Category: Moms

Iron in energy generation

Iron in energy generation

The effect of cysteine-containing peptides released ennergy meat Iron in energy generation on iron Iron in energy generation in humans. Eneergy IEEE SIGN IN. Even a small amount of nanoparticles makes the burner technology inefficient because a fraction of iron is lost with each cycle. Materials provided by California Institute of Technology. Once absorbed, most binds to transferrin, a transport protein that delivers it to the bone marrow for erythropoiesis.

Iron in energy generation -

Since non-heme iron is less bioavailable than heme iron, people consuming a plant-based vegan or vegetarian diet require a higher intake than individuals consuming animal products.

The recommended daily allowance RDA for iron ranges from 0. The RDA for menstruating females is 18 mg daily and 8 mg for women past reproductive age and adult men. Breast milk contains relatively low levels of highly bioavailable iron, and controversy exists regarding the need to supplement with fortified foods or iron drops and at what age to do so.

Iron needs are higher during the adolescent growth spurt, as girls begin menstruation and boys markedly increase their hemoglobin concentration. Hormone replacement therapy HRT that causes ongoing uterine bleeding and blood donation can also increase dietary iron requirements.

On the other hand, oral contraceptives often decrease menstrual blood loss and lower iron needs. People from less economically developed regions may experience food insecurity, resulting in iron deficiency and malnutrition. In addition, these geographic areas may experience a higher prevalence of parasitic intestinal infections, leading to malabsorption and gastrointestinal blood loss, resulting in increased dietary iron requirements.

Medical conditions that interfere with iron absorption include celiac disease, Crohn disease, malabsorptive disorders, and a history of gastric bypass surgery. Patients with these disorders also have increased iron needs and often require supplementation to maintain adequate stores.

Whole foods, fortified foods, and supplements are the primary sources of iron. The RDA depends on an individual's age, stage of life, and sex. See Table below. Red meat, poultry especially the darker meat in thighs and drumsticks , fish, and shellfish are rich sources of heme iron.

These foods also contain some non-heme iron. Foods containing exclusively non-heme iron include legumes, dark leafy greens, nuts, seeds, whole grains, and dried fruits. Most non-heme iron is found in plant-based foods; eggs are the exception, as they are an animal-based food containing non-heme iron.

Many processed foods, like bread, cereal, and nutritional drinks, are fortified with non-heme iron, representing about half of daily dietary iron intake in the US. Eating vitamin C-rich foods such as citrus fruits, bell peppers, and tomatoes with foods plentiful in non-heme iron can increase iron absorption.

Avoiding beverages like tea and coffee that contain phytates and calcium-containing milk during meals improves the absorption of non-heme iron. Table 2 illustrates the iron content of common foods and the percent of daily value DV from the United States Department of Agriculture USDA National Nutrient Database.

Cooking food in iron cookware significantly improves the iron content of foods. Studies have shown that iron content and absorption are 1. Iron supplements are widely available when dietary intake does not meet nutritional needs.

These include multivitamins with iron and iron-only preparations, which can usually be purchased without a prescription. These are reviewed in detail in the StatPearls companion topics " Iron Supplementation " and " Iron Deficiency Anemia.

The table below lists the iron content of selected animal and plant-based foods from the US National Institutes of Health Iron Factsheet. Dietary iron deficiency affects more than 1. Iron deficiency progresses through stages, from asymptomatic iron depletion to iron deficiency anemia.

Iron toxicity occurs only with excess supplement consumption, including accidental ingestions, and in genetic conditions causing iron overload, such as hemochromatosis.

Because the body absorbs less iron when stores are adequate, toxicity is unlikely to occur solely by eating iron-rich foods. See StatPearls companion topics " Iron Deficiency Anemia ," " Chronic Iron Deficiency ," and " Iron Supplementation.

The trace mineral iron facilitates oxygen transport via hemoglobin, is a component of myoglobin, and is involved in energy production, DNA synthesis, and immune function. Maintaining adequate iron levels is essential for optimal physiological functioning and overall health.

Since the body cannot synthesize iron, it must be derived from dietary sources or supplements. Non-heme iron, mainly found in plant sources such as beans, nuts, dark chocolate, legumes, spinach, and fortified grains, has about half the bioavailability of heme iron. However, it represents a significant percentage of iron absorbed when a diet includes diverse plant-based foods.

The interprofessional healthcare team can collaborate to improve patient care by educating patients on the importance of consuming adequate amounts of iron-containing foods. Primary care nurses and clinicians can work together to prevent nutritional deficiencies by obtaining patient dietary histories, providing specific advice about iron needs for different stages of life, answering patient questions, and advocating for healthy foods in local school meals, food pantries, and nutrition programs.

They can coordinate services with registered dieticians, nutritionists, subspecialists, and pharmacists for patients with medical conditions requiring consultation. All team members can counsel patients about iron-rich foods, interactions with foods and beverages that enhance or hinder absorption, and when indicated, how to take supplements safely.

The healthcare team must be aware of special populations at risk for iron deficiency due to inadequate nutrition or physiologic needs and screen and educate them to ensure optimal health. Recommended Dietary requirement of IRON. Iron Content in different food sources described by American department of Agriculture.

Contributed by Dr. Fady Moustarah, MD. Disclosure: Fady Moustarah declares no relevant financial relationships with ineligible companies. Disclosure: Sharon Daley declares no relevant financial relationships with ineligible companies. This book is distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivatives 4.

You are not required to obtain permission to distribute this article, provided that you credit the author and journal. Turn recording back on. National Library of Medicine Rockville Pike Bethesda, MD Web Policies FOIA HHS Vulnerability Disclosure.

Help Accessibility Careers. Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation. Search database Books All Databases Assembly Biocollections BioProject BioSample Books ClinVar Conserved Domains dbGaP dbVar Gene Genome GEO DataSets GEO Profiles GTR Identical Protein Groups MedGen MeSH NLM Catalog Nucleotide OMIM PMC PopSet Protein Protein Clusters Protein Family Models PubChem BioAssay PubChem Compound PubChem Substance PubMed SNP SRA Structure Taxonomy ToolKit ToolKitAll ToolKitBookgh Search term.

StatPearls [Internet]. Treasure Island FL : StatPearls Publishing; Jan-. Show details Treasure Island FL : StatPearls Publishing ; Jan-. Search term. Dietary Iron Fady Moustarah ; Sharon F. Author Information and Affiliations Authors Fady Moustarah 1 ; Sharon F.

Affiliations 1 Central Michigan University. Introduction The trace element iron is essential for optimal physiological functioning and overall health and must be derived from dietary food sources and supplements. Function Iron is absorbed in the small intestine, primarily in the duodenum and upper jejunum.

Issues of Concern Iron absorption involves heme iron from animal-based foods and non-heme iron from plant-based foods and supplements. Iron Content of Selected Foods Table. Clinical Significance Dietary iron deficiency affects more than 1.

Enhancing Healthcare Team Outcomes The trace mineral iron facilitates oxygen transport via hemoglobin, is a component of myoglobin, and is involved in energy production, DNA synthesis, and immune function. Review Questions Access free multiple choice questions on this topic.

Comment on this article. Figure Recommended Dietary requirement of IRON. References 1. Coad J, Pedley K. Iron deficiency and iron deficiency anemia in women. Scand J Clin Lab Invest Suppl. Hunt JR, Zito CA, Johnson LK. Body iron excretion by healthy men and women. Am J Clin Nutr. Abbaspour N, Hurrell R, Kelishadi R.

Review on iron and its importance for human health. J Res Med Sci. von Drygalski A, Adamson JW. Iron metabolism in man.

JPEN J Parenter Enteral Nutr. Han O. Molecular mechanism of intestinal iron absorption. Hunt JR, Roughead ZK. Adaptation of iron absorption in men consuming diets with high or low iron bioavailability.

Hurrell R, Egli I. Iron bioavailability and dietary reference values. Carpenter CE, Mahoney AW. Contributions of heme and nonheme iron to human nutrition. Crit Rev Food Sci Nutr. However, excess iron can also generate ROS. B Iron uptake, redistribution, and export at the cellular level.

Endocytosis of Tf-TfR1 complex releases iron to the cytosol LIP by DMT1 or directly to mitochondria via DMT1 and mitoferrin. Iron from LIP can be transported to mitochondria by DMT1, mitoferrin, and SFXN1, exported by FPN, or stored in ferritin.

After degradation of ferritin in the lysosomes, iron is released to replenish LIP. Given the importance of iron in the context of fundamental physiology, the levels of iron both within cells and in the body as a whole are stringently regulated, and the processes of iron absorption, transfer, storage, and retrieval by multiple mechanisms are finely balanced Fig.

In mammals, iron is absorbed through the enterocytes of the duodenal mucosa in the gastrointestinal tract by divalent metal transporter 1 DMT1 , which provides the primary pathway for the entry of dietary iron into enterocytes [ 7 , 8 ]. After being transferred from the enterocytes into bloodstream via ferroportin FPN [ 9 ], iron binds to transferrin Tf which delivers it to cells throughout the body.

These cells take up iron-laden Tf via transferrin receptor 1 TfR1. After internalization of the Tf-TfR1 complex into endosome, iron is released from Tf, and transferred to the cytosol by DMT1 to join the labile iron pool LIP.

Endosomal iron can be directly delivered to the mitochondria via the interaction between DMT1 and mitoferrin. Excess iron from LIP is stored in ferritin, which can be delivered to and degraded in lysosomes, and this, in turn, replenishes LIP. Furthermore, iron from LIP can be transferred in mitochondria by DMT1, mitoferrin, and siderofexin SFXN1.

Iron efflux is mediated by FPN Fig. Certain cells such as immature erythroid cells in the bone exhibit particularly high iron requirements, while other cells, particularly hepatocytes and splenic macrophages, play a major role in iron storage [ 11 ].

Systemic iron metabolism is finely regulated through multiple mechanisms, including transcriptional, translational, post-translational e. Hepcidin-mediated FPN internalization and degradation the hepcidin-FPN axis is the most important regulatory mechanism for systemic iron metabolism and regulates both dietary iron intake and iron recycling from senescent red blood cells by macrophages Fig.

Quantitatively, the latter is significantly more important [ 12 ]. To this end, iron is vital for a wide range of physiological functions and processes, including DNA replication, the TCA cycle, ETC-driven ATP production, and signal transduction.

In view of this, systemic and cellular iron homeostasis is finely tuned to avoid iron overload or iron deficiency through multiple regulatory mechanisms. Both iron deficiency and iron overload have been associated with dysregulated glucose metabolism Table 1.

Animals with iron deficiency exhibit hyperinsulinemia, hyperglycemia, and hyperlipidemia, ultimately leading to their preferential fuel usage being changed from fat to glucose [ 13 ].

Additionally, cardiomyocytes and muscle cells that are treated with iron chelators markedly increase their glucose uptake and transport, and this is associated with an increased expression of GLUT1 [ 14 ].

Conversely, iron overload decreases insulin sensitivity and induces insulin resistance, which is associated with reduced glucose uptake, and this occurs either by promoting ROS production or impairing autophagy [ 15 ].

However, an in vivo study demonstrated that mice fed a high-iron diet exhibited enhanced glucose uptake and elevated AMP-activated protein kinase AMPK activity in skeletal muscle and the liver [ 16 ]. These reports suggest that glucose is the preferred metabolic fuel when iron homeostasis is disturbed.

Hepatic production of glucosephosphatase G6Pase , an enzyme that catalyzes the last step in gluconeogenesis, is known to be inhibited by both insulin and AMPK [ 17 ].

AMPK can be activated by iron overload, thus supporting the view that iron acts as a suppressor of gluconeogenesis. Furthermore, the transcription of gluconeogenic genes, including G6Pase, can be downregulated by heme or heme-derived iron [ 18 ].

Carnitine palmitoyl transferase 1 CPT-1 is the rate-limiting enzyme in fatty acid oxidation and conjugates fatty acids with carnitine [ 19 ]. In the fetal liver, iron deficiency markedly decreases the abundance of CPT-1 mRNA, thus suggesting that fatty acid oxidation is impaired [ 20 ].

Peroxisome proliferator-activated receptors are key transcription factors that regulate the expression of enzymes involved in fatty acid oxidation. Hepatic expression of peroxisome proliferator-activated receptor α is dramatically inhibited by iron overload.

Iron deficiency and iron chelation promote fatty acid synthesis and cytosolic lipid droplet accumulation, which is accompanied by a rapid increase in intracellular citrate concentrations [ 22 ], leading to non-autophagic and non-apoptotic cell death in human breast cancer cells [ 23 ].

Both in vitro and in vivo studies have demonstrated that hepatic lipogenesis is enhanced by iron deficiency [ 24 ]. As a result, hepatic phospholipids in iron-depleted rats possessed lower proportions of palmitoleic and oleic acids and a higher proportion of stearic acid [ 26 ], thus indicating impaired desaturation of saturated and essential fatty acids.

An important amino acid, 4-hydroxyproline, in collagen is synthesized from proline by the iron-containing dioxygenase prolylhydroxylase [ 27 ].

Cysteine dioxygenase, another iron-containing enzyme, is vital for cysteine catabolism [ 28 ]. BOLA3, a ISC biogenesis protein, is required for glycine cleavage, and BOLA3 deficiency leads to increased glycine accumulation and promotes endothelial proliferation [ 29 ]. Furthermore, the iron-driven Fenton reaction catalyzes the oxidative deamination-decarboxylation of all amino acids, with [Fe III salen ]Cl serving as an active and selective catalyst for the oxidation of amino acids [ 30 ].

Importantly, NH 4 , α-ketoacids, CO 2 , aldehydes, and carboxylic acids are generated by the oxidation of amino acids via the Fenton reaction [ 31 ]. The oxidation of amino acids is also promoted by iron chelators [ 31 ].

Although our knowledge of the involvement of iron in amino acid metabolism is relatively limited, it is clearly an area that warrants further investigation Table 1. Energy production from glucose catabolism is conducted by two major metabolic programs, anaerobic glycolysis in the cytoplasm and aerobic OXPHOS in mitochondria.

Glycolysis links the metabolism of glucose, lipids, and amino acids. Glucose is converted into pyruvate by enzyme-catalyzed reactions in the cytoplasm, and it then enters the mitochondria and is decarboxylated to form acetyl CoA.

Under aerobic conditions, acetyl-CoA enters the TCA cycle and is oxidized to water and carbon dioxide, ultimately producing a large amount of ATP through OXPHOS Fig.

Iron deficiency can increase intracellular glucose levels by promoting glucose uptake and gluconeogenesis. Lipid metabolism is also altered in response to iron deficiency through increased lipogenesis and lipid droplet formation, and the inhibition of fatty acid desaturation.

The metabolism of glucose, lipids, and amino acids converges in mitochondria at the point of acetyl-CoA, which enters the TCA cycle for energy production.

Iron levels can also modulate the synthesis of several key enzymes in the TCA cycle, including aconitase, SDH, and fumarase, and complexes I, II, III, and IV in the ETC. Iron is essential for metabolic activity in all living organisms due to its catalytic role.

In the TCA cycle, ISCs are crucial cofactors for three enzymes, aconitase, succinate dehydrogenase SDH , and fumarase Fig. Intriguingly, there is also a cytosolic form of aconitase, and when it loses its ISC, it becomes iron regulatory protein 1 that acts as an important regulator of cellular iron uptake and storage [ 33 ].

Given the importance of iron for these various enzymes, modulating iron levels has the potential to alter the expression of enzymes involved in glycolysis and the TCA cycle, including citrate synthase, aconitase, isocitrate dehydrogenase, and SDH and also their intermediates [ 34 ].

ISCs are also essential for OXPHOS efficiency as a key component in several complexes in the respiratory chain, including complex I NADH-dehydrogenase , complex II SDH , and complex III ubiquinol: cytochrome c -oxidoreductase [ 32 ]. Impaired ISC biogenesis and assembly lead to deficiencies in multiple respiratory chain complexes [ 35 ].

Excess iron alters the mitochondrial oxidative enzymatic machinery, and iron-guided metabolic remodeling is gaining increasing attention. For example, iron supplementation results in pyruvate accumulation and a decrease in lactate levels in conjunction with changes in the concentrations of several other metabolites [ 34 ].

Iron overload reduces glucose oxidation in murine cardiac muscle, and this is accompanied by decreased activity of mitochondrial complexes I-IV and low ATP production [ 36 ]. In contrast, iron deprivation enhances glycolysis and abolishes OXPHOS in human macrophages, along with associated inhibition of the TCA cycle [ 22 ].

Metabolic reprogramming has also been observed in human fibroblasts and cardiac myocytes when iron status is perturbed [ 37 ]. In , a novel synthetic compound, erastin was found to initiate a new form of non-apoptotic cell death in RAS-overexpressing cancer cells [ 38 ].

Subsequently, erastin was observed to directly bind to mitochondrial voltage-dependent anion channels, leading to oxidative stress and cell death via a non-apoptotic mechanism in cancer cells with oncogenic RAS [ 39 ].

Furthermore, Ras-selective lethal small molecule 3 RSL3 was revealed to induce this type of cell death with the implication of labile iron [ 40 ]. In , this new form of RCD was formally named ferroptosis by Dixon et al.

Ferroptosis is characterized by iron-dependent ROS accumulation, glutathione GSH depletion, glutathione peroxidase 4 GPX4 suppression, and ultimately lipid peroxidation [ 3 , 6 ].

The morphological features of ferroptosis include intact nuclei and aberrant mitochondria with a decrease in the number of mitochondrial cristae, and the occurrence of inner membrane condensation, outer membrane rupture, and mitochondrial shrinkage [ 41 ].

Ferroptosis can be induced or inhibited through several metabolic pathways Fig. DMT1 is also expressed in the outer mitochondrial membrane and induces mitochondrial uptake of iron and manganese [ 45 ], thus indicating its potential role in iron influx and ferroptosis. Several metabolic pathways are involved in the regulation of ferroptosis, including: i iron-Fenton reaction black.

ROS are produced by the Fenton reaction that is driven by excessive iron, which can be derived from the import of extracellular iron and the supply of intracellular stored iron via ferritinophagy.

ii GPX4 antioxidant activity brown and blue. Suppression of lipid peroxidation and ferroptosis occurrence largely depends on GPX4 activity, which relies on GSH and IPP-derived Sec. Additionally, IPP-derived CoQ inhibits ferroptosis mediated by FSP1 in the cytosol or by DHODH in mitochondria.

iii Lipid metabolism pathway green. The oxidation of PUFA and AA is also involved in ferroptosis. The antioxidant activity of GPX4 is key to inhibiting lipid peroxidation and preventing ferroptosis by restoring cellular redox homeostasis [ 46 ].

Cystine-derived GSH is necessary for the maintenance of GPX activity. Glutamate itself can be replenished by glutamine import via solute carrier family 1 member 5 SLC1A5 [ 48 ]. Isopentenyl pyrophosphate IPP -derived Sec is also necessary for the catalytic activity of GPX4.

IPP can be produced from acetyl-CoA via the mevalonate pathway and function as the donor of Sec in the incorporation of GPX4 [ 49 ]. RSL3 directly inhibits GPX4 [ 41 ]. Furthermore, IPP-derived CoQ can be converted into ubiquinol CoQH 2 by ferroptosis suppressor protein 1 FSP1 to inhibit lipid peroxidation and ferroptosis [ 51 ], and FSP1 can act in parallel to the GPX4 pathway to inhibit ferroptosis in cancer cells [ 52 ].

Interestingly, conversion of CoQ to ubiquinol by dihydroorotate dehydrogenase DHODH in mitochondria was recently reported to suppress mitochondrial lipid peroxidation and ferroptosis [ 53 ].

More importantly, DHODH operates in parallel with mitochondrial GPX4 to inhibit ferroptosis, and this is independent of cytosolic FSP1 and GPX4 [ 53 ]. PUFA and arachidonic acid AA can be converted into PE-AA by lysophosphatidylcholine acyltransferase LPCAT and acyl-CoA synthetase long-chain family member 4 ACSL4.

PE-AA is then oxidized by lipoxygenases LOXs , leading to ROS production and lipid peroxidation [ 54 ]. Ceruloplasmin and hephaestin are two multicopper ferroxidases that play important roles in iron export.

Knockout of ceruloplasmin and hephaestin was demonstrated to induce iron deposition in mouse astrocytes and oligodendrocytes, respectively [ 55 , 56 ]. Furthermore, a copper chelator, cuprizone, was revealed to induce rapid ferroptosis-mediated loss of oligodendrocytes by mobilizing iron from ferritin [ 57 ].

Little is known regarding the physiological functions of ferroptosis. GPX4 ablation in the hematopoietic system resulted in anemia as a result of failed maturation of reticulocytes into red blood cells [ 62 ].

These reticulocytes accumulated large autophagosomes that engulfed the mitochondria [ 62 ], thus suggesting an indispensable role for GPX4 in erythropoiesis. Ferroptosis has also been demonstrated to be activated to combat the infection of mice with Plasmodium falciparum [ 63 ] and rice with the fungus Magnaporthe oryzae [ 64 ].

Additionally, ferroptosis has been revealed to play a critical role in the defense against tumorigenesis. P53 and BAP1, two important tumor suppressors, predispose nascent tumor cells to ferroptosis by downregulating SLC7A11 expression [ 65 ].

Cells carrying a p53 mutation that was defective in apoptosis induction were revealed to be capable of suppressing tumorigenesis by potentiating ferroptosis [ 65 ]. However, little is known regarding the physiological roles of ferroptosis during ageing.

In the roundworm Caenorhabditis elegans C. elegans , glutathione depletion is inversely correlated with the aging-related accumulation of ferrous iron, thus leading to the priming of ferroptosis [ 67 ]. Inhibition of ferroptosis can reduce age-related cell death, thus increasing the lifespan of C.

The contribution of ferroptosis at specific life phases rather uniformly throughout life appears to be particularly important in determining the lifespan of C.

elegans [ 67 ]. Other areas where ferroptosis may be important include placenta shedding, a process in which iron accumulation occurs [ 68 ]. In summary, ferroptosis may contribute to embryonic development, erythropoiesis, determination of lifespan, and defense against infection and tumors under physiological conditions.

How iron is involved and contributes to ferroptosis under these diverse circumstances remains poorly understood.

Thus, both the activation and inactivation of ferroptosis are indispensable in the context of physiological settings.

The LIP is a pool of non-protein-bound, chelatable and redox-active iron, that serves as a source of free iron and sits at the crossroads of iron metabolism.

LIP levels are primarily regulated by iron uptake, iron release from ferritin, and iron utilization. Iron-containing LOXs, such as arachidonatelipoxygenase, also catalyze the reaction between O 2 and LH to form LOOH [ 70 , 71 ].

Moreover, iron is a component of the catalytic subunit of LOX [ 72 ]. In general, iron-dependent LOXs initiate ferroptosis, whereas the iron-driven Fenton reaction propagates ferroptosis [ 73 ].

Ferroptosis can be inhibited by iron chelators e. Inhibition of iron uptake by knocking down TfR1 can suppress lipid ROS formation [ 74 ].

Furthermore, compounds that chelate intracellular iron, such as desferrioxamine DFO or deferiprone DFP , can suppress lipid ROS generation [ 5 ]. Iron chelators can also remove iron from LOXs, thus rescuing ferroptosis [ 75 ]. Ferritinophagy has been reported to induce ferroptosis by releasing iron from ferritin and thus increasing intracellular LIP levels Fig.

Knocking down NCOA4 inhibits ferritinophagy and subsequently suppresses lipid ROS formation [ 42 ]. Suppression of the mitochondrial protein frataxin, an iron chaperone that drives ISC biogenesis, has been demonstrated to promote cysteine deprivation-induced ferroptosis in cancer cells [ 78 ].

Iron accumulation induced by the downregulation of mitochondrial ferritin can also cause mitochondrial ROS accumulation, leading to ferroptosis [ 80 ]. Taken together, both cytosolic iron and mitochondrial iron is an essential for ferroptosis.

Iron-containing LOXs are required for initiating lipid peroxidation, whereas the iron-driven Fenton reactions are required for propagating lipid peroxidation.

The inhibition of iron uptake and chelation of intracellular iron are effective in reducing lipid peroxidation and suppressing ferroptosis.

However, the details of how intracellular iron levels, particularly the size of the LIP, are controlled and what threshold of iron concentration is required to induce ferroptosis remain elusive.

As mitochondria play an important role in ROS production, they are closely associated with ferroptosis [ 81 ]. Indeed, complete depletion of mitochondria increases the tolerance of cells to ferroptosis under cysteine deprivation conditions [ 81 , 82 ].

However, cells that are only partially depleted of mitochondria remain sensitive to ferroptosis [ 83 ], thus suggesting that residual mitochondria are capable of initiating ferroptosis. The inhibition of the TCA cycle and ETC also suppresses ferroptosis, and this is consistent with the role played by mitochondria in generating ROS [ 81 , 82 ].

In response to cysteine deprivation or erastin treatment, several enzymes in the TCA cycle, including fumarate hydratase FH , aconitase ACO , and citrate synthase CS , are required to induce ferroptosis [ 81 ]. Renal cancer cells with FH loss are resistant to cystine deprivation-induced ferroptosis [ 81 ].

Additionally, reducing the activity of the TCA cycle suppresses lipid peroxidation and ferroptosis [ 84 ]. Under cysteine deprivation, mitochondrial respiration is promoted, leading to ROS production, lipid peroxidation, and ferroptosis. However, regardless of cysteine depletion or erastin treatment, glutamine Gln is required to induce ferroptosis [ 81 ].

Moreover, fatty acid metabolism in mitochondria is an important contributor to ferroptosis by inducing lipid peroxidation [ 85 ]. There are several characteristics that distinguish ferroptosis from apoptosis, necrosis, and pyroptosis [ 5 , 41 ] Table 2.

In particular, ferroptosis induces a series of mitochondrial morphological changes, including a reduction in the number of cristae, outer membrane rupture, and membrane coagulation, whereas other forms of RCD only exhibit swollen mitochondria [ 5 ].

However, ferroptosis shares some regulators and signaling pathways with other forms of RCDs. For example, all of these pathways can be triggered by iron-related signals Table 2.

For intrinsic apoptosis, iron-dependent oxidative stress is an important inducer of mitochondrial outer membrane permeabilization. This is followed by the release of cytochrome c and activation of caspase-9 and caspase-3, leading to the induction of apoptosis [ 86 ].

Iron-induced ROS also dissociate thioredoxin from apoptosis signal-regulating kinase 1 ASK1. In extrinsic apoptosis, excess iron prevents the generation of a short form of Fas by inhibiting the alternative splicing of Fas, and thus activates caspasedependent apoptosis [ 88 ].

Similarly, suppression of alternative splicing of Fas by iron activates the mixed-lineage kinase domain-like MLKL and receptor-interacting serine-threonine kinase RIPK , thus leading to necroptosis [ 89 ].

Heme-induced tumor necrosis factor α also activates the MLKL and RIPK pathways [ 90 ]. Moreover, iron-induced ROS are involved in necroptosis [ 91 ].

Iron-induced formation of ROS causes oxidation and oligomerization of Tom20, leading to the recruitment of Bax and subsequent cytochrome c release and caspase-3 activation. Caspase-3 further triggers gasdermin E GSDME cleavage, leading to a switch from apoptosis to pyroptosis [ 92 ].

Collectively, cytochrome c release and caspase activation driven by iron-induced ROS are the common routes for the activation of intrinsic apoptosis and pyroptosis, while the suppression of short Fas by iron is a common route for the activation of extrinsic apoptosis and necroptosis.

Thus, excess iron is one of the inducers for the activation of apoptosis, necroptosis, and pyroptosis, while ferroptosis is initiated by iron-dependent LOXs and propagated by the iron-driven Fenton reaction.

More importantly, disorders of iron hemostasis and energy production frequently occur in these pathological conditions. Here, we review the literature to elucidate the complicated connections among iron homeostasis disorders, dysregulated energy production, and ferroptosis in the context of these pathological conditions, particularly in cancer, diabetes, and neurodegenerative diseases.

Many types of cancer cells appear to be intrinsically sensitive to ferroptosis. Importantly, sensitivity to ferroptosis occurs during the therapy-resistant state transitions in cancer cells [ 93 ].

Furthermore, cancer cells with a higher degree of malignancy, particularly those with high metastatic capacity, are more sensitive to ferroptosis [ 94 ]. Indeed, the levels of intracellular iron, PUFAs, oxidative stress, and lipid peroxidation are key factors in determining the susceptibility of cancer cells to ferroptosis [ 54 , 70 , 95 ].

However, cancer cells can also develop resistance to ferroptosis. The mechanisms underlying resistance to ferroptosis are not well defined, and several routes have been implicated.

FSP1 was observed to be capable of compensating for GPX4 deletion to inhibit ferroptosis in cancer cells [ 51 ].

Furthermore, FSP1 expression was positively correlated with ferroptosis resistance across hundreds of cancer cell lines and in mouse tumor xenografts [ 52 ]. Monounsaturated fatty acids MUFAs were demonstrated to potently inhibit ferroptosis in human fibrosarcoma cells after activation by ACSL3 [ 96 ].

The tumor malignancy of liver cancer is modulated by the balance between HIC1, a transcription factor controlling the expression of a set of ferroptosis-upregulated factors, and HNF4A, another transcription factor controlling the expression of a set of ferroptosis-downregulated factors [ 97 ].

Cancer cells exhibit particularly high iron requirements to support their rapid growth, and consequently, they are well adapted for acquiring iron and preventing its loss [ 98 ]. For example, TfR1 is highly expressed on the surface of cancer cells to facilitate iron uptake and support enhanced survival and resistance to chemotherapy of these cells [ 99 ].

Consistent with an increased iron content of cancer cells, levels of the iron storage protein ferritin are increased in many cancers, including breast cancer, and ferritin can be used as a prognostic marker for breast cancer progression [ ].

We have previously demonstrated that serum levels of the iron regulatory peptide hepcidin are increased, while levels of its target FPN are decreased in breast cancer tissue from patients, and this is consistent with increased iron levels in breast cancer cells [ ].

Importantly, disordered iron metabolism, dysregulated p53 expression, and mitochondrial dysfunction appear to be integrated in cancer metabolic reprogramming, and this may explain cancer resistance to ferroptosis. Cancer cells adapt to hypoxia through metabolic reprogramming [ ].

Glycolysis is enhanced in the cytosol, while the TCA cycle and OXPHOS are inhibited in the mitochondria of cancer cells under hypoxia in a response termed the Warburg effect [ ]. Reprogrammed glucose metabolism in cancer cells is coupled with increased uptake of glucose and amino acids, reduced ATP production [ ], and reduced ROS generation.

This, in turn, increases the carbon supply for synthesizing proteins, lipids, and nucleic acids, prevents ROS-triggered apoptosis [ ], and suppresses ferroptosis by oxidative stress. Adaption to hypoxia is primarily regulated by hypoxia inducible factors HIFs , which are composed of α and β subunits [ ].

Under hypoxia, ubiquitination and degradation of the HIF α subunits, a process that is mediated by prolylhydroxylase-catalyzed hydroxylation, is inhibited [ ]. Ironically, iron is required for the activity of prolylhydroxylase [ ]. P53 can activate ferroptosis by suppressing SLC7A11 [ ]. P53 can also inhibit the production of some anti-ferroptosis metabolites, such as squalene and ubiquinone, by modulating the mevalonate pathway [ ].

Whether FSP1, MUFAs, and the disrupted balance of HIC1 and HNF4A correlate with metabolic reprogramming in inducing cancer resistance to ferroptosis still needs to be investigated. Recently, energy stress was observed to inhibit ferroptosis, and human renal carcinoma cells exhibiting high basal AMPK activation were demonstrated to be resistant to ferroptosis via AMPK-mediated phosphorylation of acetyl-CoA carboxylase and biosynthesis of PUFA [ ].

Whether AMPK activation upon energy stress has a cross-talk with metabolic reprogramming to induce ferroptosis resistance also warrants investigation. Furthermore, how iron accumulation and metabolic reprogramming co-exist in cancer cells remains unknown. To this end, ferroptosis resistance can be induced directly by elevating the levels of FSP1, HNF4A, and MUFAs, or by reducing P53 levels.

Although iron accumulates in these cells, ferroptosis is suppressed and does not lead to excess ROS generation and lipid peroxidation due to metabolic reprogramming.

Loss of function of p53 , as a result of p53 mutations and excess iron, and AMPK activation contributes to ferroptosis resistance either directly or indirectly by promoting metabolic reprogramming in cancer cells Fig. Targeting ferroptosis resistance is emerging as a promising therapeutic strategy for cancer treatment.

Therapeutic strategies based on enhancing ferroptosis are being tested in some cancers, including chemical drugs and nano drugs.

In cancer cells, direct and indirect mechanisms may be involved in inducing ferroptosis resistance. Elevated levels of FSP1 and MUFAs, and a disrupted balance between HNF4A and HIC1 can directly inhibit ferroptosis in cancer cells. Furthermore, p53 mutations can directly induce ferroptosis resistance.

In contrast, metabolic reprogramming could indirectly and substantially induce ferroptosis resistance by integrating iron metabolism disorders, dysregulated p53 levels, and mitochondrial dysfunctions p53 mutations and hypoxia induce metabolic reprogramming, including enhanced glycolysis in the cytosol and inhibition of the TCA cycle and OXPHOS in the mitochondria.

This reduces ATP production and ROS generation, which can attenuate lipid peroxidation and lead to ferroptosis resistance. Accumulated iron in cancer cells can induce p53 mutations and also directly promote metabolic reprogramming.

Energy stress-induced AMPK activation also contributes to ferroptosis resistance in cancer cells. Ferroptosis has been implicated in the pathology of diabetes. A high-fat high-sucrose diet diminished the expression and activity of GPX4 in the hypothalamus relative to a normal diet, indicating that GPX4 plays an important role in regulating metabolic signals [ ].

Additionally, ferroptosis impairs islet function, and ferroptosis inhibitors can reverse this impairment.

Consequently, glucose levels normalized after bilirubin-pretreated islets were transplanted into diabetic mice [ ]. Furthermore, maternal hyperandrogenism and insulin resistance activate ferroptosis in the gravid uterus and placenta [ ].

Mitochondrial dysfunction is an important feature of diabetes mellitus and is a key component of ferroptosis. In most organs and tissues of patients with diabetes and also animal models of diabetes, excessive mitochondrial ROS production is observed, while other mitochondrial abnormalities, including impaired mitochondrial biogenesis and OXPHOS, disordered mitochondrial dynamics, and mitophagy, are tissue-specific [ ].

In both type 1 diabetes mellitus T1DM and type 2 diabetes mellitus T2DM , a switch in the energy source from glucose to fatty acids occurs. Increased uptake and utilization of fatty acids via β-oxidation further reduces glucose uptake in T2DM, thus leading to enhanced gluconeogenesis [ ] and mitochondrial uncoupling [ ].

Ultimately, the energy source switch and mitochondrial dysfunction lead to decreased ATP production and increased ROS generation [ ]. Excessive ROS production in mitochondria can induce insulin resistance by decreasing GLUT4 levels, inducing beta-cell and mitochondrial dysfunction, promoting inflammation, and inhibiting insulin signaling pathways [ ].

Diabetes mellitus is also associated with iron overload, another ferroptosis inducer. A variety of meta-analyses and systematic reviews have confirmed the relationship between iron homeostasis disorders and T2DM risk [ ].

Pregnant women with gestational diabetes were also observed to exhibit significantly higher levels of serum iron, serum ferritin, and transferrin saturation, and increased fasting plasma glucose levels compared to those of pregnant women without this condition [ ]. Additionally, increased serum ferritin levels were observed to be positively associated with the risk for T2DM in otherwise healthy women, and this was independent of known diabetic risk factors [ ].

Consistent with these observations, iron depletion through phlebotomy can increase insulin sensitivity [ ].

In streptozotocin-induced diabetic rats, an iron-restricted diet ameliorated diabetes-induced mitochondrial dysfunction and restored mitochondrial respiration and respiratory complex activity, thereby reducing oxidative stress [ ]. Diabetes-driven ferroptosis reflects a switch in cellular energy sources, mitochondrial dysfunction, and iron overload, leading to reduced insulin secretion.

Collectively, both the energy source switch and the iron overload contribute to mitochondrial dysfunction, including uncoupled respiration and reduced activities of ETC complexes.

Mitochondrial dysfunction decreases ATP production, increases ROS generation, and likely suppresses GPX expression, all of which are key inducers of ferroptosis.

Iron overload can directly promote ROS generation via the Fenton reaction as discussed above. Importantly, bilirubin can suppress ferroptosis in the context of diabetes by decreasing iron levels and restoring mitochondrial function Fig.

In diabetes mellitus, a switch of energy source from glucose to fatty acids occurs, leading to uncoupling of OXPHOS, and impaired function of ETC complexes. Consequently, Gpx expression and ATP production are both inhibited, while ROS generation is increased. These events ultimately induce ferroptosis in the islet cells.

In addition to directly promoting ROS generation and insulin resistance, iron overload in diabetes mellitus can promote uncoupling of OXPHOS and dysfunction of the components of the ETC. Disordered iron metabolism is closely associated with neurodegenerative diseases, and ferroptosis has been demonstrated in this context [ ].

For example, upregulation of the iron regulatory protein IRP2 induces ferroptosis by increasing the intracellular iron content in aging-related auditory cortical neurodegeneration [ ].

Additionally, enhanced NCOA4-mediated ferritinophagy is linked to neurodegenerative diseases, and is also linked to ferroptosis [ 77 ]. Importantly, mitochondria play an important role in regulating iron homeostasis in the brain, and are implicated in the death of neuronal cells by inducing lipid peroxidation [ ].

In patients with AD, reduced activity of various TCA cycle enzymes, including isocitrate dehydrogenase, α-ketoglutarate dehydrogenase, and pyruvate dehydrogenase, has been observed [ ]. In turn, ROS impairs the functions of complexes I, III, and IV [ ].

Furthermore, the release of cytochrome c oxidase COX and mitochondrial permeability are both increased by Aβ and alpha-synuclein oligomerization and polymerization in AD [ ]. Nevertheless, direct evidence for the involvement of mitochondrial dysfunction in ferroptosis is still lacking in neurodegenerative diseases.

In AD, tau tangles inhibit the transport of β-amyloid precursor protein APP , a protein that stabilizes FPN1, to the cell membrane, and in turn, this leads to intracellular iron accumulation and oxidative stress-induced cell death, including ferroptosis [ ].

Iron accumulation and lipid peroxidation occur in the SN, where the death of melanized neurons is the most severe degeneration event in PD, and is associated with more rapid PD progression [ ]. Furthermore, ferroptosis has been observed to be involved in the death of dopaminergic neurons, another severe degeneration event that occurs in PD [ ].

DFP and ferrostatin-1 treatments were demonstrated to be beneficial to PD by chelating iron and inhibiting ferroptosis, respectively [ , ]. Similar to AD, tau tangles in PD also inhibit the transport of APP and, in turn, they decrease the stabilization of FPN1, leading to intracellular iron accumulation and lipid peroxidation in dopaminergic neurons [ , ].

Moreover, iron deposition in central nervous system is observed in HD and is associated disease progression [ ]. Ferroptosis-related features have also been observed in patients with HD [ ]. Marked lipid peroxidation has been observed in the striatal neurons of an HD mouse model [ ], and a significant decrease in GSH level was observed in a HD rat model [ ].

Taken together, excess iron in both the cytosol and mitochondria significantly contributes to the development of ferroptosis in neurodegenerative diseases, and this occurs primarily through its promotion of oxidative stress and lipid peroxidation.

The inhibited formation of the FPN1-APP complex is the key mechanism driving iron accumulation in neurodegenerative diseases. Although neurodegenerative diseases are associated with disturbed mitochondrial function, further investigation is required to determine whether impaired mitochondria are directly involved in ferroptosis.

In the current review, we recapitulate the intricate functions of iron governing energy metabolism and modulating ferroptosis at the cellular and subcellular levels in the context of both physiology and pathology.

Thus far, a wealth of insights has been obtained to understand the complex regulatory networks that modulate the balanced energy supply and toxic effects of excess iron on cells. In contrast, deregulated regulatory networks would give rise to iron-dependent disorders including ferroptosis and closely implicated diseases, such as cancers, diabetes, and neurodegenerative diseases.

Regardless, puzzling questions and knowledge gaps still exist as follows:. As iron plays a key role in promoting ferroptosis, the rationale underlying the rapid involvement of iron and also the amount of iron in LIP that is sufficient for this process remained to be explored.

Specifically, little is known regarding the threshold regulation of iron availability. Furthermore, whether diseases related to disordered iron homeostasis are prone to ferroptosis in certain cells remains to be tested.

The difficulty in directly measuring LIP changes hinders the establishment of an iron threshold for ferroptosis. Recently, a new fluorescence resonance energy transfer iron probe FRET Iron Probe 1, FIP-1 was designed [ ] and could be used in the future to define the LIP threshold for ferroptosis.

As discussed above, disordered copper metabolism induces iron deposition and ferroptosis in oligodendrocytes. Moreover, copper has also been reported to potentiate both GSH loss and nerve cell death [ ], thus posing a question regarding the substitutes of iron in inducing ferroptosis.

Although mitochondria have been implicated in ferroptosis, current reports remain debatable, and further efforts are warranted to elucidate the biochemical reactions related to ferroptosis in different contexts and in different cell types.

Moreover, iron-mediated electron transfer, OXPHOS, and energy production converge within the mitochondria warrants to be investigated. Accordingly, it would be of great interest to untangle the unbalanced sites responsible for promoting ferroptosis. The imbalance in energy metabolism is closely associated with the occurrence of ferroptosis; however, research in this field is in its infancy, and with several questions concerning dysregulated glucose, lipid, and protein metabolism remains unresolved.

For example, the current literature detailing the role of AMPK, the central sensor in response to the cellular energy state, remains controversial. This discrepancy may be ascribed to tissue specificity. Additionally, different cancers may use distinct energy sources for their major energy supply, thus indicating that their metabolic processes are different, leading to different ferroptotic signaling pathways.

Additional work is required to determine the intricate nexus responsible for coordinating energy production and preventing ferroptosis. The implications of ferroptosis and iron dysregulation in metabolic disorders have not been clearly defined.

For example, diabetes is often observed following iron overload, such as in patients with hereditary hemochromatosis. In the real context, the co-existence of different RCD forms appears to be more frequent under physiological and particularly pathological conditions.

Thus, it is important to investigate the connections among them e. This would help to provide a combination therapy and solve the issues of drug resistance. However, one form of RCD would dominate other forms at a specific disease stage. How this is mediated is an interesting question.

Importantly, whether a superior regulatory network coordinates these forms of RCD warrants further investigation. In this regard, iron and disordered iron homeostasis may provide a breakthrough to elaborate the interplay of ferroptosis with other iron-coupled RCD forms.

Further information on research design is available in the Nature Research Reporting Summary linked to this article. Evstatiev R, Gasche C. Iron sensing and signalling. Article CAS PubMed Google Scholar. Dixon SJ, Stockwell BR.

The role of iron and reactive oxygen species in cell death. Nat Chem Biol. Doll S, Conrad M. Iron and ferroptosis: A still ill-defined liaison. IUBMB Life. Dev S, Babitt JL. Overview of iron metabolism in health and disease. Hemodial Int.

Article PubMed PubMed Central Google Scholar. Dixon Scott J, Lemberg Kathryn M, Lamprecht Michael R, Skouta R, Zaitsev Eleina M, Gleason, et al. Ferroptosis: An iron-dependent form of nonapoptotic. Cell Death Cell. CAS PubMed Google Scholar. Hirschhorn T, Stockwell BR.

The development of the concept of ferroptosis. Free Radic Biol Med. Gunshin H, Mackenzie B, Berger UV, Gunshin Y, Romero MF, Boron WF, et al.

Cloning and characterization of a mammalian proton-coupled metal-ion transporter. Wang C-Y, Knutson MD. Hepatocyte divalent metal-ion transporter-1 is dispensable for hepatic iron accumulation and non-transferrin-bound iron uptake in mice.

Drakesmith H, Nemeth E, Ganz T. Ironing out ferroportin. Cell Metab. Article CAS PubMed PubMed Central Google Scholar. Aroun A, Zhong JL, Tyrrell RM, Pourzand C. Iron, oxidative stress and the example of solar ultraviolet A radiation. Photochem Photobiol Sci: Off J Eur Photochem Assoc Eur Soc Photobiol.

Article CAS Google Scholar. Anderson GJ, Frazer DM. Current understanding of iron homeostasis. Am J Clin Nutr. Coffey R, Ganz T. Iron homeostasis: An anthropocentric perspective.

J Biol Chem. Davis MR, Rendina E, Peterson SK, Lucas EA, Smith BJ, Clarke SL. Enhanced expression of lipogenic genes may contribute to hyperglycemia and alterations in plasma lipids in response to dietary iron deficiency.

Genes Nutr. Potashnik R, Kozlovsky N, Ben-Ezra S, Rudich A, Bashan N. Regulation of glucose transport and GLUT-1 expression by iron chelators in muscle cells in culture.

Am J Physiol-Endocrinol Metab. Jahng JWS, Alsaadi RM, Palanivel R, Song E, Hipolito VEB, Sung HK, et al. Iron overload inhibits late stage autophagic flux leading to insulin resistance. EMBO Rep. Huang J, Simcox J, Mitchell TC, Jones D, Cox J, Luo B, et al.

Iron regulates glucose homeostasis in liver and muscle via AMP-activated protein kinase in mice. FASEB J. Backe MB, Moen IW, Ellervik C, Hansen JB, Mandrup-Poulsen T.

Iron regulation of pancreatic beta-cell functions and oxidative stress. Annu Rev Nutr. Weis S, Carlos AR, Moita MR, Singh S, Blankenhaus B, Cardoso S, et al. Metabolic adaptation establishes disease tolerance to sepsis. Amengual J, Petrov P, Bonet ML, Ribot J, Palou A.

Induction of carnitine palmitoyl transferase 1 and fatty acid oxidation by retinoic acid in HepG2 cells. Int J Biochem Cell Biol. Hay SM, McArdle HJ, Hayes HE, Stevens VJ, Rees WD. The effect of iron deficiency on the temporal changes in the expression of genes associated with fat metabolism in the pregnant rat.

Physiol Rep. Radi R. Oxygen radicals, nitric oxide, and peroxynitrite: Redox pathways in molecular medicine. Proc Natl Acad Sci USA. Pereira M, Chen T-D, Buang N, Olona A, Ko J-H, Prendecki M, et al.

Acute iron deprivation reprograms human macrophage metabolism and reduces inflammation in vivo. Cell Rep. De Bortoli M, Taverna E, Maffioli E, Casalini P, Crisafi F, Kumar V, et al.

Lipid accumulation in human breast cancer cells injured by iron depletors. J Exp Clin Cancer Res: CR. Sherman AR. Lipogenesis in iron-deficient adult rats. Ananda Rao G, Crane RT, Larkin EC. Reduction of hepatic stearoyl-CoA desaturase activity in rats fed iron-deficient diets.

Article Google Scholar. Stangl GI, Kirchgeßner M. Effect of different degrees of moderate iron deficiency on the activities of tricarboxylic acid cycle enzymes, and the cytochrome oxidase, and the iron, copper, and zinc concentration in rat tissues.

Z für Ernährungswissenschaft. Myllyharju J. Prolyl 4-hydroxylases, the key enzymes of collagen biosynthesis. Matrix Biol.

Buongiorno D, Straganz GD. Structure and function of atypically coordinated enzymatic mononuclear non-heme-Fe II centers.

Coord Chem Rev. Yu Q, Tai Y-Y, Tang Y, Zhao J, Negi V, Culley MK, et al. BOLA BolA Family Member 3 deficiency controls endothelial metabolism and glycine homeostasis in pulmonary hypertension. Góger S, Bogáth D, Baráth G, Simaan AJ, Speier G, Kaizer J. Bio-inspired amino acid oxidation by a non-heme iron catalyst.

J Inorg Biochem. Article PubMed Google Scholar. Stadtman ER, Berlett BS. Fenton chemistry. Amino acid oxidation. Stiban J, So M, Kaguni LS. Beinert H, Kennedy MC. Aconitase, a two-faced protein: enzyme and iron regulatory factor Telser J, Volani C, Hilbe R, Seifert M, Brigo N, Paglia G, et al.

Metabolic reprogramming of Salmonella infected macrophages and its modulation by iron availability and the mTOR pathway. Microb Cell. Lim SC, Friemel M, Marum JE, Tucker EJ, Bruno DL, Riley LG, et al. Hum Mol Genet. Huang J, Jones D, Luo B, Sanderson M, Soto J, Abel ED, et al.

Iron overload and diabetes risk: a shift from glucose to fatty acid oxidation and increased hepatic glucose production in a mouse model of hereditary hemochromatosis. Dziegala M, Josiak K, Kasztura M, Kobak K, von Haehling S, Banasiak W, et al. Iron deficiency as energetic insult to skeletal muscle in chronic diseases.

J Cachexia, Sarcopenia Muscle. Dolma S, Lessnick SL, Hahn WC, Stockwell BR. Identification of genotype-selective antitumor agents using synthetic lethal chemical screening in engineered human tumor cells.

Cancer Cell. Yagoda N, von Rechenberg M, Zaganjor E, Bauer AJ, Yang WS, Fridman DJ, et al. RAS—RAF—MEK-dependent oxidative cell death involving voltage-dependent anion channels. Yang WS, Stockwell BR. Synthetic lethal screening identifies compounds activating iron-dependent, nonapoptotic cell death in oncogenic-RAS-harboring cancer cells.

Chem Biol. Stockwell B, Pedro J, Friedmann Angeli JP, Bayir H, Bush A, Conrad M, et al. Ferroptosis: A regulated cell death nexus linking metabolism, redox biology, and disease. Hou W, Xie Y, Song X, Sun X, Lotze MT, Zeh HJ, et al.

Autophagy promotes ferroptosis by degradation of ferritin. Li N, Wang W, Zhou H, Wu Q, Duan M, Liu C, et al. Ferritinophagy-mediated ferroptosis is involved in sepsis-induced cardiac injury. Tang M, Huang Z, Luo X, Liu M, Wang L, Qi Z, et al.

Ferritinophagy activation and sideroflexin1-dependent mitochondria iron overload is involved in apelininduced cardiomyocytes hypertrophy. Wolff NA, Garrick MD, Zhao L, Garrick LM, Ghio AJ, Thévenod F. A role for divalent metal transporter DMT1 in mitochondrial uptake of iron and manganese.

Sci Rep. Han C, Liu Y, Dai R, Ismail N, Su W, Li B. Ferroptosis and its potential role in human diseases. Front Pharmacol. Dixon SJ, Patel DN, Welsch M, Skouta R, Lee ED, Hayano M, et al.

Pharmacological inhibition of cystine—glutamate exchange induces endoplasmic reticulum stress and ferroptosis. Bhutia YD, Ganapathy V. Glutamine transporters in mammalian cells and their functions in physiology and cancer. Biochim Biophys Acta BBA - Mol Cell Res.

Ingold I, Berndt C, Schmitt S, Doll S, Poschmann G, Buday K, et al. Selenium utilization by GPX4 is required to prevent hydroperoxide-induced ferroptosis.

Sato M, Kusumi R, Hamashima S, Kobayashi S, Sasaki S, Komiyama Y, et al. Doll S, Freitas FP, Shah R, Aldrovandi M, da Silva MC, Ingold I, et al. FSP1 is a glutathione-independent ferroptosis suppressor.

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. Mao C, Liu X, Zhang Y, Lei G, Yan Y, Lee H, et al. DHODH-mediated ferroptosis defence is a targetable vulnerability in cancer.

Doll S, Proneth B, Tyurina YY, Panzilius E, Kobayashi S, Ingold I, et al. ACSL4 dictates ferroptosis sensitivity by shaping cellular lipid composition. Jeong SY, David S. Age-related changes in iron homeostasis and cell death in the cerebellum of ceruloplasmin-deficient mice.

J Neurosci. Schulz K, Vulpe CD, Harris LZ, David S. Iron efflux from oligodendrocytes is differentially regulated in gray and white matter.

Jhelum P, Santos-Nogueira E, Teo W, Haumont A, Lenoël I, Stys PK, et al. Ferroptosis mediates cuprizone-induced loss of oligodendrocytes and demyelination.

Brütsch SH, Wang CC, Li L, Stender H, Neziroglu N, Richter C, et al. Expression of inactive glutathione peroxidase 4 leads to embryonic lethality, and inactivation of the Alox15 gene does not rescue such knock-in mice. Antioxid Redox Signal.

Matsushita M, Freigang S, Schneider C, Conrad M, Bornkamm GW, Kopf M. T cell lipid peroxidation induces ferroptosis and prevents immunity to infection. J Exp Med. Seiler A, Schneider M, Förster H, Roth S, Wirth EK, Culmsee C, et al. 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. Nat Cell Biol. Altamura S, Vegi N, Hoppe P, Schroeder T, Aichler M, Walch A, et al. Glutathione peroxidase 4 and vitamin E control reticulocyte maturation, stress erythropoiesis and iron homeostasis.

Google Scholar. Kain HS, Glennon EKK, Vijayan K, Arang N, Douglass AN, Fortin CL, et al. Liver stage malaria infection is controlled by host regulators of lipid peroxidation. Cell Death Differ. Dangol S, Chen Y, Hwang BK, Jwa N-S. Iron- and reactive oxygen species-dependent ferroptotic cell death in rice-magnaporthe oryzae interactions.

Plant Cell. Jiang L, Kon N, Li T, Wang S-J, Su T, Hibshoosh H, et al. Ferroptosis as a pmediated activity during tumour suppression.

Wang W, Green M, Choi JE, Gijón M, Kennedy PD, Johnson JK, et al. Jenkins NL, James SA, Salim A, Sumardy F, Speed TP, Conrad M, et al. Changes in ferrous iron and glutathione promote ferroptosis and frailty in aging Caenorhabditis elegans.

Solar and wind energies lie at Weight loss advice foundation. On Iron in energy generation come with their problems: at Iroh they produce too much, at gdneration too little energy. We need energy storage, in order to use the excess and to bridge periods of low supply. A new option for such an energy storage: iron powder. An option that comes with its strong and it weak points. Producing a lot of heat. Iron in energy generation

Author: Dozuru

1 thoughts on “Iron in energy generation

Leave a comment

Yours email will be published. Important fields a marked *

Design by ThemesDNA.com