Open access peer-reviewed chapter

The Roles of Iron and Ferroptosis in Human Chronic Diseases

Written By

Yanbo Shi, Junyong Zhang, Kaitao Luo, Sunfeng Pan, Hanqiang Shi, Lie Xiong and Shuqin Du

Submitted: 17 October 2022 Reviewed: 27 October 2022 Published: 25 January 2023

DOI: 10.5772/intechopen.108790

From the Edited Volume

Cell Death and Disease

Edited by Ke Xu

Chapter metrics overview

101 Chapter Downloads

View Full Metrics

Abstract

Ferroptosis, an iron-dependent novel type of cell death, has been characterized as an excessive accumulation of lipid peroxides and reactive oxygen species. A growing number of studies demonstrate that ferroptosis not only plays an important role in the pathogenesis and progression of chronic diseases, but also functions differently in different diseases. As a double-edged sword, activation of ferroptosis could potently inhibit tumor growth and increase sensitivity to chemotherapy and immunotherapy in various cancer settings. Therefore, the development of more efficacious ferroptosis agonists or inhibitors remains the mainstay of ferroptosis-targeting strategy for cancer therapeutics or cardiovascular and cerebrovascular diseases and neurodegenerative diseases therapeutics.

Keywords

  • iron metabolism disorder
  • ferroptosis
  • tumor
  • neurodegeneration
  • vascular diseases

1. Introduction

Chronic diseases are non-infectious, of long duration, and are persistent diseases. These diseases mainly include cardiovascular and cerebrovascular diseases, tumors, diabetes, and chronic respiratory diseases, etc. According to the statistics of WHO, approximately 41 million people worldwide die of chronic diseases, accounting for 73.6% of all deaths in 2021 [1]. The occurrence and development of chronic diseases are not only complicated in pathological mechanisms, but also affected by many factors such as genetics, understanding of the molecular mechanisms underlying the pathogenesis of chronic diseases is helpful for diagnosis and treatment. In 2012, a new way of cell death, ferroptosis, was discovered and reported, to some extent, the discovery is a milestone in the study of cell death; it provides a new perspective to study the occurrence, development, and prevention of chronic diseases.

As an essential trace element, iron is present in nearly all forms of life and involved in various of biological processes, including respiration, oxygen transport, intermediary metabolism, gene regulation, and nucleotide synthesis and repair [2, 3]. However, dysregulated iron homeostasis leads to common hematological, metabolic, and neurodegenerative diseases.

Ferroptosis is an iron-dependent cell death, it is different from apoptosis, pyroptosis, or necrosis in morphology, genetics, metabolism, and molecular biology [4]. The morphology is mainly manifested as mitochondrial swelling, increased membrane density, smaller volume, decreased number of cristae, increased lamellar phenotype, and increased autophagosomes, etc. Molecular biology is mainly manifested as glutathione (glutathione, GSH) depletion or inactivation of glutathione peroxidases (GPX4), increased intracellular free iron content, and increased production of reactive oxygen species (ROS), etc., ultimately manifested as the accumulation of toxic lipid hydroperoxides in cells [4]. Its main characteristics are excessive accumulation of lipid peroxides and reactive oxygen species [5]. Since the production of toxic lipid hydroperoxides mostly depends on ferrous iron, and specific iron chelators can inhibit iron-disturbance-mediated ferroptosis. Therefore, iron metabolism and lipid peroxidation play an important regulating role in ferroptosis pathways [6, 7], the possible molecular mechanism is shown in Figure 1.

Figure 1.

Molecular mechanism of ferroptosis.

Ferric ions in the circulation are bound to transferrin and transported into the cells through transferrin receptor 1 (TFR1), which is located on the cell membrane. After being reduced to divalent iron in the cell, ferric iron is transported by divalent metal transport1 (DMT1) and released into the cytoplasmic iron pool, excess iron is stored in ferritin. Some studies showed that ferritin selective autophagy promotes ferritin getting into autophagosomes through the nuclear receptor co-activator 4 pathway and results in the releasing of free iron [8].

Generally, it is believed that excess iron causes ferroptosis mainly through reactive oxygen species produced by the Fenton reaction, application of iron chelators can effectively inhibit ferroptosis [10]. Lipid peroxides play the role of agents in the process of ferroptosis, phosphatidyl ethanolamine (PE) is the substrate of choice for lipid oxidation. Therefore, hydrogen peroxide-PE (OOH-PE) is considered to be the main signal of ferroptosis [9]. In the process of lipid peroxidation accumulation, NADPH oxidase, lipoxygenase, Acyl-CoA long-chain family member 4(ACSL4), and lysophosphatidyl cholinyltransferase 3 may play important roles in the occurrence and development of ferroptosis [10, 11, 12]. The commonly used ferroptosis inhibitors ferrostatin-1(Fer-1), liproxstatin-1(Lip-1), and small-molecule compounds such as vitamin E inhibit ferroptosis mainly by scavenging lipid peroxides [13].

Glutathione peroxidase 4 (GPX4) is a crucial enzyme in the regulation of ferroptosis; it catalyzes the reduction of lipid peroxides, converts OOH-PE into OH-PE, further suppressing the occurrence of ferroptosis [14]. Small-molecule compounds such as RSL3 and ML162 can inhibit GPX4 activity, lead to the accumulation of fatty acid free radicals, eventually lead to ferroptosis. Reduced glutathione is coenzyme factor of GPX4, the rate-limiting step in its synthesis is the absorption of cystine [15]. Cystine/glutamate transporter consists of the membrane transporters solute carrier family 7 members of 11 (SLC7A11) and regulatory proteins across the membrane of solute carrier family 3 member 2 (SLC3A2); it can transport cystine into the cell and excrete the same amount of glutamate at the same time [16], serves as an important regulator of ferroptosis. Small molecules such as erastin can inhibit glutamic acid/cystine reverse transporter, causing ferroptosis. Wang lab first reported that gene Slc7a11 knockout can promote the occurrence of ferroptosis in mice [17]. Additionally, a novel ferroptosis inhibiting factor 1 (FSP1) was discovered independently in two labs recently, in their studies, NADPH was used to reduce ubiquinone (CoQ10) to ubiquinol (CoQ10H2), leading to the reduction of lipid peroxidation of cell membranes, thereby inhibiting ferroptosis [18, 19]; these findings provide an important basis of developing drugs targeting ferroptosis.

Although the specific mechanisms of ferroptosis are not fully clarified, with the deepening of the research, researchers gradually found that ferroptosis plays an important role in the development of major chronic diseases [20]. So far, a growing evidence showed that ferroptosis is involved in the pathophysiological process of neurodegenerative diseases, tumors, ischemia–reperfusion injury, kidney injury, and other diseases; recent studies have found that ferroptosis plays an important role in cardiovascular disease. Here, we summarized the latest research progress of ferroptosis in cancer, neurodegenerative diseases, and cerebrovascular diseases, to provide new ideas and strategies for the prevention and treatment of major chronic diseases.

Advertisement

2. Ferroptosis and tumor

Tumor cells can proliferate by avoiding cell death, and apoptosis, necrosis, autophagy also played important roles in the development of tumor. In Table 1, we summarized the latest research progress of molecular mechanism studies on ferroptosis in common tumors, to provide a series of potential new targets for tumor prevention and control.

Tumor typeRelated studiesReference
Hepatocellular carcinomaSorafenib induces ferroptosis in hepatocellular carcinoma cells[21]
Retinoblastoma protein-deficient hepatocellular carcinoma cells are more sensitive to sorafenib-induced ferroptosis[22]
Activation of p62-Keap1-NRF2 leads to ferroptosis resistance in hepatocellular carcinoma cells[23, 24, 25, 26]
The expression levels of SLC7A11, Rb, and MT1 are related to the prognosis of patients with hepatocellular carcinoma[22, 27, 28]
Pancreatic cancerArtesunate induces ferroptosis in pancreatic ductal adenocarcinoma cells[29, 30]
Piperamide induces ferroptosis in pancreatic ductal adenocarcinoma cells[31]
Combination of piperamide, Cotylenin A, and sulfasalazine effectively inhibits pancreatic cancer via ferroptosis[31]
Renal cell carcinomaCompared with other tumor cells, renal clear cell carcinoma cells are more sensitive to ferroptosis induced by glutathione peroxidase 4 inhibition[32]
HIF-2α-HILPDA pathway regulates the sensitivity of renal clear cell carcinoma cells to ferroptosis[33]
TAZ/EMP1/NOX4 pathway regulates the sensitivity of renal clear cell carcinoma cells to ferroptosis[34, 35]
Breast cancerSiramesine combined with apatinib upregulates iron levels and induces ferroptosis in breast cancer cells[36]
Mucin 1C subunit, SLC7A11, and CD44v form a complex to upregulate reduced glutathione expression and make triple-negative breast cancer cells resistant to ferroptosis[37]
Sulfasalazine inhibits the growth of glutamine auxotrophic triple-negative breast cancer cells[38]
SLC7A11 is closely related to drug resistance and metastasis of triple-negative breast cancer cells[39]
Transferrin receptor expression level is associated with breast cancer prognosis[40, 41]
Bladder tumorFerritin phagocytosis releases intracellular free iron and induces ferroptosis and inhibits bladder tumors.[44]

Table 1.

Ferroptosis in cancer.

SLC7A11: solute carrier family 7 member 11; Rb: retinoblastoma; MT1: metallothionein 1.

2.1 Ferroptosis and hepatocellular carcinoma (HCC)

Targeting ferroptosis is a potential mechanism in the treatment of hepatocellular carcinoma (HCC) [21]. Retinoblastoma (Rb)-deficient cancer cells are more sensitive to ferroptosis induced by sorafenib [22], it is likely that it enhances the oxidative stress response by affecting the concentration of reactive oxygen species in mitochondria. Furthermore, Sun et al. found that nuclear factor E2–related factor 2 (NRF2) protects hepatocellular cancer cells from ferroptosis induced with sorafenib, indicating that targeting p62-Keapl-NRF2 pathway may overcome sorafenib resistance in hepatocellular carcinoma cells [23, 24, 25, 26]. Additionally, NRF2 also induces expression of metallothionein1G (MT-1G), an important negative regulator of ferroptosis, through the cythionase pathway, leading to sorafenib resistance in cancer cells [27]. CDGSH [Fe-S]-containing domain 1 can protect the mitochondria from ferroptosis in hepatocellular cancer cells and can be upregulated by erastin in an iron-dependent manner [28]. Moreover, p53S47 mutation carried by hepatocellular carcinomas causes ferroptosis tolerance by inhibiting ASCL4 [28].

In hepatocellular carcinoma, the expression of ferroptosis-related genes is related to the prognosis of patients, the mRNA expression levels of SLC7A11 in HCC tissues and adjacent normal tissues were compared between 130 cases of hepatocellular carcinoma (HCC) tissues and adjacent normal tissues by Kinoshita et al. [29], it showed that the expression of SLC7A11 in HCC tissues was significantly higher than that in normal tissues, the survival time and disease-free survival time of liver cancer patients were significantly shorter than those of SLC7A11 low expression of hepatocellular carcinoma patients. In addition, in the process of sorafenib treatment of hepatocellular carcinoma patients, the high expression of Rb and MT1 is also associated with the poor prognosis of patients [22, 27].

2.2 Ferroptosis and pancreatic cancer

The main mechanism of pancreatic cancer is that mutated KRAS gene reprograms pancreatic ductal adenocarcinoma (PDAC) cells to a state that is highly resistant to apoptosis. Artesunate can induce tumor cell apoptosis by generating reactive oxygen species [30]. Eling et al. found that artesunate can induce ferroptosis in PDAC cells with KRAS mutations and the process can be effectively inhibited by Fer-1 [31]. Yamaguchi et al. found that the natural product piperamide can induce ferroptosis in tumor cells by promoting the generation of reactive oxygen species, and its anti-tumor effect can be inhibited by antioxidants, ferroptosis inhibitors, and iron chelators [32]. The combined use of piperamide, Cotylenin A (a plant growth regulator), and sulfasalazine has a good synergistic effect on pancreatic cancer. These results suggest that ferroptosis inducers are expected to be used in the treatment of pancreatic cancer.

2.3 Ferroptosis and renal cell carcinoma

Renal cell carcinoma originates from the renal parenchyma urothelial system and is a highly malignant tumor in the urinary system. Yang et al. [33] found that GPX4 is a key regulator of the ferroptosis signaling pathway in clear cell renal cell carcinoma. When compared with the other tumor cell from other tissues (lung cancer, colon cancer, central nervous system, melanoma, ovarian cancer, breast cancer, and leukemia), renal cells were more sensitive to ferroptosis induced with inhibition of GPX4. Renal cancer cells are induced by the hepatocyte factor -1β-1-acylglycerol-3 phosphate oxyacyltransferase 3 (AGPAT3) axis and the HIF-2α-HILPDA pathway, which can induce polyunsaturated fatty acyl lipid–enriched cells state, thereby increasing its susceptibility to ferroptosis [34]. Recently, Yang et al. [35] found that the sensitivity of renal cancer cells to ferroptosis is regulated by cell density and transcriptional regulator 1 (TAZ)-TAZ regulation of epidermal membrane protein 1 (EMP1)/NOX4 pathway [34, 35], suggesting that TAZ is a potential therapeutic target for ferroptosis.

2.4 Ferroptosis and breast cancer

Breast cancer is derived from breast epithelial tissue. Ma et al. [36] found that the lysosomal interfering agent siramesine and the tyrosine kinase inhibitor lapatinib can disrupt the iron homeostasis in breast cancer cells to generate reactive oxygen species and induce cell ferroptosis, and overexpression of trasnferrin receptor 1 (TfR1) or iron chelators can reduce siramesine and lapatinib-induced reactive oxygen species.

Some studies have shown that the formation of a complex between mucin 1C subunit and SLC7A11 can upregulate the expression of reduced glutathione and inhibit ferroptosis in triple-negative breast cancer cells [37]. Timmerman et al. [38] found a subpopulation of glutamine auxotrophic triple-negative breast cancer cells that were highly dependent on SLC7A11 acquires cystine for glutamine metabolism. The SLC7A11 inhibitor sulfasalazine can inhibit tumor growth by promoting ferroptosis. In addition, the activities of SLC7A11 and glutamate/cystine antiporter can be regulated by the Keap1/NRF2 redox pathway. Lanzardo et al. [39] believed that SLC7A11 is closely related to drug resistance and metastasis of triple-negative breast cancer cells. The increased expression of TFR1 in breast cancer cells is negatively correlated with the expression of estrogen receptor, and the high expression of TFR1 in breast cancer tissue is associated with poor prognosis of patients [40, 41].

2.5 Ferroptosis and bladder tumors

Intracellular iron concentration is closely related to the progression of bladder tumors. Martin-Sanchez et al. [42] analyzed the relationship between intracellular iron concentration and bladder tumor proliferation and found that when transferrin combined with iron, the free iron decreased in tumor cells, which was conducive to the proliferation of bladder cancer cells. When the application of gallium (Ga) to transferrin interferes with the binding of iron to transferrin, the intracellular free iron increases and thus inhibits the proliferation of bladder tumor cells. Mazdak et al. [43] found that the serum iron level of patients with bladder tumors was significantly lower than that of the normal control group, suggesting that the decreased serum iron level may be an important reason for the occurrence of bladder tumors. Tang et al. [44] proposed the phenomenon of ferritin phagocytosis, which releases intracellular free iron through ferritin and increases the content of intracellular free iron, which may play a role in inhibiting bladder tumors. The results suggest that activating ferroptosis can achieve ideal therapeutic effect on bladder tumors. These studies suggest that increasing intracellular iron concentration may help to inhibit bladder tumor progression.

2.6 Tumor-associated ferroptosis regulatory protein

2.6.1 SLC7A11 regulates ferroptosis

SLC7A11(also known as xCT) is the substrate-specific subunit of System Xc responsible for the transport of cystine from the extracellular to the intracellular. The nuclear factor erythroid-like 2 (Nrf2) and transcription factor 4 (activating factor 4, ATF4) can induce SLC7A11 expression when cells are in a state of oxidative stress and L-cysteine deficiency [45]. Studies have found that SLC7A11 is highly expressed in tumor tissues, and that high expression of SLC7A11 can inhibit ROS-induced ferroptosis. p53 leads to cystine deficiency by inhibiting the expression of SLC7A11, which in turn increases the sensitivity to ferroptosis. Importantly, the survival and proliferation of SLC7A11-overexpressing cancer cells are dependent on glucose, such tumors may be sensitive to glucose-blocking drugs, also suggesting a role for SLC7A11 in modulating nutrient dependence and demonstrating another therapeutic strategy for tumors with high SLC7A11 expression [46].

2.6.2 p53 regulates ferroptosis

p53 is a widely recognized tumor suppressor that can induce senescence and programmed cell death in human cancer. It can affect ferroptosis of tumor cells through transcriptional or posttranslational mechanisms. Also, it can inhibit tumors by regulating cell cycle arrest, apoptosis, or premature aging.

Increasing the stability of wild-type p53 can promote the expression of its transcriptional target gene CDKN1A (encoding p21 protein), which can increase the intracellular glutathione level, inhibit the accumulation of ROS, and negatively regulate ferroptosis of cancer cells. Acetylation-deficient p53 increased the sensitivity of tumor cells to ferroptosis by inhibiting SLC7A11 expression and System Xc function.

Jiang et al. [47] found that three lysines in the DNA-binding domain of p53 were mutated to arginine (K117/161/162R, namely p533KR), p533KR can further restrict cystine uptake by inhibiting SLC7A11 gene transcription, make tumor cells more sensitive to oxidative stress-induced ferroptosis. Wang et al. [48] found that the site K98 in the DNA-binding domain of p53 is particularly important for the regulation of SLC7A11. Additionally, the mutant S47 of p53 fails to inhibit the transcription of SLC7A11 and induces ferroptosis resistance in hepatocellular carcinoma cells, increasing the risk of cancer in mice [49]. Moreover, p53 can make colon cancer cells insensitive to ferroptosis by inhibiting dipeptidyl peptidase 4 activity [50].

2.6.3 NRF2 regulates ferroptosis

NRF2 is an important transcriptional regulator in oxidative reactions [51], and its overexpression can inhibit apoptosis and lead to drug resistance in some tumors [52]. NRF2 plays an important role in protecting hepatocellular carcinoma cells from ferroptosis [22]. After treatment of hepatocellular carcinoma cells with Erastin and Sorafenib, p62 inhibits the degradation of NRF2 and induces NRF2 accumulation in the nucleus through the inactivation of Keap1, thereby regulating downstream gene transcription. Inhibition of NRF2 by the alkaloid trigonelline can induce ferroptosis in hepatocellular carcinoma cells, and combined use with chemotherapeutic drugs has the application prospect of overcoming tumor drug resistance [22]. Thus, activation of the p62-Keap1-NRF2 pathway can activate ferroptosis to reverse tumor chemotherapeutic drug resistance.

2.6.4 ACSL4 regulates ferroptosis

ACSL4 is expressed on the mitochondrial outer membrane and endoplasmic reticulum and can convert long-chain fatty acids into fatty acyl-CoA, which plays an important role in lipid biosynthesis and fatty acid degradation. ACSL4 increases the sensitivity of cells to ferroptosis by accumulating long-chain polyunsaturated ω-6 fatty acids in the cell membrane [10]. Studies have shown that in basal-like breast cancer cell lines, liver cancer cells, leukemia cells, and prostate cancer cells, the expression level of ACSL4 can be used to predict the sensitivity of tumor cells to ferroptosis [10, 28, 52]. The results suggest that ACSL4 is expected to be a potential target and biological marker for targeting ferroptosis in tumor therapy.

2.6.5 GPX4

Glutathione peroxidase 4 (GPX4) is the only glutathione peroxidase that can use glutathione as the electron donor to reduce the toxic lipid hydroperoxides in biofilms to corresponding alcohols. Tumor cells with high GPX4 show impaired proliferation, decreased proliferation, and inhibition of angiogenesis. Overexpression of GPX4 in hepatocellular carcinoma cells can inhibit the formation and development of hepatocellular carcinoma by decreasing ROS level, increasing glutathione and decreasing the formation of the cytokine-cytokine IL-8, inhibiting cell cycle progression and cell migration [53]. Based on the clinicopathological study and in vitro cell death analysis, it was found that overexposure to GPX4 in diffuse large B-cell lymphoma (DLBCL) inhibited ROS-induced ferroptosis [54]. GPX4 is a major target molecule for ferroptosis inducers such as erastin and RSL3. Erastin inhibits GPX4 activity by depleting glutathione, whereas RSL3 can directly inhibit GPX4 activity. In addition, previous studies have demonstrated that GPX4 can induce ferroptosis in mouse tumor xenograft models [33].

2.6.6 FSP1 inhibits ferroptosis

FSP1 inhibits ferroptosis FSP1 was originally named mitochondrial apoptosis-inducing factor 2 (AIFM2), as the newly discovered GPX4-independent ferroptosis inhibitor, and its expression is closely related to the sensitivity of tumor cells to ferroptosis. Recently, Doll’s group and Bersuker’s group simultaneously screened independently and found that FSP1 levels are different in different cell lines, and the resistance level of various tumor cell lines to ferroptosis was positively correlated with the FSP1 level, which results in differences in the sensitivity of different tumor cell lines to ferroptosis [18, 19]. Additionally, it was also reported that FSP1 was a novel KEAP1/NRF2 target gene regulating ferroptosis and radioresistance in lung cancers [55]. These achievements provide important evidences for the development of drug-targeted ferroptosis in tumors.

Advertisement

3. Ferroptosis and neurodegenerative diseases

Iron homeostasis is critical for brain and neural development and cognitive function, especially in the fetal or early neonatal period, iron deficiency can severely affect neurodevelopment, leading to impaired memory and learning [56]. Iron accumulates gradually in the brain with age, and accumulation studies have shown that iron accumulation is related to neurodegenerative diseases such as Alzheimer’s disease, Parkinson’s disease, and amyotrophic lateral sclerosis [57]. In recent years, studies have found that there are main characteristics of ferroptosis such as increased lipid peroxidation, decreased glutathione, and GPX4 inhibition in neurodegenerative diseases and cognitive impairment. The use of ferroptosis inhibitors can effectively protect neurons and improve cognitive function. Related research progress is shown in Table 2.

DiseasesRelated studiesReferences
Alzheimer’s diseaseOverexpression or increase of phosphorylated tau protein can induce neuronal ferroptosis, and α-lipoic acid can inhibit tau protein-induced ferroptosis[63]
Hippocampal neuronal death and cognitive decline in brain Gpx4-induced knockout mice[68]
Ferritin levels in cerebrospinal fluid can predict the progression of Alzheimer’s disease[73]
Parkinson’s diseaseActivation of protein kinase C triggers ferroptosis[79]
Serine/threonine protein kinase is involved in Erastin-induced ferroptosis[85]
Astrocytes provide neurons with GSTM2 to protect neurons from oxidative damage[81, 82, 83]
Amyotrophic lateral sclerosisNeuronal Gpx4-inducible knockout mice develop symptoms of amyotrophic lateral sclerosis[68]

Table 2.

Ferroptosis in neurodegenerative diseases.

Gpx4: glutathione peroxidase 4; GSTM2:Glutathione S-transferase Mu2.

3.1 Ferroptosis and Alzheimer’s disease

Patients with Alzheimer’s disease possess destabilization of metal metabolism, inflammatory response, oxidative stress, abnormal mitochondrial function, and impaired glial function [57, 58]. Studies have shown that iron accumulation in the brain is associated with the formation of senile plaques and neurofibrillary tangles, elevated iron levels in the brain increase the risk of Alzheimer’s disease, and ferritin levels in cerebrospinal fluid predict the progression from mild cognitive impairment to Alzheimer’s disease [59, 60, 61]. The chronic inflammation, neuronal degeneration, and lack of downstream apoptosis indicators associated with Alzheimer’s disease suggest the existence of other cell death manners such as ferroptosis in Alzheimer’s disease [62, 63, 64].

An investigation on the Gpx4-specific knockout mice in cerebral cortex and hippocampal neurons exhibited cognitive decline and degeneration of hippocampal neurons in the water maze test, after feeding with a vitamin-E-rich diet or Lip-1, the neuronal degeneration of the mice was significantly alleviated, suggesting that ferroptosis plays an important role in neuronal degeneration [62]. Another study found that overexpression or hyperphosphorylation of tau protein can induce ferroptosis in neurons, while α-lipoic acid can rescue neurons by downregulating TfR1, reducing p38 phosphorylation level, and upregulating the expression of Slc7a11 and Gpx4 [63]. In addition, feeding with a deuterated polyunsaturated fatty acid in a mouse model of Alzheimer’s disease can alleviate the lipid peroxidation of tissues and reduces β-amyloid deposition [64, 65].

3.2 Ferroptosis and Parkinson’s disease

An important feature of Parkinson’s disease is iron accumulation in neurons and substantia nigra glia, and the concentration of iron accumulation is positively correlated with disease severity [66, 67]. Significant changes in iron regulatory protein 1(IRP1), divalent metal transporter 1(DMT1), and other key proteins involved in iron homeostasis have been observed in Parkinson’s disease patients and mouse models [68, 69, 70, 71, 72]. The τ knockout mice developed parkinsonism with iron accumulation in the nigra, which can be inhibited by iron chelators [73, 74, 75]. In addition to elevated iron levels in the substantia nigra pars compactus, parkinsonism is also characterized by ferroptosis, such as reduced glutathione depletion and lipid peroxidation [76], iron chelators and N-acetylcystine can alleviate and improve some of the symptoms in patients and mouse model of Parkinson’s disease [77, 78], suggesting that ferroptosis may be involved in the occurrence and development of Parkinson’s disease.

Do Van et al. [79] found that dopaminergic neurons’ ferroptosis occurred in LUHMES cell lines, brain tissue slices cultured in vitro, and 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-induced Parkinson’s disease models. However, the use of Fer-1, Lip-1, and iron chelators can alleviate or reverse the symptoms of Parkinson’s disease. Gouel et al. [80] found that human platelet lysate can make LUHMES cells resistant to Earstin-induced neuronal ferroptosis. In addition, astrocytes have a strong iron storage capacity, which prevent iron overload in neurons [81]. Astrocytes provide neurons with glutathione S-transferase Mu2 (GSTM2) and other antioxidant factors to protect neurons from oxidative damage. In conclusion, the dysregulation of the interaction between astrocytes and neurons may lead to ferroptosis in dopaminergic neurons [82, 83].

3.3 Ferroptosis and amyotrophic lateral sclerosis

Iron accumulation occurred in the brain of amyotrophic lateral sclerosis model mice [84, 85, 86], and the therapeutic effect of iron chelators confirms the role of iron in the pathogenesis of amyotrophic lateral sclerosis. Patients with amyotrophic lateral sclerosis have increased lipid peroxidation in cerebrospinal fluid and plasma and decreased levels of reduced glutathione in the motor cortex, suggesting the possibility of ferroptosis occurrence [87, 88]. Knockout of Gpx4 in mouse neurons can cause symptoms of amyotrophic lateral sclerosis, mainly characterized by rapid paralysis, severe muscle atrophy, and death, which is related to ferroptosis of spinal motor neurons [62]. However, no significant neurodegeneration was observed in the cortex of neuronal Gpx4-inducible knockout and other Gpx4-selective cortical neuron knockout mouse models. It is suggested that Gpx4 plays an important role in the process of ferroptosis of spinal motor neurons [62].

Advertisement

4. Ferroptosis and cardiovascular disease

Cardiomyocyte and neuron death in cardiovascular and cerebrovascular diseases are related to a variety of cell death manners, and ferroptosis is also involved. The relevant research progress is shown in Table 3.

DiseasesRelated studiesReferences
ischemia–reperfusionIn an isolated mouse cardiac ischemia–reperfusion model, inhibition of glutamine metabolism can attenuate ferroptosis-induced cardiac injury[93]
ferroptosis inhibitors and iron chelators can effectively alleviate myocardial injury induced by cardiac ischemia–reperfusion in mice[89]
Doxorubicin-induced myocardial injuryDoxorubicin induces ferroptosis in cardiomyocytes via heme oxygenase; iron accumulation and lipid peroxidation mainly occur in mitochondria[89]
Heme oxygenase inhibitors, ferroptosis inhibitors, mitochondrial antioxidant inhibitors, iron chelators, etc., can effectively reverse doxorubicin-induced myocardial injury[89]
Myocardial damage after heart transplantationFerroptosis regulates neutrophil recruitment after cardiac transplantation in mice[105]
Ischemic strokeHypoxia-inducible factor prolyl hydroxylase may be a potential target of iron chelators to inhibit neuronal ferroptosis[115]
Inhibition of ferroptosis protects neurons in mice with middle cerebral artery occlusion, and the interaction between iron and tau protein is pleiotropically regulated[105]
Hemorrhagic stroke(−)-Epicatechin alleviates early brain injury in hemorrhagic stroke by reducing cerebral iron accumulation and ferroptosis-related protein expression[116]
Ferroptosis inhibitors attenuate neuronal death in brain slices and in a mouse model of hemorrhagic stroke[117, 118]
Increased glutathione peroxidase 4 expression can avoid neuronal ferroptosis and improve prognosis[119]

Table 3.

Ferroptosis in cardiovascular and cerebrovascular diseases.

4.1 Ferroptosis and cardiovascular disease

In some pathological conditions, the heart exhibits excessive accumulation of iron, production of reactive oxygen species, and pathological transformation of membrane lipids, which are all important factors that constitute ferroptosis. So far, there are few studies directly linking ferroptosis with cardiovascular disease. The latest research results of the Wang group in 2019 revealed the important role of ferroptosis in cardiomyopathy and ischemia–reperfusion-induced cardiac injury for the first time [89]. This landmark discovery provides a new strategy for the prevention and treatment of cardiomyopathy and other heart diseases.

4.1.1 Ferroptosis is involved in tissue and organ induced by ischemia: reperfusion damage

During cardiac ischemia–reperfusion, excess reactive oxygen species, lipid peroxidation, and iron accumulation caused by the release of iron in heme will be produced [90, 91, 92]. Gao et al. [93] established an isolated mouse cardiac ischemia–reperfusion model and found that inhibiting glutamine metabolism could inhibit ferroptosis, thereby reducing cardiac injury. Fang et al. [89] established an in vivo myocardial ischemia–reperfusion model and found that Fer-1 and iron chelators can significantly reduce the acute and chronic cardiac injury of ischemia–reperfusion, confirming the role of ferroptosis in cardiac ischemia–reperfusion injury. In addition, ferroptosis is also involved in ischemia–reperfusion injury in the kidney [94] and liver [95].

4.1.2 Ferroptosis is involved in antitumor drug-induced myocardial injury

As a broad-spectrum antitumor drug, adriamycin was limited in clinical use due to its cardiotoxicity. Autophagy, apoptosis, necrosis, and other cell death type are involved in the myocardial injury caused by adriamycin [96, 97, 98]. Fang et al. [89] found that ferroptosis occurred in cardiomyopathy induced by doxorubicin in mice deficient in apoptosis and proposed that heme oxygenase 1 (HO-1) may be a key regulator in this procedure. They also found that iron accumulation and lipid peroxidation in cardiomyocytes occur in mitochondria and the mitochondria-targeting antioxidant MitoTEMPO can effectively inhibit ferroptosis and protect the heart.

4.1.3 Ferroptosis is involved in myocardial injury after heart transplantation

In Li’s studies [99], it was found that the recruitment of neutrophils after heart transplantation is regulated by ferroptosis. The donor heart can induce ferroptosis in cardiomyocytes due to ischemia, hypoxia, and other reasons after transplantation, and the cellular contents are released and recruit neutrophils to produce necrotic inflachannelled by TLR4 / Trif/type I interflammatory via TLR/Trif/type I interferon pathway. Fer-1 can reduce arachidyl phosphatidylethanolamine after heart transplantation and decreased cardiomyocyte death and neutrophil recruitment.

4.1.4 Ferroptosis and diabetic cardiomyopathy

In diabetes, persistent high blood glucose and insulin resistance can cause a vicious circle by altering cellular metabolism, promoting the accumulation of peroxidation and the death of cells. So far, diabetes has been verified to be associated with abnormal iron metabolism. For example, systemic iron overload can contribute to abnormal glucose metabolism and the onset of type 2 diabetes (T2DM) [100] and aggravate insulin resistance [101]. Recently, Cai group [102] identified the role of ferroptosis in DCM and reported that Nrf2 activation by sulforaphane inhibited ferroptosis and prevented DCM, suggesting that it is feasible to treat DCM by inhibiting ferroptosis. Due to the limited regenerative capacity of the myocardium in mammalian adult hearts, inhibition of cardiomyocyte death might be one of the important ways to alleviate DCM [103]. In our studies, we induced DCM models in diabetic C57BL6 mice and treated with canagliflozin and found that canagliflozin mitigates ferroptosis and improves myocardial oxidative stress in mice with diabetic cardiomyopathy [104]. Taken together, taking ferroptosis as the starting point may provide a new strategy for the prevention and control of DCM.

4.2 Ferroptosis and cerebrovascular disease

Both ischemic stroke and hemorrhagic stroke can lead to neuronal ferroptosis [105, 106].

4.2.1 Ferroptosis and ischemic stroke

Before the discovery of ferroptosis, iron accumulation in clinical and animal models of ischemic stroke has been shown to aggravate neuronal damage during reperfusion [107, 108, 109, 110]. Iron chelators can reduce the risk of post-ischemic stroke in experimental animals [111, 112, 113, 114]. Speer et al. [115] proposed that ferroptosis leads to neuronal death after cerebral ischemia, and hypoxia-inducible factor prolyl hydroxylase may be the target for the beneficial effects of iron chelators. Inhibition of ferroptosis in a mouse model can protect neurons from ischemia–reperfusion injury [105].

4.2.2 Ferroptosis and hemorrhagic stroke

Chang et al. [116] found that epicatechin could alleviate early brain injury in hemorrhagic stroke by reducing cerebral iron accumulation and ferroptosis-related protein expression. Later, they found that Fer-1 could alleviate hemoglobin-induced brain injury. Cell death in slices and alleviation of neuronal death in a mouse model of collagenase-induced hemorrhagic stroke [117]. At the same time, Zille et al. [118] found that ferroptosis inhibitors such as Fer-1 and deferoxamine can inhibit the production of ferroptosis in mice. The expression level of Gpx4 in rats with acute hemorrhagic stroke decreased sharply, and increasing the level of Gpx4 could avoid secondary ferroptosis injury in neurons and improve the prognosis of hemorrhagic stroke [119]. Therefore, the ferroptosis pathway may be involved in the process of neuronal death in stroke, and it is speculated that targeted inhibition of ferroptosis may be an effective treatment for alleviating stroke.

Advertisement

5. Conclusion

Besides tumors, neurodegenerative diseases, cardiovascular and cerebrovascular diseases, ferroptosis has also been reported in liver diseases such as non-alcoholic fatty liver disease and non-alcoholic steatohepatitis [120, 121, 122, 123, 124].

ROS-induced ferroptosis can inhibit tumor growth and increase the sensitivity of tumor cells to chemotherapy and radiotherapy. Contrary to tumor treatment strategies, ferroptosis can promote the occurrence and development of neurodegenerative diseases and cardiovascular and cerebrovascular diseases. Therefore, relevant translational medicine research mainly focuses on the discovery of small molecules that can effectively inhibit ferroptosis. These small-molecule activators targeting ferroptosis can be used directly as chemotherapeutics, or as chemosensitizers in combination with chemotherapeutics. However, ferroptosis is complex in different types of tumors and different gene mutations (such as p53 or RAS mutations), and its feasibility in preclinical and clinical research needs to be further studied. Notably, the discovery of GPX4 pathway-independent FSP1 and the discovery of new mechanisms and targets such as CD8+ T cells inducing ferroptosis in tumor cells through the release of interferon-gamma [18, 19, 125], it provides a new perspectives and strategy for tumor treatment and drug discovery.

Iron accumulation and ferroptosis in the brain and nerve tissue have been proved to be closely related to Alzheimer’s disease and Parkinson’s disease. There is a direct relationship between the pathogenesis of various neurodegenerative diseases such as Parkinson’s disease and amyotrophic lateral sclerosis. At present, various clinical trials using iron chelators to treat neurodegenerative diseases are emerging, but there is still no effective treatment for stroke. Given the important role of ferroptosis in neuronal death after stroke, effective inhibition of ferroptosis is expected to provide a new strategy for preventing neuronal death caused by stroke.

Similar to the pathogenesis of neurodegenerative diseases, many cardiac diseases share common ferroptosis features, such as iron overload, oxidative stress, endoplasmic reticulum stress, and mitochondrial dysfunction. Previous studies by the author’s team suggest that ferroptosis inhibitors can effectively prevent and treat cardiomyopathy and heart failure induced by myocardial cell iron overload, doxorubicin-induced cardiotoxicity, and cardiac ischemia–reperfusion [89]. Five different approaches, including ferroptosis inhibitors, iron chelators, mitochondria-specific antioxidants, heme oxygenase 1 inhibitors, and low-iron diets, can effectively prevent ferroptosis in cardiomyocytes, thereby protecting the heart. And these ferroptosis inhibitors are relatively safe and feasible in mice. It provides an optimistic prospect for clinical translational research on targeting ferroptosis to prevent and treat heart disease [121, 126, 127].

With a view to clinical translation, here are some issues need to be considered, e.g., which disease or tumor needs to be considered for ferroptosis-targeted therapies? In clinical or preclinical experiments, drugs targeting ferroptosis need high tissue-organ specificity and fewer adverse reactions, and nano-targeted drug delivery systems have shown some advantages [128, 129]. Although there is a growing awareness of ferroptosis, some key scientific questions related to ferroptosis still need to be resolved, such as what are the key executive molecules in ferroptosis? To what extent is lipid peroxidation related to ferroptosis? Does ferroptosis exist in physiological processes? Is ferroptosis conservative in the evolutionary process? We are well aware of the long road ahead. With the deepening and expansion of ferroptosis-related research, we believe that it will provide a basis for the clinical translation of targeting ferroptosis to prevent and treat major chronic diseases.

Advertisement

Acknowledgments

This work was supported by grants from Medicine and Health Science and Technology Plan Projects of Zhejiang Province (YS, 2020PY029), Science and Technology Innovation Special Project of Jiaxing Science and Technology Bureau (YS, 2020AY30003), Zhejiang Provincial Health Science and Technology Program of Traditional Chinese Medicine (YS, 2021ZB283), Zhejiang Basic Public Welfare Research Program (YS, LGF18H200004), and Jiaxing Key Laboratory of Diabetic Angiopathy.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. World Health Statistics. Monitoring Health for the SDGs, Sustainable Development Goals. Genève: World Health Organization; 2021
  2. 2. Hentze MW, Muckenthaler MU, Andrews NC. Balancing acts: Molecular control of mammalian iron metabolism. Cell. 2004;117:285-297
  3. 3. Rouault TA. Cell biology. An ancient gauge for iron. Science. 2009;326(5953):676-677
  4. 4. Dixon SJ, Lemberg KM, Lamprecht MR, Skouta R, Zaitsev EM, Gleason CE, et al. Ferroptosis: An iron-dependent form of nonapoptotic cell death. Cell. 2012;149(5):1060-1072
  5. 5. Cao JY, Dixon SJ. Mechanisms of ferroptosis. Cellular and Molecular Life Sciences. 2016;73(11-12):2195-2209
  6. 6. Dixon SJ, Stockwell BR. The role of iron and reactive oxygen species in cell death. Nature Chemical Biology. 2014;10(1):9-17
  7. 7. Kagan VE, Mao G, Qu F, et al. Oxidized arachidonic and adrenic PEs navigate cells to ferroptosis. Nature Chemical Biology. 2017;13(1):81-90
  8. 8. Mancias JD, Wang X, Gygi SP, Harper JW, Kimmelman AC. Quantitative proteomics identifies NCOA4 as the cargo receptor mediating ferritinophagy. Nature. 2014;509(7498):105-1090
  9. 9. Wenzel SE, Tyurina YY, Zhao J, St Croix CM, Dar HH, Mao G, et al. PEBP1 wardens ferroptosis by enabling lipoxygenase generation of lipid death signals. Cell. 2017;171(3):628-641
  10. 10. Doll S, Proneth B, Tyurina YY, Panzilius E, Kobayashi S, Ingold I, et al. ACSL4 dictates ferroptosis sensitivity by shaping cellular lipid composition. Nature Chemical Biology. 2017;13(1):91-98
  11. 11. Hangauer MJ, Viswanathan VS, Ryan MJ, Bole D, Eaton JK, Matov A, et al. Drug-tolerant persister cancer cells are vulnerable to GPX4 inhibition. Nature. 2017;9(7679):247-250
  12. 12. Dixon SJ, Winter GE, Musavi LS, Lee ED, Snijder B, Rebsamen M, et al. Human haploid cell genetics reveals roles for lipid metabolism genes in nonapoptotic cell death. ACS Chemical Biology. 2015;10(7):1604-1609
  13. 13. Gao M, Monian P, Pan Q , Zhang W, Xiang J, Jiang X. Ferroptosis is an autophagic cell death process. Cell Research. 2016;26(9):1021-1032
  14. 14. Friedmann Angeli JP, Conrad M. Selenium and GPX4, a vital symbiosis. Free Radical Biology & Medicine. 2018;1(127):153-159
  15. 15. Maiorino M, Conrad M, Ursini F. GPx4, lipid peroxidation, and cell death: Discoveries, rediscoveries, and open issues. Antioxidants & Redox Signaling. 2018;29(1):61-74
  16. 16. Lu L, Hope BT, Shaham Y. The cystine-glutamate transporter in the accumbens: A novel role in cocaine relapse. Trends in Neurosciences. 2004;27(2):74-76
  17. 17. Wang H, An P, Xie E, Wu Q , Fang X, Gao H, et al. Characterization of ferroptosis in murine models of hemochromatosis. Hepatology. 2017;66(2):449-465
  18. 18. Bersuker K, Hendricks JM, Li Z, Magtanong L, Ford B, Tang PH, et al. The CoQ oxidoreductase FSP1 acts parallel to GPX4 to inhibit ferroptosis. Nature. 2019;575(7784):688-692
  19. 19. Doll S, Freitas FP, Shah R, Aldrovandi M, da Silva MC, Ingold I, et al. FSP1 is a glutathione-independent ferroptosis suppressor. Nature. 2019;575(7784):693-698
  20. 20. Stockwell BR, Friedmann Angeli JP, Bayir H, Bush AI, Conrad M, Dixon SJ, et al. Ferroptosis: A regulated cell death Nexus linking metabolism, redox biology, and disease. Cell. 2017;171(2):273-285
  21. 21. Louandre C, Marcq I, Bouhlal H, Lachaier E, Godin C, Saidak Z, et al. The retinoblastoma (Rb) protein regulates ferroptosis induced by sorafenib in human hepatocellular carcinoma cells. Cancer Letters. 2015;356(2):971-977
  22. 22. Sun X, Ou Z, Chen R, Niu X, Chen D, Kang R, et al. Activation of the p62-Keap1-NRF2 pathway protects against ferroptosis in hepatocellular carcinoma cells. Hepatology. 2016;63(1):173-184
  23. 23. Harrison PM, Arosio P. The ferritins: Molecular properties, iron storage function and cellular regulation. Biochimica et Biophysica Acta. 1996;1275(3):161-203
  24. 24. Raghunath A, Sundarraj K, Arfuso F, Sethi G, Perumal E. Dysregulation of Nrf2 in Hepatocellular Carcinoma: Role in Cancer Progression and Chemoresistance. Cancers (Basel). 3 Dec 2018;10(12):481. doi: 10.3390/cancers10120481. PMID: 30513925; PMCID: PMC6315366
  25. 25. Arlt A, Sebens S, Krebs S, Geismann C, Grossmann M, Kruse ML, et al. Inhibition of the Nrf2 transcription factor by the alkaloid trigonelline renders pancreatic cancer cells more susceptible to apoptosis through decreased proteasomal gene expression and proteasome activity. Oncogene. 2013;32(40):4825-4835
  26. 26. Suzuki T, Motohashi H, Yamamoto M. Toward clinical application of the Keap1-Nrf2 pathway. Trends in Pharmacological Sciences. 2013;34(6):340-346
  27. 27. Sun X, Niu X, Chen R, He W, Chen D, Kang R, et al. Metallothionein-1G facilitates sorafenib resistance through inhibition of ferroptosis. Hepatology. 2016;64(2):488-500
  28. 28. Chen WC, Wang CY, Hung YH, Weng TY, Yen MC, Lai MD. Systematic analysis of gene expression alterations and clinical outcomes for long-chain acyl-coenzyme a synthetase family in Cancer. PLoS One. 2016;11(5):e0155660
  29. 29. Kinoshita H, Okabe H, Beppu T, Chikamoto A, Hayashi H, Imai K, et al. Cystine/glutamic acid transporter is a novel marker for predicting poor survival in patients with hepatocellular carcinoma. Oncology Reports. 2013;29(2):685-689
  30. 30. Efferth T, Dunstan H, Sauerbrey A, Miyachi H, Chitambar CR. The anti-malarial artesunate is also active against cancer. International Journal of Oncology. 2001;18(4):767-773
  31. 31. Eling N, Reuter L, Hazin J, Hamacher-Brady A, Brady NR. Identification of artesunate as a specific activator of ferroptosis in pancreatic cancer cells. Oncoscience. 2015;2(5):517-532
  32. 32. Yamaguchi Y, Kasukabe T, Kumakura S. Piperlongumine rapidly induces the death of human pancreatic cancer cells mainly through the induction of ferroptosis. International Journal of Oncology. 2018;52(3):1011-1022
  33. 33. Yang WS, SriRamaratnam R, Welsch ME, Shimada K, Skouta R, Viswanathan VS, et al. Regulation of ferroptotic cancer cell death by GPX4. Cell. 2014;156(1-2):317-331
  34. 34. Wu J, Minikes AM, Gao M, Bian H, Li Y, Stockwell BR, et al. Publisher correction: Intercellular interaction dictates cancer cell ferroptosis via NF2-YAP signalling. Nature. 2019;572(7770):E20
  35. 35. Yang WH, Ding CC, Sun T, Rupprecht G, Lin CC, Hsu D, et al. The hippo pathway effector TAZ regulates ferroptosis in renal cell carcinoma. Cell Reports. 2019;28(10):2501-2508
  36. 36. Ma S, Henson ES, Chen Y, Gibson SB. Ferroptosis is induced following siramesine and lapatinib treatment of breast cancer cells. Cell Death & Disease. 2016;7(7):e2307
  37. 37. Hasegawa M, Takahashi H, Rajabi H, Alam M, Suzuki Y, Yin L, et al. Functional interactions of the cystine/glutamate antiporter, CD44v and MUC1-C oncoprotein in triple-negative breast cancer cells. Oncotarget. 2016;7(11):11756-11769
  38. 38. Timmerman LA, Holton T, Yuneva M, Louie RJ, Padró M, Daemen A, et al. Glutamine sensitivity analysis identifies the xCT antiporter as a common triple-negative breast tumor therapeutic target. Cancer Cell. 2013;24(4):450-465
  39. 39. Lanzardo S, Conti L, Rooke R, Ruiu R, Accart N, Bolli E, et al. Immunotargeting of antigen xCT attenuates stem-like cell behavior and metastatic progression in breast cancer. Cancer Research. 2016;76(1):62-72
  40. 40. Habashy HO, Powe DG, Staka CM, Rakha EA, Ball G, Green AR, et al. Transferrin receptor (CD71) is a marker of poor prognosis in breast cancer and can predict response to tamoxifen. Breast Cancer Research and Treatment. 2010;119(2):283-293
  41. 41. Tonik SE, Shindelman JE, Sussman HH. Transferrin receptor is inversely correlated with estrogen receptor in breast cancer. Breast Cancer Research and Treatment. 1986;7(2):71-76
  42. 42. Martin-Sanchez D, Fontecha-Barriuso M, Sanchez-Niño MD, Ramos AM, Cabello R, Gonzalez-Enguita C, et al. Cell death-based approaches in treatment of the urinary tract-associated diseases: A fight for survival in the killing fields. Cell Death & Disease. 2018;9(2):118
  43. 43. Mazdak H, Yazdekhasti F, Movahedian A, Mirkheshti N, Shafieian M. The comparative study of serum iron, copper, and zinc levels between bladder cancer patients and a control group. International Urology and Nephrology. 2010;42(1):89-93
  44. 44. Tang M, Chen Z, Wu D, Chen L. Ferritinophagy/ferroptosis: Iron-related newcomers in human diseases. Journal of Cellular Physiology. 2018;233(12):9179-9190
  45. 45. Koppula P, Zhang Y, Zhuang L, Gan B. Amino acid transporter SLC7A11/xCT at the crossroads of regulating redox homeostasis and nutrient dependency of cancer. Cancer Communication (Lond). 2018;38(1):12
  46. 46. Koppula P, Zhang Y, Shi J, Li W, Gan B. The glutamate/cystine antiporter SLC7A11/xCT enhances cancer cell dependency on glucose by exporting glutamate. Journal of Biology Chemistry. 2017;292(34):14240-14249
  47. 47. Jiang L, Kon N, Li T, Wang SJ, Su T, Hibshoosh H, et al. Ferroptosis as a p53-mediated activity during tumour suppression. Nature. 2015;520(7545):57-62
  48. 48. Wang SJ, Li D, Ou Y, Jiang L, Chen Y, Zhao Y, et al. Acetylation is crucial for p53-mediated ferroptosis and tumor suppression. Cell Reports. 2016;17(2):366-373
  49. 49. Jennis M, Kung CP, Basu S, Budina-Kolomets A, Leu JI, Khaku S, et al. An African-specific polymorphism in the TP53 gene impairs p53 tumor suppressor function in a mouse model. Genes Development. 2016;30(8):918-930
  50. 50. Xie Y, Zhu S, Song X, Sun X, Fan Y, Liu J, et al. The tumor suppressor p53 limits ferroptosis by blocking DPP4 activity. Cell Reports. 2017;20(7):1692-1704
  51. 51. Ma Q. Role of nrf2 in oxidative stress and toxicity. Annual Review of Pharmacology and Toxicology. 2013;53:401-426
  52. 52. Yuan H, Li X, Zhang X, Kang R, Tang D. Identification of ACSL4 as a biomarker and contributor of ferroptosis. Biochemical and Biophysical Research Communications. 2016;478(3):1338-1343
  53. 53. Rohr-Udilova N, Bauer E, Timelthaler G, et al. Impact of glutathione peroxidase 4 on cell proliferation, angiogenesis and cytokine production in hepatocellular carcinoma. Oncotarget, 2018;9: 10054-10068.
  54. 54. Kinowaki Y, Kurata M, Ishibashi S, Ikeda M, Tatsuzawa A, Yamamoto M, et al. Glutathione peroxidase 4 overexpression inhibits ROS-induced cell death in diffuse large B-cell lymphoma. Laboratory Investigation. 2018;98(5):609-619
  55. 55. Emmanuel N, Li H, Chen J, Zhang Y. FSP1, a novel KEAP1/NRF2 target gene regulating ferroptosis and radioresistance in lung cancers. Oncotarget. 19 Oct 2022;13:1136-1139. doi: 10.18632/oncotarget.28301. PMID: 36264074; PMCID: PMC9584440
  56. 56. Radlowski EC, Johnson RW. Perinatal iron deficiency and neurocognitive development. Frontiers in Human Neuroscience. 2013;23(7):585
  57. 57. Belaidi AA, Bush AI. Iron neurochemistry in Alzheimer's disease and Parkinson's disease: Targets for therapeutics. Journal of Neurochemistry. 2016;139(Suppl. 1):179-197
  58. 58. Bush AI, Curtain CC. Twenty years of metallo-neurobiology: Where to now? European Biophysics Journal. 2008;37(3):241-245
  59. 59. Ayton S, Faux NG, Bush AI. Alzheimer’s disease neuroimaging initiative. Ferritin levels in the cerebrospinal fluid predict Alzheimer's disease outcomes and are regulated by APOE. Nature Communications. 2015;19(6):6760
  60. 60. Connor JR, Snyder BS, Beard JL, Fine RE, Mufson EJ. Regional distribution of iron and iron-regulatory proteins in the brain in aging and Alzheimer's disease. Journal of Neuroscience Research. 1992;31(2):327-335
  61. 61. Bilgic B, Pfefferbaum A, Rohlfing T, Sullivan EV, Adalsteinsson E. MRI estimates of brain iron concentration in normal aging using quantitative susceptibility mapping. Neuroimage. 2012;59(3):2625-2635
  62. 62. Hambright WS, Fonseca RS, Chen L, Na R, Ran Q. Ablation of ferroptosis regulator glutathione peroxidase 4 in forebrain neurons promotes cognitive impairment and neurodegeneration. Redox Biology. 2017;12:8-17
  63. 63. Zhang YH, Wang DW, Xu SF, Zhang S, Fan YG, Yang YY, et al. α-Lipoic acid improves abnormal behavior by mitigation of oxidative stress, inflammation, ferroptosis, and tauopathy in P301S Tau transgenic mice. Redox Biolology. 2018;14:535-548
  64. 64. Raefsky SM, Furman R, Milne G, Pollock E, Axelsen P, Mattson MP, et al. Deuterated polyunsaturated fatty acids reduce brain lipid peroxidation and hippocampal amyloid β-peptide levels, without discernable behavioral effects in an APP/PS1 mutant transgenic mouse model of Alzheimer's disease. Neurobiology of Aging. 2018;66:165-176
  65. 65. Yang WS, Kim KJ, Gaschler MM, Patel M, Shchepinov MS, Stockwell BR. Peroxidation of polyunsaturated fatty acids by lipoxygenases drives ferroptosis. Proceedings of the National Academy Science U S A. 2016;113:E4966-E4975
  66. 66. Dexter DT, Wells FR, Agid F, Agid Y, Lees AJ, Jenner P, et al. Increased nigral iron content in postmortem parkinsonian brain. Lancet. 1987;2(8569):1219-1220
  67. 67. Dexter DT, Wells FR, Lees AJ, Agid F, Agid Y, Jenner P, et al. Increased nigral iron content and alterations in other metal ions occurring in brain in Parkinson's disease. Journal of Neurochemistry. 1989;52(6):1830-1836
  68. 68. Ayton S, Lei P, Duce JA, Wong BX, Sedjahtera A, Adlard PA, et al. Ceruloplasmin dysfunction and therapeutic potential for Parkinson disease. Annals of Neurology. 2013;73(4):554-559
  69. 69. Khandelwal PJ, Herman AM, Moussa CE. Inflammation in the early stages of neurodegenerative pathology. Journal of Neuroimmunology. 2011;238(1-2):1-11
  70. 70. Raina AK, Hochman A, Zhu X, Rottkamp CA, Nunomura A, Siedlak SL, et al. Abortive apoptosis in Alzheimer’s disease. Acta Neuropathologica. 2001;101(4):305-310
  71. 71. Boag MK, Roberts A, Uversky VN, Ma L, Richardson DR, Pountney DL. Ferritinophagy and α-Synuclein: Pharmacological Targeting of Autophagy to Restore Iron Regulation in Parkinson’s Disease. International Journal of Molecular Sciences. 21 Feb 2022;23(4):2378. doi: 10.3390/ijms23042378. PMID: 35216492; PMCID: PMC8878351
  72. 72. Rao IY, Hanson LR, Johnson JC, Rosenbloom MH, Frey WH 2nd. Brain Glucose Hypometabolism and Iron Accumulation in Different Brain Regions in Alzheimer’s and Parkinson’s Diseases. Pharmaceuticals (Basel). 29 Apr 2022;15(5):551. doi: 10.3390/ph15050551. PMID: 35631378; PMCID: PMC9143620
  73. 73. Lei P, Ayton S, Finkelstein DI, Adlard PA, Masters CL, Bush AI. Tau protein: Relevance to Parkinson’s disease. The International Journal of Biochemistry & Cell Biology. 2010;42(11):1775-1778
  74. 74. Lei P, Ayton S, Moon S, Zhang Q , Volitakis I, Finkelstein DI, et al. Motor and cognitive deficits in aged tau knockout mice in two background strains. Molecular Neurodegeneration. 2014;14(9):29
  75. 75. Lei P, Ayton S, Appukuttan AT, Volitakis I, Adlard PA, Finkelstein DI, et al. Clioquinol rescues parkinsonism and dementia phenotypes of the tau knockout mouse. Neurobiology of Disease. 2015;81:168-175
  76. 76. Sian J, Dexter DT, Lees AJ, Daniel S, Agid Y, Javoy-Agid F, et al. Alterations in glutathione levels in Parkinson's disease and other neurodegenerative disorders affecting basal ganglia. Annals of Neurology. 1994;36(3):348-355
  77. 77. Farr AC, Xiong MP. Challenges and Opportunities of Deferoxamine Delivery for Treatment of Alzheimer’s Disease, Parkinson’s Disease, and Intracerebral Hemorrhage. Mol Pharm. 1 Feb 2021;18(2):593-609. doi: 10.1021/acs.molpharmaceut.0c00474. Epub 2020 Oct 9. PMID: 32926630; PMCID: PMC8819678
  78. 78. Monti DA, Zabrecky G, Kremens D, Liang TW, Wintering NA, Cai J, et al. N-acetyl cysteine may support dopamine neurons in Parkinson's disease: Preliminary clinical and cell line data. PLoS One. 2016;11(6):e0157602
  79. 79. Do Van B, Gouel F, Jonneaux A, Timmerman K, Gelé P, Pétrault M, et al. Ferroptosis, a newly characterized form of cell death in Parkinson's disease that is regulated by PKC. Neurobiology of Disease. 2016;94:169-178
  80. 80. Gouel F, Do Van B, Chou ML, Jonneaux A, Moreau C, Bordet R, et al. The protective effect of human platelet lysate in models of neurodegenerative disease: Involvement of the Akt and MEK pathways. Journal of Tissue Engineering and Regenerative Medicine. 2017;11:3236-3240
  81. 81. Codazzi F, Pelizzoni I, Zacchetti D, Grohovaz F. Iron entry in neurons and astrocytes: A link with synaptic activity. Frontiers in Molecular Neuroscience. 2015;3(8):18
  82. 82. Cui Z, Zhong Z, Yang Y, Wang B, Sun Y, Sun Q , et al. Ferrous Iron induces Nrf2 expression in mouse brain astrocytes to prevent neurotoxicity. Journal of Biochemical and Molecular Toxicology. 2016;30(8):396-403
  83. 83. Ishii T, Warabi E, Mann GE. Circadian control of BDNF-mediated Nrf2 activation in astrocytes protects dopaminergic neurons from ferroptosis. Free Radical Biology & Medicine. 2019;133:169-178
  84. 84. Gajowiak A, Styś A, Starzyński RR, Staroń R, Lipiński P. Misregulation of iron homeostasis in amyotrophic lateral sclerosis. Posteṃpy Higieny i Medycyny Doświadczalnej (Online). 2016;70:709-721
  85. 85. Moreau C, Danel V, Devedjian JC, Grolez G, Timmerman K, Laloux C, et al. Could conservative Iron chelation Lead to neuroprotection in amyotrophic lateral sclerosis? Antioxidants & Redox Signaling. 2018;29(8):742-748
  86. 86. Veyrat-Durebex C, Corcia P, Mucha A, Benzimra S, Mallet C, Gendrot C, et al. Iron metabolism disturbance in a French cohort of ALS patients. Biomedical Research International. 2014;2014:485723
  87. 87. Simpson EP, Henry YK, Henkel JS, Smith RG, Appel SH. Increased lipid peroxidation in sera of ALS patients: A potential biomarker of disease burden. Neurology. 2004;62(10):1758-1765
  88. 88. Neel DV, Basu H, Gunner G, Chiu IM. Catching a killer: Mechanisms of programmed cell death and immune activation in amyotrophic lateral sclerosis. Immunological Reviews. 2022;311(1):130-150
  89. 89. Fang X, Wang H, Han D, Xie E, Yang X, Wei J, et al. Ferroptosis as a target for protection against cardiomyopathy. Proceedings of the National Academy Science U S A. 2019;116(7):2672-2680
  90. 90. Cadenas S. ROS and redox signaling in myocardial ischemia-reperfusion injury and cardioprotection. Free Radical Biology & Medicine. 2018;117:76-89
  91. 91. Das DK, Engelman RM, Liu X, Maity S, Rousou JA, Flack J, et al. Oxygen-derived free radicals and hemolysis during open heart surgery. Molecular and Cellular Biochemistry. 1992;111(1-2):77-86
  92. 92. Meerson FZ, Kagan VE, Kozlov YP, Belkina LM, Arkhipenko YV. The role of lipid peroxidation in pathogenesis of ischemic damage and the antioxidant protection of the heart. Basic Research in Cardiology. 1982;77(5):465-485
  93. 93. Gao M, Monian P, Quadri N, Ramasamy R, Jiang X. Glutaminolysis and transferrin regulate ferroptosis. Molecular Cell. 2015;59(2):298-308
  94. 94. Linkermann A, Skouta R, Himmerkus N, Mulay SR, Dewitz C, De Zen F, et al. Synchronized renal tubular cell death involves ferroptosis. Proceedings of the National Academy Science U S A. 2014;111(47):16836-16841
  95. 95. Friedmann Angeli JP, Schneider M, Proneth B, Tyurina YY, Tyurin VA, Hammond VJ, et al. Inactivation of the ferroptosis regulator Gpx4 triggers acute renal failure in mice. Nature Cell Biology. 2014;16(12):1180-1191
  96. 96. Lu L, Wu W, Yan J, Li X, Yu H, Yu X. Adriamycin-induced autophagic cardiomyocyte death plays a pathogenic role in a rat model of heart failure. International Journal of Cardiology. 2009;134(1):82-90
  97. 97. Takemura G, Kanoh M, Minatoguchi S, Fujiwara H. Cardiomyocyte apoptosis in the failing heart--a critical review from definition and classification of cell death. International Journal of Cardiology. 2013;167(6):2373-2386
  98. 98. Zhang T, Zhang Y, Cui M, Jin L, Wang Y, Lv F, et al. CaMKII is a RIP3 substrate mediating ischemia- and oxidative stress-induced myocardial necroptosis. Nature Medicine. 2016;22(2):175-182
  99. 99. Li W, Feng G, Gauthier JM, Lokshina I, Higashikubo R, Evans S, et al. Ferroptotic cell death and TLR4/Trif signaling initiate neutrophil recruitment after heart transplantation. The Journal of Clinical Investigation. 2019;129(6):2293-2304
  100. 100. Guo X, Zhou D, An P, Wu Q , Wang H, Wu A, et al. Associations between serum hepcidin, ferritin and hb concentrations and type 2 diabetes risks in a han Chinese population. The British Journal of Nutrition. 2013;110(12):2180-2185
  101. 101. Altamura S, Mudder K, Schlotterer A, Fleming T, Heidenreich E, Qiu R, et al. Iron aggravates hepatic insulin resistance in the absence of inflammation in a novel db/db mouse model with iron overload. Molecular Metabolism. 2021;51:101235
  102. 102. Wang X, Chen X, Zhou W, Men H, Bao T, Sun Y, et al. Ferroptosis is essential for diabetic cardiomyopathy and is prevented by sulforaphane via AMPK/NRF2 pathways. Acta Pharmaceutica Sinica B. 2022;12(2):708-722
  103. 103. Zang H, Wu W, Qi L, Tan W, Nagarkatti P, Nagarkatti M, et al. Autophagy inhibition enables Nrf2 to exaggerate the progression of diabetic cardiomyopathy in mice. Diabetes. 2020;69(12):2720-2734. DOI: 10.2337/db19-1176
  104. 104. Du S, Shi H, Xiong L, Wang P, Shi Y. Canagliflozin mitigates ferroptosis and improves myocardial oxidative stress in mice with diabetic cardiomyopathy. Frontiers in Endocrinology. 2022. DOI: 10.3389/fendo.2022.1011669
  105. 105. Tuo QZ, Lei P, Jackman KA, Li XL, Xiong H, Li XL, et al. Tau-mediated iron export prevents ferroptotic damage after ischemic stroke. Molecular Psychiatry. 2017;22(11):1520-1530
  106. 106. Li Q , Han X, Lan X, Gao Y, Wan J, Durham F, et al. Inhibition of neuronal ferroptosis protects hemorrhagic brain. JCI Insight. 2017;2(7):e90777
  107. 107. Dietrich RB, Bradley WG Jr. Iron accumulation in the basal ganglia following severe ischemic-anoxic insults in children. Radiology. 1988;168(1):203-206
  108. 108. Lipscomb DC, Gorman LG, Traystman RJ, Hurn PD. Low molecular weight iron in cerebral ischemic acidosis in vivo. Stroke. 1998;29(2):487-492
  109. 109. Ding H, Yan CZ, Shi H, Zhao YS, Chang SY, Yu P, et al. Hepcidin is involved in iron regulation in the ischemic brain. PLoS One. 2011;6(9):e25324
  110. 110. Park UJ, Lee YA, Won SM, Lee JH, Kang SH, Springer JE, et al. Blood-derived iron mediates free radical production and neuronal death in the hippocampal CA1 area following transient forebrain ischemia in rat. Acta Neuropathologica. 2011;121(4):459-473
  111. 111. Patt A, Horesh IR, Berger EM, Harken AH, Repine JE. Iron depletion or chelation reduces ischemia/reperfusion-induced edema in gerbil brains. Journal of Pediatric Surgery. 1990;25(2):224-227
  112. 112. Davis S, Helfaer MA, Traystman RJ, Hurn PD. Parallel antioxidant and antiexcitotoxic therapy improves outcome after incomplete global cerebral ischemia in dogs. Stroke. 1997;28(1):198-204
  113. 113. Prass K, Ruscher K, Karsch M, Isaev N, Megow D, Priller J, et al. Desferrioxamine induces delayed tolerance against cerebral ischemia in vivo and in vitro. Journal of Cerebral Blood Flow and Metabolism. 2002;22(5):520-525
  114. 114. Hanson LR, Roeytenberg A, Martinez PM, Coppes VG, Sweet DC, Rao RJ, et al. Intranasal deferoxamine provides increased brain exposure and significant protection in rat ischemic stroke. The Journal of Pharmacology and Experimental Therapeutics. 2009;330(3):679-686
  115. 115. Speer RE, Karuppagounder SS, Basso M, Sleiman SF, Kumar A, Brand D, et al. Hypoxia-inducible factor prolyl hydroxylases as targets for neuroprotection by “antioxidant” metal chelators: From ferroptosis to stroke. Free Radical Biology Medical. 2013;62:26-36
  116. 116. Chang CF, Cho S, Wang J. (−)-epicatechin protects hemorrhagic brain via synergistic Nrf2 pathways. Annals of Clinical Translational Neurology. 2014;1(4):258-271
  117. 117. Rodríguez C, Sobrino T, Agulla J, Bobo-Jiménez V, Ramos-Araque ME, Duarte JJ, et al. Neovascularization and functional recovery after intracerebral hemorrhage is conditioned by the Tp53 Arg72Pro single-nucleotide polymorphism. Cell Death Differentiation. Jan 2017;24(1):144-154. doi: 10.1038/cdd.2016.109. Epub 2016 Oct 21. PMID: 27768124; PMCID: PMC5260494
  118. 118. Zille M, Karuppagounder SS, Chen Y, Gough PJ, Bertin J, Finger J, et al. Neuronal death after hemorrhagic stroke in vitro and in vivo shares features of ferroptosis and necroptosis. Stroke. 2017;48(4):1033-1043
  119. 119. Zhang Z, Wu Y, Yuan S, Zhang P, Zhang J, Li H, et al. Glutathione peroxidase 4 participates in secondary brain injury through mediating ferroptosis in a rat model of intracerebral hemorrhage. Brain Research. 2018;1701:112-125
  120. 120. Handa P, Thomas S, Morgan-Stevenson V, Maliken BD, Gochanour E, Boukhar S, et al. Iron alters macrophage polarization status and leads to steatohepatitis and fibrogenesis. Journal of Leukocyte Biology. 2019;105(5):1015-1026
  121. 121. Woodhoo A, Iruarrizaga-Lejarreta M, Beraza N, García-Rodríguez JL, Embade N, Fernández-Ramos D, et al. Human antigen R contributes to hepatic stellate cell activation and liver fibrosis. Hepatology. 2012;56(5):1870-1882
  122. 122. Zhang Z, Yao Z, Wang L, Ding H, Shao J, Chen A, et al. Activation of ferritinophagy is required for the RNA-binding protein ELAVL1/HuR to regulate ferroptosis in hepatic stellate cells. Autophagy. 2018;14(12):2083-2103
  123. 123. Atarashi M, Izawa T, Kuwamura M, Yamate J. The role of iron overload in the progression of nonalcoholic steatohepatitis. Nihon Yakurigaku Zasshi. 2019;154(2):61-65
  124. 124. Tsurusaki S, Tsuchiya Y, Koumura T, Nakasone M, Sakamoto T, Matsuoka M, et al. Hepatic ferroptosis plays an important role as the trigger for initiating inflammation in nonalcoholic steatohepatitis. Cell Death & Disease. 2019;10(6):449
  125. 125. Wang W, Green M, Choi JE, Gijón M, Kennedy PD, Johnson JK, et al. CD8+ T cells regulate tumour ferroptosis during cancer immunotherapy. Nature. 2019;569(7755):270-274
  126. 126. Fan X, Li A, Yan Z, Geng X, Lian L, Lv H, et al. From Iron metabolism to ferroptosis: Pathologic changes in coronary heart disease. Oxidative Medicine and Cellular Longevity. 2022;10(2022):6291889
  127. 127. Li Q , Zhao Z, Zhou X, Yan Y, Shi L, Chen J, et al. Ferroptosis: The potential target in heart failure with preserved ejection fraction. Cell. 2022;11(18):2842
  128. 128. Zheng DW, Lei Q , Zhu JY, Fan JX, Li CX, Li C, et al. Switching apoptosis to ferroptosis: Metal-organic network for high-efficiency anticancer therapy. Nano Letters. 2017;17(1):284-291
  129. 129. You L, Wang J, Liu T, Zhang Y, Han X, Wang T, et al. Targeted brain delivery of rabies virus glycoprotein 29-modified deferoxamine-loaded nanoparticles reverses functional deficits in parkinsonian mice. ACS Nano. 2018;12(5):4123-4139

Written By

Yanbo Shi, Junyong Zhang, Kaitao Luo, Sunfeng Pan, Hanqiang Shi, Lie Xiong and Shuqin Du

Submitted: 17 October 2022 Reviewed: 27 October 2022 Published: 25 January 2023