Next Article in Journal
Synthesis and Initial In Vivo Evaluation of [11C]AZ683—A Novel PET Radiotracer for Colony Stimulating Factor 1 Receptor (CSF1R)
Next Article in Special Issue
Iron Regulation: Macrophages in Control
Previous Article in Journal
Dissipative Particle Dynamics Investigation of the Transport of Salicylic Acid through a Simulated In Vitro Skin Permeation Model
Previous Article in Special Issue
Iron as a Therapeutic Target in HFE-Related Hemochromatosis: Usual and Novel Aspects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Established and Emerging Concepts to Treat Imbalances of Iron Homeostasis in Inflammatory Diseases

1
Department of Internal Medicine II, Medical University of Innsbruck, 6020 Innsbruck, Austria
2
Christian Doppler Laboratory for Iron Metabolism and Anemia Research, Medical University of Innsbruck, 6020 Innsbruck, Austria
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2018, 11(4), 135; https://doi.org/10.3390/ph11040135
Submission received: 10 November 2018 / Revised: 5 December 2018 / Accepted: 6 December 2018 / Published: 11 December 2018
(This article belongs to the Special Issue Iron as Therapeutic Targets in Human Diseases)

Abstract

:
Inflammation, being a hallmark of many chronic diseases, including cancer, inflammatory bowel disease, rheumatoid arthritis, and chronic kidney disease, negatively affects iron homeostasis, leading to iron retention in macrophages of the mononuclear phagocyte system. Functional iron deficiency is the consequence, leading to anemia of inflammation (AI). Iron deficiency, regardless of anemia, has a detrimental impact on quality of life so that treatment is warranted. Therapeutic strategies include (1) resolution of the underlying disease, (2) iron supplementation, and (3) iron redistribution strategies. Deeper insights into the pathophysiology of AI has led to the development of new therapeutics targeting inflammatory cytokines and the introduction of new iron formulations. Moreover, the discovery that the hormone, hepcidin, plays a key regulatory role in AI has stimulated the development of several therapeutic approaches targeting the function of this peptide. Hence, inflammation-driven hepcidin elevation causes iron retention in cells and tissues. Besides pathophysiological concepts and diagnostic approaches for AI, this review discusses current guidelines for iron replacement therapies with special emphasis on benefits, limitations, and unresolved questions concerning oral versus parenteral iron supplementation in chronic inflammatory diseases. Furthermore, the review explores how therapies aiming at curing the disease underlying AI can also affect anemia and discusses emerging hepcidin antagonizing drugs, which are currently under preclinical or clinical investigation.

1. Introduction

Iron has a crucial role in all living organisms. In humans, iron is essential for many biochemical processes, including electron transfer reactions in mitochondria, the citric acid cycle, gene expression, binding and transport of oxygen, regulation of cell growth and differentiation as well as the cellular immune response [1]. From a systemic point of view, hepcidin, a liver-derived hormone, has been found to be the master regulator of iron homeostasis, controlling cellular iron efflux [2]. Hepcidin binds to the sole known iron exporter, ferroportin (FPN), mediating internalization and degradation of this transporter [3,4]. As a further consequence, dietary iron absorption as well as iron release from cells, such as macrophages, is prevented [5]. Hepcidin expression is regulated by different stimuli, such as anemia, hypoxia, and inflammation [6]. Different molecular pathways involved in hepcidin expression have been uncovered [7,8]. Among these, the bone morphogenetic protein (BMP)-SMAD signaling pathway is the most critical. Liver endothelial cell-derived BMP6 and BMP2 have non-redundant roles to induce hepcidin expression. However, BMP6 is the dominant ligand and a threshold signaling of BMP6 via the BMP-SMAD pathway is indispensable for sufficient and appropriate hepcidin induction [9,10,11,12]. During inflammation, hepcidin expression is induced via the interleukin (IL)6-JAK-STAT and Activin B-SMAD1/5/8 signaling pathways [2,13,14,15]. As FPN regulates iron release from absorptive enterocytes in the duodenum and from iron recycling macrophages of the mononuclear phagocyte system (MPS), elevated hepcidin levels during inflammation cause diminished systemic iron availability [16]. While iron retention in the MPS appears to be beneficial for host responses during infections, as it withholds this metal from invading extracellular pathogens, anemia is an undesired, ultimate consequence of iron restriction in patients suffering from chronic diseases [17,18,19]. Consequently, anemia of inflammation (AI) or anemia of chronic disease (ACD) represents the most common disease-related complication in patients suffering from rheumatoid arthritis (RA) inflammatory bowel diseases (IBD), cancer, infectious diseases, and chronic kidney disease (CKD) [1,20,21,22,23,24,25,26].
Whereas the development and persistence of anemia in several diseases, including infections and cancer, has been associated with a poor prognosis, the true impact of alterations in iron homeostasis or anemia on the pathology of the underlying disease remains largely elusive [27]. However, anemia negatively impacts on many aspects of the patients [28]. Moreover, iron exerts multiple effects on immune cell differentiation, functionality, and plasticity, which has been studied in depth toward the interconnection of iron homeostasis with the biology of M1 and M2 macrophages [29,30]. M1 macrophages, being either activated upon pathogen recognition or stimulated by cytokines (e.g., interferon-gamma (IFNγ), tumor necrosis factor-alpha (TNFα), or IL1, IL6 and IL10), induce subtle changes of transcellular iron fluxes, aiming to limit the availability of the essential nutrient iron for circulating pathogens. Therefore, the term, “nutritional immunity”, has been proposed [31]. Thereafter, cytokines directly or cytokine-inducible products, such as oxygen radicals and nitric oxide, regulate the expression of critical iron transport and storage proteins [19,32,33,34,35,36,37,38]. Consequently, the accessibility of iron for microbes is modulated and their growth and pathogenicity is impacted. Moreover, bacteria can acquire iron from holo-transferrin. Therefore, limitation of circulating transferrin-bound iron levels and mutations of the iron binding sites of transferrin were shown to be protective against infections with circulating bacteria [39,40,41].
Of note, iron per se also affects immune effector pathways of macrophages and, subsequently, T-cell differentiation by regulating IFNγ activity, nitric oxide formation, or T-helper cell plasticity [42,43,44,45,46]. Thus, local and systemic iron availability determines not only microbial growth, but also the efficacy of anti-microbial immune effector pathways. It appears that the alterations of systemic and macrophage-responsible iron fluxes are specifically regulated depending on the nature and localization of the pathogen [19,32,33,39,47]. The hepcidin–FPN axis has attracted specific attention regarding alterations of iron fluxes during infections. Hepatocytes produce large amounts of hepcidin following challenge with circulating bacteria, resulting in iron retention with the MPS and low circulating iron levels [17,18]. Moreover, autocrine formation of hepcidin by macrophages further reduces cellular iron availability for circulating pathogens [19,39,48,49]. In contrast, invasion of cells and macrophages with bacteria, such as Listeria, Mycobacteria, or Salmonella, induces alternative mechanisms. Specifically, the upregulation of FPN by different mechanisms results in macrophage iron efflux and limitation of bacterial growth [50,51,52,53,54]. In addition, M1 macrophages and other immune cells produce several factors, such as lipocalin-2, lactoferrin, and calprotectin, which limit the bacterial access to iron [55,56,57]. Of note, in addition to iron flux regulation by the FPN-hepcidin axis, several hepcidin-independent mechanisms have been identified that control iron trafficking during infection [33,50,58,59,60,61].
In contrast, M2 macrophages exert anti-inflammatory effects and these cells are highly specialized for iron recycling from senescent erythrocytes via erythrophagocytosis, yielding approximately 90% of the daily needs of iron for erythropoiesis [30]. Iron is released from heme by the anti-inflammatory enzyme, heme oxygenase-1 [62,63]. The latter enzyme has attracted specific interest because it exerts immune regulatory effects, but, importantly, it also exerts disease tolerance during certain infections by limiting tissue damage, thereby improving the outcome from sepsis [64,65].
While it has long been known that iron is essential for the production of hemoglobin of red blood cells, our knowledge on the regulation of iron homeostasis under steady state conditions and in association with different pathologies has dramatically expanded over the past centuries thanks to the identification and characterization of numerous iron genes and associated regulatory molecules [1]. Indeed, unbiased iron supplementation or withdrawal therapy via phlebotomy dates to the middle ages. However, due to our expanding knowledge on iron metabolism regulation during inflammation, targeted modulation of specific iron metabolic pathways, including the hepcidin-FPN axis, has emerged only recently [3,66,67]. Although we have several established and novel iron therapies at hand, there are still many unresolved questions and unmet needs when treating imbalances of iron homeostasis in patients with inflammatory diseases. This includes lack of gold-standard tests to properly distinguish between absolute versus functional ID, lack of knowledge regarding safe and efficient therapeutic start and end points as well as complications of iron redistribution and supplementation strategies towards the course of the diseases underlying AI.

2. Diagnosis

The diagnosis of AI is based on several laboratory markers. Classically, hemoglobin levels are decreased; markers of inflammation, such as C-reactive protein (CRP) or IL6, are increased; and iron homeostasis is altered as follows: Circulating iron levels are low, transferrin saturation (Tf-Sat) is reduced, and ferritin concentrations are normal or increased (Table 1) [68]. Diagnosis becomes challenging if AI is associated with true ID (AI/ID), as there is still a lack of a gold standard for differentiation between AI and AI/ID. However, as therapies to overcome anemia differ, proper diagnosis and understanding of underlying pathophysiological regulations are necessary [69]. While ferritin strongly correlates with the body’s iron stores in IDA, ferritin levels are not reliable during inflammation. Thus, low ferritin levels (<30 mg/mL) in any case indicate true ID, but ferritin levels are upregulated during inflammation largely independently of iron availability [70]. This fact has led to corrections towards elevated cut-off values for ferritin during concomitant inflammation [71,72]. Until now, the gold standard for diagnosis of ID is still the microscopic evaluation of iron-stained bone marrow aspirates, which is not routinely used due to its high invasiveness [73]. A recent study in heart failure patients proposed to use serum iron and Tf-Sat instead of ferritin to diagnose true ID, which was evaluated by bone marrow staining [74]. Compared to ferritin-based definition of ID (with a sensitivity and specificity of 82% and 72%, respectively), the diagnosis of ID based on reduced Tf-Sat (cut-off: ≤19.8%) and low serum iron (cut-off: ≤13 μmol/L) had an improved sensitivity (94%) and specificity (84% and 88%) in this specific group of patients. Although these findings need further confirmation among other disease entities, it highlights that ferritin-based definitions of ID appear to be suboptimal.
As erythrocytes are the main consumers of iron and thus most affected by ID, efforts have been undertaken to establish markers that are related to red blood cell morphology and iron content of these cells. Alongside the well-established classical hematological indices of the mean corpuscular volume (MCV) and mean corpuscular hemoglobin (MCH), new parameters, such as the hemoglobin content of reticulocytes, percentage of hypochromic red blood cells, and the soluble TfR (sTfR), were introduced as indicators of iron availability for the erythron and/or efficacy of erythropoiesis [68,75,76]. Some studies recommend the sTfR as an alternative biomarker to distinguish between absolute (or true) and functional ID. In general, absolute ID and higher rates of erythroid output causes an up-regulation of the TfR on erythrocytes, which then concomitantly leads to higher detectable forms of its cleaved monomer, the sTfR, in the plasma [77]. As inflammation negatively impacts erythropoiesis and TfR expression, sTfR values are also altered during inflammation [26,36]. Therefore, the use of this marker led to unsatisfactory sensitivity and specificity (83% and 50%, respectively) for the detection of ID compared to bone marrow findings in a cohort of 180 anemic children in Mozambique [78]. Attempts to correct this marker for inflammation, using a sTfR versus log ferritin ratio, did not classify patients properly, thus limiting the diagnostic potential of this test [79]. While these parameters add additional information on true iron availability for erythropoiesis in patients with AI, none of these measurements are adjudged as efficient for distinguishing between AI and AI/ID.
A number of reports indicating that hepcidin is competent to distinguish between IDA and AI in several diseases, including RA, anemia of cancer, anemia of critical illness, and IBD, have suggested hepcidin to be a promising biomarker in the future [16,80,81,82,83]. Moreover, other reports also exist suggesting that plasma hepcidin levels could predict the response to oral iron in different settings [84,85,86,87]. However, there are also reports from studies in hemodialysis patients to the contrary, highlighting the need for further detailed investigations [88,89,90]. Further discussion on hepcidin is presented in Section 3.2.2. The measurements of molecules that affect hepcidin expression under different conditions may turn out to be of diagnostic benefit. Erythroferron, hypoxia inducible factors (HIFs), and platelet derived growth factor BB are all signaling peptides induced by hypoxia and were found to impact directly or via modulation of hepcidin on iron availability for erythropoiesis [91,92,93]. The biomarkers of hypoxia thus hold promise to better identify subjects suffering from AI/ID and to predict the erythroid response in patients with AI with and without ID, once commercially available ELISAs are available [94,95,96]. Of importance, none of these tests is currently standardized, which is a necessity to make them a reliable routine biomarker for the evaluation of iron status. Consequently, trials investigating these parameters cannot be easily compared, making interpretations even more difficult. However, according to a recent report, a hepcidin reference standard allows equivalence and comparability between hepcidin measurement results [97].
Despite ongoing efforts to find and establish new biomarkers, a recently published study conducted in a cohort of IBD patients revealed that differentiation between AI and iron deficiency anemia (IDA) and the combination thereof was only possible in 22% of all anemic patients, because only CRP, hemoglobin, and ferritin levels were available as diagnostic markers. [98]. This highlights that improvement of diagnostic approaches to identify patients with true ID in the setting of inflammation is urgently needed and is still a challenging field of investigation.

3. Treatment Strategies

Treatment of ID and IDA is paramount as it is associated with several detrimental effects on quality of life, exercise capacity, mental status, and activity of patients [99,100]. To this end, two strategies can be pursued. First, treatment of the underlying disease; second, if a cure cannot be achieved, therapies directly or indirectly addressing imbalances of iron homeostasis are indicated.

3.1. “First line”: Treatment of the Underlying Inflammation

If possible, treatment of the underlying disease is decidedly the pivotal approach to treat AI. Resolution of inflammation results in the normalization of hepcidin levels, leading to the correction of macrophage iron retention and normalization of duodenal iron uptake. In addition, the negative cytokine-mediated proliferative effects on hematopoiesis are abrogated, overall leading to anemia improvement. One approach, which has been shown to be effective, is the neutralization of inflammatory cytokines. Accordingly, targeted therapy using an anti-IL6 receptor antibody (Tocilizumab) improved anemia in patients suffering from multicentric Castleman’s disease (MCD), a lymphoproliferative disorder where IL6 was found to be the main cytokine contributing to its pathogenesis [101,102]. Of note, IL6 is one major driver for hypoferremia in patients suffering from AI [2]. Further work-up revealed that anemia amelioration due to IL6 receptor blockade is related to down-regulation of hepcidin levels [103,104]. In parallel, a monoclonal anti-IL6 antibody (Siltuximab) has also been evaluated for its potential to decrease hepcidin plasma levels and consequently improved anemia not only in patients suffering from MCD, but also in subjects with multiple myeloma and solid tumors [105,106,107].
Of interest, not only systemic, but also autocrine hepcidin expression in macrophages has been found to be of importance in AI and possibly also for iron distribution in cancer cells [48,108]. In patients with ovarian cancer, polarization towards an M1 phenotype and high IL6 levels were associated with more profound anemia. Treatment with Tocilizumab resulted in the reversion of iron restriction and improvement of anemia, supporting previous evidence that anti IL6-directed therapy may be effective for anemia in cancer [108,109].
TNFα is also a target to treat the underlying complications and ameliorate anemia. Monoclonal antibodies directed against TNFα (e.g., Infliximab, Adalimumab, Golimumab) are routinely applied in patients suffering from RA and IBD. As TNFα’s contribution to AI is different from IL6, the beneficial effect on anemia was ascribed to discontinuation of TNFα’s negative impact on bone marrow erythropoiesis or, likewise, erythrocyte’s half-life, without having direct effects on hepcidin levels [15,110,111,112,113,114]. However, a study investigating two different TNFα inhibitors in IBD patients found that the beneficial effect of anti-TNFα is indirect and it is mediated via down-regulation of IL6 [115]. Anti-TNF therapy may also reduce intravascular radical formation, thereby preventing the radical-mediated damage of erythrocyte membranes and increasing their circulating half-life. Moreover, comparative evaluation of TNFα inhibitors and Tocilizumab revealed that IL6-mediated therapy, directly affecting hepcidin levels, is more effective than TNFα inhibitors in respect to anemia correction [116]. Furthermore, hematological response after one year of anti-TNFα treatment was only observed in 34% of patients, even with oral iron supplementation [117].
Patients suffering from myeloproliferative neoplasms (MPN) have been shown to develop anemia, in part as a consequence of elevated hepcidin levels [118]. As mutations related to the activity of Janus kinase 2 (Jak2), resulting in constant activation, were found to be central to the pathogenesis of MPN, Jak2 inhibitors became one treatment option However, erythropoietin (EPO) is an essential hormone for sufficient production of red blood cells and also signals via the JAK2 pathway [119]. Consequently, anemia dose-dependently developed in patients who were treated with a JAK2 inhibitor (Ruxolitinib) and this was a dose-limiting adverse event [120,121]. In contrast, results from a phase II study for the treatment of myelofibrosis with a different Jak2 inhibitor (Momelotinib) surprisingly resulted even in an improvement of anemia [122]. Further dissection of the underlying mechanisms demonstrated that Momelotinib not only effectively inhibited Jak2 signaling, but also blocked ACVR1/ALK2-driven induction of hepcidin, resulting in an egress of iron from macrophages to sites of erythropoiesis [123].
Although these therapies are effective in lowering hepcidin levels and therefore ameliorate the anemia seen in chronic diseases, these therapies are probably not be suitable for sole treatment of AI because of potential side effects of these therapies, such as increased risk of infections due to impaired host responses [124]. A compromise might be a combinatorial therapeutic approach to target both the improvement of iron status and the treatment of infections.

3.2. Iron Supplementation and Iron Redistribution Therapies

Despite ongoing development of new treatment strategies and efforts towards personalized-based medicine, diseases, such as cancer, chronic heart failure, autoimmune diseases, and end stage kidney diseases, are proving unattainable because of persistent chronic inflammation. This being the case, anemia must be addressed via different approaches. Besides direct iron supplementation, iron redistribution strategies are emerging. The choice of the most appropriate therapy depends on the categorization of anemia whether there is pure AI with functional ID versus AI in combination with true ID. While iron replacement therapy appears to be mandatory in the latter setting, iron supplementation is questionable in patients with pure AI and strategies aiming at iron redistribution from macrophages to the circulation may be the more pragmatic approach.

3.2.1. Iron Supplementation

In general, iron can be directly supplemented either via the oral or intravenous (i.v.) route. However, this decision is based on several factors, including the availability and cost of drugs, the underlying disease, the degree of inflammation, therapeutic efficacy, and side effects, but also on patients’ compliance and convenience (Table 2). Oral iron may be used in ID and mild to moderate anemia, specifically among patients with a stable disease or only a low grade of inflammation [14,15,69,125]. Oral iron may also be effective in patients with AI and combined true ID due to the fact that ID-mediated inhibition of hepcidin expression dominates over inflammation-driven hepcidin induction [126,127]. Indications when i.v. iron therapy should be initiated are not that straight forward, based on the low grade of available evidence, and heterogeneity between guidelines for different disease entities [128]. However, i.v. iron may be used if oral iron therapy is ineffective, causes therapy-related side effects, and in patients with impaired oral iron absorption (Table 2). Of note, guidelines for recommendations whether to use oral or iv iron supplementation vary in different countries, in particular with regard to CKD. Examples include the Canadian guidelines, the Caring for Australians with Renal Impairment (CARI), the National Institute for Health and Care Excellence (NICE), and the Kidney Disease: Improving Global Outcomes (KDIGO), with each of them having their own guidelines and diagnostic algorithms, as well as choice of preferred administration route (oral vs. i.v.) [129,130,131,132]. This situation is far from being satisfactory as it causes deterrence and confusion among physicians and highlights the necessity of prospective clinical outcome data from rigorously conducted randomized controlled trials.
Nevertheless, the importance of i.v. iron supplementation among CKD patients became clear when the first human EPO preparation was licensed for use in dialysis-associated anemia nearly 30 years ago. Patients who suffered from EPO hypo-responsiveness experienced resolution of this condition with concomitant administration of i.v. iron. Hence, KDIGO guidelines propose that iron therapy should be aimed to treat ID, increase iron stores prior to initiation of therapy with erythropoiesis stimulating asgents (ESA), and enhance the response to these drugs [129].
In addition, within the last few years, concerns regarding the use of ESA (including EPO) for the treatment of anemia in CKD patients have been raised [133,134]. This was because of increased risk of adverse clinical outcomes, such as stroke and venous thromboembolic disease, culminating in high mortality [113,114,115,116]. Indeed, the US Food and Drug Administration (FDA) released a black box warning on the use of high EPO doses. Consequently, iron supplementation, either alone or in combination with ESA agents, are recommended as front-line options [129,130,131,132]. The latter is also related to findings of the TREAT (Trial to Reduce Cardiovascular Events with Aranesp® Therapy) study, which not only emphasized the possible risks related to ESA therapy, but also revealed that iron therapy increases hemoglobin levels and is capable of delaying the initiation need of ESA therapy [133,134]. Furthermore, another multicenter, prospective, and randomized study, FIND-CKD (Ferinject® assessment in patients with IDA and non-dialysis-dependent chronic kidney disease), reported that both i.v. and oral iron supplementation were capable of maintaining hemoglobin levels, thus reducing the dosages of ESA [135]. However, a recent randomized trial in non-dialyzed patients with CKD found that the use of i.v. iron was associated with an increased risk for adverse cardiovascular events and infections when compared to oral iron treatment [136].
Iron supplementation in patients with IBD is also still far from being consistent and many questions are still open, including the value of iron supplementation in subjects without anemia, or the preferred route of iron supplementation. Anemia seen in IBD is unique, as most patients suffer from AI together with ID, which is the consequence of continuous blood loss by the inflammatory mucosa and impaired iron intake as a consequence of malnutrition [125,137,138]. I.v. iron, as a sole treatment, has been shown to correct anemia in more than 80% of patients [139]. According to the European Crohn’s and Colitis Organisation (ECCO) guidelines published in 2015, iron supplementation is recommended whenever IDA is present. In contrast to recommendations made for CKD patients, iron supplementation aims to normalize hemoglobin levels in patients with IBD [140]. I.v. iron is recommended in patients with a clinically active disease, previous intolerance to oral iron, severe anemia (Hb < 10 g/dL), and who have initiated combination therapy with an ESA. Otherwise, oral iron therapy may be applied. However, several comparative studies, where i.v. versus oral iron supplementation was investigated, revealed that in IBD patients with AI and true ID and low disease activity, oral iron is as effective as i.v. iron to correct anemia [141,142,143,144,145,146]. Although AI is listed as a common cause for “non-IDA” in IBD subjects, no guidelines are provided regarding the practice of iron supplementation for these patients. Indeed, data from clinical trials on this issue are scarce. However, in view of published evidence that IBD patients with anemia have higher CRP values and a more active disease status, the necessity for further evaluation of this eventual relationship and its clinical management is evident [147,148].
In conclusion, no matter which subtype of AI is present, today’s evidence and treatment recommendations are based on altered biomarkers of haematology and inflammation and their correction, while end-point data on the effects of iron therapies (e.g., death, survival and disease resolution or progression) are almost completely lacking. Moreover, hardly any information from prospective trials is available regarding optimal therapeutic targets (e.g., hemoglobin or ferritin levels), which, however, may be different according to the underlying disease. One pioneering study (PIVOTAL for Proactive IV Iron Therapy in Hemodialysis Patients) addressed some of these end-points (risk of death, major adverse cardiovascular events, and infection) in patients undergoing hemodialysis and has just been published [149,150]. A high-dose regimen of i.v. iron (400 mg of iron sucrose per month, administered in a proactive fashion) was compared to a low-dose regimen (0 mg–400 mg of iron sucrose per month, administered in a reactive fashion). While the high iron regimen led to a reduced cumulative dose of administered ESA, there was no association with any of the end-points. How this study will influence existing guidelines and iron supplementation strategies remains to be seen.

3.2.2. Hepcidin Modulation

As mentioned above, hepcidin is the master regulator of systemic iron homeostasis, as this hormone is decisive for FPN expression, regulating iron efflux [151]. Thus, circulating levels of hepcidin determine the transfer of iron from the diet via the duodenum and release of iron from macrophages of the MPS. Since hepcidin is central to the pathophysiology of AI, several strategies that either modulate the synthesis of hepcidin or neutralize its activity have been developed [152,153]. The purpose of hepcidin modulation is to reverse iron retention in the MPS, thus enhancing iron availability for erythropoiesis in AI. As multiple causes lead to AI (e.g., negative impact of cytokines on erythropoiesis, impaired EPO activity), it needs to be investigated if increase of iron availability on its own is sufficient to effectively restore hemoglobin levels. Alternatively, a combination with an ESA could lead to a more favorable outcome [25,154,155,156,157,158].
The first approach that has been used were antibodies directed against hepcidin, which initially were only effective to reverse anemia in animal models when combined with ESA, whereas a subsequently developed human antibody modulated iron homeostasis in mice and cynomolgus monkeys without concomitant ESA administration [159].
Another approach for hepcidin neutralization is based on the use of antichalins (bioengineered lipocalin; small ligand-binding protein) or aptamers (also called Spiegelmer or Lexaptepid pegol L-stereoisomeric RNA aptamer). Indeed, these compounds have also been proven to be effective in preclinical models, and phase I trials have been successfully completed. In detail, positive data were obtained from a phase I study for the anticalin PRS-080 thus a phase II study was initiated, which is evaluating the effect of PRS-080 administration in anemic hemodialysis CKD patients (https://clinicaltrials.gov/ct2/show/NCT03325621) [160]. The outcomes of this clinical study on anti-hepcidin treatment are awaited. In addition, details on the impact of this compound on iron metabolism in cynomolgus monkeys has just recently been published [161]. In parallel, a placebo-controlled study on the safety, pharmacokinetics, and pharmacodynamics of the spiegelmer NOX-H94 in healthy humans demonstrated that hepcidin was inhibited dose-dependently, thus causing an increase in serum iron and Tf-Sat [162]. Furthermore, clinical phase II studies for the treatment of AI in patients suffering from multiple myeloma, low grade non-, or Hodgkin lymphoma, and ESA-hypo-responsive chronic hemodialysis patients have shown favorable effects, but cohorts were small, so further assessment is warranted [163,164].
As BMPs, specifically BMP2 and BMP6, are potent inducers of hepcidin, inhibition of the BMP-SMAD pathway is an attractive therapeutic approach to control hepcidin production [10,11,12,165]. Since this pathway is highly complex, involving different players, many possible targets can be contemplated [166]: First, BMP sequestration is one strategy. Therefore, BMP6 antibodies, a soluble hemojuvelin–Fc fusion protein, and modified heparins have been developed [167,168,169,170]. A phase I clinical trial of such a latter compound (Roneparstat, SST0001), which has competitive heparanase inhibitor properties, has been conducted in patients suffering from multiple myeloma, regarding its anti-myeloma effect, dosing, and safety profile (https://clinicaltrials.gov/ct2/show/record/NCT01764880) [171]. Impacts on hepcidin and iron metabolism have not been published yet.
Representing one step further down the BMP/SMAD pathway, efforts have been undertaken to target the BMP receptor (BMPR). TP-0184, a small-molecule inhibitor of ALK2 activity, has entered a phase I study this year (https://clinicaltrials.gov/ct2/show/NCT03429218), after having shown promising effects on hepcidin suppression in vitro and in preclinical mouse models [172]. Not only BMPR, but also BMP co-receptors, have been investigated as hepcidin lowering strategies. Two monoclonal antibodies targeting hemojuvelin have been developed, and are still in preclinical development [173].
A third reasonable approach to counteract hepcidin activity is to block hepcidin-induced internalization of FPN. Even though a phase II trial for such a stabilizing FPN antibody has been successfully completed in 2015, its further development has been stopped [174,175].
Notably, EPO at high doses can decrease hepcidin levels [176]. This effect is only of short duration and seems to be indirect, as signals derived from expanding erythroid progenitors in the bone marrow mediate this suppression [177,178]. Indeed, among CKD patients, no long-term effects of EPO on hepcidin levels have been observed, which, however, may also be partly related to impaired urinary hepcidin excretion [179]. In addition, HIF-prolyl hydroxylase inhibitors (HIF-PHIs), stabilizing HIFs, and thus activating HIF-controlled pathways, such as intrinsic EPO expression, have been reported to impact on iron homeostasis [93,180]. However, these effects can be traced back to transcriptional regulation, resulting in enhanced expression of specific iron transporters in the intestine (such as FPN and divalent metal transporter 1), thereby promoting iron absorption. Table 3 gives an overview of the drugs that directly or indirectly modulate hepcidin levels.

4. Perspectives

Anemia, being the final consequence of imbalances in iron homeostasis, in the setting of chronic diseases must be recognized as a clinical condition contributing to the morbidity of patients and awareness for ID must be improved. Indeed, due to the knowledge gap in clearly defining and diagnosing this condition, IDA, AI, and combined AI/ID are often used mutually. However, ID itself precedes anemia and should be detected, even outside the context of anemia. Efforts to counteract this common misconception have been made and a group of experts has proposed the following overarching definition for ID: “Iron deficiency is a health-related condition in which iron availability is insufficient to meet the body’s needs and which can be present with or without anemia” [181]. Of importance, ID is not only associated with anemia, but also higher morbidity and mortality among certain chronic diseases, which could be alleviated after treatment [100,182,183,184,185]. For example, in patients suffering from chronic heart failure, ID has been shown to adversely impact performance status and quality of life, including prolonged hospitalizations independent of anemia [186,187,188]. Another significant upcoming challenge will be the management of anemia of the elderly, which often has a heterogeneous and multifactorial etiology, but is also specifically related to age-related changes [189].
Although AI is a condition that should be treated, there is also an evolutionary rational for iron restriction during inflammation: Iron restriction is beneficial during acute infections, especially to withhold iron from circulating microbes [19,39]. Malaria represents one of the best studied examples in this context [65,190,191]. The fact that red blood cells are the host for plasmodia highlights their dependency on iron metabolism. An important finding was just recently made to better understand underlying pathomechanisms: FPN expression on red blood cells is critical to prevent detrimental intracellular iron accumulation and hemolysis, all in all leading to a more severe course of malaria. Of interest, these authors found that a human mutation in FPN (Q248H), which is unresponsive to hepcidin-mediated degradation, has been positively selected in sub-Saharan African populations [192]. Thus, anti-hepcidin treatment strategies as listed above could be discussed as a treatment option for malaria in the future. Despite anemia being associated with this infection, iron supplementation has been shown to be detrimental. This is also in line with studies showing that iron supplementation in children of developing countries resulted in higher morbidity and mortality from infections [193,194].
Moreover, there is increasing evidence for the role of iron availability for the gut microbiome and oral versus i.v. iron have different effects on the composition of the microbiome [143,195,196]. This is of interest, because the composition of the gut microbiome was found to play decisive roles for the progression of IBD and carcinogenesis in different mouse models [197]. Further workup in vitro showed that certain iron formulations (ferric citrate and ferric ethylenediaminetetraacetic acid) also bear the risk of exacerbation of colon cancer advancement in an amphiregulin-dependent fashion, however, it needs to be defined whether or not the dosages used in such models are relevant for humans [198].
Another issue of general importance are the effects of iron supplementation or hepcidin targeting strategies on immune regulation. This is based on the observation that iron impacts on the differentiation and proliferation of immune cells, but also directly impacts on immune effector pathways either by promoting oxygen radical formation or inhibiting pro-inflammatory cytokine production or anti-microbial immune effector pathways of macrophages [29,199,200]. Pre-clinical and clinical models have shown that iron supplementation reduces TNFα formation in CKD patients while negatively impacting on the host response in mammalian models of invasive fungal infection [44,201]. Thus, depending on the underlying disease, iron supplementation could have disease modifying effects through its regulatory effects on the immune function [44,202].

5. Conclusions

Anemia and ID in the setting of chronic inflammatory diseases are leading causes of morbidity worldwide. While we have gained significant knowledge on the mechanism underlying iron misdistribution and development of AI, highlighting the role of immune mediators and the iron hormone hepcidin, there is still the need for reliable biomarkers to evaluate iron homeostasis in patients suffering from inflammatory diseases and to choose the best therapy or to predict its efficacy. Specifically, distinction between AI versus AI combined with true iron deficiency is of importance because these groups of patients may likewise need different iron redistribution therapies. The development of new drugs (e.g., hepcidin antagonists) and the improvement of old drugs (novel formulation for oral and intravenous iron preparations) are the subject of future investigations. Although there is good evidence that iron supplementation improves quality of life, the effect of iron supplementation on the course of an underlying disease or associated co-morbidities are poorly understood. There is only limited information on therapeutic start- and end-points for iron supplementation and anemia correction in such patients. However, negligence of anemia and iron deficiency may also exacerbate the underlying disease state and cause clinical deterioration [203,204]. Thus, there is still a lot to learn to optimize and personalize treatment in subjects with AI. Therefore, investigations through pre-clinical models, but also through prospective randomized trials, are urgently needed to gain more detailed insights into this clinically very frequent, but poorly understood condition.

Author Contributions

Conceptualization, V.P.; I.T. and G.W.; writing—original draft preparation, V.P.; writing—review and editing, I.T. and G.W.; supervision, G.W.; project administration, G.W.

Funding

This research was funded by the doctoral college project [W1253 HOROS to V.P. and G.W.], the Austrian Research Funds (FWF) project [P 28302 to I.T.], the “Verein zur Förderung von Forschung und Weiterbildung in Infektiologie und Immunologie, Innsbruck” [to G.W.], the Christian Doppler Society, Austria [to G.W.]

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ACDAnemia of chronic disease
ACVRActivin A receptor
AIAnemia of inflammation
ALKActivin receptor-like kinase
BMPBone morphogenic protein
BMPRBone morphogenic protein receptor
CKDChronic kidney disease
CRPC-reactive protein
EPOErythropoietin
EPORErythropoietin receptor
ERFEErythroferrone
ESAErythropoiesis stimulating agent
FDAFood and drug administration
FPNFerroportin-1 AKA SLC40A1
HbHemoglobin
HIFsHypoxia inducible factors
HIF-PHDHypoxia inducible factor prolyl hydroxylase inhibitors
IBDInflammatory bowel disease
IDIron deficiency
IDAIron deficiency anemia
IFNγInterferon gamma
ILInterleukin
IL6RInterleukin 6 receptor
i.v.Intravenous
JAKJanus kinase
MCDMulticentric Castleman’s disease
MCHMean corpuscular hemoglobin
MCVMean corpuscular volume
MPNMyeloproliferative neoplasms
MPSMononuclear Phagocyte system
RARheumatoid arthritis
RBCRed blood cell
SMADHomologues of Sma and Mad (mothers against decapentaplegic) proteins
STATSignal transducer and activator of transcription
sTfRSoluble transferrin receptor
TfRTransferrin receptor
Tf-SatTransferrin saturation
TNFαTumor necrosis factor alpha

References

  1. Muckenthaler, M.U.; Rivella, S.; Hentze, M.W.; Galy, B. A Red Carpet for Iron Metabolism. Cell 2017, 168, 344–361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Nemeth, E.; Rivera, S.; Gabayan, V.; Keller, C.; Taudorf, S.; Pedersen, B.K.; Ganz, T. IL-6 mediates hypoferremia of inflammation by inducing the synthesis of the iron regulatory hormone hepcidin. J. Clin. Investig. 2004, 113, 1271–1276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Nemeth, E.; Tuttle, M.S.; Powelson, J.; Vaughn, M.B.; Donovan, A.; Ward, D.M.; Ganz, T.; Kaplan, J. Hepcidin regulates cellular iron efflux by binding to ferroportin and inducing its internalization. Science 2004, 306, 2090–2093. [Google Scholar] [CrossRef] [PubMed]
  4. Donovan, A.; Brownlie, A.; Zhou, Y.; Shepard, J.; Pratt, S.J.; Moynihan, J.; Paw, B.H.; Drejer, A.; Barut, B.; Zapata, A.; et al. Positional cloning of zebrafish ferroportin1 identifies a conserved vertebrate iron exporter. Nature 2000, 403, 776–781. [Google Scholar] [CrossRef] [PubMed]
  5. Donovan, A.; Lima, C.A.; Pinkus, J.L.; Pinkus, G.S.; Zon, L.I.; Robine, S.; Andrews, N.C. The iron exporter ferroportin/Slc40a1 is essential for iron homeostasis. Cell Metab. 2005, 1, 191–200. [Google Scholar] [CrossRef] [PubMed]
  6. Nicolas, G.; Chauvet, C.; Viatte, L.; Danan, J.L.; Bigard, X.; Devaux, I.; Beaumont, C.; Kahn, A.; Vaulont, S. The gene encoding the iron regulatory peptide hepcidin is regulated by anemia, hypoxia, and inflammation. J. Clin. Investig. 2002, 110, 1037–1044. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Verga Falzacappa, M.V.; Spasic, M.V.; Kessler, R.; Stolte, J.; Hentze, M.W.; Muckenthaler, M.U. STAT3 mediates hepatic hepcidin expression and its inflammatory stimulation. Blood 2007, 109, 353–358. [Google Scholar] [CrossRef] [Green Version]
  8. Meynard, D.; Kautz, L.; Darnaud, V.; Canonne-Hergaux, F.; Coppin, H.; Roth, M.-P. Lack of the bone morphogenetic protein BMP6 induces massive iron overload. Nat. Genet. 2009, 41, 478–481. [Google Scholar] [CrossRef]
  9. Fillebeen, C.; Wilkinson, N.; Charlebois, E.; Katsarou, A.; Wagner, J.; Pantopoulos, K. Hepcidin-mediated hypoferremic response to acute inflammation requires a threshold of Bmp6/Hjv/Smad signaling. Blood 2018, 132, 1829–1841. [Google Scholar] [CrossRef]
  10. Koch, P.-S.; Olsavszky, V.; Ulbrich, F.; Sticht, C.; Demory, A.; Leibing, T.; Henzler, T.; Meyer, M.; Zierow, J.; Schneider, S.; et al. Angiocrine Bmp2 signaling in murine liver controls normal iron homeostasis. Blood 2017, 129, 415–419. [Google Scholar] [CrossRef]
  11. Canali, S.; Wang, C.-Y.; Zumbrennen-Bullough, K.B.; Bayer, A.; Babitt, J.L. Bone morphogenetic protein 2 controls iron homeostasis in mice independent of Bmp6. Am. J. Hematol. 2017, 92, 1204–1213. [Google Scholar] [CrossRef] [PubMed]
  12. Canali, S.; Zumbrennen-Bullough, K.B.; Core, A.B.; Wang, C.-Y.; Nairz, M.; Bouley, R.; Swirski, F.K.; Babitt, J.L. Endothelial cells produce bone morphogenetic protein 6 required for iron homeostasis in mice. Blood 2017, 129, 405–414. [Google Scholar] [CrossRef] [PubMed]
  13. Armitage, A.E.; Eddowes, L.A.; Gileadi, U.; Cole, S.; Spottiswoode, N.; Selvakumar, T.A.; Ho, L.P.; Townsend, A.R.M.; Drakesmith, H. Hepcidin regulation by innate immune and infectious stimuli. Blood 2011, 118, 4129–4139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Camaschella, C. New insights into iron deficiency and iron deficiency anemia. Blood Rev. 2017, 31, 225–233. [Google Scholar] [CrossRef] [PubMed]
  15. Weiss, G.; Goodnough, L.T. Anemia of chronic disease. N. Engl. J. Med. 2005, 352, 1011–1023. [Google Scholar] [CrossRef] [PubMed]
  16. Theurl, I.; Aigner, E.; Theurl, M.; Nairz, M.; Seifert, M.; Schroll, A.; Sonnweber, T.; Eberwein, L.; Witcher, D.R.; Murphy, A.T.; et al. Regulation of iron homeostasis in anemia of chronic disease and iron deficiency anemia: Diagnostic and therapeutic implications. Blood 2009, 113, 5277–5286. [Google Scholar] [CrossRef] [PubMed]
  17. Stefanova, D.; Raychev, A.; Arezes, J.; Ruchala, P.; Gabayan, V.; Skurnik, M.; Dillon, B.J.; Horwitz, M.A.; Ganz, T.; Bulut, Y.; et al. Endogenous hepcidin and its agonist mediate resistance to selected infections by clearing non-transferrin-bound iron. Blood 2017, 130, 245–257. [Google Scholar] [CrossRef] [PubMed]
  18. Michels, K.R.; Zhang, Z.; Bettina, A.M.; Cagnina, R.E.; Stefanova, D.; Burdick, M.D.; Vaulont, S.; Nemeth, E.; Ganz, T.; Mehrad, B. Hepcidin-mediated iron sequestration protects against bacterial dissemination during pneumonia. JCI Insight 2017, 2, e92002. [Google Scholar] [CrossRef] [PubMed]
  19. Soares, M.P.; Weiss, G. The Iron age of host-microbe interactions. EMBO Rep. 2015, 16, 1482–1500. [Google Scholar] [CrossRef] [Green Version]
  20. Van Assche, G.; Dignass, A.; Bokemeyer, B.; Danese, S.; Gionchetti, P.; Moser, G.; Beaugerie, L.; Gomollón, F.; Häuser, W.; Herrlinger, K.; et al. European Crohn’s and Colitis Organisation Second European evidence-based consensus on the diagnosis and management of ulcerative colitis Part 3: Special situations. J. Crohns Colitis 2013, 7, 1–33. [Google Scholar] [CrossRef]
  21. Khan, A.N.; Hameed, A.; Naeem, M.; Murtaza, G.; Shah, N.A. Iron status and hemoglobin level in chronic renal insufficiency. Med. Forum Mon. 2008, 19, 11–15. [Google Scholar] [CrossRef]
  22. Macdougall, I.C.; Bircher, A.J.; Eckardt, K.-U.; Obrador, G.T.; Pollock, C.A.; Stenvinkel, P.; Swinkels, D.W.; Wanner, C.; Weiss, G.; Chertow, G.M.; et al. Iron management in chronic kidney disease: Conclusions from a “Kidney Disease: Improving Global Outcomes” (KDIGO) Controversies Conference. Kidney Int. 2016, 89, 28–39. [Google Scholar] [CrossRef] [PubMed]
  23. Del Vecchio, L.; Locatelli, F. Anemia in chronic kidney disease patients: Treatment recommendations and emerging therapies. Expert Rev. Hematol. 2014, 7, 495–506. [Google Scholar] [CrossRef] [PubMed]
  24. Heath, J.; Weiss, J.; Lavau, C.; Wechsler, D. Iron Deprivation in Cancer––Potential Therapeutic Implications. Nutrients 2013, 5, 2836–2859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Weiss, G.; Schett, G. Anaemia in inflammatory rheumatic diseases. Nat. Rev. Rheumatol. 2013, 9, 205–215. [Google Scholar] [CrossRef] [PubMed]
  26. Weiss, G.; Ganz, T.; Goodnough, L.T. Anemia of inflammation. Blood 2018. [Google Scholar] [CrossRef] [PubMed]
  27. Busti, F.; Marchi, G.; Ugolini, S.; Castagna, A.; Girelli, D. Anemia and Iron Deficiency in Cancer Patients: Role of Iron Replacement Therapy. Pharmaceuticals 2018, 11, 94. [Google Scholar] [CrossRef]
  28. Camaschella, C. Iron-Deficiency Anemia. N. Engl. J. Med. 2015, 373, 484–486. [Google Scholar] [CrossRef]
  29. Weiss, G.; Schaible, U.E. Macrophage defense mechanisms against intracellular bacteria. Immunol. Rev. 2015, 264, 182–203. [Google Scholar] [CrossRef] [Green Version]
  30. Recalcati, S.; Locati, M.; Gammella, E.; Invernizzi, P.; Cairo, G. Iron levels in polarized macrophages: Regulation of immunity and autoimmunity. Autoimmun. Rev. 2012, 11, 883–889. [Google Scholar] [CrossRef]
  31. Nairz, M.; Theurl, I.; Swirski, F.K.; Weiss, G. “Pumping iron”—How macrophages handle iron at the systemic, microenvironmental, and cellular levels. Pflügers Arch. Eur. J. Physiol. 2017, 469, 397–418. [Google Scholar] [CrossRef] [PubMed]
  32. Drakesmith, H.; Prentice, A.M. Hepcidin and the iron-infection axis. Science 2012, 338, 768–772. [Google Scholar] [CrossRef] [PubMed]
  33. Ludwiczek, S.; Aigner, E.; Theurl, I.; Weiss, G. Cytokine-mediated regulation of iron transport in human monocytic cells. Blood 2003, 101, 4148–4154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Weiss, G.; Bogdan, C.; Hentze, M.W. Pathways for the regulation of macrophage iron metabolism by the anti-inflammatory cytokines IL-4 and IL-13. J. Immunol. 1997, 158, 420–425. [Google Scholar] [PubMed]
  35. Pantopoulos, K.; Weiss, G.; Hentze, M.W. Nitric oxide and oxidative stress (H2O2) control mammalian iron metabolism by different pathways. Mol. Cell. Biol. 1996, 16, 3781–3788. [Google Scholar] [CrossRef] [PubMed]
  36. Byrd, T.F.; Horwitz, M.A. Regulation of transferrin receptor expression and ferritin content in human mononuclear phagocytes. Coordinate upregulation by iron transferrin and downregulation by interferon gamma. J. Clin. Investig. 1993, 91, 969–976. [Google Scholar] [CrossRef] [PubMed]
  37. Mulero, V.; Brock, J.H. Regulation of iron metabolism in murine J774 macrophages: Role of nitric oxide-dependent and -independent pathways following activation with gamma interferon and lipopolysaccharide. Blood 1999, 94, 2383–2389. [Google Scholar] [CrossRef]
  38. Tilg, H.; Ulmer, H.; Kaser, A.; Weiss, G. Role of IL-10 for Induction of Anemia During Inflammation. J. Immunol. 2002, 169, 2204–2209. [Google Scholar] [CrossRef] [Green Version]
  39. Ganz, T.; Nemeth, E. Iron homeostasis in host defence and inflammation. Nat. Rev. Immunol. 2015, 15, 500–510. [Google Scholar] [CrossRef] [Green Version]
  40. Nairz, M.; Schroll, A.; Haschka, D.; Dichtl, S.; Tymoszuk, P.; Demetz, E.; Moser, P.; Haas, H.; Fang, F.C.; Theurl, I.; et al. Genetic and Dietary Iron Overload Differentially Affect the Course of Salmonella Typhimurium Infection. Front. Cell. Infect. Microbiol. 2017, 7, 110. [Google Scholar] [CrossRef]
  41. Barber, M.F.; Elde, N.C. Escape from bacterial iron piracy through rapid evolution of transferrin. Science 2014, 346, 1362–1366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Weiss, G.; Fuchs, D.; Hausen, A.; Reibnegger, G.; Werner, E.R.; Werner-Felmayer, G.; Wachter, H. Iron modulates interferon-gamma effects in the human myelomonocytic cell line THP-1. Exp. Hematol. 1992, 20, 605–610. [Google Scholar] [PubMed]
  43. Weiss, G.; Werner-Felmayer, G.; Werner, E.R.; Grunewald, K.; Wachter, H.; Hentze, M.W. Iron regulates nitric oxide synthase activity by controlling nuclear transcription. J. Exp. Med. 1994, 180, 969–976. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Mencacci, A.; Cenci, E.; Boelaert, J.R.; Bucci, P.; Mosci, P.; Fè d’Ostiani, C.; Bistoni, F.; Romani, L. Iron overload alters innate and T helper cell responses to Candida albicans in mice. J. Infect. Dis. 1997, 175, 1467–1476. [Google Scholar] [CrossRef] [PubMed]
  45. Oexle, H.; Kaser, A.; Möst, J.; Bellmann-Weiler, R.; Werner, E.R.; Werner-Felmayer, G.; Weiss, G. Pathways for the regulation of interferon-gamma-inducible genes by iron in human monocytic cells. J. Leukoc. Biol. 2003, 74, 287–294. [Google Scholar] [CrossRef] [PubMed]
  46. Cherayil, B.J.; Ellenbogen, S.; Shanmugam, N.N. Iron and intestinal immunity. Curr. Opin. Gastroenterol. 2011, 27, 523–528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Cassat, J.E.; Skaar, E.P. Iron in Infection and Immunity. Cell Host Microbe 2013, 13, 509–519. [Google Scholar] [CrossRef] [Green Version]
  48. Theurl, I.; Theurl, M.; Seifert, M.; Mair, S.; Nairz, M.; Rumpold, H.; Zoller, H.; Bellmann-Weiler, R.; Niederegger, H.; Talasz, H.; et al. Autocrine formation of hepcidin induces iron retention in human monocytes. Blood 2008, 111, 2392–2399. [Google Scholar] [CrossRef]
  49. Peyssonnaux, C.; Zinkernagel, A.S.; Datta, V.; Lauth, X.; Johnson, R.S.; Nizet, V. TLR4-dependent hepcidin expression by myeloid cells in response to bacterial pathogens. Blood 2006, 107, 3727–3732. [Google Scholar] [CrossRef] [Green Version]
  50. Nairz, M.; Schleicher, U.; Schroll, A.; Sonnweber, T.; Theurl, I.; Ludwiczek, S.; Talasz, H.; Brandacher, G.; Moser, P.L.; Muckenthaler, M.U.; et al. Nitric oxide-mediated regulation of ferroportin-1 controls macrophage iron homeostasis and immune function in Salmonella infection. J. Exp. Med. 2013, 210, 855–873. [Google Scholar] [CrossRef] [Green Version]
  51. Wu, A.; Tymoszuk, P.; Haschka, D.; Heeke, S.; Dichtl, S.; Petzer, V.; Seifert, M.; Hilbe, R.; Sopper, S.; Talasz, H.; et al. Salmonella Utilizes Zinc To Subvert Antimicrobial Host Defense of Macrophages via Modulation of NF-κB Signaling. Infect. Immun. 2017, 85. [Google Scholar] [CrossRef] [PubMed]
  52. Paradkar, P.N.; De Domenico, I.; Durchfort, N.; Zohn, I.; Kaplan, J.; Ward, D.M. Iron depletion limits intracellular bacterial growth in macrophages. Blood 2008, 112, 866–874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Ben-Othman, R.; Flannery, A.R.; Miguel, D.C.; Ward, D.M.; Kaplan, J.; Andrews, N.W. Leishmania-mediated inhibition of iron export promotes parasite replication in macrophages. PLoS Pathog. 2014, 10, e1003901. [Google Scholar] [CrossRef] [PubMed]
  54. Wu, Q.; Shen, Y.; Tao, Y.; Wei, J.; Wang, H.; An, P.; Zhang, Z.; Gao, H.; Zhou, T.; Wang, F.; et al. Hemojuvelin regulates the innate immune response to peritoneal bacterial infection in mice. Cell Discov. 2017, 3, 17028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Neumann, W.; Hadley, R.C.; Nolan, E.M. Transition metals at the host–pathogen interface: How Neisseria exploit human metalloproteins for acquiring iron and zinc. Essays Biochem. 2017, 61, 211–223. [Google Scholar] [CrossRef]
  56. Weiss, G.; Carver, P.L. Role of divalent metals in infectious disease susceptibility and outcome. Clin. Microbiol. Infect. 2018, 24, 16–23. [Google Scholar] [CrossRef] [PubMed]
  57. Skaar, E.P.; Raffatellu, M. Metals in infectious diseases and nutritional immunity. Metallomics 2015, 7, 926–928. [Google Scholar] [CrossRef] [PubMed]
  58. Kim, A.; Fung, E.; Parikh, S.G.; Valore, E.V.; Gabayan, V.; Nemeth, E.; Ganz, T. A mouse model of anemia of inflammation: Complex pathogenesis with partial dependence on hepcidin. Blood 2014, 123, 1129–1136. [Google Scholar] [CrossRef] [PubMed]
  59. Nairz, M.; Ferring-Appel, D.; Casarrubea, D.; Sonnweber, T.; Viatte, L.; Schroll, A.; Haschka, D.; Fang, F.C.; Hentze, M.W.; Weiss, G.; et al. Iron Regulatory Proteins Mediate Host Resistance to Salmonella Infection. Cell Host Microbe 2015, 18, 254–261. [Google Scholar] [CrossRef] [PubMed]
  60. Guida, C.; Altamura, S.; Klein, F.A.; Galy, B.; Boutros, M.; Ulmer, A.J.; Hentze, M.W.; Muckenthaler, M.U. A novel inflammatory pathway mediating rapid hepcidin-independent hypoferremia. Blood 2015, 125, 2265–2275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Moreira, A.C.; Neves, J.V.; Silva, T.; Oliveira, P.; Gomes, M.S.; Rodrigues, P.N. Hepcidin-(in)dependent mechanisms of iron metabolism regulation during infection by Listeria and Salmonella. Infect. Immun. 2017, 85, IAI-00353. [Google Scholar] [CrossRef] [PubMed]
  62. Mitterstiller, A.-M.; Haschka, D.; Dichtl, S.; Nairz, M.; Demetz, E.; Talasz, H.; Soares, M.P.; Einwallner, E.; Esterbauer, H.; Fang, F.C.; et al. Heme oxygenase 1 controls early innate immune response of macrophages to Salmonella Typhimurium infection. Cell. Microbiol. 2016, 18, 1374–1389. [Google Scholar] [CrossRef] [PubMed]
  63. Weis, S.; Carlos, A.R.; Moita, M.R.; Singh, S.; Blankenhaus, B.; Cardoso, S.; Larsen, R.; Rebelo, S.; Schäuble, S.; Del Barrio, L.; et al. Metabolic Adaptation Establishes Disease Tolerance to Sepsis. Cell 2017, 169, 1263.e14–1275.e14. [Google Scholar] [CrossRef] [PubMed]
  64. Gozzelino, R.; Jeney, V.; Soares, M.P. Mechanisms of cell protection by heme oxygenase-1. Annu. Rev. Pharmacol. Toxicol. 2010, 50, 323–354. [Google Scholar] [CrossRef] [PubMed]
  65. Portugal, S.; Carret, C.; Recker, M.; Armitage, A.E.; Gonçalves, L.A.; Epiphanio, S.; Sullivan, D.; Roy, C.; Newbold, C.I.; Drakesmith, H.; et al. Host-mediated regulation of superinfection in malaria. Nat. Med. 2011, 17, 732–737. [Google Scholar] [CrossRef] [PubMed]
  66. Stockman, R. The Treatment of Chlorosis by Iron and some Other Drugs. Br. Med. J. 1893, 1, 942–944. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Nairz, M.; Haschka, D.; Demetz, E.; Weiss, G. Iron at the interface of immunity and infection. Front. Pharmacol. 2014, 5, 152. [Google Scholar] [CrossRef] [PubMed]
  68. Weiss, G. Anemia of Chronic Disorders: New Diagnostic Tools and New Treatment Strategies. Semin. Hematol. 2015, 52, 313–320. [Google Scholar] [CrossRef]
  69. Nairz, M.; Theurl, I.; Wolf, D.; Weiss, G. Iron deficiency or anemia of inflammation?: Differential diagnosis and mechanisms of anemia of inflammation. Wien. Med. Wochenschr. 2016, 166, 411–423. [Google Scholar] [CrossRef]
  70. Torti, F.M.; Torti, S.V. Regulation of ferritin genes and protein. Blood 2002, 99, 3505–3516. [Google Scholar] [CrossRef] [Green Version]
  71. Garcia-Casal, M.N.; Peña-Rosas, J.P.; Pasricha, S.-R. Rethinking ferritin cutoffs for iron deficiency and overload. Lancet Haematol. 2014, 1, e92–e94. [Google Scholar] [CrossRef]
  72. Suchdev, P.S.; Namaste, S.M.; Aaron, G.J.; Raiten, D.J.; Brown, K.H.; Flores-Ayala, R. BRINDA Working Group Overview of the Biomarkers Reflecting Inflammation and Nutritional Determinants of Anemia (BRINDA) Project. Adv. Nutr. 2016, 7, 349–356. [Google Scholar] [CrossRef] [PubMed]
  73. Gale, E.; Torrance, J.; Bothwell, T. The quantitative estimation of total iron stores in human bone marrow. J. Clin. Investig. 1963, 42, 1076–1082. [Google Scholar] [CrossRef] [PubMed]
  74. Grote Beverborg, N.; Klip, I.T.; Meijers, W.C.; Voors, A.A.; Vegter, E.L.; van der Wal, H.H.; Swinkels, D.W.; van Pelt, J.; Mulder, A.B.; Bulstra, S.K.; et al. Definition of Iron Deficiency Based on the Gold Standard of Bone Marrow Iron Staining in Heart Failure Patients. Circ. Heart Fail. 2018, 11, e004519. [Google Scholar] [CrossRef] [PubMed]
  75. Speeckaert, M.M.; Speeckaert, R.; Delanghe, J.R. Biological and clinical aspects of soluble transferrin receptor. Crit. Rev. Clin. Lab. Sci. 2010, 47, 213–228. [Google Scholar] [CrossRef] [PubMed]
  76. Brugnara, C.; Mohandas, N. Red cell indices in classification and treatment of anemias. Curr. Opin. Hematol. 2013, 20, 222–230. [Google Scholar] [CrossRef] [PubMed]
  77. Harms, K.; Kaiser, T. Beyond soluble transferrin receptor: Old challenges and new horizons. Best Pract. Res. Clin. Endocrinol. Metab. 2015, 29, 799–810. [Google Scholar] [CrossRef] [PubMed]
  78. Aguilar, R.; Moraleda, C.; Quintó, L.; Renom, M.; Mussacate, L.; Macete, E.; Aguilar, J.L.; Alonso, P.L.; Menéndez, C. Challenges in the Diagnosis of Iron Deficiency in Children Exposed to High Prevalence of Infections. PLoS ONE 2012, 7, e50584. [Google Scholar] [CrossRef]
  79. Castel, R.; Tax, M.G.H.M.; Droogendijk, J.; Leers, M.P.G.; Beukers, R.; Levin, M.-D.; Sonneveld, P.; Berendes, P.B. The transferrin/log(ferritin) ratio: A new tool for the diagnosis of iron deficiency anemia. Clin. Chem. Lab. Med. 2012, 50, 1343–1349. [Google Scholar] [CrossRef]
  80. Lasocki, S.; Baron, G.; Driss, F.; Westerman, M.; Puy, H.; Boutron, I.; Beaumont, C.; Montravers, P. Diagnostic accuracy of serum hepcidin for iron deficiency in critically ill patients with anemia. Intensive Care Med. 2010, 36, 1044–1048. [Google Scholar] [CrossRef]
  81. Shu, T.; Jing, C.; Lv, Z.; Xie, Y.; Xu, J.; Wu, J. Hepcidin in tumor-related iron deficiency anemia and tumor-related anemia of chronic disease: Pathogenic mechanisms and diagnosis. Eur. J. Haematol. 2015, 94, 67–73. [Google Scholar] [CrossRef] [PubMed]
  82. Bergamaschi, G.; Di Sabatino, A.; Albertini, R.; Costanzo, F.; Guerci, M.; Masotti, M.; Pasini, A.; Massari, A.; Campostrini, N.; Corbella, M.; et al. Serum Hepcidin in Inflammatory Bowel Diseases. Inflamm. Bowel Dis. 2013, 19, 2166–2172. [Google Scholar] [CrossRef] [PubMed]
  83. Van Santen, S.; van Dongen-Lases, E.C.; de Vegt, F.; Laarakkers, C.M.M.; van Riel, P.L.C.M.; van Ede, A.E.; Swinkels, D.W. Hepcidin and hemoglobin content parameters in the diagnosis of iron deficiency in rheumatoid arthritis patients with anemia. Arthritis Rheum. 2011, 63, 3672–3680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Bregman, D.B.; Morris, D.; Koch, T.A.; He, A.; Goodnough, L.T. Hepcidin levels predict nonresponsiveness to oral iron therapy in patients with iron deficiency anemia. Am. J. Hematol. 2013, 88, 97–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Steensma, D.P.; Sasu, B.J.; Sloan, J.A.; Tomita, D.K.; Loprinzi, C.L. Serum hepcidin levels predict response to intravenous iron and darbepoetin in chemotherapy-associated anemia. Blood 2015, 125, 3669–3671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Theurl, M.; Nairz, M.; Schroll, A.; Sonnweber, T.; Asshoff, M.; Haschka, D.; Seifert, M.; Willenbacher, W.; Wilflingseder, D.; Posch, W.; et al. Hepcidin as a predictive factor and therapeutic target in erythropoiesis-stimulating agent treatment for anemia of chronic disease in rats. Haematologica 2014, 99, 1516–1524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Takasawa, K.; Takaeda, C.; Maeda, T.; Ueda, N.; Takasawa, K.; Takaeda, C.; Maeda, T.; Ueda, N. Hepcidin-25, Mean Corpuscular Volume, and Ferritin as Predictors of Response to Oral Iron Supplementation in Hemodialysis Patients. Nutrients 2014, 7, 103–118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Gaillard, C.A.; Bock, A.H.; Carrera, F.; Eckardt, K.-U.; Van Wyck, D.B.; Bansal, S.S.; Cronin, M.; Meier, Y.; Larroque, S.; Roger, S.D.; et al. Hepcidin Response to Iron Therapy in Patients with Non-Dialysis Dependent CKD: An Analysis of the FIND-CKD Trial. PLoS ONE 2016, 11, e0157063. [Google Scholar] [CrossRef]
  89. Ford, B.A.; Eby, C.S.; Scott, M.G.; Coyne, D.W. Intra-individual variability in serum hepcidin precludes its use as a marker of iron status in hemodialysis patients. Kidney Int. 2010, 78, 769–773. [Google Scholar] [CrossRef]
  90. Tessitore, N.; Girelli, D.; Campostrini, N.; Bedogna, V.; Pietro Solero, G.; Castagna, A.; Melilli, E.; Mantovani, W.; De Matteis, G.; Olivieri, O.; et al. Hepcidin is not useful as a biomarker for iron needs in haemodialysis patients on maintenance erythropoiesis-stimulating agents. Nephrol. Dial. Transplant. 2010, 25, 3996–4002. [Google Scholar] [CrossRef] [Green Version]
  91. Kautz, L.; Jung, G.; Nemeth, E.; Ganz, T. Erythroferrone contributes to recovery from anemia of inflammation. Blood 2014, 124, 2569–2574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Sonnweber, T.; Nachbaur, D.; Schroll, A.; Nairz, M.; Seifert, M.; Demetz, E.; Haschka, D.; Mitterstiller, A.M.; Kleinsasser, A.; Burtscher, M.; et al. Hypoxia induced downregulation of hepcidin is mediated by platelet derived growth factor BB. Gut 2014, 63, 1951–1959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Haase, V.H. Regulation of erythropoiesis by hypoxia-inducible factors. Blood Rev. 2013, 27, 41–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Hanudel, M.R.; Rappaport, M.; Chua, K.; Gabayan, V.; Qiao, B.; Jung, G.; Salusky, I.B.; Ganz, T.; Nemeth, E. Levels of the erythropoietin-responsive hormone erythroferrone in mice and humans with chronic kidney disease. Haematologica 2018, 103, e141–e142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Ganz, T.; Jung, G.; Naeim, A.; Ginzburg, Y.; Pakbaz, Z.; Walter, P.B.; Kautz, L.; Nemeth, E. Immunoassay for human serum erythroferrone. Blood 2017, 130, 1243–1246. [Google Scholar] [CrossRef]
  96. Mastrogiannaki, M.; Matak, P.; Mathieu, J.R.R.; Delga, S.; Mayeux, P.; Vaulont, S.; Peyssonnaux, C. Hepatic hypoxia-inducible factor-2 down-regulates hepcidin expression in mice through an erythropoietin-mediated increase in erythropoiesis. Haematologica 2012, 97, 827–834. [Google Scholar] [CrossRef]
  97. Van der Vorm, L.N.; Hendriks, J.C.M.; Laarakkers, C.M.; Klaver, S.; Armitage, A.E.; Bamberg, A.; Geurts-Moespot, A.J.; Girelli, D.; Herkert, M.; Itkonen, O.; et al. Toward Worldwide Hepcidin Assay Harmonization: Identification of a Commutable Secondary Reference Material. Clin. Chem. 2016, 62, 993–1001. [Google Scholar] [CrossRef] [Green Version]
  98. Burisch, J.; Vegh, Z.; Katsanos, K.H.; Christodoulou, D.K.; Lazar, D.; Goldis, A.; O’Morain, C.; Fernandez, A.; Pereira, S.; Myers, S.; et al. Occurrence of Anaemia in the First Year of Inflammatory Bowel Disease in a European Population-based Inception Cohort—An ECCO-EpiCom Study. J. Crohns Colitis 2017, 11, 1213–1222. [Google Scholar] [CrossRef] [Green Version]
  99. Ershler, W.B.; Chen, K.; Reyes, E.B.; Dubois, R. Economic Burden of Patients with Anemia in Selected Diseases. Value Heal. 2005, 8, 629–638. [Google Scholar] [CrossRef]
  100. Wouters, H.J.C.M.; van der Klauw, M.M.; de Witte, T.; Stauder, R.; Swinkels, D.W.; Wolffenbuttel, B.H.R.; Huls, G. Association of anemia with health-related quality of life and survival: A large population-based cohort study. Haematologica 2018. [Google Scholar] [CrossRef]
  101. Nishimoto, N.; Sasai, M.; Shima, Y.; Nakagawa, M.; Matsumoto, T.; Shirai, T.; Kishimoto, T.; Yoshizaki, K. Improvement in Castleman’s disease by humanized anti-interleukin-6 receptor antibody therapy. Blood 2000, 95, 56–61. [Google Scholar] [PubMed]
  102. Yoshizaki, K.; Matsuda, T.; Nishimoto, N.; Kuritani, T.; Taeho, L.; Aozasa, K.; Nakahata, T.; Kawai, H.; Tagoh, H.; Komori, T. Pathogenic significance of interleukin-6 (IL-6/BSF-2) in Castleman’s disease. Blood 1989, 74, 1360–1367. [Google Scholar] [CrossRef] [PubMed]
  103. Song, S.-N.J.; Tomosugi, N.; Kawabata, H.; Ishikawa, T.; Nishikawa, T.; Yoshizaki, K. Down-regulation of hepcidin resulting from long-term treatment with an anti-IL-6 receptor antibody (tocilizumab) improves anemia of inflammation in multicentric Castleman disease. Blood 2010, 116, 3627–3634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Kawabata, H.; Tomosugi, N.; Kanda, J.; Tanaka, Y.; Yoshizaki, K.; Uchiyama, T. Anti-interleukin 6 receptor antibody tocilizumab reduces the level of serum hepcidin in patients with multicentric Castleman’s disease. Haematologica 2007, 92, 857–858. [Google Scholar] [CrossRef] [PubMed]
  105. Kurzrock, R.; Voorhees, P.M.; Casper, C.; Furman, R.R.; Fayad, L.; Lonial, S.; Borghaei, H.; Jagannath, S.; Sokol, L.; Usmani, S.Z.; et al. A phase I, open-label study of siltuximab, an anti-IL-6 monoclonal antibody, in patients with B-cell non-hodgkin lymphoma, multiple myeloma, or castleman disease. Clin. Cancer Res. 2013, 19, 3659–3670. [Google Scholar] [CrossRef] [PubMed]
  106. Casper, C.; Chaturvedi, S.; Munshi, N.; Wong, R.; Qi, M.; Schaffer, M.; Bandekar, R.; Hall, B.; van de Velde, H.; Vermeulen, J.; et al. Analysis of Inflammatory and Anemia-Related Biomarkers in a Randomized, Double-Blind, Placebo-Controlled Study of Siltuximab (Anti-IL6 Monoclonal Antibody) in Patients With Multicentric Castleman Disease. Clin. Cancer Res. 2015, 21, 4294–4304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Angevin, E.; Tabernero, J.; Elez, E.; Cohen, S.J.; Bahleda, R.; van Laethem, J.-L.; Ottensmeier, C.; Lopez-Martin, J.A.; Clive, S.; Joly, F.; et al. A phase I/II, multiple-dose, dose-escalation study of siltuximab, an anti-interleukin-6 monoclonal antibody, in patients with advanced solid tumors. Clin. Cancer Res. 2014, 20, 2192–2204. [Google Scholar] [CrossRef]
  108. Madeddu, C.; Gramignano, G.; Kotsonis, P.; Coghe, F.; Atzeni, V.; Scartozzi, M.; Macciò, A. Microenvironmental M1 tumor-associated macrophage polarization influences cancer-related anemia in advanced ovarian cancer: Key role of Interleukin-6. Haematologica 2018. [Google Scholar] [CrossRef]
  109. Noguchi-Sasaki, M.; Sasaki, Y.; Shimonaka, Y.; Mori, K.; Fujimoto-Ouchi, K. Treatment with anti-IL-6 receptor antibody prevented increase in serum hepcidin levels and improved anemia in mice inoculated with IL-6–producing lung carcinoma cells. BMC Cancer 2016, 16, 270. [Google Scholar] [CrossRef]
  110. Bergamaschi, G.; Di Sabatino, A.; Albertini, R.; Ardizzone, S.; Biancheri, P.; Bonetti, E.; Cassinotti, A.; Cazzola, P.; Markopoulos, K.; Massari, A.; et al. Prevalence and pathogenesis of anemia in inflammatory bowel disease. Influence of anti-tumor necrosis factor- treatment. Haematologica 2010, 95, 199–205. [Google Scholar] [CrossRef]
  111. Bes, C.; Yazici, A.; Soy, M. Monoclonal anti-TNF antibodies can elevate hemoglobin level in patients with ankylosing spondylitis. Rheumatol. Int. 2013, 33, 1415–1418. [Google Scholar] [CrossRef] [PubMed]
  112. Davis, D.; Charles, P.J.; Potter, A.; Feldmann, M.; Maini, R.N.; Elliott, M.J. Anaemia of chronic disease in rheumatoid arthritis: In vivo effects of tumour necrosis factor alpha blockade. Br. J. Rheumatol. 1997, 36, 950–956. [Google Scholar] [CrossRef] [PubMed]
  113. Doyle, M.K.; Rahman, M.U.; Frederick, B.; Birbara, C.A.; de Vries, D.; Toedter, G.; Wu, X.; Chen, D.; Ranganath, V.K.; Westerman, M.E.; et al. Effects of subcutaneous and intravenous golimumab on inflammatory biomarkers in patients with rheumatoid arthritis: Results of a phase 1, randomized, open-label trial. Rheumatology 2013, 52, 1214–1219. [Google Scholar] [CrossRef] [PubMed]
  114. Papadaki, H.A.; Kritikos, H.D.; Valatas, V.; Boumpas, D.T.; Eliopoulos, G.D. Anemia of chronic disease in rheumatoid arthritis is associated with increased apoptosis of bone marrow erythroid cells: Improvement following anti-tumor necrosis factor-α antibody therapy. Blood 2002, 100, 474–482. [Google Scholar] [CrossRef] [PubMed]
  115. Cavallaro, F.; Duca, L.; Pisani, L.F.; Rigolini, R.; Spina, L.; Tontini, G.E.; Munizio, N.; Costa, E.; Cappellini, M.D.; Vecchi, M.; et al. Anti-TNF-Mediated Modulation of Prohepcidin Improves Iron Availability in Inflammatory Bowel Disease, in an IL-6-Mediated Fashion. Can. J. Gastroenterol. Hepatol. 2017, 2017, 1–12. [Google Scholar] [CrossRef] [PubMed]
  116. Song, S.-N.J.; Iwahashi, M.; Tomosugi, N.; Uno, K.; Yamana, J.; Yamana, S.; Isobe, T.; Ito, H.; Kawabata, H.; Yoshizaki, K. Comparative evaluation of the effects of treatment with tocilizumab and TNF-α inhibitors on serum hepcidin, anemia response and disease activity in rheumatoid arthritis patients. Arthritis Res. Ther. 2013, 15, R141. [Google Scholar] [CrossRef] [PubMed]
  117. Koutroubakis, I.E.; Ramos-Rivers, C.; Regueiro, M.; Koutroumpakis, E.; Click, B.; Schwartz, M.; Swoger, J.; Baidoo, L.; Hashash, J.G.; Barrie, A.; et al. The Influence of Anti–tumor Necrosis Factor Agents on Hemoglobin Levels of Patients with Inflammatory Bowel Disease. Inflamm. Bowel Dis. 2015, 21, 1587–1593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Pardanani, A.; Finke, C.; Abdelrahman, R.A.; Lasho, T.L.; Tefferi, A. Associations and prognostic interactions between circulating levels of hepcidin, ferritin and inflammatory cytokines in primary myelofibrosis. Am. J. Hematol. 2013, 88, 312–316. [Google Scholar] [CrossRef] [Green Version]
  119. Parganas, E.; Wang, D.; Stravopodis, D.; Topham, D.J.; Marine, J.C.; Teglund, S.; Vanin, E.F.; Bodner, S.; Colamonici, O.R.; van Deursen, J.M.; et al. Jak2 is essential for signaling through a variety of cytokine receptors. Cell 1998, 93, 385–395. [Google Scholar] [CrossRef]
  120. Pardanani, A.; Gotlib, J.R.; Jamieson, C.; Cortes, J.E.; Talpaz, M.; Stone, R.M.; Silverman, M.H.; Gilliland, D.G.; Shorr, J.; Tefferi, A. Safety and Efficacy of TG101348, a Selective JAK2 Inhibitor, in Myelofibrosis. J. Clin. Oncol. 2011, 29, 789–796. [Google Scholar] [CrossRef] [Green Version]
  121. Verstovsek, S.; Mesa, R.A.; Gotlib, J.; Levy, R.S.; Gupta, V.; DiPersio, J.F.; Catalano, J.V.; Deininger, M.; Miller, C.; Silver, R.T.; et al. A Double-Blind, Placebo-Controlled Trial of Ruxolitinib for Myelofibrosis. N. Engl. J. Med. 2012, 366, 799–807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Pardanani, A.; Laborde, R.R.; Lasho, T.L.; Finke, C.; Begna, K.; Al-Kali, A.; Hogan, W.J.; Litzow, M.R.; Leontovich, A.; Kowalski, M.; et al. Safety and efficacy of CYT387, a JAK1 and JAK2 inhibitor, in myelofibrosis. Leukemia 2013, 27, 1322–1327. [Google Scholar] [CrossRef] [Green Version]
  123. Asshoff, M.; Petzer, V.; Warr, M.R.; Haschka, D.; Tymoszuk, P.; Demetz, E.; Seifert, M.; Posch, W.; Nairz, M.; Maciejewski, P.; et al. Momelotinib inhibits ACVR1/ALK2, decreases hepcidin production, and ameliorates anemia of chronic disease in rodents. Blood 2017, 129, 1823–1830. [Google Scholar] [CrossRef] [PubMed]
  124. Cornez, I.; Yajnanarayana, S.P.; Wolf, A.M.; Wolf, D. JAK/STAT disruption induces immuno-deficiency: Rationale for the development of JAK inhibitors as immunosuppressive drugs. Mol. Cell. Endocrinol. 2017, 451, 88–96. [Google Scholar] [CrossRef] [PubMed]
  125. Nielsen, O.; Soendergaard, C.; Vikner, M.; Weiss, G. Rational Management of Iron-Deficiency Anaemia in Inflammatory Bowel Disease. Nutrients 2018, 10, 82. [Google Scholar] [CrossRef] [PubMed]
  126. Theurl, I.; Schroll, A.; Nairz, M.; Seifert, M.; Theurl, M.; Sonnweber, T.; Kulaksiz, H.; Weiss, G. Pathways for the regulation of hepcidin expression in anemia of chronic disease and iron deficiency anemia in vivo. Haematologica 2011, 96, 1761–1769. [Google Scholar] [CrossRef] [PubMed]
  127. Lasocki, S.; Millot, S.; Andrieu, V.; Lettéron, P.; Pilard, N.; Muzeau, F.; Thibaudeau, O.; Montravers, P.; Beaumont, C. Phlebotomies or erythropoietin injections allow mobilization of iron stores in a mouse model mimicking intensive care anemia. Crit. Care Med. 2008, 36, 2388–2394. [Google Scholar] [CrossRef]
  128. Del Vecchio, L.; Locatelli, F. Clinical practice guidelines on iron therapy: A critical evaluation. Hemodial. Int. 2017, 21, S125–S131. [Google Scholar] [CrossRef]
  129. Drüeke, T.B.; Parfrey, P.S. Summary of the KDIGO guideline on anemia and comment: Reading between the (guide)line(s). Kidney Int. 2012, 82, 952–960. [Google Scholar] [CrossRef]
  130. Locatelli, F.; Bárány, P.; Covic, A.; De Francisco, A.; Del Vecchio, L.; Goldsmith, D.; Hörl, W.; London, G.; Vanholder, R.; Van Biesen, W.; et al. Kidney Disease: Improving Global Outcomes guidelines on anaemia management in chronic kidney disease: A European Renal Best Practice position statement. Nephrol. Dial. Transplant. 2013, 28, 1346–1359. [Google Scholar] [CrossRef]
  131. The National Collaborating Centre for Chronic Conditions. Anaemia Management in Chronic Kidney Disease: National Clinical Guideline for Management in Adults and Children; Royal College of Physicians: London, UK, 2006; ISBN 9781860162930. [Google Scholar]
  132. Kliger, A.S.; Foley, R.N.; Goldfarb, D.S.; Goldstein, S.L.; Johansen, K.; Singh, A.; Szczech, L. KDOQI US Commentary on the 2012 KDIGO Clinical Practice Guideline for Anemia in CKD. Am. J. Kidney Dis. 2013, 62, 849–859. [Google Scholar] [CrossRef] [PubMed]
  133. Singh, A.K.; Szczech, L.; Tang, K.L.; Barnhart, H.; Sapp, S.; Wolfson, M.; Reddan, D. CHOIR Investigators Correction of Anemia with Epoetin Alfa in Chronic Kidney Disease. N. Engl. J. Med. 2006, 355, 2085–2098. [Google Scholar] [CrossRef] [PubMed]
  134. Drüeke, T.B.; Locatelli, F.; Clyne, N.; Eckardt, K.-U.; Macdougall, I.C.; Tsakiris, D.; Burger, H.-U.; Scherhag, A. CREATE Investigators Normalization of hemoglobin level in patients with chronic kidney disease and anemia. N. Engl. J. Med. 2006, 355, 2071–2084. [Google Scholar] [CrossRef] [PubMed]
  135. Macdougall, I.C.; Bock, A.H.; Carrera, F.; Eckardt, K.-U.; Gaillard, C.; Van Wyck, D.; Roubert, B.; Nolen, J.G.; Roger, S.D. FIND-CKD Study Investigators FIND-CKD: A randomized trial of intravenous ferric carboxymaltose versus oral iron in patients with chronic kidney disease and iron deficiency anaemia. Nephrol. Dial. Transplant. 2014, 29, 2075–2084. [Google Scholar] [CrossRef] [PubMed]
  136. Agarwal, R.; Kusek, J.W.; Pappas, M.K. A randomized trial of intravenous and oral iron in chronic kidney disease. Kidney Int. 2015, 88, 905–914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Gasche, C.; Lomer, M.C.E.; Cavill, I.; Weiss, G. Iron, anaemia, and inflammatory bowel diseases. Gut 2004, 53, 1190–1197. [Google Scholar] [CrossRef] [PubMed]
  138. Nielsen, O.H.; Ainsworth, M.; Coskun, M.; Weiss, G. Management of Iron-Deficiency Anemia in Inflammatory Bowel Disease. Medicine 2015, 94, e963. [Google Scholar] [CrossRef]
  139. Gasche, C.; Dejaco, C.; Reinisch, W.; Tillinger, W.; Waldhoer, T.; Fueger, G.F.; Lochs, H.; Gangl, A. Sequential Treatment of Anemia in Ulcerative Colitis with Intravenous Iron and Erythropoietin. Digestion 1999, 60, 262–267. [Google Scholar] [CrossRef]
  140. Dignass, A.U.; Gasche, C.; Bettenworth, D.; Birgegård, G.; Danese, S.; Gisbert, J.P.; Gomollon, F.; Iqbal, T.; Katsanos, K.; Koutroubakis, I.; et al. European Consensus on the Diagnosis and Management of Iron Deficiency and Anaemia in Inflammatory Bowel Diseases. J. Crohns Colitis 2015, 9, 211–222. [Google Scholar] [CrossRef] [Green Version]
  141. Gisbert, J.P.; Bermejo, F.; Pajares, R.; Pérez-Calle, J.-L.; Rodríguez, M.; Algaba, A.; Mancenido, N.; de la Morena, F.; Carneros, J.A.; McNicholl, A.G.; et al. Oral and intravenous iron treatment in inflammatory bowel disease: Hematological response and quality of life improvement. Inflamm. Bowel Dis. 2009, 15, 1485–1491. [Google Scholar] [CrossRef]
  142. Reinisch, W.; Staun, M.; Tandon, R.K.; Altorjay, I.; Thillainayagam, A.V.; Gratzer, C.; Nijhawan, S.; Thomsen, L.L. A Randomized, Open-Label, Non-Inferiority Study of Intravenous Iron Isomaltoside 1000 (Monofer) Compared with Oral Iron for Treatment of Anemia in IBD (PROCEED). Am. J. Gastroenterol. 2013, 108, 1877–1888. [Google Scholar] [CrossRef] [PubMed]
  143. Lee, T.; Clavel, T.; Smirnov, K.; Schmidt, A.; Lagkouvardos, I.; Walker, A.; Lucio, M.; Michalke, B.; Schmitt-Kopplin, P.; Fedorak, R.; et al. Oral versus intravenous iron replacement therapy distinctly alters the gut microbiota and metabolome in patients with IBD. Gut 2017, 66, 863–871. [Google Scholar] [CrossRef] [PubMed]
  144. Onken, J.E.; Bregman, D.B.; Harrington, R.A.; Morris, D.; Acs, P.; Akright, B.; Barish, C.; Bhaskar, B.S.; Smith-Nguyen, G.N.; Butcher, A.; et al. A multicenter, randomized, active-controlled study to investigate the efficacy and safety of intravenous ferric carboxymaltose in patients with iron deficiency anemia. Transfusion 2014, 54, 306–315. [Google Scholar] [CrossRef] [PubMed]
  145. Kulnigg, S.; Stoinov, S.; Simanenkov, V.; Dudar, L.V.; Karnafel, W.; Garcia, L.C.; Sambuelli, A.M.; D’Haens, G.; Gasche, C. A Novel Intravenous Iron Formulation for Treatment of Anemia in Inflammatory Bowel Disease: The Ferric Carboxymaltose (FERINJECT®) Randomized Controlled Trial. Am. J. Gastroenterol. 2008, 103, 1182–1192. [Google Scholar] [CrossRef] [PubMed]
  146. Lindgren, S.; Wikman, O.; Befrits, R.; Blom, H.; Eriksson, A.; Grännö, C.; Ung, K.-A.; Hjortswang, H.; Lindgren, A.; Unge, P. Intravenous iron sucrose is superior to oral iron sulphate for correcting anaemia and restoring iron stores in IBD patients: A randomized, controlled, evaluator-blind, multicentre study. Scand. J. Gastroenterol. 2009, 44, 838–845. [Google Scholar] [CrossRef] [PubMed]
  147. Koutroubakis, I.E.; Ramos-Rivers, C.; Regueiro, M.; Koutroumpakis, E.; Click, B.; Schwartz, M.; Swoger, J.; Baidoo, L.; Hashash, J.G.; Barrie, A.; et al. Five-Year Period Prevalence and Characteristics of Anemia in a Large US Inflammatory Bowel Disease Cohort. J. Clin. Gastroenterol. 2016, 50, 638–643. [Google Scholar] [CrossRef] [PubMed]
  148. Iqbal, T.; Stein, J.; Sharma, N.; Kulnigg-Dabsch, S.; Vel, S.; Gasche, C. Clinical Significance of C-Reactive Protein Levels in Predicting Responsiveness to Iron Therapy in Patients with Inflammatory Bowel Disease and Iron Deficiency Anemia. Dig. Dis. Sci. 2015, 60, 1375–1381. [Google Scholar] [CrossRef] [PubMed]
  149. Macdougall, I.C.; White, C.; Anker, S.D.; Bhandari, S.; Farrington, K.; Kalra, P.A.; McMurray, J.J.V.; Murray, H.; Steenkamp, R.; Tomson, C.R.V.; et al. Randomized Trial Comparing Proactive, High-Dose versus Reactive, Low-Dose Intravenous Iron Supplementation in Hemodialysis (PIVOTAL): Study Design and Baseline Data. Am. J. Nephrol. 2018, 48, 260–268. [Google Scholar] [CrossRef]
  150. Macdougall, I.C.; White, C.; Anker, S.D.; Bhandari, S.; Farrington, K.; Kalra, P.A.; McMurray, J.J.V.; Murray, H.; Tomson, C.R.V.; Wheeler, D.C.; et al. Intravenous Iron in Patients Undergoing Maintenance Hemodialysis. N. Engl. J. Med. 2018. [Google Scholar] [CrossRef]
  151. Drakesmith, H.; Nemeth, E.; Ganz, T. Ironing out Ferroportin. Cell Metab. 2015, 22, 777–787. [Google Scholar] [CrossRef]
  152. Crielaard, B.J.; Lammers, T.; Rivella, S. Targeting iron metabolism in drug discovery and delivery. Nat. Rev. Drug Discov. 2017, 16, 400–423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Sebastiani, G.; Wilkinson, N.; Pantopoulos, K. Pharmacological Targeting of the Hepcidin/Ferroportin Axis. Front. Pharmacol. 2016, 7, 160. [Google Scholar] [CrossRef] [PubMed]
  154. Means, R.T.; Krantz, S.B. Inhibition of human erythroid colony-forming units by tumor necrosis factor requires beta interferon. J. Clin. Investig. 1993, 91, 416–419. [Google Scholar] [CrossRef] [PubMed]
  155. La Ferla, K.; Reimann, C.; Jelkmann, W.; Hellwig-Bürgel, T. Inhibition of erythropoietin gene expression signaling involves the transcription factors GATA-2 and NF-kappaB. FASEB J. 2002, 16, 1811–1813. [Google Scholar] [CrossRef] [PubMed]
  156. Cazzola, M.; Ponchio, L.; de Benedetti, F.; Ravelli, A.; Rosti, V.; Beguin, Y.; Invernizzi, R.; Barosi, G.; Martini, A. Defective iron supply for erythropoiesis and adequate endogenous erythropoietin production in the anemia associated with systemic-onset juvenile chronic arthritis. Blood 1996, 87, 4824–4830. [Google Scholar] [CrossRef] [PubMed]
  157. Khalil, S.; Delehanty, L.; Grado, S.; Holy, M.; White, Z.; Freeman, K.; Kurita, R.; Nakamura, Y.; Bullock, G.; Goldfarb, A. Iron modulation of erythropoiesis is associated with Scribble-mediated control of the erythropoietin receptor. J. Exp. Med. 2018, 215, 661–679. [Google Scholar] [CrossRef] [PubMed]
  158. Weiss, G.; Houston, T.; Kastner, S.; Jöhrer, K.; Grünewald, K.; Brock, J.H. Regulation of cellular iron metabolism by erythropoietin: Activation of iron-regulatory protein and upregulation of transferrin receptor expression in erythroid cells. Blood 1997, 89, 680–687. [Google Scholar]
  159. Cooke, K.S.; Hinkle, B.; Salimi-Moosavi, H.; Foltz, I.; King, C.; Rathanaswami, P.; Winters, A.; Steavenson, S.; Begley, C.G.; Molineux, G.; et al. A fully human anti-hepcidin antibody modulates iron metabolism in both mice and nonhuman primates. Blood 2013, 122, 3054–3061. [Google Scholar] [CrossRef] [Green Version]
  160. Rothe, C.; Skerra, A. Anticalin® Proteins as Therapeutic Agents in Human Diseases. BioDrugs 2018, 32, 233–243. [Google Scholar] [CrossRef]
  161. Hohlbaum, A.M.; Gille, H.; Trentmann, S.; Kolodziejczyk, M.; Rattenstetter, B.; Laarakkers, C.M.; Katzmann, G.; Christian, H.J.; Andersen, N.; Allersdorfer, A.; et al. Sustained plasma hepcidin suppression and iron elevation by Anticalin-derived hepcidin antagonist in cynomolgus monkey. Br. J. Pharmacol. 2018, 175, 1054–1065. [Google Scholar] [CrossRef]
  162. Boyce, M.; Warrington, S.; Cortezi, B.; Zöllner, S.; Vauléon, S.; Swinkels, D.W.; Summo, L.; Schwoebel, F.; Riecke, K. Safety, pharmacokinetics and pharmacodynamics of the anti-hepcidin Spiegelmer lexaptepid pegol in healthy subjects. Br. J. Pharmacol. 2016, 173, 1580–1588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Macdougall, I.C.; Rumjon, A.; Cinco, J.; Goldstein, L.; Summo, L.; Vauleon, S.; Riecke, K. FP660 Pharmacokinetics and Pharmacodynamics of Lexaptepid, a Novel Anti-Hepcidin Molecule, in ESA-Resistant Haemodialysis Patients. Nephrol. Dial. Transplant. 2015, 30, iii294–iii295. [Google Scholar] [CrossRef]
  164. Georgiev, P.; Lazaroiu, M.; Ocroteala, L.; Grudeva-Popova, J.; Gheorghita, E.; Vasilica, M.; Popescu, S.M.; Cucuianu, A.; Summo, L.; Schwoebel, F.; et al. Abstract 3847: The anti-hepcidin Spiegelmer® Lexaptepid Pegol (NOX-H94) as treatment of anemia of chronic disease in patients with multiple myeloma, low grade lymphoma, and CLL: A phase II pilot study. Cancer Res. 2014, 74, 3847. [Google Scholar] [CrossRef]
  165. Andriopoulos, B.; Corradini, E.; Xia, Y.; Faasse, S.A.; Chen, S.; Grgurevic, L.; Knutson, M.D.; Pietrangelo, A.; Vukicevic, S.; Lin, H.Y.; et al. BMP6 is a key endogenous regulator of hepcidin expression and iron metabolism. Nat. Genet. 2009, 41, 482–487. [Google Scholar] [CrossRef] [Green Version]
  166. Parrow, N.L.; Fleming, R.E. Bone Morphogenetic Proteins as Regulators of Iron Metabolism. Annu. Rev. Nutr. 2014, 34, 77–94. [Google Scholar] [CrossRef] [PubMed]
  167. Theurl, I.; Schroll, A.; Sonnweber, T.; Nairz, M.; Theurl, M.; Willenbacher, W.; Eller, K.; Wolf, D.; Seifert, M.; Sun, C.C.; et al. Pharmacologic inhibition of hepcidin expression reverses anemia of chronic inflammation in rats. Blood 2011, 118, 4977–4984. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Babitt, J.L.; Huang, F.W.; Xia, Y.; Sidis, Y.; Andrews, N.C.; Lin, H.Y. Modulation of bone morphogenetic protein signaling in vivo regulates systemic iron balance. J. Clin. Investig. 2007, 117, 1933–1939. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Poli, M.; Asperti, M.; Ruzzenenti, P.; Mandelli, L.; Campostrini, N.; Martini, G.; Di Somma, M.; Maccarinelli, F.; Girelli, D.; Naggi, A.; et al. Oversulfated heparins with low anticoagulant activity are strong and fast inhibitors of hepcidin expression in vitro and in vivo. Biochem. Pharmacol. 2014, 92, 467–475. [Google Scholar] [CrossRef] [PubMed]
  170. Petzer, V.; Tymoszuk, P.; Wake, M.; Bayliss, L.; Papworth, J.; Carvalho, J.; Deantonio, C.; Asshoff, M.F.; Seifert, M.; Berger, S.; et al. A Fully Human Anti-BMP6 Antibody Reduces the Need for Erythropoietin Stimulating Agent in Two Rodent Anemia of Chronic Disease Models. Blood 2018, 132, 1045. [Google Scholar] [CrossRef]
  171. Galli, M.; Chatterjee, M.; Grasso, M.; Specchia, G.; Magen, H.; Einsele, H.; Celeghini, I.; Barbieri, P.; Paoletti, D.; Pace, S.; et al. Phase I study of the heparanase inhibitor Roneparstat: An innovative approach for multiple myeloma therapy. Haematologica 2018. [Google Scholar] [CrossRef]
  172. Peterson, P.; Whatcott, C.; Siddiqui-Jain, A.; Weitman, S.; Kieran, M.; Bearss, D.J.; Warner, S.L. TP-0184 Inhibits ALK2/ACVR1, Decreases Hepcidin Levels, and Demonstrates Activity in Preclinical Mouse Models of Functional Iron Deficiency. Blood 2017, 130, 937. [Google Scholar]
  173. Kovac, S.; Böser, P.; Cui, Y.; Ferring-Appel, D.; Casarrubea, D.; Huang, L.; Fung, E.; Popp, A.; Mueller, B.K.; Hentze, M.W. Anti-hemojuvelin antibody corrects anemia caused by inappropriately high hepcidin levels. Haematologica 2016, 101, e173–e176. [Google Scholar] [CrossRef] [PubMed]
  174. Barrington, P.; Sheetz, M.J.; Callies, S.; Waters, D.G.; Berg, P.H.; Pappas, D.; Marbury, T.C.; Decker, B.S.; Berg, J.K. Safety, Tolerability, Pharmacokinetics and Pharmacodynamics of an Anti-Ferroportin Antibody in Patients with Anemia Due to Chronic Renal Failure. Blood 2016, 128, 1280. [Google Scholar]
  175. Leung, D.; Hill, K.A.; De Rosa, D.C.; Xu, J.; Manetta, J.; Wroblewski, V.J.; Benschop, R.J. LY2928057, An Antibody Targeting Ferroportin, Is a Potent Inhibitor Of Hepcidin Activity and Increases Iron Mobilization In Normal Cynomolgus Monkeys. Blood 2013, 122, 3433. [Google Scholar]
  176. Ashby, D.R.; Gale, D.P.; Busbridge, M.; Murphy, K.G.; Duncan, N.D.; Cairns, T.D.; Taube, D.H.; Bloom, S.R.; Tam, F.W.K.; Chapman, R.; et al. Erythropoietin administration in humans causes a marked and prolonged reduction in circulating hepcidin. Haematologica 2010, 95, 505–508. [Google Scholar] [CrossRef] [PubMed]
  177. Kautz, L.; Jung, G.; Valore, E.V.; Rivella, S.; Nemeth, E.; Ganz, T. Identification of erythroferrone as an erythroid regulator of iron metabolism. Nat. Genet. 2014, 46, 678–684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Camaschella, C.; Pagani, A. Advances in understanding iron metabolism and its crosstalk with erythropoiesis. Br. J. Haematol. 2018, 182, 481–494. [Google Scholar] [CrossRef]
  179. Provenzano, R.; Besarab, A.; Wright, S.; Dua, S.; Zeig, S.; Nguyen, P.; Poole, L.; Saikali, K.G.; Saha, G.; Hemmerich, S.; et al. Roxadustat (FG-4592) Versus Epoetin Alfa for Anemia in Patients Receiving Maintenance Hemodialysis: A Phase 2, Randomized, 6- to 19-Week, Open-Label, Active-Comparator, Dose-Ranging, Safety and Exploratory Efficacy Study. Am. J. Kidney Dis. 2016, 67, 912–924. [Google Scholar] [CrossRef]
  180. Peyssonnaux, C.; Zinkernagel, A.S.; Schuepbach, R.A.; Rankin, E.; Vaulont, S.; Haase, V.H.; Nizet, V.; Johnson, R.S. Regulation of iron homeostasis by the hypoxia-inducible transcription factors (HIFs). J. Clin. Investig. 2007, 117, 1926–1932. [Google Scholar] [CrossRef] [Green Version]
  181. Cappellini, M.D.; Comin-Colet, J.; de Francisco, A.; Dignass, A.; Doehner, W.; Lam, C.S.; Macdougall, I.C.; Rogler, G.; Camaschella, C.; Kadir, R.; et al. Iron deficiency across chronic inflammatory conditions: International expert opinion on definition, diagnosis, and management. Am. J. Hematol. 2017, 92, 1068–1078. [Google Scholar] [CrossRef] [Green Version]
  182. Avni, T.; Leibovici, L.; Gafter-Gvili, A. Iron supplementation for the treatment of chronic heart failure and iron deficiency: Systematic review and meta-analysis. Eur. J. Heart Fail. 2012, 14, 423–429. [Google Scholar] [CrossRef] [PubMed]
  183. Çekiç, C.; İpek, S.; Aslan, F.; Akpınar, Z.; Arabul, M.; Topal, F.; Sarıtaş Yüksel, E.; Alper, E.; Ünsal, B. The Effect of Intravenous Iron Treatment on Quality of Life in Inflammatory Bowel Disease Patients with Nonanemic Iron Deficiency. Gastroenterol. Res. Pract. 2015, 2015, 1–5. [Google Scholar] [CrossRef] [PubMed]
  184. Kovesdy, C.P.; Trivedi, B.K.; Kalantar-Zadeh, K.; Anderson, J.E. Association of anemia with outcomes in men with moderate and severe chronic kidney disease. Kidney Int. 2006, 69, 560–564. [Google Scholar] [CrossRef] [PubMed]
  185. Okonko, D.O.; Mandal, A.K.J.; Missouris, C.G.; Poole-Wilson, P.A. Disordered iron homeostasis in chronic heart failure: Prevalence, predictors, and relation to anemia, exercise capacity, and survival. J. Am. Coll. Cardiol. 2011, 58, 1241–1251. [Google Scholar] [CrossRef] [PubMed]
  186. Enjuanes, C.; Bruguera, J.; Grau, M.; Cladellas, M.; Gonzalez, G.; Meroño, O.; Moliner-Borja, P.; Verdú, J.M.; Farré, N.; Comín-Colet, J. Iron Status in Chronic Heart Failure: Impact on Symptoms, Functional Class and Submaximal Exercise Capacity. Rev. Española Cardiol. 2016, 69, 247–255. [Google Scholar] [CrossRef]
  187. Anker, S.D.; Comin Colet, J.; Filippatos, G.; Willenheimer, R.; Dickstein, K.; Drexler, H.; Lüscher, T.F.; Bart, B.; Banasiak, W.; Niegowska, J.; et al. Ferric Carboxymaltose in Patients with Heart Failure and Iron Deficiency. N. Engl. J. Med. 2009, 361, 2436–2448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Ponikowski, P.; van Veldhuisen, D.J.; Comin-Colet, J.; Ertl, G.; Komajda, M.; Mareev, V.; McDonagh, T.; Parkhomenko, A.; Tavazzi, L.; Levesque, V.; et al. Beneficial effects of long-term intravenous iron therapy with ferric carboxymaltose in patients with symptomatic heart failure and iron deficiency. Eur. Heart J. 2015, 36, 657–668. [Google Scholar] [CrossRef]
  189. Stauder, R.; Valent, P.; Theurl, I. Anemia at older age: Etiologies, clinical implications, and management. Blood 2018, 131, 505–514. [Google Scholar] [CrossRef]
  190. Mabeza, G.F.; Loyevsky, M.; Gordeuk, V.R.; Weiss, G. Iron chelation therapy for malaria: A review. Pharmacol. Ther. 1999, 81, 53–75. [Google Scholar] [CrossRef]
  191. Zlotkin, S.; Newton, S.; Aimone, A.M.; Azindow, I.; Amenga-Etego, S.; Tchum, K.; Mahama, E.; Thorpe, K.E.; Owusu-Agyei, S. Effect of Iron Fortification on Malaria Incidence in Infants and Young Children in Ghana. JAMA 2013, 310, 938–947. [Google Scholar] [CrossRef]
  192. Zhang, D.-L.; Wu, J.; Shah, B.N.; Greutélaers, K.C.; Ghosh, M.C.; Ollivierre, H.; Su, X.-Z.; Thuma, P.E.; Bedu-Addo, G.; Mockenhaupt, F.P.; et al. Erythrocytic ferroportin reduces intracellular iron accumulation, hemolysis, and malaria risk. Science 2018, 359, 1520–1523. [Google Scholar] [CrossRef] [PubMed]
  193. Sazawal, S.; Black, R.E.; Ramsan, M.; Chwaya, H.M.; Stoltzfus, R.J.; Dutta, A.; Dhingra, U.; Kabole, I.; Deb, S.; Othman, M.K.; et al. Effects of routine prophylactic supplementation with iron and folic acid on admission to hospital and mortality in preschool children in a high malaria transmission setting: Community-based, randomised, placebo-controlled trial. Lancet 2006, 367, 133–143. [Google Scholar] [CrossRef]
  194. Soofi, S.; Cousens, S.; Iqbal, S.P.; Akhund, T.; Khan, J.; Ahmed, I.; Zaidi, A.K.; Bhutta, Z.A. Effect of provision of daily zinc and iron with several micronutrients on growth and morbidity among young children in Pakistan: A cluster-randomised trial. Lancet 2013, 382, 29–40. [Google Scholar] [CrossRef]
  195. Kortman, G.A.M.; Reijnders, D.; Swinkels, D.W. Oral iron supplementation: Potential implications for the gut microbiome and metabolome in patients with CKD. Hemodial. Int. 2017, 21, S28–S36. [Google Scholar] [CrossRef] [PubMed]
  196. Paganini, D.; Uyoga, M.; Zimmermann, M. Iron Fortification of Foods for Infants and Children in Low-Income Countries: Effects on the Gut Microbiome, Gut Inflammation, and Diarrhea. Nutrients 2016, 8, 494. [Google Scholar] [CrossRef] [PubMed]
  197. Moschen, A.R.; Gerner, R.R.; Wang, J.; Klepsch, V.; Adolph, T.E.; Reider, S.J.; Hackl, H.; Pfister, A.; Schilling, J.; Moser, P.L.; et al. Lipocalin 2 Protects from Inflammation and Tumorigenesis Associated with Gut Microbiota Alterations. Cell Host Microbe 2016, 19, 455–469. [Google Scholar] [CrossRef] [PubMed]
  198. Scheers, N.M.; Pereira, D.I.A.; Faria, N.; Powell, J.J. Ferric citrate and ferric EDTA but not ferrous sulfate drive amphiregulin-mediated activation of the MAP kinase ERK in gut epithelial cancer cells. Oncotarget 2018, 9, 17066–17077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Porto, G.; De Sousa, M. Iron overload and immunity. World J. Gastroenterol. 2007, 13, 4707–4715. [Google Scholar] [CrossRef]
  200. Hara, N.; Okamoto, N.; Imanishi, K.; Takahashi, T.; Makiyama, K. Improvement in reliability of InP-based HEMTs by suppressing impact ionization. Electron. Commun. Jpn. Part II Electron. 2007, 90, 33–38. [Google Scholar] [CrossRef]
  201. Weiss, G.; Meusburger, E.; Radacher, G.; Garimorth, K.; Neyer, U.; Mayer, G. Effect of iron treatment on circulating cytokine levels in ESRD patients receiving recombinant human erythropoietin. Kidney Int. 2003, 64, 572–578. [Google Scholar] [CrossRef]
  202. Ibrahim, A.S.; Gebermariam, T.; Fu, Y.; Lin, L.; Husseiny, M.I.; French, S.W.; Schwartz, J.; Skory, C.D.; Edwards, J.E.; Spellberg, B.J. The iron chelator deferasirox protects mice from mucormycosis through iron starvation. J. Clin. Investig. 2007, 117, 2649–2657. [Google Scholar] [CrossRef] [Green Version]
  203. Jankowska, E.A.; Rozentryt, P.; Witkowska, A.; Nowak, J.; Hartmann, O.; Ponikowska, B.; Borodulin-Nadzieja, L.; Banasiak, W.; Polonski, L.; Filippatos, G.; et al. Iron deficiency: An ominous sign in patients with systolic chronic heart failure. Eur. Heart J. 2010, 31, 1872–1880. [Google Scholar] [CrossRef]
  204. Macdougall, I.C.; Canaud, B.; De Francisco, A.L.M.; Filippatos, G.; Ponikowski, P.; Silverberg, D.; Van Veldhuisen, D.J.; Anker, S.D. Beyond the cardiorenal anaemia syndrome: Recognizing the role of iron deficiency. Eur. J. Heart Fail. 2012, 14, 882–886. [Google Scholar] [CrossRef]
Table 1. Diagnostic markers for the diagnosis of different types of inflammatory anemia.
Table 1. Diagnostic markers for the diagnosis of different types of inflammatory anemia.
MarkerAnemia of InflammationAnemia of Inflammation plus Iron Deficiency AnemiaLimitations/Comments
Bone marrow iron stainingNormal–ElevatedNormal–Reduced
  • Gold standard
  • Invasive method, not routinely used
Serum IronLowLowUnderlies diurnal variations
FerritinElevatedReduced–Normal–Elevated
  • Most commonly used marker
  • Ferritin is an acute phase protein and does not accurately reflect iron status during inflammation
  • Ferritin < 30 ng/mL always associated with true iron deficiency
TransferrinNormal–ReducedNormal–High
Tf-SatLowLowDependent on iron and transferrin levels
sTfRNormal–ElevatedElevated
  • Good marker for needs of iron for erythropoiesis in absence of inflammation
  • Values affected by inflammation and ESA application
sTfR/log FerritinNormalElevatedUsed for differentiation, but there is a lack of a prospective study
HepcidinElevatedNormal–Reduced
  • Expression is more affected by iron deficiency (suppressing) than by inflammation
  • Not standardized
  • Weak correlations in CKD patients
  • Possible predictive parameter for success of iron and/or ESA treatment
ErythroferronNot knownNot known
  • Not standardized
  • Higher ERFE levels in CKD patients
  • Positively correlated with serum erythropoietin and negatively with hemoglobin
MCV/MCHNormalNormal–ReducedIf reduced, indication of iron deficiency
Reticulocyte Hb contentNormal–ReducedReducedIndicated insufficient iron availability for erythropoiesis, not prospectively studied
Hypochromic RBCNormalNormal–Elevated
  • Related to MCV, as a sensitive marker for iron availability for erythroid progenitors
  • Cut-off values are different between different machines
CRPIncreasedIncreased
  • Non-specific inflammatory marker
  • Iron–independent parameter
  • Correlation with severity of anemia
IL6IncreasedIncreased
  • Non-specific inflammatory marker
  • Iron–independent parameter
Table 2. Characteristics of oral and intravenous iron therapy.
Table 2. Characteristics of oral and intravenous iron therapy.
Indication(s)BenefitsLimitationsUncertainties/Comments
Oral iron
  • True iron deficiency
  • Combined true and functional iron deficiency with low grade inflammation
  • Low costs
  • Easy to apply
  • Effective if applied appropriately
  • High pill burden
  • Low bioavailability
  • High rate of non-responders
  • Ineffective in the presence of high hepcidin levels
  • Gastro-intestinal side effects
  • Low compliance
  • Identification of the underlying cause
  • Absorption defect must be excluded
  • No predictor for response
  • Oral iron as a trigger for cancer or intestinal inflammation
  • Effects on gut microbiome
  • Disease specific therapeutic start and endpoints
Intravenous iron
  • True and functional iron deficiency
  • Absorption defects
  • Severe anemia
  • Intolerance to oral iron therapy
  • Lack of efficacy of oral iron therapy
  • Faster replacement of iron stores than with oral iron
  • Fewer gastro-intestinal side effects
  • New i.v. iron formulations allowing high single dose administration
  • Effective in the presence of inflammation
  • Better control of compliance
  • Rare but possible life threatening anaphylactic reactions
  • Route of application requires consultation of a physician
  • Higher costs
  • Hypophosphatemia
  • Long-term outcome on underlying disease unclear
  • No predictor of response
  • Possible iron-induced oxidative/nitrosative stress
  • Unknown efficacy in patients with more advanced inflammation and/or high hepcidin levels
  • Disease specific therapeutic start and endpoints
Table 3. Drugs impacting on hepcidin-mediated alteration of iron homeostasis.
Table 3. Drugs impacting on hepcidin-mediated alteration of iron homeostasis.
Name(s)Primary Indication(s)TargetDrug TypeMechanism
Tocilizumab
  • Rheumatoid arthritis
  • Systemic juvenile idiopathic arthritis
  • Giant cell arteritis
  • MCD
  • Cytokine release syndrome
IL6RHumanized monoclonal antibodyIL6 signaling inhibition
SiltuximabMCDIL6Chimeric monoclonal AntibodyIL6 binding
Infliximab
  • IBD (Crohn’s disease, Ulcerative colitis)
  • Rheumatoid arthritis
  • Psoriatic arthritis
  • Ankylosing spondylitis
  • Psoriasis
TNFαChimeric monoclonal antibodyTNFα binding/blocker
Adalimumab
  • IBD (Crohn’s disease, Ulcerative colitis)
  • Rheumatoid arthritis
  • Psoriatic arthritis
  • Ankylosing spondylitis
  • Psoriasis
  • Hidradenitis suppurativa Juvenile idiopathic arthritis
TNFαHumanized monoclonal antibodyTNFα binding/blocker
Momelotinib GS-0387 CYT-387MyelofibrosisJAK1 and JAK2Small molecule
  • Jak1 and Jak2 inhibition
  • Blockig of hepcidin production via ALK2 inhibition
CSJ137
  • Hepcidin modulation
  • Anemia amelioration
BMP6AntibodyBMP6 binding/blocking
SST0001
RO-82
RO-68
NAc-91
NAcRO-00
  • Myeloma therapy
  • Hepcidin modulation
BMP6Modified heparinBMP6 binding
TP-0184
  • Antitumor activity in advanced solid tumors
  • Hepcidin modulation
  • Anemia amelioration
ALK2Small moleculeALK2 inhibition
h5F9.23,
h5F9-AM8
  • Hepcidin modulation
  • Anemia amelioration
HJV/RGMcAntibodyBMP Co-receptor binding binding
Spiegelmer
Aptamers
NOX-H94H
  • Hepcidin modulation
  • Anemia amelioration
HepcidinLexaptepid pegol L-stereoisomeric RNA aptamerHepcidin binding
PRS-080
  • Hepcidin modulation
  • Anemia amelioration
HepcidinAntichalin, bioengineered lipocalinHepcidin binding
ErythropoetinAnemiaEpoRProteinInduction of Erythroferron and blockage of hepcidin

Share and Cite

MDPI and ACS Style

Petzer, V.; Theurl, I.; Weiss, G. Established and Emerging Concepts to Treat Imbalances of Iron Homeostasis in Inflammatory Diseases. Pharmaceuticals 2018, 11, 135. https://doi.org/10.3390/ph11040135

AMA Style

Petzer V, Theurl I, Weiss G. Established and Emerging Concepts to Treat Imbalances of Iron Homeostasis in Inflammatory Diseases. Pharmaceuticals. 2018; 11(4):135. https://doi.org/10.3390/ph11040135

Chicago/Turabian Style

Petzer, Verena, Igor Theurl, and Günter Weiss. 2018. "Established and Emerging Concepts to Treat Imbalances of Iron Homeostasis in Inflammatory Diseases" Pharmaceuticals 11, no. 4: 135. https://doi.org/10.3390/ph11040135

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop