Abstract
CD44 is a transmembrane glycoprotein linked to various biological processes reliant on epigenetic plasticity, including development, inflammation, immune responses, wound healing and cancer progression. Functional regulatory roles of this so-called ‘cell surface marker’ remain elusive. Here, we report the discovery that CD44 mediates endocytosis of iron-bound hyaluronates in tumorigenic cell lines, primary cancer cells and tumors. This glycan-mediated iron endocytosis mechanism is enhanced during epithelial-mesenchymal transitions, where iron operates as a metal catalyst to demethylate repressive histone marks that govern the expression of mesenchymal genes. CD44 itself is transcriptionally regulated by nuclear iron, demonstrating a positive feedback loop, which is in contrast to the negative regulation of transferrin receptor by excess iron. Finally, we show that epigenetic plasticity can be altered by interfering with iron homeostasis using small molecules. This study reveals an alternative iron uptake mechanism that prevails in the mesenchymal state of cells, illuminating a central role of iron as a rate-limiting regulator of epigenetic plasticity.
CD44 is a transmembrane glycoprotein implicated in many physiological and pathological processes including development, inflammation, immune responses, wound healing and cancer progression 1,2 . These processes involve cells able to dynamically and reversibly shift between cell states through epithelial-mesenchymal transitions (EMT) 3–5 . For instance, epithelial-mesenchymal plasticity can give rise to cells exhibiting elevated levels of CD44 across lineages in cancers and healthy tissues. Furthermore, cancer cells of a mesenchymal origin, cancer associated fibroblasts and activated T-cells involved in immune responses can also display CD44 at the plasma membrane. In the context of cancer, diseased mesenchymal cells contribute to drug resistance 4 , and CD44 itself has been shown to confer metastatic potential to these cells 6 . However, molecular mechanisms underlying a functional role of CD44 are not fully understood. CD44 has been proposed to act as a ligand-binding surface protein regulating cell adhesion and migration through the extracellular matrix, and as a membrane co-receptor mediating intracellular signal transduction 1 . Importantly, CD44 interacts with and mediates endocytosis of hyaluronates (Hyal) 7,8 , a class of glycosaminoglycan biopolymers that can promote EMT and tumor development 9–11 .
Hyal exhibit negatively charged carboxylates on every disaccharide unit. This raises the question of what counterions these biopolymers interact with at physiological pH during endocytosis, to enable the uptake of Hyal without altering the intracellular ionic balance and membrane potentials. The ability of Hyal to interact with metals including iron (Fe) 12 , together with the iron dependency of mesenchymal cancer cells in vivo 13–16 and their vulnerability to ferroptotic cell death 15,17,18 , prompted us to search for a putative CD44-mediated iron uptake mechanism. Here, we report the discovery of a hyaluronate-dependent iron endocytosis pathway mediated by CD44, which is upregulated during EMT. We found that increase of iron uptake is required for the activity of iron-dependent demethylases regulating the expression of specific genes in the mesenchymal state of cells. This work illuminates a central role of iron in the regulation of epigenetic plasticity, functionally linking the membrane glycoprotein CD44 to the nucleus.
Results
CD44 mediates hyaluronate-dependent iron endocytosis
We first employed nuclear magnetic resonance (NMR) spectroscopy to study interactions between Hyal and iron. Addition of Fe(III) to a low-molecular-mass (LMM) Hyal in water led to line broadening of proton signals of LMM-Hyal, indicating that this organic substrate interacts with iron at neutral pH (Fig. 1a). Acidifying the sample to protonate the carboxylates of LMM-Hyal and to disrupt organometallic complexes released the metal, thereby rescuing the proton signals of free LMM-Hyal (Fig. 1a). These data demonstrate the dynamic and reversible nature of metal coordination by LMM-Hyal in near-physiological conditions of solvent, temperature and pH.
Figure 1|. CD44 mediates hyaluronate-dependent iron endocytosis.
a, Molecular structure (top) and 1H NMR spectra of LMM-Hyal (bottom). Functional groups susceptible to reversibly interact with iron are highlighted in red. TFA, trifluoroacetic acid. b, Western blot analysis of CD44 and TfR1 levels in CD44 and TFRC ko clones generated using CRISPR-Cas9. c, Fluorescence microscopy analysis of internalized Hyal-FITC. N = 3 biologically independent experiments. d, Fluorescence microscopy analysis of RhoNox-M-positive vesicles colocalizing with internalized Hyal-FITC or TF-647. Dotted lines delineate the cell contours. N = 3 biologically independent experiments. e, Fluorescence microscopy analysis of RhoNox-M-positive vesicles colocalizing with Hyal-FITC in a CD44 ko clone complemented with CD44. N = 3 biologically independent experiments. f, SIMS imaging of cellular iron in TFRC ko cells. Yellow lines delineate the cell contours. N = 1. At least 50 cells were quantified per condition. FAC, ferric ammonium citrate. g, ICP-MS measurements of cellular iron. N = 5 biologically independent experiments. h, Western blot analysis of ferritin levels representative of 8 biologically independent experiments. i, Schematic illustration of CD44high/low cell sorting from a PDX by flow cytometry. Western blot analysis of iron homeostasis proteins in CD44high/low cells sorted from breast cancer tumors. N = 1. Parental HMLER CD44high and corresponding ko cell lines were used throughout panels b-h. Scale bars, 10 μm. DAPI, 4',6-diamidino-2-phenylindole; Unt., untreated. HMM-Hyal (0.6-1 MDa) was used in f-h. Data representative of three independent experiments unless stated otherwise. Bars and error bars, mean values +/− SD. One-way ANOVA with Bonferroni correction for c, d, e. Unpaired t-test for g, h. Statistical tests throughout the figure two-sided. Full western blots see Supplementary Information.
To evaluate the potential of high-molecular-mass (HMM) Hyal-based organometallic complexes to be internalized by CD44, we knocked out (ko) CD44 from mesenchymal human mammary HMLER cells stably expressing high levels of CD44 (HMLER CD44high) 19 (Fig. 1b). For comparison, we independently knocked out TFRC, the gene coding for transferrin receptor protein 1 (TfR1) (Fig. 1b), which mediates endocytosis of transferrin (TF) and plays a central role in the regulation of cellular iron homeostasis.
We then evaluated the capacity of parental cells and ko clones to take up fluorescently labeled HMM-Hyal (0.8 MDa). Internalization of Hyal was reduced in CD44 ko and increased in TFRC ko clones compared to parental cells (Fig. 1c), validating the functional role of CD44 in mediating this process. Hyal and TF, which interacts with CD44 and TfR1, respectively, predominantly localized in distinct vesicles in parental cells (Fig. 1d), allowing for a comparative analysis of metal uptake mediated either by CD44 or TfR1. Thus, we evaluated whether endocytic vesicles contained Fe(II) and Hyal, independent of TF. To this end, we used the Fe(II)-specific turn-on fluorescent probe RhoNox-M, which can detect the presence of Fe(II) in lysosomes 20 . Parental HMLER cells exhibited RhoNox-M/Hyal-positive vesicles free of TF as observed by fluorescence microscopy. CD44 ko cells exhibited a lower proportion of RhoNox-M/Hyal-positive vesicles with an increased number of RhoNox-M/TF-positive vesicles compared to parental cells, while TFRC ko cells showed the opposite trend (Fig. 1d). Consistently, complementing CD44 ko cells with CD44 restored a higher proportion of RhoNox-M/Hyal-positive vesicles (Fig. 1e). RhoNox-M/Hyal-positive vesicles were also observed in other cancer cell lines, primary cancer cells including circulating tumor cells (CTC) of lung and colon cancers and in primary human T-cells (Extended Data Fig. 1a,b). Consistently, LNCaP prostate cancer cells, which do not show detectable levels of CD44 protein, exhibited a low proportion of RhoNox-M/Hyal-positive vesicles (Extended Data Fig. 1a).
In line with CD44 mediating iron endocytosis, supplementing cells with HMM-Hyal (0.6-1 MDa) increased levels of cellular iron in parental CD44high and TFRC ko but not in CD44 ko HMLER and MCF7 cells as defined by the fluorescence of RhoNox-M quantified by flow cytometry (Extended Data Fig. 2a). This was observed across cell types (Extended Data Fig. 2a,b), indicating this to be a general feature of cells expressing CD44 protein. In CD44-expressing MDA-MB-468 breast cancer cells, iron uptake was more pronounced when cells were treated with Hyal of higher molecular mass (Extended Data Fig. 2c). Conversely, treating CD44-expressing cells with hyaluronidase or blocking antibodies against CD44 or TfR1 reduced iron uptake (Extended Data Fig. 2d,e). Additionally, supplementing HMLER CD44high cells with Hyal increased cellular iron as detected by secondary ion mass spectrometry (SIMS) imaging 21 and quantified by inductively coupled plasma mass spectrometry (ICP-MS) (Fig. 1f,g). Supplementing HMLER CD44high cells with Hyal also led to increased levels of the iron storage protein ferritin in a CD44-dependent manner (Fig. 1h), further demonstrating increased iron uptake in these conditions.
To evaluate the physiological significance of this pathway, we sorted CD44high/low cell populations from two patient-derived xenografts (PDX) of breast tumors refractory to conventional chemotherapy, prone to form metastasis and associated with poor outcome. We found that CD44high cells contained higher levels of ferritin with reduced levels of TfR1 compared to the CD44low cell population (Fig. 1i). Collectively, these data provide evidence that CD44 mediates iron endocytosis in a Hyal-dependent manner.
Prevalence of CD44-mediated iron endocytosis in the mesenchymal state of cells
The mesenchymal state of cells is characterized by elevated levels of CD44 in several lineages 22 . To investigate the cellular changes reliant on CD44 levels, we triggered EMT in triple-negative breast cancer MDA-MB-468, luminal breast cancer MCF7 and HMLER CD44low cell lines, primary breast cancer cells and primary lung CTC using epidermal growth factor (EGF), oncostatin M (OSM) or transforming growth factor beta 1 (TGF-β). Upon treatment, cells shifted from an epithelial to a mesenchymal phenotype according to cell morphology, reduced levels of the epithelial marker E-cadherin and increased levels of CD44 (Fig. 2a). Consistent with EMT induction, levels of the transcription factor snail increased along with that of the mesenchymal markers vimentin and fibronectin (Fig. 2b). Remarkably, CD44 levels increased together with ferritin, whereas levels of TfR1 decreased at a later time point (Fig. 2b). In line with this, CD44 loading at the plasma membrane was enhanced during EMT, which was in contrast to a reduced loading of TfR1 (Fig. 2c and Extended Data Fig. 3a), and the number of RhoNox-M/Hyal-positive vesicles increased together with cellular iron (Fig. 2d,e and Extended Data Fig. 3b). Importantly, knocking down CD44 antagonized the production of ferritin in cells treated with EGF (Fig. 2f).
Figure 2|. CD44-mediated iron endocytosis prevails in the mesenchymal state of cells.
a, Fluorescence microscopy analysis of EMT markers (scale bar, 10 μm) and bright field microscopy analysis of cell morphology (scale bar, 100 μm). b, Time course western blot analysis of levels of iron homeostasis and EMT markers. c, Time course flow cytometry analysis of CD44 and TfR1 loading at the plasma membrane. d, Fluorescence microscopy analysis of RhoNox-M-positive vesicles colocalizing with internalized Hyal-FITC or TF-647. Scale bar, 10 μm. N = 3 biologically independent experiments. e, ICP-MS measurements of cellular iron. N = 11 biologically independent experiments. f, Western blot analysis of ferritin levels in CD44 knock down conditions. g, Time course RT-qPCR analysis of CD44, TFRC and FTH1 mRNA levels. N = 3 biologically independent experiments. MDA-MB-468 cells were used throughout the figure and were treated with EGF for 72 h, unless stated otherwise. Data representative of three independent experiments unless stated otherwise. Bars and error bars, mean values +/− SD. Unpaired t-test for d, e. One-way ANOVA with Bonferroni correction for g. Statistical tests throughout the figure two-sided. Full western blots see Supplementary Information.
Furthermore, RT-qPCR showed that mRNA levels of CD44, and FTH1 coding for ferritin, increased during EMT, whereas that of TFRC decreased (Fig. 2g). This data was in agreement with the well-established negative feedback loop regulating the biosynthesis of TfR1 at the translational level by excess cytosolic iron 23,24 . It also indicated that in contrast to TFRC, CD44 is not subjected to this regulatory loop. Together, these results advocate for an alternative glycan-mediated iron endocytosis pathway reliant on CD44 that prevails in the mesenchymal state of cells.
EMT is characterized by a redox signature implicating iron
To identify iron-dependent cellular functions required during EMT or for the maintenance of mesenchymal states of cells, we employed a combination of quantitative proteomics and metabolomics. Out of 5,574 proteins that were detected by mass spectrometry, 255 proteins were downregulated, while 276 were upregulated in EGF-treated MDA-MB-468 cells (Fig. 3a, and Supplementary Table 1). Gene ontology (GO) revealed a bias towards proteins with oxidoreductase activity having an increased expression in the mesenchymal state (Extended Data Fig. 4a). The response to EGF was further characterized by changes in EMT markers, proteins linked to metastasis, proteins involved in the regulation of endocytosis and iron homeostasis as well as metabolic enzymes and an iron-dependent demethylase. For example, EMT was characterized by reduced levels of the epithelial proteins BCAM and CADM4, whereas vimentin was upregulated (Fig. 3a). The pro-metastatic protein CD109 was upregulated 25 , which was in agreement with the notion that mesenchymal cancer cells are capable of dissemination. Consistent with increased iron uptake during EMT, levels of ferritin increased together with sorting nexin 9 (SNX9), a protein that regulates endocytosis of CD44, which has also been linked to metastasis 26 . Conversely, levels of TF and TfR1 were reduced, further supporting the prevalent functional role of CD44 in mediating iron endocytosis during EMT. Importantly, we detected an increase of the iron-dependent histone demethylase PHF8, which is involved in active transcription, brain development, cell cycle progression as well as the regulation of EMT in vivo and in breast tumor growth 27–31 . In contrast, levels of other iron-dependent demethylases that could be detected by mass spectrometry did not change significantly. Western blotting further supported protein changes identified by mass spectrometry (Extended Data Fig. 4b).
Figure 3|. EMT is characterized by a redox signature implicating iron.
a, Quantitative label-free proteomics. Volcano plot representing protein expression fold changes. Upregulated (right) and downregulated (left) proteins are shown. Dashed blue line defines a fold change of 1.42, continuous blue line defines a P-value of 0.05. The mitochondrial iron-sulfur cluster-containing protein ACO2 and the nuclear iron-dependent demethylase PHF8 are highlighted in red. N = 3 biologically independent experiments. b, Schematic illustration of metabolic pathways involved in the production of αKG. Enzymes that are upregulated at the protein level during EMT are highlighted in red. c, Quantitative metabolomics. Heatmap of upregulated (red) and downregulated (green) metabolites in cells treated with EGF for 60 h. N = 4 technical replicates. d, Quantification of αKG levels in cells treated as indicated using an αKG assay. Cells were treated as indicated for 72 h. N = 3 technical replicates. e, Quantification of αKG levels in CD44 knock down conditions using an αKG assay. N = 3 biologically independent experiments. MDA-MB-468 cells were used in a, b, c, e and were treated with EGF for 72 h, unless stated otherwise. Bars and error bars, mean values +/− SD. Statistical analysis for a see Supplementary Information. Unpaired t-test for d. One-way ANOVA with Bonferroni correction for e. Statistical tests throughout the figure two-sided.
Additionally, metabolic enzymes involved in the conversion of glutamate and pyruvate to α-ketoglutarate (αKG), including glutamate dehydrogenase 1 (GLUD1), isocitrate dehydrogenase 2 (IDH2) and the iron-sulfur cluster-containing aconitase 2 (ACO2), were upregulated along with other enzymes of the Krebs cycle (Fig. 3a,b). αKG is a co-substrate required for the iron-catalyzed oxidative demethylation of histone and nucleic acid methyl marks, and is implicated in the maintenance of pluripotency 32 . Consistent with the upregulation of these mitochondrial enzymes, quantitative metabolomics indicated a marked increase of αKG in various cell lines and primary cells undergoing EMT (Fig. 3c,d, Extended Data Fig. 4c and Supplementary Table 2). We then investigated whether reducing iron uptake could potentially alter mitochondrial metabolism in cells treated with EGF. Knocking down CD44 antagonized the upregulation of αKG in cells treated with EGF (Fig. 3e), which was consistent with previous findings showing that iron restriction inhibits the activity of ACO2 33 , supporting a role of iron in mitochondria during EMT.
These biochemical changes, including enhanced cellular uptake of iron together with the upregulation of PHF8 and its co-substrate αKG, further hinted toward a specific regulation of EMT occurring at the chromatin level 34 .
Nuclear iron is a rate-limiting regulator of epigenetic plasticity
Metals are ubiquitous in cell biology, acting as protein cofactors. Due to its unique electron configuration, iron distinguishes itself by its capacity to catalyze oxidative demethylation of protein residues, including those of histones, and methylated nucleobases. Thus, iron is a rate-limiting factor of these processes (Fig. 4a). In line with a role of iron in the nucleus of cells undergoing EMT, subcellular fractionation indicated higher levels of ferritin in the nucleus of MDA-MB-468 cells undergoing EMT (Fig. 4b) 35 .
Figure 4|. Nuclear iron is a rate-limiting regulator of epigenetic plasticity.
a, Catalytic cycle of iron-mediated oxidative demethylation of a methylated lysine residue. b, Subcellular fractionation and western blot analysis of nuclear ferritin levels. NE, nuclear extract; CE, cytoplasmic extract; WCE, whole cell extract. c, Quantitative mass spectrometry. Heatmap of H3K9me2, H3K9me3 and H3K27me3 levels. N = 5 biologically independent experiments. d, Western blot analysis of H3K9me2 levels in PHF8 knock down conditions. e, H3K9me2 ChIP-seq profiles for selected genes. f, Scatter plot correlation of H3K9me2 ChIP-seq reads count in genes (N = 3 biologically independent experiments) and RNA-seq (N = 3 biologically independent experiments). g, Western blot analysis of proteins whose genes are regulated by H3K9me2 in PHF8 knock down conditions. h, Western blot analysis of H3K9me2 levels in CD44 knock down conditions. i, Western blot analysis of proteins whose genes are regulated by H3K9me2 in CD44 knock down conditions. j, Western blot analysis of fibronectin, PHF8 and H3K9me2 in CD44high/low cells sorted from breast cancer tumors. N = 1. MDA-MB-468 cells were used for b-i and treated with EGF for 72 h. Data representative of three independent experiments unless stated otherwise. Two-sided associated t-tests for c. Statistical analysis for f see Supplementary Information. Full western blots see Supplementary Information.
The iron- and αKG-dependent demethylase PHF8 has been shown to promote demethylation of H3K9me2 36 , a repressive histone mark whose reduction has been observed in solid tumors 37–39 . Increase of PHF8 in MDA-MB-468 cells undergoing EMT correlated with the preferential reduction of H3K9me2 compared to the other repressive histone marks H3K9me3 and H3K27me3, as defined by quantitative mass spectrometry (Fig. 4c and Supplementary Table 3). Additionally, knocking down PHF8 prevented demethylation of H3K9me2 in MDA-MB-468 cells treated with EGF (Fig. 4d), validating a functional role of this demethylase during EMT.
Next, we examined the genome-wide distribution of H3K9me2 by chromatin immunoprecipitation sequencing (ChIP-seq) to identify PHF8-regulated genes whose expressions are iron-dependent. ChIP-seq revealed large organized heterochromatin K9 modifications, previously termed LOCKs 37,39 . These data showed a reduction of H3K9me2 ChIP-seq reads count in the body of specific genes in cells treated with EGF, including CD44, CD109, VIM and FN1 coding for vimentin and fibronectin, respectively (Fig. 4e, Extended Data Fig. 5a and Supplementary Table 4), which was further validated by ChIP-qPCR (Extended Data Fig. 5b). These data indicate that CD44 positively regulates its own expression at the transcriptional level by mediating iron endocytosis, unlike TfR1 that is negatively regulated by excess iron at the translational level via iron responsive elements (IRE) 23,24 . Furthermore, analysis of the transcriptome by RNA-seq indicated that the reduction of H3K9me2 reads count correlated with increased gene expression in agreement with the repressive nature of H3K9me2 (Fig. 4F and Supplementary Table 4). For instance, out of 284 genes where reduction of H3K9me2 was observed, 258 genes were transcriptionally upregulated. In addition to CD44, we identified genes involved in cancer, regulation of EMT, development, immune responses, inflammation, and wound healing, which are biological processes CD44 is linked to (Supplementary Table 5). In comparison, we did not observe a significant reduction of H3K9me2 in genes coding for EMT transcription factors. However, we identified long non-coding (lnc) RNA genes previously reported to be involved in cancer (Supplementary Table 5). In agreement with PHF8 playing a crucial role in regulating these genes through oxidative demethylation of H3K9me2, knocking down this demethylase partly blocked EGF-induced expression of iron-regulated genes, including CD44 (Fig. 4g). Consistently, ChIP-seq indicated that H3K4me3 was enriched at active promoters, which is required for the recruitment and selective activity of PHF8 at these genomic locations (Extended Data Fig. 5c). Furthermore, ChIP-seq also revealed a reduction of the repressive histone mark H3K27me3 in a more restricted set of genes (Extended Data Fig. 5d), which was consistent with increased mRNA levels of the H3K27-specific iron-dependent demethylases KDM6A and KDM6B in cells undergoing EMT as monitored by RNA-seq (Supplementary Table 4). Conversely, we did not observe changes of the heterochromatin mark H3K9me3 (Figures 4c and Extended Data Fig. 5e). While the reduction of H3K27me3 was less pronounced compared to that of H3K9me2 according to mass spectrometry and ChIP-seq reads count (Fig. 4c and Extended Data Fig. 5d), these data support the central role of iron in the regulation of gene expression, reflecting a complex scenario occurring at the chromatin level during EMT, potentially involving other iron-dependent demethylases and histone marks. In line with this, knocking down CD44 partly blocked H3K9me2 demethylation and the expression of genes regulated by PHF8 in MDA-MB-468 cells (Fig. 4h,i). Consistently, CD44high cells from tumors exhibited higher levels of PHF8 and the iron-regulated gene fibronectin with reduced H3K9me2 compared to the CD44low population, supporting the physiological relevance of this mechanism (Fig. 4j). Taken together, these data establish a direct functional link between CD44-mediated iron endocytosis and the regulation of gene expression involving iron-catalyzed histone demethylation.
Targeting iron homeostasis interferes with the maintenance of a mesenchymal state of cells
Small molecules provide a means to manipulate cellular processes with valuable spatial and temporal resolution 40 . To directly address the implication of the nuclear iron pool, we set out to identify small molecule chelators of iron that could specifically target the nucleus. To this end, we functionalized known iron chelators with alkyne functional groups that can be chemically labeled in cells by means of click chemistry (Fig. 5a) 41 . Fluorescent labeling of a clickable surrogate of deferoxamine (cDFO) in cells revealed a predominant nuclear localization of this small molecule (Fig. 5a). Thus, deferoxamine (DFO) represents a suitable tool to directly interrogate functional roles of nuclear iron. For instance, DFO prevented the demethylation of H3K9me2 in cancer cell lines and primary cancer cells that were treated either with EGF, OSM or TGF-β (Fig. 5b and Extended Data Fig. 6a). Consistently, DFO antagonized the upregulation of proteins whose genes were found to be regulated by PHF8 in these conditions (Fig. 5c and Extended Data Fig. 6b). These data provide direct evidence of the functional role of nuclear iron in the regulation of the expression of these genes. Moreover, the iron chelator deferasirox (DFX), whose scaffold exhibits a preference for the mitochondrial compartment as defined by the staining of the clickable derivative cDFX (Extended Data Fig. 6c), blocked αKG production in MDA-MB-468 cells treated with EGF (Extended Data Fig. 6d) and interfered with H3K9me2 demethylation and CD44 expression (Extended Data Fig. 6e). Similarly, metformin, which has been shown to target metals in the mitochondria 42 and to inhibit the Krebs cycle 43 , antagonized the effect of EGF (Fig. 5d,e). These data, consistent with αKG being required for iron-dependent demethylation of H3K9me2, validate a functional role of mitochondrial iron during EMT and show that targeting mitochondrial metabolism provides another means to control epigenetic plasticity.
Figure 5|. Targeting iron homeostasis interferes with the maintenance of mesenchymal cells.
a, Molecular structure of cDFO (top), schematic illustration of the chemical labeling of cDFO in cells (bottom) and fluorescence microscopy images of labeled cDFO and DAPI (right). 488 represents the Alexa-Fluor-488 fluorophore. Scale bar, 10 μm. b, Western blot analysis of H3K9me2 levels in cells cotreated with EGF and DFO. N = 1 for primary cells. c, Western blot analysis of proteins whose genes are regulated by H3K9me2 in cells cotreated with EGF and DFO. N = 1 for primary cells. d, Quantification of αKG using an αKG assay. N = 4 biologically independent experiments. e, Western blot analysis of H3K9me2 levels in cells cotreated with EGF and metformin. f, Cell viability curves of cells cotreated with EGF and ironomycin. N = 4 biologically independent experiments. g, Flow cytometry analysis of C11-BODIPY treated as indicated. N = 1. Data representative of three independent experiments unless stated otherwise. Bars and error bars, mean values +/− SD. One-way ANOVA with Bonferroni correction for d. Two-way ANOVA with Bonferroni correction for f. Statistical tests throughout the figure two-sided. Full western blots see Supplementary Information.
Finally, the lysosomal iron-sequestering small molecule ironomycin has previously been shown to effectively kill mesenchymal and tumor-initiating cells in vitro and in vivo due to a higher iron content in these cells 15 . Consistently, EGF potentiated the cytotoxic effect of ironomycin against MDA-MB-468 cells (Fig. 5f) with increased lipid peroxidation as defined by C11-BODIPY fluorescence that was prevented by co-treatment with liproxtatin-1 (Fig. 5g) 44 . This data is in line with the observed increase of iron endocytosis in these conditions, providing a powerful strategy to eradicate tumorigenic cells undergoing EMT through induction of ferroptosis 45 . Together, these data support a mechanism whereby iron is a rate-limiting regulator of epigenetic plasticity 46 , a general pathway prone to small molecule intervention with topological resolution.
Discussion
This investigation illuminates a unifying mechanism involving CD44, Hyal and iron in the regulation of epigenetic plasticity (Fig. 6). Our findings link iron endocytosis to enhanced mitochondrial metabolism and molecular changes occurring at the chromatin level. The free carboxylates of Hyal, which is a principal ligand of CD44, are prone to interact with iron outside of cells at physiological pH, thereby enabling CD44 to mediate endocytosis of these organometallic complexes. Upon acidification of endocytic vesicles and ensuing protonation of Hyal, iron is released and translocates into the cytosol to traffic towards specific cellular compartments. In particular, increased levels of nuclear ferritin and upregulation of the iron-dependent demethylase PHF8 together with the reduction of the repressive histone mark H3K9me2 in an iron-dependent manner provide solid evidence for a functional role of nuclear iron during EMT. Upregulation of other demethylases and detectable alterations of other histone marks reflect a complex regulation at the chromatin level that is iron-dependent.
Figure 6|. Reciprocal endocytic-epigenetic regulation involving iron.
TF or iron-bound Hyal enters cells by means of TfR1-or CD44-mediated endocytosis. Excess cellular iron inhibits the canonical TfR1 pathway. Nuclear iron, αKG produced in mitochondria and iron-dependent enzymes mediate oxidative demethylation of chromatin marks to unlock the expression of specific genes including CD44. CD44 positively regulates its own expression at the transcriptional level by mediating iron endocytosis and this pathway prevails in the mesenchymal state of cells. Iron homeostasis can be targeted at the plasma membrane, the endosomal/lysosomal compartment, the mitochondria and in the nucleus using specific antibodies or small molecules.
In this context, two independent iron endocytosis pathways co-exist, one reliant on the protein TF and the other on the glycan Hyal (Fig. 6). Interestingly, these biomolecules represent distinct classes of organic metal carriers of different biosynthetic origins. The CD44-dependent pathway involves Hyal and is switched on at the onset of EMT progressively taking over as the prevalent mechanism of iron endocytosis compared to the TfR1-dependent pathway. Importantly, CD44 is an iron-regulated gene, indicating a positive feedback loop whereby CD44 regulates its own expression. In contrast, TfR1 is negatively regulated at the translational level by excess iron. This provides a rationale underlying the global increase of CD44 and the reduction of TfR1 during the course of EMT. In epithelial cells, the iron-demand is moderate making the self-regulated TfR1-dependent pathway sufficient to maintain a basal level of iron. However, the higher needs of iron in the mesenchymal state of cells to unlock the expression of mesenchymal genes cannot be solely fulfilled by this pathway. Thus, upregulation of CD44 triggered by growth factors, cytokines and other signaling molecules represent a powerful alternative to increase the cellular iron load.
In the nucleus, using αKG and taking advantage of demethylases including PHF8, iron operates as a metal catalyst mediating oxidative demethylation of histone residues. Iron is therefore a rate-limiting regulator of the expression of specific genes (Fig. 6). Hence, to unlock these genes, cells concomitantly upregulate the production of demethylases and increase iron uptake. The repressive histone mark H3K9me2, which is a direct substrate of iron and PHF8, was identified as a key post-translational modification regulating the expression of mesenchymal genes. However, other enzymes involved in methylation, acetylation and deacetylation are expected to work in concert with this iron-dependent demethylation of H3K9me2. Our analysis revealed that H3K9me2 governs the expression of genes involved in cancer, development, immune responses, inflammation, and wound healing. Interestingly, CD44 has previously been linked to these processes, consistent with the regulatory role of iron.
Cellular iron homeostasis can be altered at the plasma membrane by interfering with iron endocytosis using specific antibodies, or alternatively by controlling the chemical reactivity of this metal in selected cellular compartments using appropriate small molecules (Fig. 6). These pharmacological tools provide a means to dissect the processes reliant on iron in various settings and can be further developed for therapeutic intervention.
Intriguingly, very high-molecular-mass Hyal (6-12 MDa) confer cancer resistance to naked mole rats and these Hyal have been shown to be refractory to endocytosis due to their larger size compared to that found in other organisms 47 . It is conceivable that such large biopolymers prohibit EMT by sequestering iron outside of cells. Furthermore, while we have shown that Hyal can reversibly interact with iron, other metal ions can potentially be endocytosed using a similar glycosaminoglycan-mediated pathway and may also contribute to the regulation of epigenetic plasticity at different levels 42 . While our study illustrates a functional role of CD44, Hyal and iron in the context of tumorigenic cells, we anticipate that other physiological and pathological processes that rely on the status of distinct histone marks or modified DNA and RNAs, are under the control of similar mechanisms involving iron-dependent demethylases 48–51 .
Methods
All methods can be found in the Supplementary Information under ‘Materials and Methods’ on www.nature.com.
Extended Data
Extended data figure 1.
Extended data figure 2.
Extended data figure 3.
Extended data figure 4.
Extended data figure 5.
Extended data figure 6.
Supplementary Material
Acknowledgments
We thank A. Puisieux, C. Hivroz and S. Dogniaux for providing us with HMLER and primary human T-cells, the PICT-IBiSA@Pasteur Imaging Facility of Institut Curie, member of the France-BioImaging national research infrastructure for the use of microscopes and SIMS, C. Gaillet for assistance with NMR spectroscopy, J.-L. Guerquin-Kern for assistance with SIMS sample preparation, the ICP-MS platform at the Institut de Physique du Globe de Paris, G. Arras for assistance with mass spectrometry data analysis, S. Durand and G. Kroemer for providing access to the metabolomics platform and P. Legoix for NGS sample preparation. We thank R. Vale for fruitful discussions. The R.R. research group is funded by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No [647973]), the Fondation Charles Defforey-Institut de France and Ligue Contre le Cancer (Equipe Labellisee). R.R. and D.L. are supported by Region IdF for NMR and MS infrastructures. The Institut de Physique du Globe de Paris is supported by IPGP multidisciplinary program PARI and Paris-Region IdF (SESAME grant agreement No [12015908]). High-throughput sequencing was performed by the ICGex NGS platform of Institut Curie, supported by ANR-10-EQPX-03 (Equipex), ANR-10-INBS-09-08 (France Genomique Consortium) from the Agence Nationale de la Recherche (Investissements d’Avenir program) and by the Canceropole IdF and the SiRIC-Curie program – SiRIC Grant (INCa-DGOS-4654). Correspondence and requests for materials should be addressed to R.R.
Footnotes
Competing financial interests
The authors declare no competing interests.
Author contributions
R.R. conceptualized the study and directed the research. R.R., S.M. and F.S. designed the experiments. T.C. performed NMR spectroscopy and synthesized the clickable iron chelators. A.V. synthesized the iron(II)-specific fluorescent probes. S.M. produced knock out cell lines and performed the experiments in relation to iron endocytosis including western blotting, cell imaging, RNA interference, flow cytometry and ICP-MS. T.-D.W. performed SIMS imaging. A.L. performed RT-qPCR and subcellular fractionation experiments. F.S. prepared the samples for quantitative proteomics, metabolomics and next generation sequencing. B.L. and D.L. carried out quantitative proteomics. E.C.-J. and C.G. provided tumor samples and performed cell sorting. A.D., C.V. and S.B provided assistance with NGS library preparation. N.S. performed bioinformatics analysis. R.R., S.M., and F.S. interpreted the data and wrote the article.
Data availability
All data are available in the manuscript or the supplementary materials. Mass spectrometry data have been deposited at the ProteomeXchange Consortium (PRIDE Archive) with identifiers PXD011447 and PXD012862.
ChlP-seq and RNA-seq data are available on the National Center for Biotechnology Information website with accession reference GSE121664.
Code availability
Code employed for ChIP-seq and RNA-seq data analyses are available on Github at https://github.com/nservant/EMTiron.
References
- 1.Ponta H, Sherman L, Herrlich PA. CD44: from adhesion molecules to signalling regulators. Nat Rev Mol Cell Biol. 2003;4:33–45. doi: 10.1038/nrm1004. [DOI] [PubMed] [Google Scholar]
- 2.Zöller M. CD44: can a cancer-initiating cell profit from an abundantly expressed molecule? Nat Rev Cancer. 2011;11:254–267. doi: 10.1038/nrc3023. [DOI] [PubMed] [Google Scholar]
- 3.Nieto MA, Huang RY, Jackson RA, Thiery JP. EMT: 2016. Cell. 2016;166:21–45. doi: 10.1016/j.cell.2016.06.028. [DOI] [PubMed] [Google Scholar]
- 4.Brabletz T, Kalluri R, Nieto MA, Weinberg RA. EMT in cancer. Nat Rev Cancer. 2018;18:128134. doi: 10.1038/nrc.2017.118. [DOI] [PubMed] [Google Scholar]
- 5.Pastushenko I, et al. Identification of the tumour transition states occurring during EMT. Nature. 2018;556:463–468. doi: 10.1038/s41586-018-0040-3. [DOI] [PubMed] [Google Scholar]
- 6.Günthert U, et al. A new variant of glycoprotein CD44 confers metastatic potential to rat carcinoma cells. Cell. 1991;65:13–24. doi: 10.1016/0092-8674(91)90403-l. [DOI] [PubMed] [Google Scholar]
- 7.Aruffo A, Stamenkovic I, Melnick M, Underhill CB, Seed B. CD44 is the principal cell surface receptor for hyaluronate. Cell. 1990;61:1303–1313. doi: 10.1016/0092-8674(90)90694-a. [DOI] [PubMed] [Google Scholar]
- 8.Hua Q, Knudson CB, Knudson W. Internalization of hyaluronan by chondrocytes occurs via receptor-mediated endocytosis. J Cell Sci. 1993;106:365–375. doi: 10.1242/jcs.106.1.365. [DOI] [PubMed] [Google Scholar]
- 9.Bartolazzi A, Peach R, Aruffo A, Stamenkovic I. Interaction between CD44 and hyaluronate is directly implicated in the regulation of tumor development. J Exp Med. 1994;180:53–66. doi: 10.1084/jem.180.1.53. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Zoltan-Jones A, Huang L, Ghatak S, Toole BP. Elevated hyaluronan production induces mesenchymal and transformed properties in epithelial cells. J Biol Chem. 2003;278:45801–45810. doi: 10.1074/jbc.M308168200. [DOI] [PubMed] [Google Scholar]
- 11.Toole BP. Hyaluronan: from extracellular glue to pericellular cue. Nat Rev Cancer. 2004;4:528–539. doi: 10.1038/nrc1391. [DOI] [PubMed] [Google Scholar]
- 12.Mercê AL, Marques Carrera LC, Santos Romanholi LK, Lobo Recio MA. Aqueous and solid complexes of iron(III) with hyaluronic acid Potentiometric titrations and infrared spectroscopy studies. J Inorg Biochem. 2002;89:212–218. doi: 10.1016/s0162-0134(01)00422-6. [DOI] [PubMed] [Google Scholar]
- 13.Torti SV, Torti FM. Iron and cancer: more ore to be mined. Nat Rev Cancer. 2013;13:342–355. doi: 10.1038/nrc3495. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Schonberg DL, et al. Preferential iron trafficking characterizes glioblastoma stem-like cells. Cancer Cell. 2015;28:441–455. doi: 10.1016/j.ccell.2015.09.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Mai TT, et al. Salinomycin kills cancer stem cells by sequestering iron in lysosomes. Nat Chem. 2017;9:1025–1033. doi: 10.1038/nchem.2778. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Basuli D, et al. Iron addiction: a novel therapeutic target in ovarian cancer. Oncogene. 2017;36:4089–4099. doi: 10.1038/onc.2017.11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Viswanathan VS, et al. Dependency of a therapy-resistant state of cancer cells on a lipid peroxidase pathway. Nature. 2017;547:453–457. doi: 10.1038/nature23007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Hangauer MJ, et al. Drug-tolerant persister cancer cells are vulnerable to GPX4 inhibition. Nature. 2017;551:247–250. doi: 10.1038/nature24297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Morel AP, et al. Generation of breast cancer stem cells through epithelial-mesenchymal transition. PLoS One. 2008;3:e2888. doi: 10.1371/journal.pone.0002888. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Niwa M, Hirayama T, Okuda K, Nagasawa H. A new class of high-contrast Fe(II) selective fluorescent probes based on spirocyclized scaffolds for visualization of intracellular labile iron delivered by transferrin. Org Biomol Chem. 2014;12:6590–6597. doi: 10.1039/c4ob00935e. [DOI] [PubMed] [Google Scholar]
- 21.Guerquin-Kern JL, Wu TD, Quintana C, Croisy A. Progress in analytical imaging of the cell by dynamic secondary ion mass spectrometry (SIMS microscopy) Biochim Biophys Acta. 2005;1724:228238. doi: 10.1016/j.bbagen.2005.05.013. [DOI] [PubMed] [Google Scholar]
- 22.Stuelten CH, et al. Complex display of putative tumor stem cell markers in the NCI60 tumor cell line panel. Stem Cells. 2010;28:649–660. doi: 10.1002/stem.324. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Pantopoulos K, Porwal SK, Tartakoff A, Devireddy L. Mechanisms of mammalian iron homeostasis. Biochemistry. 2012;51:5705–5724. doi: 10.1021/bi300752r. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Muckenthaler MU, Rivella S, Hentze MW, Galy B. A red carpet for iron metabolism. Cell. 2017;168:344–361. doi: 10.1016/j.cell.2016.12.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Chuang CH, et al. Molecular definition of a metastatic lung cancer state reveals a targetable CD109-Janus kinase-Stat axis. Nat Med. 2017;23:291–300. doi: 10.1038/nm.4285. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Bendris N, et al. SNX9 promotes metastasis by enhancing cancer cell invasion via differential regulation of RhoGTPases. Mol Biol Cell. 2016;27:1409–1419. doi: 10.1091/mbc.E16-02-0101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Feng W, Yonezawa M, Ye J, Jenuwein T, Grummt I. PHF8 activates transcription of rRNA genes through H3K4me3 binding and H3K9me1/2 demethylation. Nat Struct Mol Biol. 2010;17:445–450. doi: 10.1038/nsmb.1778. [DOI] [PubMed] [Google Scholar]
- 28.Fortschegger K, et al. PHF8 targets histone methylation and RNA polymerase II to activate transcription. Mol Cell Biol. 2010;30:3286–3298. doi: 10.1128/MCB.01520-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Liu W, et al. PHF8 mediates histone H4 lysine 20 demethylation events involved in cell cycle progression. Nature. 2010;466:508–512. doi: 10.1038/nature09272. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Qi HH, et al. Histone H4K20/H3K9 demethylase PHF8 regulates zebrafish brain and craniofacial development. Nature. 2010;466:503–507. doi: 10.1038/nature09261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Shao P, et al. Histone demethylase PHF8 promotes epithelial to mesenchymal transition and breast tumorigenesis. Nucleic Acids Res. 2017;45:1687–1702. doi: 10.1093/nar/gkw1093. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Carey BW, Finley LW, Cross JR, Allis CD, Thompson CB. Intracellular alpha-ketoglutarate maintains the pluripotency of embryonic stem cells. Nature. 2015;518:413–416. doi: 10.1038/nature13981. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Bullock GC, et al. Iron control of erythroid development by a novel aconitase-associated regulatory pathway. Blood. 2010;116:97–108. doi: 10.1182/blood-2009-10-251496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Li X, Egervari G, Wang Y, Berger SL, Lu Z. Regulation of chromatin and gene expression by metabolic enzymes and metabolites. Nat Rev Mol Cell Biol. 2018;19:563–578. doi: 10.1038/s41580-018-0029-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Smith AG, Carthew P, Francis JE, Edwards RE, Dinsdale D. Characterization and accumulation of ferritin in hepatocyte nuclei of mice with iron overload. Hepatology. 1990;12:1399–1405. doi: 10.1002/hep.1840120622. [DOI] [PubMed] [Google Scholar]
- 36.Yu L, et al. Structural insights into a novel histone demethylase PHF8. Cell Res. 2010;20:166–173. doi: 10.1038/cr.2010.8. [DOI] [PubMed] [Google Scholar]
- 37.McDonald OG, Wu H, Timp W, Doi A, Feinberg AP. Genome-scale epigenetic reprogramming during epithelial-to-mesenchymal transition. Nat Struct Mol Biol. 2011;18:867–874. doi: 10.1038/nsmb.2084. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Zhao QY, et al. Global histone modification profiling reveals the epigenomic dynamics during malignant transformation in a four-stage breast cancer model. Clin Epigenetics. 2016;8:34. doi: 10.1186/s13148-016-0201-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.McDonald OG, et al. Epigenomic reprogramming during pancreatic cancer progression links anabolic glucose metabolism to distant metastasis. Nat Genet. 2017;49:367–376. doi: 10.1038/ng.3753. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Gerry CJ, Schreiber SL. Chemical probes and drug leads from advances in synthetic planning and methodology. Nat Rev Drug Discov. 2018;17:333–352. doi: 10.1038/nrd.2018.53. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Caneque T, Müller S, Rodriguez R. Visualizing biologically active small molecules in cells using click chemistry. Nat Rev Chem. 2018;2:202–215. [Google Scholar]
- 42.Müller S, et al. Metformin reveals a mitochondrial copper addiction of mesenchymal cancer cells. PLoS One. 2018;13:e0206764. doi: 10.1371/journal.pone.0206764. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Janzer A, et al. Metformin and phenformin deplete tricarboxylic acid cycle and glycolytic intermediates during cell transformation and NTPs in cancer stem cells. Proc Natl Acad Sci USA. 2014;111:10574–10579. doi: 10.1073/pnas.1409844111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Zilka O, et al. On the mechanism of cytoprotection by ferrostatin-1 and liproxstatin-1 and the role of lipid peroxidation in ferroptotic cell death. ACS Cent Sci. 2017;3:232–243. doi: 10.1021/acscentsci.7b00028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Conrad M, Pratt DA. The chemical basis of ferroptosis. Nat Chem Biol. 2019;15:1137–1147. doi: 10.1038/s41589-019-0408-1. [DOI] [PubMed] [Google Scholar]
- 46.Flavahan WA, Gaskell E, Bernstein BE. Epigenetic plasticity and the hallmarks of cancer. Science. 2017;357:eaal2380. doi: 10.1126/science.aal2380. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Tian X, et al. High-molecular-mass hyaluronan mediates the cancer resistance of the naked mole rat. Nature. 2013;499:346–349. doi: 10.1038/nature12234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Gerken T, et al. The obesity-associated FTO gene encodes a 2-oxoglutarate-dependent nucleic acid demethylase. Science. 2007;318:1469–1472. doi: 10.1126/science.1151710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Jia G, et al. N6-methyladenosine in nuclear RNA is a major substrate of the obesity-associated FTO. Nat Chem Biol. 2011;7:885–887. doi: 10.1038/nchembio.687. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Greer EL, Shi Y. Histone methylation: a dynamic mark in health, disease and inheritance. Nat Rev Genet. 2012;13:343–357. doi: 10.1038/nrg3173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Tsai YP, et al. TET1 regulates hypoxia-induced epithelial-mesenchymal transition by acting as a co-activator. Genome Biol. 2014;15:513. doi: 10.1186/s13059-014-0513-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All data are available in the manuscript or the supplementary materials. Mass spectrometry data have been deposited at the ProteomeXchange Consortium (PRIDE Archive) with identifiers PXD011447 and PXD012862.
ChlP-seq and RNA-seq data are available on the National Center for Biotechnology Information website with accession reference GSE121664.
Code employed for ChIP-seq and RNA-seq data analyses are available on Github at https://github.com/nservant/EMTiron.












