related copper hyperaccumulation, especially in the brain. 68 Therefore, investigating CTR2‐dependent cleavage of CTR1 is crucial for understanding the regulation of copper absorption in various tissues and under different physiological or pathological conditions.
In addition to copper absorption, copper export plays a crucial role in preventing copper accumulation and cytotoxicity. The gallbladder is the main organ involved in copper excretion, and small amounts of copper are eliminated from the body through sweat. 70 , 71 , 72 The core molecular mechanisms of this process involves Cu‐ATP7A and ATP7B, whose localisation and activity are heavily influenced by copper levels. Under physiological conditions, these transporters are located in the trans‐Golgi network (TGN) where they transfer copper ions from the cytoplasm to the TGN lumen. When intracellular copper concentrations are elevated to toxic levels, ATP7A and ATP7B translocate from the TGN to the vesicular region and merge with the plasma membrane, thereby promoting copper export. 73 Moreover, ATP7A and ATP7B exhibit distinct expression patterns. ATP7A is ubiquitously expressed in almost all cell types, except hepatocytes. In contrast, ATP7B is predominantly expressed in the liver 74 and plays a key role in the transport of copper ions into the bloodstream, especially during excretion through the liver and gallbladder. These mechanisms cumulatively contribute to the harmonious and stable balance of copper in the body, including intestinal absorption, hepatic storage and biliary excretion (Figure 2).
FIGURE 2.

The schematic illustrates the systemic and cellular metabolism of copper. Copper is absorbed primarily through the small intestine and then transported via the bloodstream to the liver for storage and subsequent excretion into the bile. At the cellular level, copper uptake occurs predominantly through CTR1, and to a lesser extent DMT1. CTR1 is chiefly responsible for transporting monovalent copper ions, while STEAP and DCYTB are copper reductases. In the bloodstream, copper ions are primarily bound to CP and HSA. In the biliary system, copper ions are mainly excreted through ATP7B. STEAP, Six‐Transmembrane Epithelial Antigen of the Prostate; DCYTB, Duodenal Cytochrome b; CTR1/2, Copper Transport Protein 1 and 2; DMT1, divalent metal transporter 1; ATP7A and ATP7B, ATPase copper transporter 7A and 7B; CP, ceruloplasmin; HSA, human serum albumin.
2.2. Copper utilisation
To avoid cytotoxicity and ensure precise copper homeostasis, copper ions in the blood are predominantly bound to proteins. Approximately 75% of nonexchangeable copper ions bind to ceruloplasmin (CP), and approximately about 25% of exchangeable copper ions bind to human serum albumin (HSA). Only a marginal 0.2% of copper ions binds to histidine. 75 , 76
Under physiological conditions, copper is transported to various cellular substructures including the Golgi apparatus, cytoplasm, mitochondria, and nucleus (Figure 3). The cytoplasm contains high concentrations of GSH and metallothionein (MT), natural chelators of copper ions. 77 , 78 These proteins, which greatly outnumber copper ions, have two main functions: (1) maintaining a negative concentration gradient across the cytoplasmic membrane, thereby promoting CTR1‐mediated copper uptake 79 ; and (2) acting as cytoplasmic copper buffers, binding free copper ions to prevent cytotoxicity, and producing ROS. 6 , 80 , 81 , 82 MT contains metal–thiolate clusters that bind heavy metals and can chelate large amounts of copper ions. Interestingly, increased intracellular copper ion levels also stimulate MT expression. 83 Consequently, MT and GSH form intrinsic defence mechanisms against copper‐induced cytotoxicity. Additionally, copper ions can bind to chaperones, especially the copper chaperonin of superoxide dismutase (CCS) which interacts with copper ions and transports them to superoxide dismutase 1 (SOD1), facilitating the formation of disulphide bonds that are essential for its proper structure and enzymatic activity. 84 , 85 Furthermore, CCS regulates SOD1 distribution in the intermembrane space and cytoplasm in an oxygen‐dependent manner. This regulatory mechanism is essential for maintaining ROS levels in vivo and mitigating ROS generated by the electron transport chain, thereby preventing oxidative damage caused by copper overloading. 86 , 87
FIGURE 3.

The key molecular mechanisms of copper metabolism. Extracellular Cu2+ are reduced to Cu+ by the reductase STEAP and then imported into the cell via the CTR1. Inside the cell, Cu+ is distributed to cytosolic copper chaperones, including MT, GSH, CCS and SOD1, which in turn deliver it to specific subcellular locations such as the mitochondria, TGN and nucleus. CP, ceruloplasmin; CTR1, Copper Transport Protein 1; MT, Metallothionein; GSH, glutathione; SOD1, superoxide dismutase 1; CCS, copper chaperone for superoxide dismutase; ATOX1, antioxidant protein; MEMO1, mediator of cell motility 1; ATP7A and ATP7B, ATPase copper transporter 7A and 7B, respectively; TGN, trans‐Golgi network; CCO, cytochrome c oxidase; COX17, cytochrome c oxidase copper chaperone17; COX11, cytochrome c oxidase copper chaperone 11; SLC25A3, solute carrier family 25 member 3; SCO1/2, synthesis of cytochrome c oxidase 1 and 2; COA6, cytochrome c oxidase assembly factor 6; MT‐CO1/2, mitochondrially encoded cytochrome c oxidase 1 and 2; LOX, lysyl oxidase; LOXL, lysyl oxidase like.
Furthermore, copper ion chaperones in the mitochondria play a crucial role in the function of cytochrome c oxidase (COX), a key component of oxidative phosphorylation. The entry of copper ions into the mitochondria is mainly dependent on COX17, 88 which transports Cu+ from the cytoplasm to mitochondrial membrane proteins, such as synthesis cytochrome C oxidase 1 and cytochrome C oxidase 2, 89 facilitating copper insertion into mitochondrial encoded cytochrome c oxidase II (MT‐CO2/COX2). 90 COX17 also transports copper to mitochondria‐encoded cytochrome c oxidase I (MT‐CO1/COX1) via the cytochrome c oxidase copper chaperone 11 (COX11), 90 which drives the electrochemical production of ATP. 91 Therefore, the cellular copper pool is closely associated with mitochondrial oxidative phosphorylation, in which COX17 is essential for MT‐CO1 and MT‐CO2 activity in the mitochondrial respiratory chain. 92
Copper is transported to the nucleus by the CCS and ATOX1 proteins. Here, it initiates the activation of transcription factors such as hypoxia‐inducible factor 1 via the CCS 26 and acts as a copper‐dependent transcription factor via ATOX1. 93 Additionally, the copper chaperone ATOX1 transports copper to the trans‐Golgi network and delivers it to the copper‐dependent ATPases, ATP7A and ATP7B. 94 MEMO1 (mediator of cell motility 1) also plays a crucial role in enhancing Cu(I) binding to ATOX1, thereby alleviating the excess of Cu‐induced ROS production. 95 Moreover, copper metalloenzymes are essential for several critical cellular biological processes. These include promoting pigmentation via tyrosinase, 96 influencing neural signalling via dopamine β‐hydroxylase, 97 facilitating the cross‐linking of elastin and collagen primarily via lysyl oxidase (LOX) and lysyl oxidase‐like 2 (LOXL2), 98 and aiding leukocyte trafficking via amine‐containing oxidase copper 3 (AOC3). 99 Copper is also an important component of the myelin sheath that protects neurons. 61 Taken together, these are the main mechanisms by which copper ions function in the human body.
3. ROLE OF COPPER AND CUPROPTOSIS IN CRC
3.1. Copper and carcinogenesis in CRC
Elevated copper ion levels have been associated with tumourigenesis and malignant progression, both in vivo and in vitro. 5 , 6 , 27 This phenomenon is known as ‘cuproplasia’ and refers to copper‐dependent cell growth, including enzymatic and nonenzymatic activities regulated by copper. 27 Specifically, copper ions function as allosteric modulators, activating the RAF‐MEK‐ERK signalling pathway by binding to specific enzymes such as MEK1 and MEK2, thereby improving their capacity to phosphorylate ERK1 and ERK2. 100 , 101 , 102 , 103 This activation promotes cancer cell proliferation and growth. Additionally, copper ions stimulate the activity of the E2‐binding enzymes UBE2D1‐UBE2D4 through positive allosteric activation, thereby facilitating protein degradation. 104 In the context of cancer, this mechanism may lead to ES becoming a prominent copper ionophore in cancer therapy. Its primary function is to induce oxidative stress ion and subsequent degradation of key oncogenic proteins, including p53, affecting cell cycle regulation, inhibiting programmed cell death, and thereby enhancing the viability of malignant cells. Recent studies have found higher copper levels in CRC tissues and patient sera (mainly detected by plasma CP levels) than in healthy individuals, 105 , 106 , 107 indicating that copper plays an important role in CRC development and cell proliferation. Notably, the serum copper/zinc ratio increased progressively with the advancing stages of CRC progression. 107 Strikingly, elevated serum copper levels are strongly correlated with the degree of malignancy in CRC. 108 A significant increase in the expression of the copper translocation regulator, ATP7A, was observed in KRAS mutation‐specific CRC cells. This increase facilitates a high influx of copper into the cells, supporting CRC proliferation. 109 This indicates that copper signalling may play a crucial role in CRC development. Specifically, copper ions modulate signalling pathways associated with nutrient sensing by positively regulating the activities of two kinases, ULK1 and ULK2, and this regulation affects cell growth and proliferation and supports the survival of cancer cells. 7 , 110 Inflammatory cytokines, such as interleukin (IL)−17, have been shown to induce STEAP4‐dependent cytosolic copper uptake. This uptake activates the E3 ligase XIAP, which in turn enhances IL‐17‐induced nuclear factor NF‐kappaB (NF‐κB) activation and inhibits caspase 3 activity, thereby promoting intestinal tumourigenesis in mice. 111
3.2. Copper and metastasis in CRC
Elevated copper levels are critical not only for the development of CRC but also for its metastasis, particularly to the liver, which is the most common site of distant CRC metastasis and a storage organ for copper. 14 Recent studies have shown that copper levels in CRC liver metastases are significantly higher than those in normal liver tissues. 112 Additionally, copper and copper‐binding proteins are vital for epithelial‐mesenchymal transition (EMT) and angiogenesis in CRC metastasis. Copper is essential for LOX and LOXL proteins, which are involved in collagen and elastin crosslinking. LOX secretion by cancer cells remodels the extracellular matrix to form premetastatic niches, thereby recruiting bone marrow‐derived cells, and promoting EMT in CRC. 113 Moreover, copper‐mediated interactions between hypoxia response elements and HIF‐1α activate the transcription factors ZEB1, ZEB2 and Snail via CCS, thereby promoting EMT in CRC. 26 , 98 , 114 In addition to copper ions and copper‐binding proteins, ATOX1, a key molecule that maintains copper homeostasis, plays a significant role in CRC proliferation and metastasis. ATOX1 also serves as a transcription factor that regulates the expression of Cyclin D1 and promotes cell proliferation by binding to copper. 93 Additionally, ATOX1 is markedly overexpressed in metastatic CRC tissues and cell lines, with a significant increase in its nuclear localisation. The entry of ATOX1 into the nucleus can activate transcription and promote the expression of CyclinD1 and the NADPH oxidase subunit p47 phox. Notably, this effect disappeared when the copper‐binding structural domain of ATOX1 is knocked out, suggesting that ATOX1‐mediated CRC metastasis is copper‐dependent. 115 In summary, copper and copper‐binding proteins are inextricably linked to CRC proliferation and metastasis (Figure 4).
FIGURE 4.

Increased copper levels and copper‐binding proteins promote CRC proliferation and metastasis. In CRC proliferation, three key mechanisms are involved. First, the proliferation of CRC cells is associated with increased levels of ATOX1. Second, an increase in copper influx into cancer cells further promotes CRC proliferation. Finally, the inflammatory cytokine IL‐17 elevates intracellular copper levels by stimulating STEAP4 expression. In the context of CRC metastasis, first, ATOX1 facilitates this process by reducing p47phox expression, which in turn decreases ROS production and enhances ATP synthesis. Second, it significantly influences the EMT process. Third, copper is vital in angiogenesis. It binds to and modulates proteins essential to angiogenic processes. ATOX1, Antioxidant‐protein 1; STEAP4, Six‐Transmembrane Epithelial Antigen of the Prostate 4; ROS, reactive oxygen species; XIAP, X‐linked inhibitor of apoptosis; CTR1, Copper Transport Protein 1; ATP7A, ATPase copper transporter 7A; eNOS, endothelial nitric oxide synthase; NO, nitric oxide; HIF1, hypoxia inducible factor 1; CP, ceruloplasmin; VEGF, vascular endothelial growth factor; CCS, copper chaperone for SOD1; HRE, hypoxia response element; ZEB1/2, zinc finger E‐box binding homeobox 1 and 2; Snail, snail family transcriptional repressor 1; LOX, lysyl oxidase; EMT, epithelial‐to‐mesenchymal transition; ECM, extracellular matrix.
3.3. Copper, angiogenesis and cell metabolism in CRC
Angiogenesis, the process of creating new capillaries from preexisting vasculature, involves an intricate sequence of events that includes endothelial cell proliferation, migration, differentiation and extracellular matrix remodelling. This process is crucial for normal physiological functions and certain pathological conditions such as malignant cancers. 116 Research indicates that cancer cells exhibit an angiogenic phenotype marked by the presence of numerous angiogenic molecules. This signifies a disturbance in the equilibrium of endogenous anti‐angiogenic factors, thereby instigating cancer‐associated angiogenesis. 117 In this progression, copper ions play a critical role in the early stages of cancer vasculature formation by stimulating endothelial cell proliferation and migration. Angiogenic factors that have been identified to be activated by copper include basic fibroblast growth factor, vascular endothelial growth factor (VEGF), tumour necrosis factor‐alpha and IL‐1, which bind to endothelial cells, facilitating their transition from the G0 phase to the G1 phase and promoting proliferation. Quintin et al. demonstrated that a copper‐deficient diet or the use of tetrathiomolybdate (a copper chelator) leaded to a reduction in copper levels, which impeded angiogenesis by suppressing the expression of the aforementioned angiogenic factors and inducing endothelial cell regression to the G0 phase or apoptosis. 118 Furthermore, copper is essential for the activation of HIF‐1, a key transcription factor regulating VEGF expression. The presence of copper ensures the formation of an HIF‐1 transcription complex, which activates the expression of target genes, including VEGF. However, excess copper can lead to the accumulation of the rate‐limiting component of HIF‐1, HIF‐1α, in the cytoplasm, thereby activating HIF‐1. 119 Furthermore, copper ions mediate the activity of endothelial nitric oxide synthase and increase the production of the vasodilator nitric oxide, thereby promoting angiogenesis. 120
The precise mechanism through which copper stimulates angiogenesis in CRC remains unclear. Most studies have focused on other forms of cancer, in which copper activates numerous angiogenic factors. For example, Damiano et al. observed that copper ions, through the activation of the EGFR/ERK/c‐fos signalling pathway, induced the expression of G protein‐coupled oestrogen receptor 1, HIF‐1α, and VEGF in breast and liver cancer cells, thereby promoting tumour angiogenesis. 121 In patients with malignant ovarian cancer, increased copper level has been observed, suggesting that the angiogenic effects of copper in ovarian cancer and endothelial cells might influence the production of ascites. Therefore, strategies to reduce copper levels in ascites may be beneficial for downregulating VEGF expression and improving the prognosis of malignant ovarian tumours. 122 Genes associated with copper ion transport, including CTR1, ATP7A and LOX, also promote angiogenesis. Archita et al. showed that CTR1 acts as a redox sensor that enhances angiogenesis in endothelial cells. The absence of CTR1 in cardiovascular cells weakens the response to VEGF‐induced VEGFR2 signalling and angiogenesis. Mechanistically, upon VEGF stimulation, the cytoplasmic C‐terminal Cys189 of CTR1 undergoes rapid sulphenylation, inducing the formation of a CTR1‐VEGFR2 disulphide bond and co‐internalisation into early endosomes, thereby sustaining VEGFR2 signalling. 123 MiR‐495 combats cisplatin resistance and inhibits angiogenesis in oesophageal cancer cells by targeting ATP7A expression. 124 LOX overexpression and copper‐induced LOX activity promote angiogenesis and proliferation in oral squamous cell carcinoma cells. 125
In terms of cellular metabolism, copper is an essential cofactor for many enzymes. 126 In cancer cells, changes in copper ion levels can prompt a metabolic shift from mitochondria‐dependent oxidative phosphorylation to a more glycolysis‐dependent state, a phenomenon known as the ‘Warburg effect’. Metabolic reprogramming is associated with the regulation of the activities of many key enzymes by copper ions. 101 , 127 These include cytochrome c oxidase (COX, also known as Complex IV), a crucial enzyme complex in the mitochondrial electron transport chain that directly uses copper ions as cofactors. 91 , 128 SOD, especially SOD1 (cytoplasmic) and SOD3 (extracellular), use copper and zinc as cofactors. SOD is essential for regulating intracellular ROS levels. Changes in copper ions can affect SOD activity, thereby influencing the cellular response and regulation of ROS, which plays an important role in regulating the metabolic reprogramming of cancer cells. Other enzymes include the ubiquitin ligase system, lactate dehydrogenase (LDH) and glucose transport proteins. For example, studies have reported high expression of COX in CRC, 129 , 130 , 131 where COX subunit 1 plays a prognostic role in distinguishing between CRC and adenoma. 132 Sabrina et al. reported the involvement of the mitochondrial copper chaperone protein COX19 in the biosynthesis of human COX. 133 Furthermore, studies have indicated that the suppression of COX19 expression in nonsmall cell lung cancer cells could enhance apoptosis. 134 Moreover, Gao et al. reported that COX19 was also associated with the prognosis of patients with CRC, primarily through MACC1 transcriptional upregulation of COX19, and COX19 promoting intracellular copper transport and enhancing mitochondrial activity to exert its carcinogenic effects. 135 , 136
3.4. Copper and the antitumour immune response in CRC
In the 1950s, researchers discovered that deficiency of copper led to hypocupremia, hypoceruloplasminemia and neutropenia in infants, which subsequently affected the ability of their immune system to kill bacteria. 137 , 138 , 139 The influence of copper on the immune system has garnered growing research interest following these findings. In 2009, studies by White et al. further revealed that the effect of copper ions on macrophages was closely associated with the expression of ATP7A, and proinflammatory factors such as LPS or IFN‐γ could enhance the absorption of copper by macrophages, increasing the expression of CTR1 and ATP7A, thereby augmenting their pathogen‐killing capabilities. 140 , 141 Moreover, copper deficiency was found to affect acquired immunity, as evidenced by the weakened response of cytotoxic T lymphocytes in mice against allogeneic antigens, highlighting the critical role of copper in immune function regulation. 142
Concurrently, emerging research has reported relationship between tumour immunity and copper, such as the association of reduced copper with CD4+ T cell infiltration in mouse mesothelioma 143 and the influence of the antitumour drug Dp44mT on T cell activity through a copper‐related mechanism. 144 Regarding CRC immune response, bioinformatics studies constructing a copper‐related death prognostic index (CPS score) have indicated a significant positive correlation between CPS scores and both microsatellite instability high (MSI‐H) and tumour mutation burden (TMB), suggesting that patients with higher CPS scores possess higher immunogenicity, thereby potentially benefiting more from immune checkpoint blockade therapy. 145 , 146 By constructing a prognostic risk model utilising copper‐associated long noncoding RNAs (lncRNAs; CRLncSig), it was uncovered that the low‐risk group demonstrated prolonged overall survival (OS) in comparison to the high‐risk group. Additionally, the low‐risk group exhibited decreased tumour purity, enhanced immune cell infiltration, elevated immune scores, enriched immune activation pathways, and a more favourable response to immunotherapy. 147 The role of cuproptosis in CRC immunotherapy has primarily focused on FDX1 and DLAT. Analysis using the TCGA database showed that FDX1 is positively correlated with immune cell infiltration in colon adenocarcinoma, particularly indicating a higher proportion of CD8+ T cells in tumour tissues compared to adjacent nontumour tissues, while the proportion of CD4+ T cells was opposite, suggesting that FDX1 is a potential target for colon adenocarcinoma immune therapy. 148 DLAT is significantly associated with immune‐related pathways in pan‐cancer, liver cancer, and CRC. Moreover, DLAT expression correlated with tumour‐associated macrophage infiltration, TMB and MSI. These studies indicate that DLAT plays a crucial role in cancer immunity and development, potentially serving as a prognostic biomarker for cancer immune therapy. 149 , 150 Furthermore, DLAT promotes antitumour immune activation by reversing T cell exhaustion and inducing apoptosis. 151
3.5. Role of cuproptosis in CRC
Few studies have addressed the molecular mechanisms of cuproptosis in CRC. In other digestive system cancers, such as in gastric cancer, copper stress can promote the site K229 lactylation of METTL16, which in turn mediates m6A‐modification on FDX1 mRNA, thereby promoting cuproptosis. 152 Furthermore, ferroptosis inducers sorafenib and erastin promote cuproptosis in primary liver cancer by suppressing the degradation of FDX1 protein and reducing synthesis of intracellular copper chelator GSH through the suppression of cystine transportation. 153 These results reveal the significance of cuproptosis induction as a promising therapeutic strategy for cancer.
Furthermore, the application of multiomics analysis has expanded our understanding of the role of cuproptosis in CRC. Research has indicated that by analysing cuproptosis‐related genes (CRGs), it is possible to classify CRC patients into different risk groups and identify different molecular subtypes with different clinical characteristics. For instance, one study categorised patients with CRC into two molecular subtypes based on 41 CRGs, with subtype C1 characterised by lower OS rates and advanced clinical staging. 154 Moreover, cuproptosis is closely linked to the tumour immune microenvironment. Patients classified in the low‐risk group exhibiting stronger immunogenicity, higher immune scores, and a greater degree of microsatellite instability, suggesting a better response to immunotherapy. 50 , 145 , 155 , 156 , 157 , 158 , 159 Further investigation have highlighted the potential role of cuproptosis‐associated microRNAs and lncRNAs in CRC. For example, six cuproptosis‐associated miRNAs including hsa‐miR‐653, hsa‐miR‐216a, hsa‐miR‐3684, hsa‐miR‐4437, hsa‐miR‐641 and hsa‐miR‐552 were used to construct prognostic models. It was found that miR‐653 was overexpressed in CRC tissues, promoting cell proliferation and inhibiting apoptosis by negatively regulating DLD expression. 160 Similarly, certain lncRNA combinations have demonstrated potential for predicting CRC prognosis, including SNHG16, LENG8‐AS1, LINC0225, and RPARP‐AS1. 118 These molecular markers not only provide new biological insights but also offer potential targets for future targeted therapy strategies. In terms of drug sensitivity, a study discovered that CRC patients with different risk scores had different sensitivities to specific drugs, such as greater sensitivity to AKT inhibitors and doxorubicin in the high‐risk group, which lays the groundwork for personalised treatment. 154 Elesclomol‐copper rapidly inhibited the growth of CRC cells and oxaliplatin‐resistant cell lines, with 4‐octyl itaconate (4‐OI) inhibiting aerobic glycolysis by targeting GAPDH, thereby promoting cuproptosis. 161 Furthermore, CRC cells treated with elesclomol‐copper showed increased E2F3 expression, which significantly enhanced the resistance of colorectal adenocarcinoma (COAD) cells to elesclomol. 162
Multiomics analysis has also revealed the role of known molecular markers associated with cuproptosis in CRC progression. For example, low expression of FDX1 in COAD indicated poor prognosis, and immune microenvironment analysis showed that the proportion of CD8+ T cells was significantly lower than that of neighbouring normal tissues, whereas the proportion of CD4+ T cells showed the opposite trend. 148 Additionally, FDX1 inhibited CRC growth and progression by suppressing EMT. 163 Single‐cell and bulk RNA sequencing analyses identified AOC3, COX11, COX17, COX19, CCS, CDKN2A, DLD, DLAT, and PDHB as CRGs predictive of CRC prognosis, revealing that the elevated expression of COX17 in CD4‐CXCL13 T cells mediated Treg infiltration and T cell exhaustion, while DLAT reversed T cell exhaustion and induced apoptosis, promoting antitumour immune activation. 151 Moreover, multiomics analysis identified the novel cuproptosis core‐related gene TIGD1, which significantly regulates CRC cuproptosis, and offers new targets for CRC treatment. 164
In summary, as a novel mechanism of cell death, cuproptosis has extensive potential applications in CRC research, encompassing areas ranging from prognostic assessment and molecular typing to the development of therapeutic strategies. A comprehensive analysis of the roles of cuproplasia and cuproptosis in CRC will provide a new perspective on the complexity of CRC. Such studies will also help identify new targets and strategies for future CRC treatment.
4. ROLE OF COPPER IN THE TREATMENT AND DIAGNOSIS OF CRC
Elevated copper levels in CRC compared with adjacent normal tissues, coupled with growing evidence that high copper levels facilitate CRC proliferation and metastasis, make targeting copper ions a promising treatment for CRC. Currently, cancer therapeutic agents involving copper primarily consist of copper chelators, copper ionophores, copper‐containing compounds and the radioisotope 64Cu. 114 , 165 We categorised and detailed the roles and mechanisms of these agents in the diagnosis and treatment of CRC (Table 2). Additionally, we systematically organised and summarised the clinical trial data pertaining to these drugs, particularly the more established copper‐targeting cytotoxic agents such as tetrathiomolybdate (TM) and disulphiram (DSF) – in cancer therapy (Table 3).
TABLE 2.
Mechanisms of different types of agents related to copper for CRC treatment.
| Name | Mechanisms | Result | Refs | |
|---|---|---|---|---|
| Copper chelators | TM | Regulated phosphorylation levels of ERK1/2; increased expression of hCTR1 protein | Inhibited CRC proliferation, survival and migration; increased oxaliplatin sensitivity | 117, 167 |
| Trientine | Inhibited the angiogenic factors, particularly VEGF and IL8 | increased CRC apoptosis | 172 | |
| TPEN | Engaged in redox cycling to generate hydroxyl radicals; increased ROS accumulation and Mcl‐1 ubiquitination | increased CRC apoptosis; overcame drug resistance | 170 | |
| JYFY‐001 | Decreased CRC extracellular acidification rate and oxygen consumption rate | Inhibited CRC proliferation; Increased CRC apoptosis | 175 | |
| D‐penicillamine | Increased expression of hCTR1 protein | increased oxaliplatin sensitivity | 169 | |
| Bathocuproinedisulphonic acid | Increased expression of hCTR1 protein | increased oxaliplatin sensitivity | 166 | |
| Copper ionophores | ES | promoted the degradation of ATP7A; increased ROS accumulation and promoted SLC7A11 degradation | promoted CRC ferroptosis | 178, 179 |
| DSF | miR‐16‐5p and 15b‐5p/ALDH1A3/PKM2 axis‐mediated aerobic glycolysis pathway; upregulated ULK1; induced molecules expression of cell ICD | inhibited CRC progression; Induced autophagic; increased apoptosis | 185, 186, 187, 188, 189, 190, 191, 192, 193 | |
| Copper‐containing compounds | copper(II) complex | Activated caspases 3/7/9, BAX and ROS accumulation | Increased apoptosis | 211 |
| copper(I) complex | Activated p53 pathway; suppressed the ubiquitin‐proteasome pathway and triggered endoplasmic reticulum stress; increased ROS accumulation | Increased apoptosis | 210 | |
| Copper nanoparticles | Triggered cell autophagy‐dependent apoptosis and ferroptosis; generated cancericidal species ·OH and 1O2 from rich H2O2 in tumours for nanocatalytic tumour therapy; reprogram the macrophages from the M2 phenotype to the M1 phenotype | Inhibited CRC proliferation and metastasis; increased apoptosis and autophagy | 225, 226, 227, 228, 229 |
TABLE 3.
Clinical trials evaluating agents that target copper for the treatment of cancer.
| Name | Cancer types | NCT Number | Phases | Enrolment | Status | |
|---|---|---|---|---|---|---|
| Copper chelators | TM | CRC | NCT00176774 | II | 24 | Completed |
| Nonsmall cell lung cancer | NCT01837329 | I | 26 | Completed | ||
| Oesophageal carcinoma | NCT00176800 | II | 69 | Completed | ||
| Prostate cancer | NCT00150995 | II | 19 | Completed | ||
| Trientine | Advanced malignancies | NCT01178112 | I | 56 | Completed | |
| Ovarian cancer | NCT03480750 | I/II | 18 | Completed | ||
| D‐penicillamine | Glioblastoma | NCT00003751 | II | 40 | Completed | |
| Copper ionophores | ES | Melanoma | NCT00522834 | III | 630 | Terminated |
| Melanoma | NCT00084214 | I/II | 103 | Completed | ||
| Prostate cancer | NCT00808418 | I | 34 | Completed | ||
| Soft tissue sarcomas | NCT00087997 | II | 80 | Completed | ||
| Ovarian cancer, fallopian tube cancer, or primary peritoneal cancer | NCT00888615 | II | 58 | Completed | ||
| Acute myeloid leukaemia | NCT01280786 | I | 36 | Active | ||
| Solid tumours | NCT00827203 | I | 30 | Suspended | ||
| DSF | Breast cancer | NCT03323346 | II | 150 | Recruiting | |
| Gastric cancer | NCT05667415 | N/A | 40 | Not yet recruiting | ||
| Glioblastoma | NCT02678975 | II/III | 88 | Completed | ||
| Nonsmall cell lung cancer | NCT00312819 | II/III | 60 | Completed | ||
| Prostate cancer | NCT01118741 | N/A | 19 | Completed | ||
| Prostate cancer | NCT02963051 | I | 9 | Terminated | ||
| Pancreas cancer | NCT03714555 | II | 1 | Completed | ||
| Refractory solid tumours involving the liver | NCT00742911 | I | 21 | Completed | ||
| Melanoma | NCT00256230 | I/II | 7 | Completed | ||
| Glioblastoma | NCT03034135 | II | 24 | Completed |
4.1. Metal chelating agents that can chelate copper in CRC treatment
Copper chelators are compounds that bind to monovalent or divalent copper and play a crucial role because of the lack of free copper ions in the body and the varying affinity of copper ions for proteins. These chelators modulate the redox activity of copper and have the potential to redistribute copper between proteins. 82 , 166 There are many types of copper chelators, including polyaminocarboxylates, acyclic amino, macrocyclic amino, D‐penicillamine, thiomolybdate, hydroxyquinoline chelators, dithiocarbamates, diamines and cuprizones. 166 Among them, the most studied copper chelators for CRC mainly include TM, D‐penicillamine, trientine, bathocuproinedisulphonic acid, and N, N, N’, N’‐tetrakis‐[2‐pyridylmethyl]‐ethylenediamine (TPEN). Extensive studies have established the crucial role of TM in overcoming chemoresistance in CRC. Specifically, TM augments the cytotoxic effects of chemotherapeutic agents such as oxaliplatin, irinotecan, and 5‐fluorouracil (5‐FU) on CRC cells. Notably, TM enhanced oxaliplatin sensitivity by upregulating CTR1 expression in DLD1 and SW620 cells, which, in turn, increased copper ion entry into the cells, thereby promoting the killing of CRC cells. 167 With irinotecan and 5‐FU, TM impeded CRC metastasis by reducing angiogenesis in 24 patients with metastatic CRC, primarily by a reducing relevant factors, including VEGF, IL‐8 and IL‐6. 168 Additionally, TM was effective against BRAF‐mutated CRC cells, inhibiting cell proliferation by targeting the Ras‐RAF‐MEK‐ERK signalling pathway and reversing the chemoresistance of BRAF V600E CRC cells. 169 Notably, D‐penicillamine and bathocuproinedisulphonic acid have analogous functions to TM. 170 Furthermore, TPEN can induce HCT116 cell death through redox cycling and ROS generation. 171 When combined with 5‐FU, TPEN enhanced apoptosis by increasing ROS production and Mcl‐1 ubiquitination in CRC cell lines, thereby augmenting anticancer activity and overcoming chemoresistance. 172
Other copper chelators such as trientine have also shown promise in CRC treatment. When combined with methotrexate, trientine inhibits CRC development and angiogenesis and is a potential new strategy for CRC treatment. 173 However, current copper chelators are associated with a series of toxic side effects owing to their chelation of various other metal ions. 174 , 175 Shi et al. identified a novel copper chelator, N‐(2‐(pyridin‐3‐yl)thieno[3,2‐c]pyridine‐4‐yl)acetamide (JYFY‐001). This chelator enhanced the efficacy of programmed cell death protein 1 (PD‐1) inhibitors, increased apoptosis in HCT116 and SW620 cells and impaired glucose metabolism in these cells. 176 In summary, the primary function of copper chelators in CRC is to inhibit cell proliferation and promote cell death. Importantly, these effects are particularly pronounced in drug‐resistant CRC cells, addressing a major challenge in clinical CRC therapy.
4.2. Copper ionophores in CRC treatment
Copper ionophores primarily function by dysregulating metal flux and increasing metal import beyond the metal export capacity of cells. 177 Unlike copper chelators, which mainly inhibit cuproplasia, copper ionophores predominantly induce cuproptosis and play a distinct role in cancer therapy. 178 Many copper ionophores exist, including elesclomol (ES), DSF, diethyldithiocarbamate, diacetylbis (N4‐methylthiosemicarbazone) and glyoxal‐bis(N4‐methylthiosemicarbazone). 178 Among these, ES and DSF have been the most extensively studied in CRC. ES is an important copper ionophore in cancer therapy that induces cytotoxicity by transporting Cu2+ into cells through the formation of ES‐Cu complexes. Once inside the cell, copper ions can lead to cancer cell apoptosis by inducing oxidative stress, 179 or target the TCA cycle proteins and FDX1, inducing cell death through cuproptosis. ES alone cannot cause cuproptosis; however, it mainly promotes apoptosis by increasing ROS levels and activating caspase enzymes. 49 , 180 However, recent studies have identified several potential targets. In CRC, high ATP7A and ATP7B expression contributes to oxaliplatin resistance. 181 , 182 , 183 ES countered this resistance by degrading ATP7A, causing copper retention and ROS accumulation, which in turn enhanced ferroptosis. 44 Furthermore, 4‐OI boosts cuproptosis by targeting GAPDH to inhibit aerobic glycolysis in CRC cells. The combination of elesclomol‐Cu and 4‐OI improved antitumour effects in CRC cells. 161
DSF is another copper ionophore that effectively induces CRC cells death. DSF, a derivative of thiuram, was approved by the United States Food and Drug Administration in 1951 for the treatment of alcohol dependency. 184 Its therapeutic efficacy is attributed to the irreversible inhibition of aldehyde dehydrogenase (ALDH) activity. 185 However, emerging evidence has revealed anticancer properties of DSF. DSF and its primary metabolite, diethyldithiocarbamate, induce ROS accumulation and trigger apoptosis. 186 , 187 , 188 Given that ALDH serves as an important biomarker for cancer stem cells and plays a pivotal role in tumour progression, the targeting of ALDH by DSF results in diminished tumourigenic potential. 189 Furthermore, DSF significantly suppresses the expression of stemness‐related transcription factors, such as SOX2, Nanog and OCT3/4. 190 In terms of tumour‐related signalling pathways, existing studies have demonstrated that DSF can attenuate the activity of NF‐κB, effectively inhibiting osteoclast differentiation. 191 Chiba et al. confirmed that DSF exerts its antitumour effects by activating p38, thereby modulating the mitogen‐activated protein kinases MAPK signalling pathway. 192 As both a proteasome inhibitor and divalent metal ionophore, DSF, through its active sulphhydryl groups, can inhibit the 26S proteasome, thereby impeding the ubiquitin‐proteasome system in a copper ion‐dependent manner. 193 Upon binding to copper, it induces tumour cell death. This anticancer activity is primarily mediated by the formation of diethyldithiocarbamate‐copper complex (CuET), a compound resulting from the interaction between DSF metabolites in vivo and copper ions. 194 In CRC, the primary molecular mechanisms include CuET‐mediated reduction of aerobic glycolysis via the miR‐16‐5p and 15b‐5p/ALDH1A3/PKM2 axes. 195 Disulphiram/copper complex targets ULK1 to induce autophagy, 196 induces immunogenic death to inhibit proliferation, 197 and synergistically enhances its cytotoxic effects in the pH‐acidic TME by altering cell metabolism, Akt kinase and NF‐κB activity and increasing ROS production in CRC cells. 198 During chemotherapy for CRC, DSF/Cu amplifies the efficacy of gemcitabine by inhibiting NF‐κB activity, and similarly enhances the effects of 5‐FU. 199 , 200 In other digestive tract tumours, such as hepatocellular carcinoma, DSF/Cu triggers ER stress by altering ER structures and raising Ca2+ levels. This effect is coupled with GSH depletion, heightened lipid peroxides, and a compensatory increase in solute carrier family 7 member 11 and activated expression of ATF4, thereby inducing both ferroptosis and cuproptosis. 43 , 201 Moreover, the encapsulation of DSF in nanoparticles or liposomes has been reported to augment its therapeutic effects. 202 , 203 Currently, studies on copper ionophores in CRC mainly focused on DSF and ES, and other copper ionophores need to be further explored to enhance their therapeutic effect on CRC.
4.3. Copper‐based complexes and nanoparticles in CRC treatment
Since cancer and normal cells respond differently to copper, copper complexes have been developed as potential anticancer agents. Copper compounds have attracted interest because of their specific therapeutic properties targeting particular sites. 204 These copper‐based compounds affect the function of critical cellular organelles, including the mitochondria and endoplasmic reticulum, thereby disrupting their function and ultimately leading to cell death. 205 In CRC treatment, copper‐based compounds are the most extensively researched form of copper intervention. These predominantly include ligands, liposomes and nanoparticles. The following sections summarise and categorise these developments: copper‐based complexes mainly include copper (I) complexes, copper (II) complexes and ternary copper (III) complexes, and copper (II) is mainly used for these complexes due to its stability. Briefly, these complexes exhibit anticancer properties primarily by inhibiting the proliferation of CRC cells. 206 , 207 The key molecular mechanisms involve the inhibition of topoisomerase IIa activity 208 and interaction with DNA to induce double‐stranded pDNA cleavage. 209 Additionally, copper‐based complexes also promote cancer cell apoptosis through various mechanisms, 210 , 211 including influencing ROS production, 212 , 213 , 214 , 215 inhibiting the activity of the NF‐κB 215 and JAK2/STAT5 signalling pathway, 216 activating the p53 signalling pathway, 217 , 218 , 219 and activating caspase‐related proteins. 220 Moreover, copper‐based complexes can reverse chemoresistance in CRC treatment, particularly against drugs such as oxaliplatin, cisplatin and doxorubicin. In conjugation with oxaliplatin, copper‐based complexes inhibit the activity of the 26S proteasome, leading to the accumulation of polyubiquitinated proteins, suppression of the ubiquitin‐proteasome pathway, and triggering of endoplasmic reticulum stress. 221 Notably, cisplatin itself induces apoptosis through mechanisms such as caspase‐3/7 activation, p53 induction and Poly(ADP‐ribose) polymerase cleavage, whereas copper‐based complexes induce apoptosis in CRC cells. 36 For doxorubicin, these complexes trigger cell death via apoptosis, as evidenced by increased BAX protein expression relative to BCL‐2, depolarisation of the mitochondrial membrane potential, and autophagy. 222 Therefore, copper‐based complexes show great potential for CRC therapy, primarily by inducing immunogenic cell death.
In addition to copper‐based complexes, copper‐based nanoparticles have shown promising advancements in antitumour applications. In CRC treatment using nanomedicines, delivering lethal doses of copper to cancer cells is crucial for the induction of cell death. The use of inorganic nanoparticles for copper delivery is a viable strategy. Recently, our group developed two advanced CRC treatment strategies specifically for patients with low‐lying rectal cancer following radiotherapy. We designed a copper‐ion‐loaded oxygen generator and sodium alginate hydrogel containing elesclomol‐Cu and galactose. Our study indicated that an increased concentration of copper ions in CRC cells could efficiently trigger cuproptosis. Furthermore, these innovative treatments counteracted the radiotherapy‐induced upregulation of PD‐L1 expression. Consequently, these novel therapeutic strategies exhibit promising potential for treating CRC by promoting cuproptosis and reversing tumour immune escape. 223 , 224 Chang et al. highlighted the development of release‐based nanomedicine by designing a core‐shell nanostructure composed of Cu2O@CaCO3, which was further modified with hyaluronic acid. This design aims to provide synergistic therapies targeting CRC and triggered by the TME, including photothermal, photodynamic, chemodynamic and calcium overload‐mediated approaches. 225 Through biomimetic modifications, the blood circulation and tumour targeting abilities of these nanomedicines were significantly enhanced. The underlying mechanisms primarily involve the induction of apoptosis and autophagy in CRC cells through the upregulation of Bax, p53 and caspase expression. 226 , 227 , 228 , 229 , 230 Zhou et al. revealed that microwave dynamic therapy mediated by copper‐based nanoparticle could improve cancer treatment by inducing ferroptosis in CRC cells. 48 Furthermore, copper‐based nanomedicines can suppress distant metastasis and recurrence of CRC by inhibiting COX‐2 expression, reprogramming macrophages from M2 phenotype to M1 phenotype, and stimulating an immune response akin to vaccination upon primary tumour elimination. 225 , 231 For addressing CRC chemoresistance, copper‐based nanomedicines could be used as drug delivery system for oxaliplatin and 5‐FU‐resistant CRC cells. 232 , 233 In mice models, a copper polypyridine complex encapsulated in natural nanocarriers triggered autophagy‐dependent apoptosis, exhibiting potent tumour growth inhibition. 234 In conclusion, copper‐based compounds are primarily utilised in the treatment of CRC by inducing cell death and exhibit notable efficacy in chemotherapy‐resistant situations. However, these compounds predominantly target divalent copper ions. The stabilisation of monovalent copper ions and the development of compounds that minimise intracellular reduction while augmenting their therapeutic effects are potential research directions.
4.4. Radioisotope 64Cu in CRC treatment
Copper‐64 (64Cu) is a radioisotope with a half‐life of approximately 12.7 h, which can be simultaneously used for both positron emission tomography (PET) imaging and potential therapy due to its specific decay properties (β+ and β−). Emerging evidence supports the use of 64Cu radio‐pharmaceuticals for PET imaging and treatment of hypoxic (low tissue oxygenation) solid tumours. The diagnostic accuracy of CRC has been greatly improved by 64Cu, which not only enhances the uptake of radioactive copper in CRC lesions in vivo 235 , 236 , 237 but is also a promising PET tracer for noninvasive imaging of VEGF expression in CRC xenografts. 238 Additionally, it facilitates the detection of small CRC lesions in the pelvic and abdominal cavities, enabling precise assessment of CRC infiltration in a noninvasive manner. 239
In addition to its imaging capabilities, 64Cu significantly improves cell death and reduces chemo‐resistance in CRC. Studies have shown that ATOX1 and wild‐type p53 promote the entry of 64Cu into the nucleus, thereby enhancing its cytotoxic effects on CRC cells. 240 In addition, copper‐67 (67Cu) also plays an important role in advancing CRC radio‐immunotherapy by synergistically enhancing the eradication of CRC cells and metastatic liver foci. 241 , 242 When conjugated to platinum, 64Cu showed extreme efficacy in destroying CRC cells. 243 , 244 Furthermore, when 64Cu was conjugated to cetuximab, it killed both CRC cells and wild‐type KRAS CRC cells. 245 , 246 Therefore, radioisotope 64Cu has shown promising applications in the theranostics of CRC.
5. FUTURE PERSPECTIVES AND CONCLUSIONS
Copper, as a trace element, plays a crucial role in copper homeostasis, signalling pathways and cellular metabolism, which in turn influence cell proliferation, angiogenesis and distant metastasis in numerous cancers. Disruption of copper homeostasis can lead to excess copper accumulation, resulting in elevated ROS levels and promotion of proliferation, a process known as cuproplasia. Conversely, an excessive copper levels in the body can trigger cuproptosis. Therefore, it is vital to maintain a balanced fluctuation of copper ions in the body.
This review investigated the role of copper ions in CRC by focusing on two primary aspects. First, excess copper ions promote CRC proliferation, metastasis, angiogenesis and antitumour immune response, primarily through mechanisms such as the activation of the NF‐κB signalling pathway, CyclinD1 transcriptional activation, angiogenesis promotion, and EMT. Second, elevated copper ion concentrations can trigger cuproptosis, which is usually higher in CRC tissues than in neighbouring tissues. Therefore, this distinction is crucial for targeted copper ion therapy for CRC, which currently involves copper ion chelates, ionophores and compounds. These therapies target copper ions to inhibit cancer cell proliferation or induce cell death. Additionally, radioisotope 64Cu is significant for the theranostics of CRC. This review not only highlights the functional roles and mechanisms of copper ions in physiological states, providing valuable insights for future research, but also underscores the role of the intestinal tract in the absorption and utilisation of copper ions, offering theoretical support for future CRC treatments. Current CRC treatments targeting copper ions are notable, particularly considering the elevated serum copper levels (mainly detected by plasma CP levels) in patients with CRC, which could facilitate noninvasive, real‐time monitoring of CRC progression and recurrence. However, the molecular mechanisms underlying the effects of copper on CRC remain unclear. A key challenge is effectively differentiating between cuproplasia and cuproptosis in the human body. The development of reliable methods to quantify total copper levels in plasma and cells, including their labile forms and bioavailable, is critical for the creation of companion diagnostics. These diagnoses help identify patient populations suitable for copper‐targeted therapy. Additionally, evaluating patients’ functional ‘copper status’ is imperative to reduce adverse effects and assess treatment effectiveness. Furthermore, it remains critical to determine whether there are differences in the range of copper concentrations for cuproplasia and cuproptosis mutations in different cancer types.
In the future, homeostasis research will be poised for significant breakthroughs, especially in converting insights from copper homeostasis mechanisms into effective CRC treatments. From a molecular perspective, the investigation of copper‐dependent signalling pathways, along with the control of cuproplasia and cuproptosis via copper‐centric molecular mechanisms, is the most promising area of study. Given the significant impact of immunotherapy on CRC treatment in recent years, combined with extensive bioinformatics research on the role of copper ions in CRC, future studies should focus on exploring the interactions between copper ions and the TME and their effects on immune responses. This includes investigating the multiomics characteristics of CRC under specific conditions such as different microsatellite stability statuses, resistance to drug treatments, patients with KRAS gene mutations, and variations in copper ion levels in tissues. By analysing of a large number of clinical patient samples, this research will provide a deeper theoretical and practical foundation for future immunotherapy strategies targeting copper ions.
In clinical trials, strategies for treating tumours by targeting copper ions have primarily focused on the use of approved copper chelators (e.g. TM) and copper ionophores (e.g. DSF). Although these agents have demonstrated the potential to inhibit tumour growth in existing clinical models, there is a need to more precise identification of patient populations that would benefit from them, particularly in terms of specifying the types of tumours and understanding the genetic background of the patients. From a clinical perspective, copper is an important trace element involved in numerous biological processes, and comprehensive inhibition mediated by targeting copper can negatively affect normal cellular functions, especially during long‐term treatment. Additionally, determining the correct dosage and the bioavailability of drug is challenging and requires a balance between effective antitumour activity and reduced toxicity to normal tissues. Finally, clinical experiments targeting copper ions will also require the development of biomarkers to predict and monitor copper levels and their bioactivity, as well as the development of combination therapies with other therapeutic modalities to improve efficacy and reduce adverse effects on normal tissues.
Notably, copper nanoparticles represent a potential anticancer material with diverse applications. As a drug carrier, they can enhance the efficiency of drug delivery into tumour cells, increase therapeutic specificity and reduce damage to healthy cells. Utilising their optical properties, copper nanoparticles can kill cancer cells by generating local high temperatures or ROS under laser irradiation. Moreover, these treatments can enhance the efficacy of chemotherapy and radiotherapy. Despite these advantages, the application of copper nanoparticles faces several challenges, including ensuring biocompatibility, controlling toxicity, improving tumour targeting, and studying their metabolism and clearance mechanisms to prevent potential side effects. Future research should focus on improving the design of copper nanoparticles by adjusting their size, shape and surface modifications to enhance therapeutic outcomes; developing multifunctional copper nanoparticles that integrate drug delivery, photothermal, and photodynamic therapy functions for a unified treatment approach; and designing personalised treatment systems for specific tumour types or patients by integrating genomic and proteomic data.
Therefore, in this review, acknowledging the intestine as the primary site for copper ion absorption and CRC as the most common type of cancer within the gastrointestinal tract, we systematically explored the physiological roles and utilisation processes of copper ions in the body. Considering the narrow physiological range of copper ion concentrations, this review comprehensively assessed the role of copper ions in mediating disease under pathological conditions, and clarified how abnormal copper ion concentrations drive CRC proliferation and metastasis. Furthermore, this review outlined the role of targeted copper ion strategies in CRC diagnosis and treatment, emphasising the cytotoxic effects of copper ions inducing multiple forms of cell death as well as mechanisms for reversing chemoresistance, reducing angiogenesis, enhancing immunotherapy and increasing ROS levels (Figure 5). Through this comprehensive summary, we aim to deepen our understanding of the physiological roles of copper ions and their mechanisms of action in CRC, providing a theoretical foundation for the future diagnosis and treatment of CRC and other malignancies.
FIGURE 5.

The schematic illustration delineates the multifaceted role of copper ions in CRC malignancy, alongside their therapeutic and diagnostic applications based on copper. (A) The normal transport mechanisms of copper ions within the body and their utilisation at the intracellular level. (B) The detailed molecular mechanisms through which copper ions contribute to CRC proliferation and metastasis. (C) The mode of cell death induced by copper ion toxicity. (D) The mechanisms by which copper ions are involved in the treatment and diagnosis of CRC.
AUTHOR CONTRIBUTIONS
YL, LCG and KY designed and conducted this review. KY critically revised the final version of the manuscript. YHW conceived and drafted this manuscript. YHW and PP drew the figures. YHW drew the tables. All the authors approved the final version of the manuscript.
CONFLICT OF INTEREST STATEMENT
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
ETHICS STATEMENT
Not applicable.
ACKNOWLEDGEMENTS
We extend our profound gratitude to Professor Ledong Wan (State University of New York) for his meticulous revision and invaluable guidance throughout the development of this paper. The figures were created with BioRender. This work was supported by Suzhou Basic Research Key Project (No. SKY2023009), Suzhou health youth backbone talent ‘national tutorial system’ training project (Qngg2023005) and Beijing Xisike Clinical Oncology Research Foundation (Y‐tongshu2021/qn‐0366).
Wang Y, Pei P, Yang K, Guo L, Li Y. Copper in colorectal cancer: From copper‐related mechanisms to clinical cancer therapies. Clin Transl Med. 2024;e1724. 10.1002/ctm2.1724
Contributor Information
Kai Yang, Email: kyang@suda.edu.cn.
Lingchuan Guo, Email: szglc@hotmail.com.
Yuan Li, Email: lumoxuan2009@163.com.
REFERENCES
- 1. Siegel RL, Miller KD, Wagle NS, Jemal A. Cancer statistics, 2023. CA Cancer J Clin. 2023;73:17‐48. 10.3322/caac.21763 [DOI] [PubMed] [Google Scholar]
- 2. Dekker E, Tanis PJ, Vleugels JLA, Kasi PM, Wallace MB. Colorectal cancer. Lancet. 2019;394:1467‐1480. 10.1016/S0140-6736(19)32319-0 [DOI] [PubMed] [Google Scholar]
- 3. Li J, Ma X, Chakravarti D, Shalapour S, Depinho RA. Genetic and biological hallmarks of colorectal cancer. Genes Dev. 2021;35:787‐820. 10.1101/gad.348226.120 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. Festa RA, Thiele DJ. Copper: an essential metal in biology. Curr Biol. 2011;21:R877‐R883. 10.1016/j.cub.2011.09.040 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5. Brady DC, Crowe MS, Turski ML, et al. Copper is required for oncogenic BRAF signalling and tumorigenesis. Nature. 2014;509:492‐496. 10.1038/nature13180 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6. Shanbhag V, Jasmer‐Mcdonald K, Zhu S, et al. ATP7A delivers copper to the lysyl oxidase family of enzymes and promotes tumorigenesis and metastasis. Proc Natl Acad Sci U S A. 2019;116:6836‐6841. 10.1073/pnas.1817473116 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7. Tsang T, Posimo JM, Gudiel AA, Cicchini M, Feldser DM, Brady DC. Copper is an essential regulator of the autophagic kinases ULK1/2 to drive lung adenocarcinoma. Nat Cell Biol. 2020;22:412‐424. 10.1038/s41556-020-0481-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8. Xiong C, Ling H, Hao Q, Zhou X. Cuproptosis: p53‐regulated metabolic cell death? Cell Death Differ. 2023;30:876‐884. 10.1038/s41418-023-01125-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Matassa DS, Amoroso MR, Lu H, Avolio R, et al. Oxidative metabolism drives inflammation‐induced platinum resistance in human ovarian cancer. Cell Death Differ. 2016;23:1542‐1554. 10.1038/cdd.2016.39 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Porporato PE, Filigheddu N, Pedro JMB‐S, Kroemer G, Galluzzi L. Mitochondrial metabolism and cancer. Cell Res. 2018;28:265‐280. 10.1038/cr.2017.155 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11. Grochowski C, Blicharska E, Baj J, et al. Serum iron, magnesium, copper, and manganese levels in alcoholism: a systematic review. Molecules. 2019;24:1361. 10.3390/molecules24071361 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Linder Mc, Wooten L, Cerveza P, Cotton S, Shulze R, Lomeli N. Copper transport. Am J Clin Nutr. 1998;67:965S‐971S. 10.1093/ajcn/67.5.965S [DOI] [PubMed] [Google Scholar]
- 13. 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. 10.1016/j.jtemb.2020.126629 [DOI] [PubMed] [Google Scholar]
- 14. Wu Z, Lv G, Xing F, et al. Copper in hepatocellular carcinoma: a double‐edged sword with therapeutic potentials. Cancer Lett. 2023;571:216348. 10.1016/j.canlet.2023.216348 [DOI] [PubMed] [Google Scholar]
- 15. Moriya M, Ho Y‐H, Grana A, et al. Copper is taken up efficiently from albumin and alpha2‐macroglobulin by cultured human cells by more than one mechanism. Am J Physiol Cell Physiol. 2008;295:C708‐C721. 10.1152/ajpcell.00029.2008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Ramos D, Mar D, Ishida M, et al. Mechanism of copper uptake from blood plasma ceruloplasmin by mammalian cells. PLoS One. 2016;11:e0149516. 10.1371/journal.pone.0149516 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Sharp PA. Ctr1 and its role in body copper homeostasis. Int J Biochem Cell Biol. 2003;35:288‐291. 10.1016/s1357-2725(02)00134-6 [DOI] [PubMed] [Google Scholar]
- 18. Weiss KC, Linder MC. Copper transport in rats involving a new plasma protein. Am J Physiol. 1985;249:E77‐E88. 10.1152/ajpendo.1985.249.1.E77 [DOI] [PubMed] [Google Scholar]
- 19. Klevay LM. Cardiovascular disease from copper deficiency – a history. J Nutr. 2000;130:489S‐492S. 10.1093/jn/130.2.489S [DOI] [PubMed] [Google Scholar]
- 20. Myint ZW, Oo TH, Thein KZ, Tun AM, Saeed H. Copper deficiency anemia: review article. Ann Hematol. 2018;97:1527‐1534. 10.1007/s00277-018-3407-5 [DOI] [PubMed] [Google Scholar]
- 21. Strain JJ. A reassessment of diet and osteoporosis–possible role for copper. Med Hypotheses. 1988;27:333‐338. 10.1016/0306-9877(88)90016-3 [DOI] [PubMed] [Google Scholar]
- 22. Tadini‐Buoninsegni F, Smeazzetto S. Mechanisms of charge transfer in human copper ATPases ATP7A and ATP7B. IUBMB Life. 2017;69:218‐225. 10.1002/iub.1603 [DOI] [PubMed] [Google Scholar]
- 23. Tümer Z, Møller LB. Menkes disease. Eur J Hum Genet. 2010;18:511‐518. 10.1038/ejhg.2009.187 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Cobine PA, Moore SA, Leary SC. Getting out what you put in: copper in mitochondria and its impacts on human disease. Biochim Biophys Acta Mol Cell Res. 2021;1868:118867. 10.1016/j.bbamcr.2020.118867 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Feng J‐F, Lu L, Zeng P, et al. Serum total oxidant/antioxidant status and trace element levels in breast cancer patients. Int J Clin Oncol. 2012;17:575‐583. 10.1007/s10147-011-0327-y [DOI] [PubMed] [Google Scholar]
- 26. Feng W, Ye F, Xue W, Zhou Z, Kang YJ. Copper regulation of hypoxia‐inducible factor‐1 activity. Mol Pharmacol. 2009;75:174‐182. 10.1124/mol.108.051516 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Ge EJ, Bush AI, Casini A, et al. Connecting copper and cancer: from transition metal signalling to metalloplasia. Nat Rev Cancer. 2022;22:102‐113. 10.1038/s41568-021-00417-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Ishida S, Andreux P, Poitry‐Yamate C, Auwerx J, Hanahan D. Bioavailable copper modulates oxidative phosphorylation and growth of tumors. Proc Natl Acad Sci U S A. 2013;110:19507‐19512. 10.1073/pnas.1318431110 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. Gu M, Cooper J, Butler P, et al. Oxidative‐phosphorylation defects in liver of patients with Wilson's disease. Lancet. 2000;356:469‐474. 10.1016/s0140-6736(00)02556-3 [DOI] [PubMed] [Google Scholar]
- 30. Bremner I. Manifestations of copper excess. Am J Clin Nutr. 1998;67:1069S‐1073S. 10.1093/ajcn/67.5.1069S [DOI] [PubMed] [Google Scholar]
- 31. Kawakami M, Inagawa R, Hosokawa T, Saito T, Kurasaki M. Mechanism of apoptosis induced by copper in PC12 cells. Food Chem Toxicol. 2008;46:2157‐2164. 10.1016/j.fct.2008.02.014 [DOI] [PubMed] [Google Scholar]
- 32. Sutton HC, Winterbourn CC. On the participation of higher oxidation states of iron and copper in Fenton reactions. Free Radic Biol Med. 1989;6:53‐60. 10.1016/0891-5849(89)90160-3 [DOI] [PubMed] [Google Scholar]
- 33. Li G‐N, Zhao X‐J, Wang Z, et al. Elaiophylin triggers paraptosis and preferentially kills ovarian cancer drug‐resistant cells by inducing MAPK hyperactivation. Signal Transduct Target Ther. 2022;7:317. 10.1038/s41392-022-01131-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Zhou Y, Huang F, Yang Y, et al. Paraptosis‐Inducing nanomedicine overcomes cancer drug resistance for a potent cancer therapy. Small. 2018;14:1702446. 10.1002/smll.201702446 [DOI] [PubMed] [Google Scholar]
- 35. Dixon SJ, Lemberg KM, Lamprecht MR, et al. Ferroptosis: an iron‐dependent form of nonapoptotic cell death. Cell. 2012;149:1060‐1072. 10.1016/j.cell.2012.03.042 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36. Kabolizadeh P, Ryan J, Farrell N. Differences in the cellular response and signaling pathways of cisplatin and BBR3464 ([[trans‐PtCl(NH3)(2)]2mu‐(trans‐Pt(NH3)(2)(H2N(CH2)(6)‐NH2)2)]4+) influenced by copper homeostasis. Biochem Pharmacol. 2007;73:1270‐1279. 10.1016/j.bcp.2006.12.016 [DOI] [PubMed] [Google Scholar]
- 37. Yang F, Pei R, Zhang Z, et al. Copper induces oxidative stress and apoptosis through mitochondria‐mediated pathway in chicken hepatocytes. Toxicol In Vitro. 2019;54:310‐316. 10.1016/j.tiv.2018.10.017 [DOI] [PubMed] [Google Scholar]
- 38. Shen W‐Y, Jia C‐P, Liao L‐Y, et al. Copper(II) complex enhanced chemodynamic therapy through GSH depletion and autophagy flow blockade. Dalton Trans. 2023;52:3287‐3294. 10.1039/d2dt04108a [DOI] [PubMed] [Google Scholar]
- 39. Galdino ACM, Viganor L, Pereira MM, et al. Copper(II) and silver(I)‐1,10‐phenanthroline‐5,6‐dione complexes interact with double‐stranded DNA: further evidence of their apparent multi‐modal activity towards Pseudomonas aeruginosa . J Biol Inorg Chem. 2022;27:201‐213. 10.1007/s00775-021-01922-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Liu N, Liu C, Li X, et al. A novel proteasome inhibitor suppresses tumor growth via targeting both 19S proteasome deubiquitinases and 20S proteolytic peptidases. Sci Rep. 2014;4:5240. 10.1038/srep05240 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41. Tardito S, Bassanetti I, Bignardi C, et al. Copper binding agents acting as copper ionophores lead to caspase inhibition and paraptotic cell death in human cancer cells. J Am Chem Soc. 2011;133:6235‐6242. 10.1021/ja109413c [DOI] [PubMed] [Google Scholar]
- 42. Barilli A, Atzeri C, Bassanetti I, et al. Oxidative stress induced by copper and iron complexes with 8‐hydroxyquinoline derivatives causes paraptotic death of HeLa cancer cells. Mol Pharm. 2014;11:1151‐1163. 10.1021/mp400592n [DOI] [PubMed] [Google Scholar]
- 43. Ren X, Li Y, Zhou Y, et al. Overcoming the compensatory elevation of NRF2 renders hepatocellular carcinoma cells more vulnerable to disulfiram/copper‐induced ferroptosis. Redox Biol. 2021;46:102122. 10.1016/j.redox.2021.102122 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Gao W, Huang Z, Duan J, Nice EC, Lin J, Huang C. Elesclomol induces copper‐dependent ferroptosis in colorectal cancer cells via degradation of ATP7A. Mol Oncol. 2021;15:3527‐3544. 10.1002/1878-0261.13079 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. Xue Q, Yan D, Chen X, et al. Copper‐dependent autophagic degradation of GPX4 drives ferroptosis. Autophagy. 2023;19:1982‐1996. 10.1080/15548627.2023.2165323 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. Li F, Wu X, Liu H, et al. Copper depletion strongly enhances ferroptosis via mitochondrial perturbation and reduction in antioxidative mechanisms. Antioxidants (Basel). 2022;11:2084. 10.3390/antiox11112084 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47. Martin F, Linden T, Katschinski DM, et al. Copper‐dependent activation of hypoxia‐inducible factor (HIF)‐1: implications for ceruloplasmin regulation. Blood. 2005;105:4613‐4619. 10.1182/blood-2004-10-3980 [DOI] [PubMed] [Google Scholar]
- 48. Zhou H, Liu Z, Zhang Z, et al. Copper‐cysteamine nanoparticle‐mediated microwave dynamic therapy improves cancer treatment with induction of ferroptosis. Bioact Mater. 2023;24:322‐330. 10.1016/j.bioactmat.2022.12.023 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49. Tsvetkov P, Coy S, Petrova B, et al. Copper induces cell death by targeting lipoylated TCA cycle proteins. Science. 2022;375:1254‐1261. 10.1126/science.abf0529 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50. Li L, Sun F, Kong F, et al. Characterization of a cuproptosis‐related signature to evaluate immune features and predict prognosis in colorectal cancer. Front Oncol. 2023;13:1083956. 10.3389/fonc.2023.1083956 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51. Tsvetkov P, Detappe A, Cai K, et al. Mitochondrial metabolism promotes adaptation to proteotoxic stress. Nat Chem Biol. 2019;15:681‐689. 10.1038/s41589-019-0291-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Mason KE. A conspectus of research on copper metabolism and requirements of man. J Nutr. 1979;109:1979‐2066. 10.1093/jn/109.11.1979 [DOI] [PubMed] [Google Scholar]
- 53. Dancis A, Roman DG, Anderson GJ, Hinnebusch AG, Klausner RD. Ferric reductase of Saccharomyces cerevisiae: molecular characterization, role in iron uptake, and transcriptional control by iron. Proc Natl Acad Sci U S A. 1992;89:3869‐3873. 10.1073/pnas.89.9.3869 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54. Georgatsou E, Mavrogiannis LA, Fragiadakis GS, Alexandraki D. The yeast Fre1p/Fre2p cupric reductases facilitate copper uptake and are regulated by the copper‐modulated Mac1p activator. J Biol Chem. 1997;272:13786‐13792. 10.1074/jbc.272.21.13786 [DOI] [PubMed] [Google Scholar]
- 55. Lane D, Bae DH, Merlot AM, Sahni S, Richardson DR. Duodenal cytochrome b (DCYTB) in iron metabolism: an update on function and regulation. Nutrients. 2015;7:2274‐2296. 10.3390/nu7042274 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. McKie AT, Barrow D, Latunde‐Dada GO, et al. An iron‐regulated ferric reductase associated with the absorption of dietary iron. Science. 2001;291:1755‐1759. 10.1126/science.1057206 [DOI] [PubMed] [Google Scholar]
- 57. Gomes IM, Maia CJ, Santos CR. STEAP proteins: from structure to applications in cancer therapy. Mol Cancer Res. 2012;10:573‐587. 10.1158/1541-7786.MCR-11-0281 [DOI] [PubMed] [Google Scholar]
- 58. Ohgami RS, Campagna DR, Greer EL, et al. Identification of a ferrireductase required for efficient transferrin‐dependent iron uptake in erythroid cells. Nat Genet. 2005;37:1264‐1269. 10.1038/ng1658 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59. Finegold AA, Shatwell KP, Segal AW, Klausner RD, Dancis A. Intramembrane bis‐heme motif for transmembrane electron transport conserved in a yeast iron reductase and the human NADPH oxidase. J Biol Chem. 1996;271:31021‐31024. 10.1074/jbc.271.49.31021 [DOI] [PubMed] [Google Scholar]
- 60. Arnesano F, Natile G. Interference between copper transport systems and platinum drugs. Semin Cancer Biol. 2021;76:173‐188. 10.1016/j.semcancer.2021.05.023 [DOI] [PubMed] [Google Scholar]
- 61. Colombo E, Triolo D, Bassani C, et al. Dysregulated copper transport in multiple sclerosis may cause demyelination via astrocytes. Proc Natl Acad Sci U S A. 2021;118:e2025804118. 10.1073/pnas.2025804118 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Ilyechova E, Bonaldi E, Orlov I, Skomorokhova E, Puchkova L, Broggini M. CRISP‐R/Cas9 mediated deletion of copper transport genes CTR1 and DMT1 in NSCLC cell line H1299. Biological and pharmacological consequences. Cells. 2019;8:322. 10.3390/cells8040322 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Lin C, Zhang Z, Wang T, Chen C, James Kang Y. Copper uptake by DMT1: a compensatory mechanism for CTR1 deficiency in human umbilical vein endothelial cells. Metallomics. 2015;7:1285‐1289. 10.1039/c5mt00097a [DOI] [PubMed] [Google Scholar]
- 64. Lönnerdal B. Intestinal regulation of copper homeostasis: a developmental perspective. Am J Clin Nutr. 2008;88:846S‐850S. 10.1093/ajcn/88.3.846S [DOI] [PubMed] [Google Scholar]
- 65. Liang ZD, Tsai W‐B, Lee M‐Y, Savaraj N, Kuo MT. 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. 10.1124/mol.111.076422 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. Lee J, Petris MJ, Thiele DJ. Characterization of mouse embryonic cells deficient in the ctr1 high affinity copper transporter. Identification of a Ctr1‐independent copper transport system. J Biol Chem. 2002;277:40253‐40259. 10.1074/jbc.M208002200 [DOI] [PubMed] [Google Scholar]
- 67. Nose Y, Kim B‐E, Thiele DJ. Ctr1 drives intestinal copper absorption and is essential for growth, iron metabolism, and neonatal cardiac function. Cell Metab. 2006;4:235‐244. 10.1016/j.cmet.2006.08.009 [DOI] [PubMed] [Google Scholar]
- 68. Öhrvik H, Nose Y, Wood LK, et al. Ctr2 regulates biogenesis of a cleaved form of mammalian Ctr1 metal transporter lacking the copper‐ and cisplatin‐binding ecto‐domain. Proc Natl Acad Sci U S A. 2013;110:E4279‐E4288. 10.1073/pnas.1311749110 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69. Öhrvik H, Thiele DJ. The role of Ctr1 and Ctr2 in mammalian copper homeostasis and platinum‐based chemotherapy. J Trace Elem Med Biol. 2015;31:178‐182. 10.1016/j.jtemb.2014.03.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70. Aggett PJ. An overview of the metabolism of copper. Eur J Med Res. 1999;4:214‐216. [PubMed] [Google Scholar]
- 71. Peña MMO, Lee J, Thiele DJ. A delicate balance: homeostatic control of copper uptake and distribution. J Nutr. 1999;129:1251‐1260. 10.1093/jn/129.7.1251 [DOI] [PubMed] [Google Scholar]
- 72. Thiele DJ. Integrating trace element metabolism from the cell to the whole organism. J Nutr. 2003;133:1579S‐1580S. 10.1093/jn/133.5.1579S [DOI] [PubMed] [Google Scholar]
- 73. La Fontaine S, Mercer JFB. Trafficking of the copper‐ATPases, ATP7A and ATP7B: role in copper homeostasis. Arch Biochem Biophys. 2007;463:149‐167. 10.1016/j.abb.2007.04.021 [DOI] [PubMed] [Google Scholar]
- 74. Lutsenko S, Barnes NL, Bartee MY, Dmitriev OY. Function and regulation of human copper‐transporting ATPases. Physiol Rev. 2007;87:1011‐1046. 10.1152/physrev.00004.2006 [DOI] [PubMed] [Google Scholar]
- 75. Cabrera A, Alonzo E, Sauble E, et al. Copper binding components of blood plasma and organs, and their responses to influx of large doses of (65)Cu, in the mouse. Biometals. 2008;21:525‐543. 10.1007/s10534-008-9139-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76. Kirsipuu T, Zadorožnaja A, Smirnova J, et al. Copper(II)‐binding equilibria in human blood. Sci Rep. 2020;10:5686. 10.1038/s41598-020-62560-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77. Jiang X, Chen J, Bajić A, et al. Quantitative real‐time imaging of glutathione. Nat Commun. 2017;8:16087. 10.1038/ncomms16087 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Juárez‐Rebollar D, Rios C, Nava‐Ruíz C, Méndez‐Armenta M. Metallothionein in brain disorders. Oxid Med Cell Longev. 2017;2017:5828056. 10.1155/2017/5828056 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Maryon EB, Molloy SA, Kaplan JH. Cellular glutathione plays a key role in copper uptake mediated by human copper transporter 1. Am J Physiol Cell Physiol. 2013;304:C768‐C779. 10.1152/ajpcell.00417.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80. Bonaccorsi M, Knight MJ, Le Marchand T, et al. Multimodal response to copper binding in superoxide dismutase dynamics. J Am Chem Soc. 2020;142:19660‐19667. 10.1021/jacs.0c09242 [DOI] [PubMed] [Google Scholar]
- 81. Freedman JH, Ciriolo MR, Peisach J. The role of glutathione in copper metabolism and toxicity. J Biol Chem. 1989;264:5598‐5605. [PubMed] [Google Scholar]
- 82. Rae TD, Schmidt PJ, Pufahl RA, Culotta VC, O'Halloran TV. Undetectable intracellular free copper: the requirement of a copper chaperone for superoxide dismutase. Science. 1999;284:805‐808. 10.1126/science.284.5415.805 [DOI] [PubMed] [Google Scholar]
- 83. Gudekar N, Shanbhag V, Wang Y, Ralle M, Weisman GA, Petris MJ. Metallothioneins regulate ATP7A trafficking and control cell viability during copper deficiency and excess. Sci Rep. 2020;10:7856. 10.1038/s41598-020-64521-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84. Casareno RLB, Waggoner D, Gitlin JD. The copper chaperone CCS directly interacts with copper/zinc superoxide dismutase. J Biol Chem. 1998;273:23625‐23628. 10.1074/jbc.273.37.23625 [DOI] [PubMed] [Google Scholar]
- 85. Furukawa Y, Torres AS, O'halloran TV. Oxygen‐induced maturation of SOD1: a key role for disulfide formation by the copper chaperone CCS. EMBO J. 2004;23:2872‐2881. 10.1038/sj.emboj.7600276 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86. Dong X, Zhang Z, Zhao J, et al. The rational design of specific SOD1 inhibitors via copper coordination and their application in ROS signaling research. Chem Sci. 2016;7:6251‐6262. 10.1039/c6sc01272h [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. Sturtz LA, Diekert K, Jensen LT, Lill R, Culotta VC. A fraction of yeast Cu,Zn‐superoxide dismutase and its metallochaperone, CCS, localize to the intermembrane space of mitochondria. A physiological role for SOD1 in guarding against mitochondrial oxidative damage. J Biol Chem. 2001;276:38084‐38089. 10.1074/jbc.M105296200 [DOI] [PubMed] [Google Scholar]
- 88. Punter FA, Adams DL, Glerum DM. Characterization and localization of human COX17, a gene involved in mitochondrial copper transport. Hum Genet. 2000;107:69‐74. 10.1007/s004390000339 [DOI] [PubMed] [Google Scholar]
- 89. Morgada MN, Abriata LA, Cefaro C, Gajda K, Banci L, Vila AJ. Loop recognition and copper‐mediated disulfide reduction underpin metal site assembly of CuA in human cytochrome oxidase. Proc Natl Acad Sci U S A. 2015;112:11771‐11776. 10.1073/pnas.1505056112 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90. Cobine PA, Pierrel F, Winge DR. Copper trafficking to the mitochondrion and assembly of copper metalloenzymes. Biochim Biophys Acta. 2006;1763:759‐772. 10.1016/j.bbamcr.2006.03.002 [DOI] [PubMed] [Google Scholar]
- 91. Timón‐Gómez A, Nývltová E, Abriata LA, Vila AJ, Hosler J, Barrientos A. Mitochondrial cytochrome c oxidase biogenesis: recent developments. Semin Cell Dev Biol. 2018;76:163‐178. 10.1016/j.semcdb.2017.08.055 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92. Suzuki C, Daigo Y, Kikuchi T, Katagiri T, Nakamura Y. Identification of COX17 as a therapeutic target for non‐small cell lung cancer. Cancer Res. 2003;63:7038‐7041. [PubMed] [Google Scholar]
- 93. Itoh S, Kim HW, Nakagawa O, et al. Novel role of antioxidant‐1 (Atox1) as a copper‐dependent transcription factor involved in cell proliferation. J Biol Chem. 2008;283:9157‐9167. 10.1074/jbc.M709463200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94. Hamza I, Faisst A, Prohaska J, Chen J, Gruss P, Gitlin JD. The metallochaperone Atox1 plays a critical role in perinatal copper homeostasis. Proc Natl Acad Sci U S A. 2001;98:6848‐6852. 10.1073/pnas.111058498 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95. Zhang X, Walke GR, Horvath I, et al. Memo1 binds reduced copper ions, interacts with copper chaperone Atox1, and protects against copper‐mediated redox activity in vitro. Proc Natl Acad Sci U S A. 2022;119:e2206905119. 10.1073/pnas.2206905119 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96. Fujieda N, Umakoshi K, Ochi Y, et al. Copper‐oxygen dynamics in the tyrosinase mechanism. Angew Chem Int Ed Engl. 2020;59:13385‐13390. 10.1002/anie.202004733 [DOI] [PubMed] [Google Scholar]
- 97. Xiao T, Ackerman CM, Carroll EC, et al. Copper regulates rest‐activity cycles through the locus coeruleus‐norepinephrine system. Nat Chem Biol. 2018;14:655‐663. 10.1038/s41589-018-0062-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98. Chen W, Yang A, Jia J, Popov YV, Schuppan D, You H. Lysyl oxidase (LOX) family members: rationale and their potential as therapeutic targets for liver fibrosis. Hepatology. 2020;72:729‐741. 10.1002/hep.31236 [DOI] [PubMed] [Google Scholar]
- 99. Salmi M, Jalkanen S. VAP‐1: an adhesin and an enzyme. Trends Immunol. 2001;22:211‐216. 10.1016/s1471-4906(01)01870-1 [DOI] [PubMed] [Google Scholar]
- 100. Hwang JJ, Park M‐H, Koh J‐Y. Copper activates TrkB in cortical neurons in a metalloproteinase‐dependent manner. J Neurosci Res. 2007;85:2160‐2166. 10.1002/jnr.21350 [DOI] [PubMed] [Google Scholar]
- 101. Michniewicz F, Saletta F, Rouaen JRC, et al. Copper: an intracellular achilles' heel allowing the targeting of epigenetics, kinase pathways, and cell metabolism in cancer therapeutics. ChemMedChem. 2021;16:2315‐2329. 10.1002/cmdc.202100172 [DOI] [PubMed] [Google Scholar]
- 102. He F, Chang C, Liu B, et al. Copper (II) ions activate ligand‐independent receptor tyrosine kinase (RTK) signaling pathway. Biomed Res Int. 2019;2019:4158415. 10.1155/2019/4158415 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103. Turski ML, Brady DC, Kim HJ, et al. A novel role for copper in Ras/mitogen‐activated protein kinase signaling. Mol Cell Biol. 2012;32:1284‐1295. 10.1128/MCB.05722-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104. Opazo CM, Lotan A, Xiao Z, et al., Nutrient copper signaling promotes protein turnover by allosteric activation of ubiquitin E2D conjugases. bioRxiv. 2021. doi: 10.1101/2021.02.15.431211 [DOI]
- 105. Barresi V, Trovato‐Salinaro A, Spampinato G, et al. Transcriptome analysis of copper homeostasis genes reveals coordinated upregulation of SLC31A1,SCO1, and COX11 in colorectal cancer. FEBS Open Bio. 2016;6:794‐806. 10.1002/2211-5463.12060 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106. Baszuk P, Marciniak W, Derkacz R, et al. Blood copper levels and the occurrence of colorectal cancer in Poland. Biomedicines. 2021;9:1628. 10.3390/biomedicines9111628 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107. Gupta SK, Shukla VK, Vaidya MP, Roy SK, Gupta S. Serum and tissue trace elements in colorectal cancer. J Surg Oncol. 1993;52:172‐175. 10.1002/jso.2930520311 [DOI] [PubMed] [Google Scholar]
- 108. Sharma K, Mittal DK, Kesarwani RC, Kamboj VP, Chowdhery. Diagnostic and prognostic significance of serum and tissue trace elements in breast malignancy. Indian J Med Sci. 1994;48:227‐232. [PubMed] [Google Scholar]
- 109. Aubert L, Nandagopal N, Steinhart Z, et al. Copper bioavailability is a KRAS‐specific vulnerability in colorectal cancer. Nat Commun. 2020;11:3701. 10.1038/s41467-020-17549-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110. Polishchuk EV, Merolla A, Lichtmannegger J, 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. e1175. 10.1053/j.gastro.2018.11.032 [DOI] [PubMed] [Google Scholar]
- 111. Liao Y, Zhao J, Bulek K, et al. Inflammation mobilizes copper metabolism to promote colon tumorigenesis via an IL‐17‐STEAP4‐XIAP axis. Nat Commun. 2020;11:900. 10.1038/s41467-020-14698-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112. Gurusamy KS, Farquharson MJ, Craig C, Davidson BR. An evaluation study of trace element content in colorectal liver metastases and surrounding normal livers by X‐ray fluorescence. Biometals. 2008;21:373‐378. 10.1007/s10534-007-9126-3 [DOI] [PubMed] [Google Scholar]
- 113. Bhuvanasundar R, John A, Sulochana K, Coral K, Deepa PR, Umashankar V. A molecular model of human Lysyl Oxidase (LOX) with optimal copper orientation in the catalytic cavity for induced fit docking studies with potential modulators. Bioinformation. 2014;10:406‐412. 10.6026/97320630010406 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114. Denoyer D, Masaldan S, La Fontaine S, Cater MA. Targeting copper in cancer therapy: ‘Copper That Cancer’. Metallomics. 2015;7:1459‐1476. 10.1039/c5mt00149h [DOI] [PubMed] [Google Scholar]
- 115. Jana A, Das A, Krett NL, et al. Nuclear translocation of Atox1 potentiates activin A‐induced cell migration and colony formation in colon cancer. PLoS One. 2020;15:e0227916. 10.1371/journal.pone.0227916 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116. Lowndes SA, Harris AL. The role of copper in tumour angiogenesis. J Mammary Gland Biol Neoplasia. 2005;10:299‐310. 10.1007/s10911-006-9003-7 [DOI] [PubMed] [Google Scholar]
- 117. Finney L, Vogt S, Fukai T, Glesne D. Copper and angiogenesis: unravelling a relationship key to cancer progression. Clin Exp Pharmacol Physiol. 2009;36:88‐94. 10.1111/j.1440-1681.2008.04969.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118. Pan Q, Kleer CG, van Golen KL, et al. Copper deficiency induced by tetrathiomolybdate suppresses tumor growth and angiogenesis. Cancer Res. 2002;62:4854‐4859. [PubMed] [Google Scholar]
- 119. Xie H, Kang Y. Role of copper in angiogenesis and its medicinal implications. Curr Med Chem. 2009;16:1304‐1314. 10.2174/092986709787846622 [DOI] [PubMed] [Google Scholar]
- 120. 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. 10.1159/000438485 [DOI] [PubMed] [Google Scholar]
- 121. Rigiracciolo DC, Scarpelli A, Lappano R, et al. Copper activates HIF‐1alpha/GPER/VEGF signalling in cancer cells. Oncotarget. 2015;6:34158‐34177. doi: 10.18632/oncotarget.5779 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122. Onuma T, Mizutani T, Fujita Y, Yamada S, Yoshida Y. Copper content in ascitic fluid is associated with angiogenesis and progression in ovarian cancer. J Trace Elem Med Biol. 2021;68:126865. 10.1016/j.jtemb.2021.126865 [DOI] [PubMed] [Google Scholar]
- 123. Das A, Ash D, Fouda AY, et al. Cysteine oxidation of copper transporter CTR1 drives VEGFR2 signalling and angiogenesis. Nat Cell Biol. 2022;24:35‐50. 10.1038/s41556-021-00822-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124. Li Z, Li S, Wen Y, Chen J, Liu K, Jia J. MiR‐495 inhibits cisplatin resistance and angiogenesis in esophageal cancer by targeting ATP7A. Technol Cancer Res Treat. 2021;20:15330338211039127. 10.1177/15330338211039127 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125. Shih Y‐H, Chang K‐W, Chen MY, et al. Lysyl oxidase and enhancement of cell proliferation and angiogenesis in oral squamous cell carcinoma. Head Neck. 2013;35:250‐256. 10.1002/hed.22959 [DOI] [PubMed] [Google Scholar]
- 126. Ruiz LM, Libedinsky A, Elorza AA. Role of copper on mitochondrial function and metabolism. Front Mol Biosci. 2021;8:711227. 10.3389/fmolb.2021.711227 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127. Shanbhag VC, Gudekar N, Jasmer K, Papageorgiou C, Singh K, Petris MJ. Copper metabolism as a unique vulnerability in cancer. Biochim Biophys Acta Mol Cell Res. 2021;1868:118893. 10.1016/j.bbamcr.2020.118893 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128. Leary SC, Sasarman F, Nishimura T, Shoubridge EA. Human SCO2 is required for the synthesis of CO II and as a thiol‐disulphide oxidoreductase for SCO1. Hum Mol Genet. 2009;18:2230‐2240. 10.1093/hmg/ddp158 [DOI] [PubMed] [Google Scholar]
- 129. Attardi G, Schatz G. Biogenesis of mitochondria. Annu Rev Cell Biol. 1988;4:289‐333. 10.1146/annurev.cb.04.110188.001445 [DOI] [PubMed] [Google Scholar]
- 130. Ussakli CH, Ebaee A, Binkley J, et al. Mitochondria and tumor progression in ulcerative colitis. J Natl Cancer Inst. 2013;105:1239‐1248. 10.1093/jnci/djt167 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131. Zhang K, Chen Y, Huang X, et al. Expression and clinical significance of cytochrome c oxidase subunit IV in colorectal cancer patients. Arch Med Sci. 2016;12:68‐77. 10.5114/aoms.2016.57581 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132. Hewedi IH, Farid RM, Sidhom KF, Salman MI, Mostafa RG. Differential expression of cytochrome C oxidase subunit I along the colorectal adenoma: carcinoma progression. Appl Immunohistochem Mol Morphol. 2018;26:689‐696. 10.1097/PAI.0000000000000509 [DOI] [PubMed] [Google Scholar]
- 133. Sacconi S, Trevisson E, Pistollato F, et al. hCOX18 and hCOX19: two human genes involved in cytochrome c oxidase assembly. Biochem Biophys Res Commun. 2005;337:832‐839. 10.1016/j.bbrc.2005.09.127 [DOI] [PubMed] [Google Scholar]
- 134. Guo Q, Zhang H, Zhang L, et al. MicroRNA‐21 regulates non‐small cell lung cancer cell proliferation by affecting cell apoptosis via COX‐19. Int J Clin Exp Med. 2015;8:8835‐8841. [PMC free article] [PubMed] [Google Scholar]
- 135. Li J‐P, Liu Y‐J, Zeng S‐H, Gao H‐J, Chen Y‐G, Zou X. Identification of COX4I2 as a hypoxia‐associated gene acting through FGF1 to promote EMT and angiogenesis in CRC. Cell Mol Biol Lett. 2022;27:76. 10.1186/s11658-022-00380-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136. Gao S, Zhang H, Zhang X, Wang J, Bai W, Jiang B. COX19 is a new target of MACC1 and promotes colorectal cancer progression by regulating copper transport in mitochondria. J Nutr. 2024;154:381‐394. 10.1016/j.tjnut.2023.12.032 [DOI] [PubMed] [Google Scholar]
- 137. Cordano A, Placko RP, Graham GG. Hypocupremia and neutropenia in copper deficiency. Blood. 1966;28:280‐283. [PubMed] [Google Scholar]
- 138. Dunlap WM, James GW 3rd, Hume DM. Anemia and neutropenia caused by copper deficiency. Ann Intern Med. 1974;80:470‐476. 10.7326/0003-4819-80-4-470 [DOI] [PubMed] [Google Scholar]
- 139. Ashkenazi A, Levin S, Djaldetti M, Fishel E, Benvenisti D. The syndrome of neonatal copper deficiency. Pediatrics. 1973;52:525‐533. [PubMed] [Google Scholar]
- 140. White C, Lee J, Kambe T, Fritsche K, Petris MJ. A role for the ATP7A copper‐transporting ATPase in macrophage bactericidal activity. J Biol Chem. 2009;284:33949‐33956. 10.1074/jbc.M109.070201 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141. Stafford SL, Bokil NJ, Achard MES, et al. Metal ions in macrophage antimicrobial pathways: emerging roles for zinc and copper. Biosci Rep. 2013;33:e00049. 10.1042/BSR20130014 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142. Flynn A, Yen BR. Mineral deficiency effects on the generation of cytotoxic T‐cells and T‐helper cell factors in vitro. J Nutr. 1981;111:907‐913. 10.1093/jn/111.5.907 [DOI] [PubMed] [Google Scholar]
- 143. Crowe A, Jackaman C, Beddoes KM, Ricciardo B, Nelson DJ. 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. 10.1371/journal.pone.0073684 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144. Gundelach JH, Madhavan AA, Wettstein PJ, Bram RJ. The anticancer drug Dp44mT inhibits T‐cell activation and CD25 through a copper‐dependent mechanism. FASEB J. 2013;27:782‐792. 10.1096/fj.12-215756 [DOI] [PubMed] [Google Scholar]
- 145. He R, Zhang H, Zhao H, et al. Multiomics analysis reveals cuproptosis‐related signature for evaluating prognosis and immunotherapy efficacy in colorectal cancer. Cancers (Basel). 2023;15:387. 10.3390/cancers15020387 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146. Zhang B, Li Y, Song L, et al. Cuproplasia characterization in colon cancer assists to predict prognosis and immunotherapeutic response. Front Oncol. 2023;13:1061084. 10.3389/fonc.2023.1061084 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147. Li S, Zhu Z, Lu J, et al. Prediction of prognosis, immune infiltration, and personalized treatment of hepatocellular carcinoma by analysis of cuproptosis‐related long noncoding RNAs and verification in vitro. Front Oncol. 2023;13:1159126. 10.3389/fonc.2023.1159126 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148. Wang L, Cao Y, Guo W, Xu J. High expression of cuproptosis‐related gene FDX1 in relation to good prognosis and immune cells infiltration in colon adenocarcinoma (COAD). J Cancer Res Clin Oncol. 2023;149:15‐24. 10.1007/s00432-022-04382-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149. Xu L, Wu P, Rong A, et al. Systematic pan‐cancer analysis identifies cuproptosis‐related gene DLAT as an immunological and prognostic biomarker. Aging (Albany NY). 2023;15:4269‐4287. doi: 10.18632/aging.204728 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150. Gao F, Yuan Y, Ding Y, Li P‐Y, Chang Y, He X‐X. DLAT as a cuproptosis promoter and a molecular target of elesclomol in hepatocellular carcinoma. Curr Med Sci. 2023;43:526‐538. 10.1007/s11596-023-2755-0 [DOI] [PubMed] [Google Scholar]
- 151. Chu B, Wang Y, Yang J, Dong B. Integrative analysis of single‐cell and bulk RNA seq to reveal the prognostic model and tumor microenvironment remodeling mechanisms of cuproptosis‐related genes in colorectal cancer. Aging (Albany NY). 2023;15:14422‐14444. doi: 10.18632/aging.205324 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152. Sun L, Zhang Y, Yang B, et al. Lactylation of METTL16 promotes cuproptosis via m(6)A‐modification on FDX1 mRNA in gastric cancer. Nat Commun. 2023;14:6523. 10.1038/s41467-023-42025-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153. Wang W, Lu K, Jiang X, et al. Ferroptosis inducers enhanced cuproptosis induced by copper ionophores in primary liver cancer. J Exp Clin Cancer Res. 2023;42:142. 10.1186/s13046-023-02720-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154. Huang Y, Yin D, Wu L. Identification of cuproptosis‐related subtypes and development of a prognostic signature in colorectal cancer. Sci Rep. 2022;12:17348. 10.1038/s41598-022-22300-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155. Du Y, Lin Y, Wang B, et al. Cuproptosis patterns and tumor immune infiltration characterization in colorectal cancer. Front Genet. 2022;13:976007. 10.3389/fgene.2022.976007 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156. Jiang P‐C, Fan J, Zhang C‐D, et al. Unraveling colorectal cancer and pan‐cancer immune heterogeneity and synthetic therapy response using cuproptosis and hypoxia regulators by multi‐omic analysis and experimental validation. Int J Biol Sci. 2023;19:3526‐3543. 10.7150/ijbs.84781 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157. Wu W, Dong J, Lv Y, Chang D. Cuproptosis‐Related genes in the prognosis of colorectal cancer and their correlation with the tumor microenvironment. Front Genet. 2022;13:984158. 10.3389/fgene.2022.984158 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158. Zhu Z, Zhao Q, Song W, et al. A novel cuproptosis‐related molecular pattern and its tumor microenvironment characterization in colorectal cancer. Front Immunol. 2022;13:940774. 10.3389/fimmu.2022.940774 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159. Huang H, Long Z, Xie Y, et al. Molecular subtypes based on cuproptosis‐related genes and tumor microenvironment infiltration characterization in colorectal cancer. J Oncol. 2022;2022:5034092. 10.1155/2022/5034092 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160. Zhu Z, Guo T, Weng J, et al. Cuproptosis‐related miRNAs signature and immune infiltration characteristics in colorectal cancer. Cancer Med. 2023;12:16661‐16678. 10.1002/cam4.6270 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161. Yang W, Wang Y, Huang Y, et al. 4‐Octyl itaconate inhibits aerobic glycolysis by targeting GAPDH to promote cuproptosis in colorectal cancer. Biomed Pharmacother. 2023;159:114301. 10.1016/j.biopha.2023.114301 [DOI] [PubMed] [Google Scholar]
- 162. Zhou L, Zhang Y, Xu Y, Jiang T, Tang L. Identification of a novel prognostic signature composed of 3 cuproptosis‐related transcription factors in colon adenocarcinoma. Genes Genomics. 2023;45:1047‐1061. 10.1007/s13258-023-01406-5 [DOI] [PubMed] [Google Scholar]
- 163. Wang C, Guo J, Zhang Y, Zhou S, Jiang B. Cuproptosis‐related gene FDX1 suppresses the growth and progression of colorectal cancer by retarding EMT progress. Biochem Genet. 2024. doi: 10.1007/s10528-024-10784-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164. Wu Z, Guo J, Zhang Y, et al. TIGD1 function as a potential cuproptosis regulator following a novel cuproptosis‐related gene risk signature in colorectal cancer. Cancers (Basel). 2023;15:2286. 10.3390/cancers15082286 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165. Xue Q, Kang R, Klionsky DJ, Tang D, Liu J, Chen X. Copper metabolism in cell death and autophagy. Autophagy. 2023;19:2175‐2195. 10.1080/15548627.2023.2200554 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166. Ding X, Xie H, Kang YJ. The significance of copper chelators in clinical and experimental application. J Nutr Biochem. 2011;22:301‐310. 10.1016/j.jnutbio.2010.06.010 [DOI] [PubMed] [Google Scholar]
- 167. Cui H, Zhang AJ, Mckeage MJ, et al. Copper transporter 1 in human colorectal cancer cell lines: effects of endogenous and modified expression on oxaliplatin cytotoxicity. J Inorg Biochem. 2017;177:249‐258. 10.1016/j.jinorgbio.2017.04.022 [DOI] [PubMed] [Google Scholar]
- 168. Gartner EM, Griffith KA, Pan Q, et al. A pilot trial of the anti‐angiogenic copper lowering agent tetrathiomolybdate in combination with irinotecan, 5‐flurouracil, and leucovorin for metastatic colorectal cancer. Invest New Drugs. 2009;27:159‐165. 10.1007/s10637-008-9165-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169. Baldari S, Di Rocco G, Heffern MC, Su TA, Chang CJ, Toietta G. Effects of copper chelation on BRAF(V600E) positive colon carcinoma cells. Cancers (Basel). 2019;11:659. 10.3390/cancers11050659 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170. Möschl P, Lubec G. [5‐fluorouracil inhibits the collagenolytic activity of invasive colonic adenocarcinomas in vitro]. Onkologie. 1978;1:149‐151. 10.1159/000213940 [DOI] [PubMed] [Google Scholar]
- 171. Fatfat M, Merhi RA, Rahal O, et al. Copper chelation selectively kills colon cancer cells through redox cycling and generation of reactive oxygen species. BMC Cancer. 2014;14:527. 10.1186/1471-2407-14-527 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172. Yu N, Zhu H, Yang Y, et al. Combination of Fe/Cu‐chelators and docosahexaenoic acid: an exploration for the treatment of colorectal cancer. Oncotarget. 2017;8:51478‐51491. doi: 10.18632/oncotarget.17807 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173. Yoshiji H, Yoshii J, Kuriyama S, et al. Combination of copper‐chelating agent, trientine, and methotrexate attenuates colorectal carcinoma development and angiogenesis in mice. Oncol Rep. 2005;14:213‐218. [PubMed] [Google Scholar]
- 174. Aggarwal A, Bhatt M. Advances in treatment of wilson disease. Tremor Other Hyperkinet Mov (N Y). 2018;8:525. 10.7916/D841881D [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175. Blanusa M, Varnai VM, Piasek M, Kostial K. Chelators as antidotes of metal toxicity: therapeutic and experimental aspects. Curr Med Chem. 2005;12:2771‐2794. 10.2174/092986705774462987 [DOI] [PubMed] [Google Scholar]
- 176. Shi X, Li Y, Jia M, et al. A novel copper chelator for the suppression of colorectal cancer. Drug Dev Res. 2023;84:312‐325. 10.1002/ddr.22034 [DOI] [PubMed] [Google Scholar]
- 177. O'brien H, Davoodian T, Johnson MDL. The promise of copper ionophores as antimicrobials. Curr Opin Microbiol. 2023;75:102355. 10.1016/j.mib.2023.102355 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178. Oliveri V. Selective targeting of cancer cells by copper ionophores: an overview. Front Mol Biosci. 2022;9:841814. 10.3389/fmolb.2022.841814 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179. Hasinoff BB, Yadav AA, 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. 10.1016/j.jinorgbio.2014.04.004 [DOI] [PubMed] [Google Scholar]
- 180. Feng Yi, Wu J‐J, Sun Z‐L, et al. Targeted apoptosis of myofibroblasts by elesclomol inhibits hypertrophic scar formation. EBioMedicine. 2020;54:102715. 10.1016/j.ebiom.2020.102715 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181. Huang C, Wang M, Wang J, et al. Suppression MGP inhibits tumor proliferation and reverses oxaliplatin resistance in colorectal cancer. Biochem Pharmacol. 2021;189:114390. 10.1016/j.bcp.2020.114390 [DOI] [PubMed] [Google Scholar]
- 182. Martinez‐Balibrea E, Martínez‐Cardús A, Musulén E, et al. Increased levels of copper efflux transporter ATP7B are associated with poor outcome in colorectal cancer patients receiving oxaliplatin‐based chemotherapy. Int J Cancer. 2009;124:2905‐2910. 10.1002/ijc.24273 [DOI] [PubMed] [Google Scholar]
- 183. Zhou Y, Zhang Q, Wang M, Huang C, Yao X. Effective delivery of siRNA‐loaded nanoparticles for overcoming oxaliplatin resistance in colorectal cancer. Front Oncol. 2022;12:827891. 10.3389/fonc.2022.827891 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 184. Johansson B. A review of the pharmacokinetics and pharmacodynamics of disulfiram and its metabolites. Acta Psychiatr Scand Suppl. 1992;369:15‐26. 10.1111/j.1600-0447.1992.tb03310.x [DOI] [PubMed] [Google Scholar]
- 185. Wang Y, Li W, Patel SS, et al. Blocking the formation of radiation‐induced breast cancer stem cells. Oncotarget. 2014;5:3743‐3755. doi: 10.18632/oncotarget.1992 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186. Morrison BW, Doudican NA, Patel KR, Orlow SJ. Disulfiram induces copper‐dependent stimulation of reactive oxygen species and activation of the extrinsic apoptotic pathway in melanoma. Melanoma Res. 2010;20:11‐20. 10.1097/CMR.0b013e328334131d [DOI] [PubMed] [Google Scholar]
- 187. Li Y, Chen F, Chen J, et al. Disulfiram/copper induces antitumor activity against both nasopharyngeal cancer cells and cancer‐associated fibroblasts through ROS/MAPK and ferroptosis pathways. Cancers (Basel). 2020;12:138. 10.3390/cancers12010138 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188. Cen D, Gonzalez RI, Buckmeier JA, Kahlon RS, Tohidian NB. Disulfiram induces apoptosis in human melanoma cells: a redox‐related process. Mol Cancer Ther. 2002;1:197‐204. [PubMed] [Google Scholar]
- 189. Wang N‐N, Wang L‐H, Li Y, et al. Targeting ALDH2 with disulfiram/copper reverses the resistance of cancer cells to microtubule inhibitors. Exp Cell Res. 2018;362:72‐82. 10.1016/j.yexcr.2017.11.004 [DOI] [PubMed] [Google Scholar]
- 190. Yang Z, Guo F, Albers AE, Sehouli J, Kaufmann AM. Disulfiram modulates ROS accumulation and overcomes synergistically cisplatin resistance in breast cancer cell lines. Biomed Pharmacother. 2019;113:108727. 10.1016/j.biopha.2019.108727 [DOI] [PubMed] [Google Scholar]
- 191. Ying H, Qin A, Cheng TS, et al. Disulfiram attenuates osteoclast differentiation in vitro: a potential antiresorptive agent. PLoS One. 2015;10:e0125696. 10.1371/journal.pone.0125696 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192. Chiba T, Suzuki E, Yuki K, et al. Disulfiram eradicates tumor‐initiating hepatocellular carcinoma cells in ROS‐p38 MAPK pathway‐dependent and ‐independent manners. PLoS One. 2014;9:e84807. 10.1371/journal.pone.0084807 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193. Chen Di, Cui QC, Yang H, Dou QP. Disulfiram, a clinically used anti‐alcoholism drug and copper‐binding agent, induces apoptotic cell death in breast cancer cultures and xenografts via inhibition of the proteasome activity. Cancer Res. 2006;66:10425‐10433. 10.1158/0008-5472.CAN-06-2126 [DOI] [PubMed] [Google Scholar]
- 194. Skrott Z, Mistrik M, Andersen KK, et al. Alcohol‐abuse drug disulfiram targets cancer via p97 segregase adaptor NPL4. Nature. 2017;552:194‐199. 10.1038/nature25016 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195. Huang X, Hou Y, Weng X, et al. Diethyldithiocarbamate‐copper complex (CuET) inhibits colorectal cancer progression via miR‐16‐5p and 15b‐5p/ALDH1A3/PKM2 axis‐mediated aerobic glycolysis pathway. Oncogenesis. 2021;10:4. 10.1038/s41389-020-00295-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196. Hu Y, Qian Y, Wei J, et al. The disulfiram/copper complex induces autophagic cell death in colorectal cancer by targeting ULK1. Front Pharmacol. 2021;12:752825. 10.3389/fphar.2021.752825 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197. You S‐Y, Rui W, Chen S‐T, et al. Process of immunogenic cell death caused by disulfiram as the anti‐colorectal cancer candidate. Biochem Biophys Res Commun. 2019;513:891‐897. 10.1016/j.bbrc.2019.03.192 [DOI] [PubMed] [Google Scholar]
- 198. Navrátilová J, Hankeová T, Beneš P, Šmarda J. Acidic pH of tumor microenvironment enhances cytotoxicity of the disulfiram/Cu2+ complex to breast and colon cancer cells. Chemotherapy. 2013;59:112‐120. 10.1159/000353915 [DOI] [PubMed] [Google Scholar]
- 199. Guo X, Xu B, Pandey S, et al. Disulfiram/copper complex inhibiting NFkappaB activity and potentiating cytotoxic effect of gemcitabine on colon and breast cancer cell lines. Cancer Lett. 2010;290:104‐113. 10.1016/j.canlet.2009.09.002 [DOI] [PubMed] [Google Scholar]
- 200. Hendrych M, Říhová K, Adamová B, et al. Disulfiram increases the efficacy of 5‐fluorouracil in organotypic cultures of colorectal carcinoma. Biomed Pharmacother. 2022;153:113465. 10.1016/j.biopha.2022.113465 [DOI] [PubMed] [Google Scholar]
- 201. Zhang P, Zhou C, Ren X, et al. Inhibiting the compensatory elevation of xCT collaborates with disulfiram/copper‐induced GSH consumption for cascade ferroptosis and cuproptosis. Redox Biol. 2024;69:103007. 10.1016/j.redox.2023.103007 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202. Jiapaer Z, Zhang L, Ma W, et al. Disulfiram‐loaded hollow copper sulfide nanoparticles show anti‐tumor effects in preclinical models of colorectal cancer. Biochem Biophys Res Commun. 2022;635:291‐298. 10.1016/j.bbrc.2022.10.027 [DOI] [PubMed] [Google Scholar]
- 203. Najlah M, Said Suliman A, Tolaymat I, et al. Development of injectable PEGylated liposome encapsulating disulfiram for colorectal cancer treatment. Pharmaceutics. 2019;11:610. 10.3390/pharmaceutics11110610 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204. Jiang Y, Huo Z, Qi X, Zuo T, Wu Z. Copper‐induced tumor cell death mechanisms and antitumor theragnostic applications of copper complexes. Nanomedicine (Lond). 2022;17:303‐324. 10.2217/nnm-2021-0374 [DOI] [PubMed] [Google Scholar]
- 205. Tsymbal S, Li Ge, Agadzhanian N, et al. Recent advances in copper‐based organic complexes and nanoparticles for tumor theranostics. Molecules. 2022;27:7066. 10.3390/molecules27207066 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206. Dam J, Ismail Z, Kurebwa T, et al. Synthesis of copper and zinc 2‐(pyridin‐2‐yl)imidazo[1,2‐a]pyridine complexes and their potential anticancer activity. Eur J Med Chem. 2017;126:353‐368. 10.1016/j.ejmech.2016.10.041 [DOI] [PubMed] [Google Scholar]
- 207. Kubiak K, Malinowska K, Langer E, Dziki Ł, Dziki A, Majsterek I. Effect of Cu(II) coordination compounds on the activity of antioxidant enzymes catalase and superoxide dismutase in patients with colorectal cancer. Pol Przegl Chir. 2011;83:155‐160. 10.2478/v10035-011-0024-6 [DOI] [PubMed] [Google Scholar]
- 208. Sandhaus S, Taylor R, Edwards T, et al. A novel copper(II) complex identified as a potent drug against colorectal and breast cancer cells and as a poison inhibitor for human topoisomerase IIalpha. Inorg Chem Commun. 2016;64:45‐49. 10.1016/j.inoche.2015.12.013 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209. Mendo AS, Figueiredo S, Roma‐Rodrigues C, et al. Characterization of antiproliferative potential and biological targets of a copper compound containing 4'‐phenyl terpyridine. J Biol Inorg Chem. 2015;20:935‐948. 10.1007/s00775-015-1277-z [DOI] [PubMed] [Google Scholar]
- 210. Hajrezaie M, Paydar M, Zorofchian Moghadamtousi S, et al. A Schiff base‐derived copper (II) complex is a potent inducer of apoptosis in colon cancer cells by activating the intrinsic pathway. ScientificWorldJournal. 2014;2014:540463. 10.1155/2014/540463 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211. Pellei M, Gandin V, Cimarelli C, et al. Syntheses and biological studies of nitroimidazole conjugated heteroscorpionate ligands and related Cu(I) and Cu(II) complexes. J Inorg Biochem. 2018;187:33‐40. 10.1016/j.jinorgbio.2018.07.008 [DOI] [PubMed] [Google Scholar]
- 212. Lesiów MK, Komarnicka UK, Stokowa‐Sołtys K, et al. Relationship between copper(II) complexes with FomA adhesin fragments of F. nucleatum and colorectal cancer. Coordination pattern and ability to promote ROS production. Dalton Trans. 2018;47:5445‐5458. 10.1039/c7dt04103a [DOI] [PubMed] [Google Scholar]
- 213. Lim H, Oh C, Park M‐S, et al. Hint from an enzymatic reaction: superoxide dismutase models efficiently suppress colorectal cancer cell proliferation. J Am Chem Soc. 2023;145:16058‐16068. 10.1021/jacs.3c04414 [DOI] [PubMed] [Google Scholar]
- 214. Ng CH, Kong SM, Tiong YL, et al. Selective anticancer copper(II)‐mixed ligand complexes: targeting of ROS and proteasomes. Metallomics. 2014;6:892‐906. 10.1039/c3mt00276d [DOI] [PubMed] [Google Scholar]
- 215. Ruiz MC, Perelmulter K, Levín P, et al. Antiproliferative activity of two copper (II) complexes on colorectal cancer cell models: impact on ROS production, apoptosis induction and NF‐kappaB inhibition. Eur J Pharm Sci. 2022;169:106092. 10.1016/j.ejps.2021.106092 [DOI] [PubMed] [Google Scholar]
- 216. Liu Z, Fan L, Niu D, et al. Copper (II) complex of salicylate phenanthroline induces the apoptosis of colorectal cancer cells, including oxaliplatin‑resistant cells. Oncol Rep. 2023;50:170. 10.3892/or.2023.8607 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217. Arikrishnan S, Loh JS, Teo XW, et al. Ternary copper (II) complex induced apoptosis and cell cycle arrest in colorectal cancer cells. Anticancer Agents Med Chem. 2022;22:999‐1011. 10.2174/1871520621666210708100019 [DOI] [PubMed] [Google Scholar]
- 218. Harmse L, Gangat N, Martins‐Furness C, Dam J, De Koning CB. Copper‐imidazo[1,2‐a]pyridines induce intrinsic apoptosis and modulate the expression of mutated p53, haem‐oxygenase‐1 and apoptotic inhibitory proteins in HT‐29 colorectal cancer cells. Apoptosis. 2019;24:623‐643. 10.1007/s10495-019-01547-7 [DOI] [PubMed] [Google Scholar]
- 219. Rashid A, Ananthnag GS, Naik S, Mague JT, Panda D, Balakrishna MS. Dinuclear Cu(I) complexes of pyridyl‐diazadiphosphetidines and aminobis(phosphonite) ligands: synthesis, structural studies and antiproliferative activity towards human cervical, colon carcinoma and breast cancer cells. Dalton Trans. 2014;43:11339‐11351. 10.1039/c4dt00832d [DOI] [PubMed] [Google Scholar]
- 220. Ali A, Mishra S, Kamaal S, et al. Evaluation of catacholase mimicking activity and apoptosis in human colorectal carcinoma cell line by activating mitochondrial pathway of copper(II) complex coupled with 2‐(quinolin‐8‐yloxy)(methyl)benzonitrile and 8‐hydroxyquinoline. Bioorg Chem. 2021;106:104479. 10.1016/j.bioorg.2020.104479 [DOI] [PubMed] [Google Scholar]
- 221. Gandin V, Pellei M, Tisato F, Porchia M, Santini C, Marzano C. A novel copper complex induces paraptosis in colon cancer cells via the activation of ER stress signalling. J Cell Mol Med. 2012;16:142‐151. 10.1111/j.1582-4934.2011.01292.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222. Sequeira D, Baptista PV, Valente R, et al. Cu(I) complexes as new antiproliferative agents against sensitive and doxorubicin resistant colorectal cancer cells: synthesis, characterization, and mechanisms of action. Dalton Trans. 2021;50:1845‐1865. 10.1039/d0dt03566a [DOI] [PubMed] [Google Scholar]
- 223. Shen W, Pei P, Zhang C, et al. A polymeric hydrogel to eliminate programmed death‐ligand 1 for enhanced tumor radio‐immunotherapy. ACS Nano. 2023;17:23998‐24011. 10.1021/acsnano.3c08875 [DOI] [PubMed] [Google Scholar]
- 224. Pei P, Wang Y, Shen W, et al. Oxygen‐driven cuproptosis synergizes with radiotherapy to potentiate tumor immunotherapy. Aggregate. 2024:e484. 10.1002/agt2.484 [DOI] [Google Scholar]
- 225. Chang M, Hou Z, Jin D, et al. Colorectal tumor microenvironment‐activated bio‐decomposable and metabolizable Cu(2) O@CaCO(3) nanocomposites for synergistic oncotherapy. Adv Mater. 2020;32:e2004647. 10.1002/adma.202004647 [DOI] [PubMed] [Google Scholar]
- 226. Al‐Zharani M, Qurtam AA, Daoush WM, et al. Antitumor effect of copper nanoparticles on human breast and colon malignancies. Environ Sci Pollut Res Int. 2021;28:1587‐1595. 10.1007/s11356-020-09843-5 [DOI] [PubMed] [Google Scholar]
- 227. Ghasemi P, Shafiee G, Ziamajidi N, Abbasalipourkabir R. Copper Nanoparticles induce apoptosis and oxidative stress in SW480 human colon cancer cell line. Biol Trace Elem Res. 2023;201:3746‐3754. 10.1007/s12011-022-03458-2 [DOI] [PubMed] [Google Scholar]
- 228. Gnanavel V, Palanichamy V, Roopan SM. Biosynthesis and characterization of copper oxide nanoparticles and its anticancer activity on human colon cancer cell lines (HCT‐116). J Photochem Photobiol B. 2017;171:133‐138. 10.1016/j.jphotobiol.2017.05.001 [DOI] [PubMed] [Google Scholar]
- 229. Khan S, Ansari AA, Khan AA, Abdulla M, Al‐Obaid O, Ahmad R. In vitro evaluation of cytotoxicity, possible alteration of apoptotic regulatory proteins, and antibacterial activity of synthesized copper oxide nanoparticles. Colloids Surf B Biointerfaces. 2017;153:320‐326. 10.1016/j.colsurfb.2017.03.005 [DOI] [PubMed] [Google Scholar]
- 230. Liu Z, Xiong L, Ouyang G, et al. Investigation of copper cysteamine nanoparticles as a new type of radiosensitiers for colorectal carcinoma treatment. Sci Rep. 2017;7:9290. 10.1038/s41598-017-09375-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- 231. Li J, Zhang Z, Li J, et al. Copper‐olsalazine metal‐organic frameworks as a nanocatalyst and epigenetic modulator for efficient inhibition of colorectal cancer growth and metastasis. Acta Biomater. 2022;152:495‐506. 10.1016/j.actbio.2022.08.076 [DOI] [PubMed] [Google Scholar]
- 232. Gholami M, Darroudi M, Hekmat A, Khazaei M. Five‐FU@CuS/NH(2) ‐UiO‐66 as a drug delivery system for 5‐fluorouracil to colorectal cancer cells. J Biochem Mol Toxicol. 2022;36:e23145. 10.1002/jbt.23145 [DOI] [PubMed] [Google Scholar]
- 233. Gholami M, Hekmat A, Khazaei M, Darroudi M. OXA‐CuS@UiO‐66‐NH(2) as a drug delivery system for Oxaliplatin to colorectal cancer cells. J Mater Sci Mater Med. 2022;33:26. 10.1007/s10856-021-06574-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234. Xiong K, Zhou Y, Karges J, et al. Autophagy‐dependent apoptosis induced by apoferritin‐Cu(II) nanoparticles in multidrug‐resistant colon cancer cells. ACS Appl Mater Interfaces. 2021;13:38959‐38968. 10.1021/acsami.1c07223 [DOI] [PubMed] [Google Scholar]
- 235. Anderson CJ, Schwarz SW, Connett JM, et al. Preparation, biodistribution and dosimetry of copper‐64‐labeled anti‐colorectal carcinoma monoclonal antibody fragments 1A3‐F(ab')2. J Nucl Med. 1995;36:850‐858. [PubMed] [Google Scholar]
- 236. Liu D, Overbey D, Watkinson LD, et al. Comparative evaluation of three 64Cu‐labeled E. coli heat‐stable enterotoxin analogues for PET imaging of colorectal cancer. Bioconjug Chem. 2010;21:1171‐1176. 10.1021/bc900513u [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237. Zhou B, Wang H, Liu R, et al. PET imaging of Dll4 expression in glioblastoma and colorectal cancer xenografts using (64)Cu‐labeled monoclonal antibody 61B. Mol Pharm. 2015;12:3527‐3534. 10.1021/acs.molpharmaceut.5b00105 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 238. Paudyal B, Paudyal P, Oriuchi N, Hanaoka H, Tominaga H, Endo K. Positron emission tomography imaging and biodistribution of vascular endothelial growth factor with 64Cu‐labeled bevacizumab in colorectal cancer xenografts. Cancer Sci. 2011;102:117‐121. 10.1111/j.1349-7006.2010.01763.x [DOI] [PubMed] [Google Scholar]
- 239. Philpott GW, Schwarz SW, Anderson CJ, et al. RadioimmunoPET: detection of colorectal carcinoma with positron‐emitting copper‐64‐labeled monoclonal antibody. J Nucl Med. 1995;36:1818‐1824. [PubMed] [Google Scholar]
- 240. Beaino W, Guo Y, Chang AJ, Anderson CJ. Roles of Atox1 and p53 in the trafficking of copper‐64 to tumor cell nuclei: implications for cancer therapy. J Biol Inorg Chem. 2014;19:427‐438. 10.1007/s00775-013-1087-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 241. Delaloye AB, Delaloye B, Buchegger F, et al. Comparison of copper‐67‐ and iodine‐125‐labeled anti‐CEA monoclonal antibody biodistribution in patients with colorectal tumors. J Nucl Med. 1997;38:847‐853. [PubMed] [Google Scholar]
- 242. Nittka S, Krueger MA, Shively JE, et al. Radioimmunoimaging of liver metastases with PET using a 64Cu‐labeled CEA antibody in transgenic mice. PLoS One. 2014;9:e106921. 10.1371/journal.pone.0106921 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 243. Khosravifarsani M, Ait‐Mohand S, Paquette B, Sanche L, Guérin B. Design, synthesis, and cytotoxicity assessment of [(64)Cu]Cu‐NOTA‐terpyridine platinum conjugate: a novel chemoradiotherapeutic agent with flexible linker. Nanomaterials (Basel). 2021;11:2154. 10.3390/nano11092154 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 244. Khosravifarsani M, Ait‐Mohand S, Paquette B, Sanche L, Guérin B. In vivo behavior of [(64)Cu]NOTA‐terpyridine platinum, a novel chemo‐radio‐theranostic agent for imaging, and therapy of colorectal cancer. Front Med (Lausanne). 2022;9:975213. 10.3389/fmed.2022.975213 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 245. Achmad A, Hanaoka H, Yoshioka H, et al. Predicting cetuximab accumulation in KRAS wild‐type and KRAS mutant colorectal cancer using 64Cu‐labeled cetuximab positron emission tomography. Cancer Sci. 2012;103:600‐605. 10.1111/j.1349-7006.2011.02166.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246. Zeng D, Guo Y, White AG, et al. Comparison of conjugation strategies of cross‐bridged macrocyclic chelators with cetuximab for copper‐64 radiolabeling and PET imaging of EGFR in colorectal tumor‐bearing mice. Mol Pharm. 2014;11:3980‐3987. 10.1021/mp500004m [DOI] [PMC free article] [PubMed] [Google Scholar]
