首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
2.

Background

Senescent red blood cells (RBC) are recognized, phagocytosed and cleared by tissue macrophages. During this erythrophagocytosis (EP), RBC are engulfed and processed in special compartments called erythrophagosomes. We previously described that following EP, heme is rapidly degraded through the catabolic activity of heme oxygenase (HO). Extracted heme iron is then either exported or stored by macrophages. However, the cellular localization of the early steps of heme processing and iron extraction during EP remains to be clearly defined.

Methodology/Principal Findings

We took advantage of our previously described cellular model of EP, using bone marrow-derived macrophages (BMDM). The subcellular localization of both inducible and constitutive isoforms of HO (HO-1 and HO-2), of the divalent metal transporters (Nramp1, Nramp2/DMT1, Fpn), and of the recently identified heme transporter HRG-1, was followed by fluorescence and electron microscopy during the earliest steps of EP. We also looked at some ER [calnexin, glucose-6-phosphatase (G6Pase) activity] and lysosomes (Lamp1) markers during EP. In both quiescent and LPS-activated BMDM, Nramp1 and Lamp1 were shown to be strong markers of the erythrophagolysosomal membrane. HRG-1 was also recruited to the erythrophagosome. Furthermore, we observed calnexin labeling and G6Pase activity at the erythrophagosomal membrane, indicating the contribution of ER in this phagocytosis model. In contrast, Nramp2/DMT1, Fpn, HO-1 and HO-2 were not detected at the membrane of erythrophagosomes.

Conclusions/Significance

Our study highlights the subcellular localization of various heme- and iron-related proteins during early steps of EP, thereby suggesting a model for heme catabolism occurring outside the phagosome, with heme likely being transported into the cytosol through HRG1. The precise function of Nramp1 at the phagosomal membrane in this model remains to be determined.  相似文献   

3.
Iron is essential for the survival as well as the proliferation and maturation of developing erythroid precursors (EP) into hemoglobin-containing red blood cells. The transferrin-transferrin receptor pathway is the main route for erythroid iron uptake. Using a two-phase culture system, we have previously shown that placental ferritin as well as macrophages derived from peripheral blood monocytes could partially replace transferrin and support EP growth in a transferrin-free medium. We now demonstrate that in the absence of transferrin, ferritin synthesized and secreted by macrophages can serve as an iron source for EP. Macrophages trigger an increase in both the cytosolic and the mitochondrial labile iron pools, in heme and in hemoglobin synthesis, along with a decrease in surface transferrin receptors. Inhibiting macrophage exocytosis, binding extracellular ferritin with specific antibodies, inhibiting EP receptor-mediated endocytosis or acidification of EP lysosomes, all resulted in a decreased EP growth when co-cultured with macrophages under transferrin-free conditions. The results suggest that iron taken up by macrophages is incorporated mainly into their ferritin, which is subsequently secreted by exocytosis. Nearby EP are able to take up this ferritin probably through clathrin-dependent, receptor-mediated endocytosis into endosomes, which following acidification and proteolysis release the iron from the ferritin, making it available for regulatory and synthetic purposes. Thus, macrophages support EP development under transferrin-free conditions by delivering essential iron in the form of metabolizable ferritin.  相似文献   

4.
5.
We have recently reported that red blood cells (RBC) promote T cell growth and survival by inhibiting activation-induced T cell death. In the present study, we have examined parameters of oxidative stress and intracellular iron in activated T cells and correlated these data with the expression of ferritin, heme oxygenase-1 (HO-1), and the transferrin receptor CD71. T cells growing in the presence of RBC had reduced levels of reactive oxygen species (ROS) and oxidatively modified proteins, suggesting that RBC efficiently counteracted ROS production on the activated T cells. Flow cytometry and immunodetection demonstrated that T cells dividing in the presence of RBC had increased levels of intracellular ferritin rich in L-subunits and HO-1 along with a downmodulation in CD71 expression. Finally, using the fluorescent iron indicator calcein and flow cytometry analysis, we were able to show that a relative amount of the labile iron pool (LIP) was upregulated in T cells growing in the presence of RBC. These findings are consistent with a typical response to iron overload. However, neither heme compounds nor ferric iron reproduced the levels of expansion and survival of T cells induced by intact RBC. Altogether, these data suggest that RBC inhibit apoptosis of activated T cells by a combination of ROS scavenging and upregulation of cytoprotective proteins such as ferritin and HO-1, which may counteract a possible toxic effect of the increased intracellular free iron.  相似文献   

6.
The catabolism of heme is carried out by members of the heme oxygenase (HO) family. The products of heme catabolism by HO-1 are ferrous iron, biliverdin (subsequently converted to bilirubin), and carbon monoxide. In addition to its function in the recycling of hemoglobin iron, this microsomal enzyme has been shown to protect cells in various stress models. Implicit in the reports of HO-1 cytoprotection to date are its effects on the cellular handling of heme/iron. However, the limited amount of uncommitted heme in non-erythroid cells brings to question the source of substrate for this enzyme in non-hemolytic circumstances. In the present study, HO-1 was induced by either sodium arsenite (reactive oxygen species producer) or hemin or overexpressed in the murine macrophage-like cell line, RAW 264.7. Both of the inducers elicited an increase in active HO-1; however, only hemin exposure caused an increase in the synthesis rate of the iron storage protein, ferritin. This effect of hemin was the direct result of the liberation of iron from heme by HO. Cells stably overexpressing HO-1, although protected from oxidative stress, did not display elevated basal ferritin synthesis. However, these cells did exhibit an increase in ferritin synthesis, compared with untransfected controls, in response to hemin treatment, suggesting that heme levels, and not HO-1, limit cellular heme catabolism. Our results suggest that the protection of cells from oxidative insult afforded by HO-1 is not due to the catabolism of significant amounts of cellular heme as thought previously.  相似文献   

7.
The ultraviolet A (UVA, 320–400 nm) component of sunlight has the potential to generate an oxidative stress in cells and tissue so that antioxidants (both endogenous and exogenous) strongly influence the biological effects of UVA. The expression of several genes (including heme oxygenase-1, HO-1; collagenase; the CL100 phosphatase and the nuclear oncogenes, c-fos and c-jun) is induced following physiological doses of UVA to cells and this effect can be strongly enhanced by removing intracellular glutathione or enhancing singlet oxygen lifetime. We have observed that heme is released from microsomal heme-containing proteins by UVA and other oxidants and that activation of HO-1 expression by UVA correlates with levels of heme release. UVA radiation also leads to an increase in labile iron pools (either directly or via HO-1) and eventual increases in ferritin levels. The role of heme oxygenase in protection of skin fibroblasts is probably an emergency inducible defense pathway to remove heme liberated by oxidants. The slower increase in ferritin levels is an adaptive response which serves to keep labile iron pools low and thereby reduce Fenton chemistry and oxidant-induced chain reactions involving lipid peroxidation. In keratinocytes, the primary target of UVA radiation, heme oxygenase levels are constitutively high (because of HO-2 expression). Since there is a corresponding increase in basal levels of ferritin the epidermis appears to be well protected constitutively against the oxidative stress generated by UVA.  相似文献   

8.
Several lines of evidence suggest that antioxidant processes and (or) endogenous antioxidants inhibit proatherogenic events in the blood vessel wall. Heme oxygenase (HO), which catabolizes heme to biliverdin, carbon monoxide, and catalytic iron, has been shown to have such antioxidative properties. The HO-1 isoform of heme oxygenase is ubiquitous and can be increased several fold by stimuli that induce cellular oxidative stress. Products of the HO reaction have important effects: carbon monoxide is a potent vasodilator, which is thought to play a role in modulation of vascular tone; biliverdin and its by-product bilirubin are potent antioxidants. Although HO induction results in an increase in catalytic free iron release, the enhancement of intracellular ferritin protein through HO-1 has been reported to decrease the cytotoxic effects of iron. Oxidized LDL has been shown to increase HO-1 expression in endothelial and smooth muscle cell cultures, and during atherogenesis. Further evidence of HO-1 expression associated with atherogenesis has been demonstrated in human, murine and rabbit atherosclerotic lesions. Moreover, genetic models of HO deficiency suggest that the actions of HO-1 are important in modulating the severity of atherosclerosis. Recent experiments in gene therapy using the HO gene suggest that interventions aimed at HO in the vessel wall could provide a novel therapeutic approach for the treatment or prevention of atherosclerotic disease.  相似文献   

9.
The splenic toxicity of aniline is characterized by vascular congestion, hyperplasia, fibrosis, and the development of a variety of sarcomas in rats. However, the underlying mechanisms by which aniline elicits splenotoxic response are not well understood. Previously we have shown that aniline exposure causes oxidative damage to the spleen. To further explore the oxidative mechanism of aniline toxicity, we evaluated the potential contribution of heme oxygenase-1 (HO-1), which catalyzes heme degradation and releases free iron. Male SD rats were given 1 mmol/kg/day aniline in water by gavage for 1, 4, or 7 days, and respective controls received water only. Aniline exposure led to significant increases in HO-1 mRNA expression in the spleen (2-and 2.4-fold at days 4 and 7, respectively) with corresponding increases in protein expression, as confirmed by ELISA and Western blot analysis. Furthermore, immunohistochemical assessment of spleen showed stronger immunostaining for HO-1 in the spleens of rats treated for 7 days, confined mainly to the red pulp areas. No changes were observed in mRNA and protein levels of HO-1 after 1 day exposure. The increase in HO-1 expression was associated with increases in total iron (2.4-and 2.7-fold), free iron (1.9-and 3.5-fold), and ferritin levels (1.9-and 2.1-fold) at 4 and 7 days of aniline exposure. Our data suggest that HO-1 up-regulation in aniline-induced splenic toxicity could be a contributing pro-oxidant mechanism, mediated through iron release, and leading to oxidative damage.  相似文献   

10.
In being both, a modifier of cellular immune effector pathways and an essential nutrient for microbes, iron is a critical determinant in host-pathogen interaction. Here, we investigated the metabolic changes of macrophage iron homeostasis and immune function following the infection of RAW264.7 murine macrophages with Salmonella typhimurium. We observed an enhanced expression of the principal iron export protein, ferroportin 1, and a subsequent increase of iron efflux in Salmonella-infected phagocytes. In parallel, the expression of haem oxygenase 1 and of the siderophore-binding peptide lipocalin 2 was markedly enhanced following pathogen entry. Collectively, these modulations reduced both the cytoplasmatic labile iron and the ferritin storage iron pool within macrophages, thus restricting the acquisition of iron by intramacrophage Salmonella. Correspondingly, limitation of macrophage iron decreased microbial survival, whereas iron supplementation impaired immune response pathways in Salmonella-infected macrophages (nitric oxide formation and tumour necrosis factor-alpha production) and promoted intracellular bacterial proliferation. Our findings suggest that the enhancement of ferroportin 1-mediated iron efflux, the upregulation of the haem-degrading enzyme haem oxygenase 1 and the induction of lipocalin 2 following infection concordantly aim at withholding iron from intracellular S. typhimurium and to increase antimicrobial immune effector pathways thus limiting pathogen proliferation.  相似文献   

11.
Earlier observations indicate that free heme is selectively toxic to cells lacking heme oxygenase-1 (HO-1) but how this enzyme prevents heme toxicity remains unexplained. Here, using A549 (human lung cancer) and immortalized human bronchial epithelial cells incubated with exogenous heme, we find knock-down of HO-1 using siRNA does promote the accumulation of cell-associated heme and heme-induced cell death. However, it appears that the toxic effects of heme are exerted by “loose” (probably intralysosomal) iron because cytotoxic effects of heme are lessened by pre-incubation of HO-1 deficient cells with desferrioxamine (which localizes preferentially in the lysosomal compartment). Desferrioxamine also decreases lysosomal rupture promoted by intracellularly generated hydrogen peroxide. Supporting the importance of endogenous oxidant production, both chemical and siRNA inhibition of catalase activity predisposes HO-1 deficient cells to heme-mediated killing. Importantly, it appears that HO-1 deficiency somehow blocks the induction of ferritin; control cells exposed to heme show ~10-fold increases in ferritin heavy chain expression whereas in heme-exposed HO-1 deficient cells ferritin expression is unchanged. Finally, overexpression of ferritin H chain in HO-1 deficient cells completely prevents heme-induced cytotoxicity. Although two other products of HO-1 activity–CO and bilirubin–have been invoked to explain HO-1-mediated cytoprotection, we conclude that, at least in this experimental system, HO-1 activity triggers the induction of ferritin and the latter is actually responsible for the cytoprotective effects of HO-1 activity.  相似文献   

12.
13.
Oxidative stress, deposition of non-transferrin iron, and mitochondrial insufficiency occur in the brains of patients with Alzheimer disease (AD) and Parkinson disease (PD). We previously demonstrated that heme oxygenase-1 (HO-1) is up-regulated in AD and PD brain and promotes the accumulation of non-transferrin iron in astroglial mitochondria. Herein, dynamic secondary ion mass spectrometry (SIMS) and other techniques were employed to ascertain (i) the impact of HO-1 over-expression on astroglial mitochondrial morphology in vitro , (ii) the topography of aberrant iron sequestration in astrocytes over-expressing HO-1, and (iii) the role of iron regulatory proteins (IRP) in HO-1-mediated iron deposition. Astroglial hHO-1 over-expression induced cytoplasmic vacuolation, mitochondrial membrane damage, and macroautophagy. HO-1 promoted trapping of redox-active iron and sulfur within many cytopathological profiles without impacting ferroportin, transferrin receptor, ferritin, and IRP2 protein levels or IRP1 activity. Thus, HO-1 activity promotes mitochondrial macroautophagy and sequestration of redox-active iron in astroglia independently of classical iron mobilization pathways. Glial HO-1 may be a rational therapeutic target in AD, PD, and other human CNS conditions characterized by the unregulated deposition of brain iron.  相似文献   

14.
Redox regulation and oxidant activation of heme oxygenase-1   总被引:4,自引:0,他引:4  
The ultraviolet A (UVA, 320-400 nm) component of sunlight has the potential to generate an oxidative stress in cells and tissue so that antioxidants (both endogenous and exogenous) strongly influence the biological effects of UVA. The expression of several genes (including heme oxygenase-1, HO-1; collagenase; the CL100 phosphatase and the nuclear oncogenes, c-fos and c-jun) is induced following physiological doses of UVA to cells and this effect can be strongly enhanced by removing intracellular glutathione or enhancing singlet oxygen lifetime. We have observed that heme is released from microsomal heme-containing proteins by UVA and other oxidants and that activation of HO-1 expression by UVA correlates with levels of heme release. UVA radiation also leads to an increase in labile iron pools (either directly or via HO-1) and eventual increases in ferritin levels. The role of heme oxygenase in protection of skin fibroblasts is probably an emergency inducible defense pathway to remove heme liberated by oxidants. The slower increase in ferritin levels is an adaptive response which serves to keep labile iron pools low and thereby reduce Fenton chemistry and oxidant-induced chain reactions involving lipid peroxidation. In keratinocytes, the primary target of UVA radiation, heme oxygenase levels are constitutively high (because of HO-2 expression). Since there is a corresponding increase in basal levels of ferritin the epidermis appears to be well protected constitutively against the oxidative stress generated by UVA.  相似文献   

15.
Iron metabolism in mammals requires a complex and tightly regulated molecular network. The classical view of iron metabolism has been challenged over the past ten years by the discovery of several new proteins, mostly Fe (II) iron transporters, enzymes with ferro-oxydase (hephaestin or ceruloplasmin) or ferri-reductase (Dcytb) activity or regulatory proteins like HFE and hepcidin. Furthermore, a new transferrin receptor has been identified, mostly expressed in the liver, and the ability of the megalin-cubilin complex to internalise the urinary Fe (III)-transferrin complex in renal tubular cells has been highlighted. Intestinal iron absorption by mature duodenal enterocytes requires Fe (III) iron reduction by Dcytb and Fe (II) iron transport through apical membranes by the iron transporter Nramp2/DMT1. This is followed by iron transfer to the baso-lateral side, export by ferroportin and oxidation into Fe (III) by hephaestin prior to binding to plasma transferrin. Macrophages play also an important role in iron delivery to plasma transferrin through phagocytosis of senescent red blood cell, heme catabolism and recycling of iron. Iron egress from macrophages is probably also mediated by ferroportin and patients with heterozygous ferroportin mutations develop progressive iron overload in liver macrophages. Iron homeostasis at the level of the organism is based on a tight control of intestinal iron absorption and efficient recycling of iron by macrophages. Signalling between iron stores in the liver and both duodenal enterocytes and macrophages is mediated by hepcidin, a circulating peptide synthesized by the liver and secreted into the plasma. Hepcidin expression is stimulated in response to iron overload or inflammation, and down regulated by anemia and hypoxia. Hepcidin deficiency leads to iron overload and hepcidin overexpression to anemia. Hepcidin synthesis in response to iron overload seems to be controlled by the HFE molecule. Patients with hereditary hemochromatosis due to HFE mutation have impaired hepcidin synthesis and forced expression of an hepcidin transgene in HFE deficient mice prevents iron overload. These results open new therapeutic perspectives, especially with the possibility to use hepcidin or antagonists for the treatment of iron overload disorders.  相似文献   

16.
Intestinal epithelial cells and reticuloendothelial macrophages are, respectively, involved in diet iron absorption and heme iron recycling from senescent erythrocytes, two critical processes of iron homeostasis. These cells appear to use the same transporter, ferroportin (Slc40a1), to export iron. The aim of this study was to compare the localization, expression, and regulation of ferroportin in both duodenal and macrophage cells. Using a high-affinity purified polyclonal antibody, we analyzed the localization and expression of ferroportin protein in the spleen, liver, and duodenum isolated from normal mice as well as from well-characterized mouse models of altered iron homeostasis. Ferroportin was found to be predominantly expressed in enterocytes of the duodenum, in splenic macrophages, and in liver Kupffer cells. Interestingly, the protein species detected in these cells migrated differently on SDS-PAGE. These differences in apparent molecular masses were partly explained by posttranslational complex N-linked glycosylations. In addition, in enterocytes, the transporter was mostly expressed at the basolateral membrane, whereas in bone marrow-derived macrophages, ferroportin was found predominantly localized in the intracellular vesicular compartment. However, some microdomains positive for ferroportin were also detected at the plasma membrane of macrophages. Despite these differences, we observed a parallel upregulation of ferroportin expression in tissue macrophages and enterocytes in response to iron-restricted erythropoiesis, suggesting that iron homeostasis is likely maintained through coordinate expression of the iron exporter in both intestinal and phagocytic cells. Our data also confirm a predominant regulation of ferroportin through systemic regulator(s) likely including hepcidin.  相似文献   

17.
来源于出血后血红蛋白或衰老细胞释放的血红素能够诱导血红素氧合酶-1(HO-1,HSP-1)的表达。血红素氧合酶-1催化血红素生成气体介质一氧化碳,铁和胆绿素。胆绿素和它的代谢产物胆红素都是有效的抗氧化剂;同时铁诱导的铁蛋白和CO也发挥着各自的保护作用。因此,HO-1的表达被看作一种重要的保护机制。在各种不同的脑病理改变发生后,如蛛网膜下腔出血,脑梗死,创伤性脑损伤及神经变性疾病,HO-1明显表达于小胶质细胞,星形细胞和神经元细胞,从而发挥其重要脑保护作用。  相似文献   

18.
BACKGROUND: Expression and enzymatic activity of heme oxygenase (HO) has been implicated in the development, as well as in the resolution, of inflammatory conditions. Because inflammation is central to tissue repair, we investigated the presence and potential functions of HO in an excisional model of normal and diabetes-impaired wound repair in mice. MATERIALS AND METHODS: Expression of HO-1 during cutaneous healing was analyzed by RNase protection assay, Western blot, and immunohistochemical techniques in a murine model of excisional repair. Furthermore, we determined HO-1-dependent release of proinflammatory cytokines from RAW 264.7 macrophages by enzyme-linked immunosorbent assay (ELISA). RESULTS: Upon injury, we observed a rapid and strong increase in HO-1 mRNA and protein levels at the wound site. By contrast to normal repair, late stages of diabetes-impaired repair were associated with elevated HO-1 expression. Besides a few keratinocytes of the hyperproliferative epithelium, immunohistochemistry revealed infiltrating macrophages as the predominant and major source of HO-1 at the wound site. In vitro studies demonstrated the potency of exogenous and also endogenous nitric oxide (NO) to strongly induce HO-1 expression in RAW 264.7 macrophages. However, L-NIL-mediated enzymatic inhibition of inducible NO-synthase (iNOS) at the wound site in vivo was not paralleled by decreased HO-1 levels. In vitro inhibition of HO-1 enzymatic activity by tin protoporphyrin IX (SnPPIX) in RAW 264.7 macrophages markedly attenuated tumor necrosis factor-alpha (TNF-alpha), but strongly increased interleukin-1beta (IL-1beta) release in RAW 264.7 macrophages in vitro. CONCLUSIONS: The observed injury-mediated increase in HO-1 mRNA and protein at the wound site was due to infiltrating HO-1 expressing monocytic cells. Macrophage-derived HO-1 expression was not under regulatory control by NO in skin repair. We provide evidence that HO-1 might exert a regulatory role in macrophage-derived cytokine release.  相似文献   

19.
We previously found that ANG II infusion into rats causes iron deposition in the kidney and heart, which may have a role in the regulation of profibrotic gene expression and tissue fibrosis. In the present study, we have investigated whether ANG II can also induce iron accumulation in the liver. Prussian blue staining detected frequent iron deposition in the interstitium of the liver of rats treated with pressor dose ANG II for 7 days, whereas iron deposition was absent in the livers of control rats. Immunohistochemical and histological analyses showed that some iron-positive nonparenchymal cells were positive for ferritin and heme oxygenase-1 (HO-1) protein and TGF-beta1 mRNA and were judged to be monocytes/macrophages. It was shown that ANG II infusion caused about a fourfold increase in ferritin and HO-1 protein expression by Western blot analysis and about a twofold increase in TGF-beta1 mRNA expression by Northern blot analysis, which were both suppressed by treating ANG II-infused rats with losartan and deferoxamine. In addition, mild interstitial fibrosis was observed in the liver of rats that had been treated with pressor dose ANG II for 7 days or with nonpressor dose ANG II for 30 days, the latter of which also caused loss of hepatocytes and intrahepatic hemorrhage in the liver. Taken together, our data suggest that ANG II infusion induces aberrant iron homeostasis in the liver, which may have a role in the ANG II-induced upregulation of profibrotic gene expression in the liver.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号