Next Article in Journal
Inhibitory Effects and Mechanisms of Luteolin on Proliferation and Migration of Vascular Smooth Muscle Cells
Next Article in Special Issue
Iron in Child Obesity. Relationships with Inflammation and Metabolic Risk Factors
Previous Article in Journal
Sources of Vitamin A in the Diets of Pre-School Children in the Avon Longitudinal Study of Parents and Children (ALSPAC)
Previous Article in Special Issue
Iron Deficiency and Bariatric Surgery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Iron Absorption in Drosophila melanogaster

1
School of Biological and Chemical Sciences, Queen Mary, University of London, Mile End Road, London, E1 4NS, UK
2
Department of Physiology, Biophysics and Neuroscience, CINVESTAV-IPN, IPN Avenue 2508, Zacatenco, 07360, Mexico City, Mexico
*
Author to whom correspondence should be addressed.
Nutrients 2013, 5(5), 1622-1647; https://doi.org/10.3390/nu5051622
Submission received: 12 April 2013 / Revised: 3 May 2013 / Accepted: 7 May 2013 / Published: 17 May 2013
(This article belongs to the Special Issue Dietary Iron and Human Health)

Abstract

:
The way in which Drosophila melanogaster acquires iron from the diet remains poorly understood despite iron absorption being of vital significance for larval growth. To describe the process of organismal iron absorption, consideration needs to be given to cellular iron import, storage, export and how intestinal epithelial cells sense and respond to iron availability. Here we review studies on the Divalent Metal Transporter-1 homolog Malvolio (iron import), the recent discovery that Multicopper Oxidase-1 has ferroxidase activity (iron export) and the role of ferritin in the process of iron acquisition (iron storage). We also describe what is known about iron regulation in insect cells. We then draw upon knowledge from mammalian iron homeostasis to identify candidate genes in flies. Questions arise from the lack of conservation in Drosophila for key mammalian players, such as ferroportin, hepcidin and all the components of the hemochromatosis-related pathway. Drosophila and other insects also lack erythropoiesis. Thus, systemic iron regulation is likely to be conveyed by different signaling pathways and tissue requirements. The significance of regulating intestinal iron uptake is inferred from reports linking Drosophila developmental, immune, heat-shock and behavioral responses to iron sequestration.

Graphical Abstract

1. Introduction

Iron is an indispensible micronutrient for the development of Drosophila melanogaster [1,2]. Enzymes that bind iron, heme or iron-sulfur clusters carry out numerous physiological functions, including respiration [3] and the synthesis of DNA [4,5], ecdysone [6,7], dopamine [8] and lipids [9]. Mitochondria are the site of respiration and synthesis of heme and iron-sulfur clusters and respond to the cellular sensing systems for oxygen and iron [10,11,12,13,14]. Despite the elucidation of key biochemical requirements for iron, our knowledge of how iron is acquired from the diet of Drosophila larvae or adults and distributed to its various target tissues and proteins in a regulated manner remains at a rudimentary level [15]. Here, following a summary of how iron absorption occurs in mammals [16], we describe early studies of iron homeostasis in Drosophila that used histochemical and radioactive methods [17], atomic absorption spectrometry [18] and electron microscopy [19] to detect iron. Then, more recent studies of particular genes involved in iron absorption are described. We also identify key genes that are conserved between Drosophila and mammals and are predicted to function in iron absorption. Despite many similarities, some of the players with known roles in mammals are not conserved in Drosophila. Therefore, significant gaps remain in our present knowledge of how iron is acquired from the insect’s diet. Yet, the immune response, the maintenance of circadian rhythms and a number of developmental and aging-related processes are known to depend on iron, meaning that further research into iron homeostasis in the Drosophila model is required.

2. Brief Overview of Iron Absorption in Mammals

A number of comprehensive reviews have been published recently that describe iron absorption in mammals [16,20,21,22,23], so only a very brief summary is presented here as a means to introduce the key proteins involved in the process. First we describe the transport of iron through the epithelial cell in the duodenum [24] and then we discuss the regulation of this process by systemic signals (Figure 1).

2.1. Iron Trafficking through the Enterocyte

Iron absorption is complete when the metal ion has crossed the duodenal enterocyte and has been delivered to transferrin in the circulatory system [25]. Divalent metal transporter-1 (DMT1) is currently the only known transporter for the cellular uptake (import) of non-heme iron [26]. The duodenal lumen is an oxidizing environment where most iron is present in the ferric state, yet only the reduced (ferrous) form of iron is transported through DMT1. To facilitate iron absorption, duodenal cytochrome b (Dcytb) reduces ferric iron [27]. Iron is also absorbed in the form of heme, which is internalized through the heme carrier protein-1 (HCP1) [28], which also transports folate [29]. Heme oxygenases breakdown heme into CO, ferrous iron and biliverdin [30]; biliverdin is further converted to bilirubin by Biliverdin Reductase. The Multidrug Resistant Protein-2 (MRP2) is localized in the villi [31] and can export bilirubin from the cell [32]. Currently, the mechanism of iron trafficking through the cell is not fully understood. Specifically, it remains to be determined whether Poly(C) binding proteins (PCBPs), which can bind three ferrous irons and deliver them to ferritin [33], have a function in iron absorption or whether iron traffics in a labile form. Ferritin sequesters iron in the enterocyte and blocks its release to the circulation in a regulated manner [34]. Iron is released from enterocytes by ferroportin [35], from which it is released only after oxidation by the multicopper oxidase (MCO) Hephaestin [36]. The ferric iron is then bound by transferrin, which is the main source of iron for cells in peripheral tissues expressing the transferrin receptor-1 [37]. Copper is also required for iron absorption because it is a prosthetic factor in Hephaestin (reviewed in [38]).
Figure 1. Simplified scheme of iron absorption in mammals. A typical enterocyte of the duodenum of the mammalian intestine has uptake transporters for iron (DMT1) and heme (HCP1) localized in the apical membrane. An iron export transporter (ferroportin) is localized in the basolateral membrane. Ferric iron is reduced by Dcytb prior to import and oxidized by Hephaestin upon export. Iron is stored locally in the enterocyte in ferritin. Whether the iron chaperone PCBP has a role in iron absorption remains to be determined (indicated by a question mark). Heme oxygenases release iron from heme. The large byproduct of this reaction (biliverdin) is modified and secreted into the gut lumen though the Multidrug Resistant Protein-2 (MRP2) transporter. Iron absorption is regulated at the systemic level by hepcidin, which is secreted by the liver hepatocytes in response to various physiologic stimuli. Local cellular regulation also occurs via the Hypoxia Inducible Factors (HIFs) and Iron Regulatory Proteins (IRPs) and may be influenced by circulating levels of erythropoietin (EPO).
Figure 1. Simplified scheme of iron absorption in mammals. A typical enterocyte of the duodenum of the mammalian intestine has uptake transporters for iron (DMT1) and heme (HCP1) localized in the apical membrane. An iron export transporter (ferroportin) is localized in the basolateral membrane. Ferric iron is reduced by Dcytb prior to import and oxidized by Hephaestin upon export. Iron is stored locally in the enterocyte in ferritin. Whether the iron chaperone PCBP has a role in iron absorption remains to be determined (indicated by a question mark). Heme oxygenases release iron from heme. The large byproduct of this reaction (biliverdin) is modified and secreted into the gut lumen though the Multidrug Resistant Protein-2 (MRP2) transporter. Iron absorption is regulated at the systemic level by hepcidin, which is secreted by the liver hepatocytes in response to various physiologic stimuli. Local cellular regulation also occurs via the Hypoxia Inducible Factors (HIFs) and Iron Regulatory Proteins (IRPs) and may be influenced by circulating levels of erythropoietin (EPO).
Nutrients 05 01622 g001

2.2. Regulation of Iron Absorption

Iron absorption depends both upon cellular Iron Regulatory Proteins (IRPs) and Hypoxia Inducible Factors (HIFs) and upon systemic regulation through the iron hormone hepcidin [39]. Hepcidin is secreted by liver hepatocytes in response to iron-related stimuli [39]. Hepcidin binds directly to ferroportin on enterocytes and other cells, which promotes the internalization and degradation of the protein [40]. IRPs respond to low iron by repressing the translation of ferritin, ferroportin, mitochondrial aconitase and HIF-2α, while they increase expression of DMT1 and Transferrin Receptor-1 (TfR1) via Iron Responsive Elements (IREs) present in the 5′UTR and 3′UTR, of the respective mRNAs [41,42]. Oxygen levels also have an impact on IRP activation [43]. HIF-2α is a major player in regulating iron absorption by directly controlling the transcription of iron transporters in the intestine [44,45,46,47]. Hypoxia is also sensed in the kidney, which releases erythropoietin (EPO) into circulation [48,49,50]. As erythropoiesis requires high amounts of iron for the production of new red blood cells, it is no surprise that duodenal enterocytes have responsive erythropoietin receptors (EPO-Rs) [48,49,50]. Thus, peripheral tissues can systemically regulate iron absorption by signaling to the intestinal mucosa to absorb iron more actively and in addition the mucosal cells can also actively sense iron and oxygen levels.

3. Early Studies of Iron Homeostasis in Drosophila

This review focuses on studies that involved Drosophila melanogaster, however it is important to highlight at the outset a number of excellent review articles that describe relevant studies in other insects [51,52,53], including recent publications focusing on insect iron-storage ferritin complexes [54] and high-affinity, iron-binding transferrin proteins [55].
In 1952, Poulson and Bowen presented an elegant demonstration of their histochemical findings on ferric iron and copper present in the intestine of Drosophila [17]. These authors used ferrocyanide (Prussian Blue Stain) to visualize iron while modifying the amounts of iron in the diet. In addition, they used radioactive Fe59 to trace the element in the larval tissues. They detected an iron region in the middle midgut and showed that iron added in the diet accumulated in the anterior midgut. Their speculation that iron was present in the form of ferritin was proven correct later [19]. In high doses of dietary iron, ferritin iron was excreted from calycocytes (also known as copper cells) and also made its way to the posterior midgut. Despite significant progress in sixty years, the opening remark by Poulson and Bowen continues to resonate today [17]: “Although the importance of the inorganic constituents of cells has long been recognized, remarkably little is actually known concerning their specific localizations and functions in the structure and economy of the cell.” We have modified and redrawn their introductory figure adding new molecular findings to corroborate their classic work (Figure 2).
The next seminal paper presenting a significant advance in the field describes work in another insect, the lepidopteran Calpodes ethlius. Larvae were again reared with diets differing in iron concentration and observed under the electron microscope using fixation conditions that permit direct visualization of the endogenous ferritin [19]. Locke and Leung confirmed that iron was stored in ferritin particles that were abundant in the secretory pathway of intestinal cells. This was in contrast to the situation in mammals, where ferritin resides in the cytosol. However, many controls were used to prove this point beyond doubt [19,56]. These authors also reported that ferritin was secreted to the gut lumen in iron overload conditions.
Figure 2. Iron absorption likely takes place in the anterior midgut where all transport proteins studied to date are found. Copper cells are acid-secreting cells, which have also been shown to secrete iron-loaded ferritin in the intestinal lumen. The iron cells express ferritin constitutively and likely serve an iron storage function. Iron cells regulate iron homeostasis independently of IRP-1A and ferritin independently of iron. The posterior midgut appears to be involved with iron homeostasis only in conditions of iron overload. How the different intestinal domains interact with each other remains unknown. For higher resolution images of the different cell types in this diagram the readers are referred to a beautiful representation based on ultrastructure studies performed recently for this tissue [57]. Our diagram is also based on the now 61-year-old study of iron and copper localization in insects [17]. The most posterior part of the midgut (not shown) has no involvement in iron homeostasis, but is a compartment specialized in zinc storage [58,59].
Figure 2. Iron absorption likely takes place in the anterior midgut where all transport proteins studied to date are found. Copper cells are acid-secreting cells, which have also been shown to secrete iron-loaded ferritin in the intestinal lumen. The iron cells express ferritin constitutively and likely serve an iron storage function. Iron cells regulate iron homeostasis independently of IRP-1A and ferritin independently of iron. The posterior midgut appears to be involved with iron homeostasis only in conditions of iron overload. How the different intestinal domains interact with each other remains unknown. For higher resolution images of the different cell types in this diagram the readers are referred to a beautiful representation based on ultrastructure studies performed recently for this tissue [57]. Our diagram is also based on the now 61-year-old study of iron and copper localization in insects [17]. The most posterior part of the midgut (not shown) has no involvement in iron homeostasis, but is a compartment specialized in zinc storage [58,59].
Nutrients 05 01622 g002
Prior to the era of molecular cloning, Massie et al. [18] demonstrated that iron accumulates during the ageing process in Drosophila. These authors also suggested that inhibiting iron absorption could enhance the lifespan of fruit flies [60]; however, their statement should be qualified since their experimental set up used dietary tea extracts. Looking at how iron was sensed in cells of Drosophila, Rothenberger et al. [61] demonstrated that Drosophila also carried an active IRP/IRE system.

4. Genes with a Known Function in Iron Absorption in Drosophila melanogaster

The first iron-related genes to be cloned in the fly were those encoding the subunits of the iron-storage ferritin protein [62,63], transferrin (Tsf1; [64]) and two IRP1 homologs [65]. Reverse genetic approaches, i.e., the discovery of mutations in genes that caused specific phenotypic alterations led to the characterization of the DMT1 homolog Malvolio (Mvl; [66]) and of the melanotransferrin homolog Tsf2 [67]. The Drosophila genome sequence [68] informed studies on a mitochondrial form of ferritin [69] and on MCOs [70,71]. Table 1 summarizes the key studies on these genes and this section describes what is known about the role of these proteins in relation to dietary iron absorption.
Table 1. Proteins with known functions in Drosophila iron absorption.
Table 1. Proteins with known functions in Drosophila iron absorption.
Protein name (mammals)Protein name (Drosophila)Key roleReferences
DMT1MalvolioMalvolio mutants are iron deficient and show behavioral defects.[66,70,72,73]
FerritinFerritinDrosophila ferritin is a secreted protein required for iron storage and also iron absorption.[2,15,54,62,63,74,75,76]
TransferrinTsf1Tsf1 is an immune-responsive gene. Whether it traffics iron between cells remains unclear.[53,55,64,77,78]
MelanotransferrinTsf2Tsf2 is required for the assembly of septate junctions in epithelial cells.[67]
HephaestinMCO1, MCO3MCO1 and MCO3 are putative ferroxidases. Both show loss-of-function phenotypes with respect to iron homeostasis.[70,71]
IRP1, IRP2IRP-1AIRP-1A regulates ferritin and succinate dehydrogenase translation via IREs.[61,79,80,81,82,83,84,85,86]

4.1. Mvl, the Drosophila Homolog of DMT-1

Mvl is the Drosophila DMT1 homolog, originally identified in a mutagenesis screen for genes affecting taste behavior in flies [66]. Mvl mutants lost a characteristic sugar-preference trait of wild type flies and this phenotype could be rescued by exposure to excess dietary iron [72]. Tissue staining with a specific antibody revealed that the Mvl protein resides in the anterior and posterior parts of the middle intestine, as well as in the Malpighian tubules, brain and testis [73]. Iron stores are depleted in Mvl mutants [70,87]. The Mvl mutant also suppressed iron accumulation in the intestine, caused by RNA interference (RNAi) of ferritin in this tissue [74]. Hence, sufficient evidence supports the conclusion that Mvl is an iron import protein of Drosophila.

4.2. Ferritin

Drosophila makes two types of ferritin: the testis-specific mitochondrial protein encoded by the X-chromosome Ferritin-3-Heavy-Chain-Homolog (Fer3HCH) gene, which has no known contribution to iron absorption and a minor (or testis-specific) role in overall iron homeostasis [69], and the major secretory type responsible for systemic iron storage [2,15] and iron absorption [74]. Drosophila ferritin is the product of two genes, Fer1HCH and Fer2LCH, located adjacent to each other on the left arm of the 3rd chromosome [75]. Fer1HCH has ferroxidase activity required for iron loading and Fer2LCH provides the iron nucleation sites required for the mineralization of the ferrihydrite iron core [2,62,63,76]. Each ferritin molecule is composed of 12 Fer1HCH and 12 Fer2LCH subunits [2,74,76]. Radioactive tracing showed that most ingested iron accumulates in Drosophila ferritin [2,69]. Analysis of single insertion mutants that disrupt Fer1HCH and Fer2LCH, respectively, showed that disruption of either gene product reduces total ferritin levels in whole flies and leads to embryonic or early larval lethality [2]. Midgut-specific RNAi of ferritin resulted in iron accumulation in the intestine but systemic iron deficiency [74].

4.3. MCOs

Drosophila melanogaster has four MCOs in its genome [88]. There is preliminary biochemical evidence that Drosophila MCO1 can oxidize ferrous iron; and RNAi of MCO1 resulted in iron-depleted flies consistent with the proposal that MCO1 is an intestinal ferroxidase implicated in iron absorption [71]. In contrast, MCO3 mutants accumulated more iron in the iron region of the intestine and restored the depleted iron stores of Mvl mutants, but with milder effects in overall iron homeostasis [70]. A comprehensive biochemical characterization of these enzymes is needed to confirm their proposed ferroxidase activities.

4.4. Transferrins

Tsf1 is an abundant protein in the hemolymph [64,78]. Whether it traffics iron between tissues remains unclear. Ventral furrow formation during the early development of the Drosophila embryo appears to require differential regulation of Tsf1 in ventral and lateral cells, but the mechanistic details of why this needs to be so, are not understood [77]. There is evidence that Tsf1 is an immune-responsive gene [64,78]. Overall, further studies are required to determine whether Drosophila Tsf1 performs a similar role to mammalian transferrin in serving as an iron transport carrier between cells.
More is known about Tsf2, the fly homolog of melanotransferrin, whose function in mammals remains unclear except for its up-regulation in melanomas [89]. In a seminal paper, Tiklová et al. [67] showed that septate junction assembly during epithelial maturation relies on endocytosis and apicolateral recycling of Tsf2. Tsf2 is a component of the epithelial septate junctions. In particular, the binding of iron to Tsf2 was shown to be required for the epithelial structure that blocks paracellular iron absorption to be formed [67]. A third homolog of transferrin, Tsf3, has not been functionally characterized to date [53,55].

4.5. IRP/IRE

A protein with IRE binding activity and IREs present within the 5´UTRs of the mRNAs of succinate dehydrogenase subunit B and Fer1HCH were described soon after the discovery of this regulatory system in mammals [13,65,79,80,81,82,83,84]. The IREs appear to have evolved independently in this lineage by convergent evolution targeting the same key players [85]. The logical implication is that under iron deficiency, ferritin (each molecule of which can bind thousands of iron atoms) and the citric acid cycle (which, coupled to oxidative phosphorylation, accounts for a significant portion of the mitochondrial iron proteins) need to be physiologically suppressed. Drosophila makes two highly homologous IRP1-like proteins (IRP-1A and IRP-1B) encoded by different genes [65]. In the evolution of the IRP/IRE system, the ancient cytosolic aconitase was duplicated in insects with one variant (IRP-1A) acquiring IRE-specific binding [86]. Studies of ferritin regulation in the intestine suggested that IRP-1A is absent from the iron region, but that it regulates ferritin in the anterior and posterior midgut (Figure 2) [15,90].

5. Genes with a Known Function in Iron Absorption in Mammals that Are Conserved but Have Not Been Studied in Drosophila melanogaster

In this section we consider other key players in mammalian iron absorption (Figure 1) that are conserved in Drosophila, but where studies are lacking either entirely or with respect to the function of these genes in iron absorption (Table 2).
Table 2. Genes with putative functions in Drosophila iron absorption.
Table 2. Genes with putative functions in Drosophila iron absorption.
Protein name (mammals)Putative homologous genes in DrosophilaCommentsReferences
DcytbCG1275; nemyOne of two fly homologs (nemy) has a function in learning and memory.[91,92]
HCP1CG30345Flybase reports low levels of expression for this gene possibly involved in cellular heme uptake.[93]
FLVCRCG1358Flybase reports low levels of expression for this gene possibly involved in cellular heme export. RNAi in clock neurons caused disrupted circadian rhythms.[93,94]
HO1, HO2HOHO is required for development; it degrades but is not inducible by heme.[95,96]
HIFα, HIFβsima; tangoHIF signaling is conserved in Drosophila but not studied in the context of iron.[97,98,99,100,101,102]

5.1. Dcytb Homologs

Dcytb reduces ferric iron and facilitates iron absorption [27], especially under hypoxic conditions [103]. Mammals also express Lcytb, a close homolog of Dcytb that localizes in the lysosome, but has not yet been studied functionally [104]. The Drosophila genome has two homolog genes, termed CG1275 and no extended memory (nemy). nemy was recovered from a genetic screen for learning and memory mutants [91] and could function in memory formation by regulating intravesicular peptidyl alpha-hydroxylating monooxygenase (PHM) activity and the formation of amidated neuropeptides [92]. Because PHM is a copper-dependent enzyme [105] and Dcytb is known to also reduce copper [106,107], the function of nemy in learning and memory may therefore be mediated through copper reduction. A recent report showed that nemy is strongly induced by hypoxia [108]. Whether nemy has a role in iron absorption remains to be elucidated. Likewise, CG1275 has not been studied to date.

5.2. HCP1 Homolog

Soon after the proposal that HCP1 is the heme import protein localized in the apical membrane of the duodenal enterocytes [28], it was shown that the same protein was undoubtedly also (or primarily) serving as a folate transporter [29]. However, further studies have suggested that HCP1 likely transports both heme and folate [109,110,111]. Drosophila has a clear homolog encoded by CG30345, which has not been studied to date.

5.3. FLVCR Homolog

The cellular receptor for feline leukemia virus subgroup C (FLVCR) has been identified as a human heme exporter that is essential for erythropoiesis [112]. Whether FLVCR plays a role in heme absorption from the diet remains unclear [113]; however a clear Drosophila homolog (CG1358) for this transporter has been identified and a putative function in the maintenance of circadian rhythms has been ascribed to the fly gene [94].

5.4. Heme Oxygenase

Drosophila HO has been biochemically characterized [95]. HO RNAi led to larval and pupal lethality, with a doubling of heme content measured in the affected individuals [96]. More experiments are required to evaluate if dietary heme represents a source of iron for Drosophila and consequently whether HO has a specific function in the intestine during the process of intestinal iron absorption or not. Interestingly, according to Flybase, HCP1 shows a developmental peak in expression at the end of the third instar larva (after the larvae cease to eat) and of HO during the first day of metamorphosis (when extensive histolysis takes place) [93].

5.5. HIF

Sima and Tango are the HIFα and HIFβ homologs in Drosophila [97]. Several laboratories are using Drosophila to study the hypoxic response, however these works are beyond the scope of this review and the reader is referred to reviews of the literature in this rapidly expanding field [98,99,100]. Iron is a cofactor for the prolyl hydroxylase Fatiga and affects Sima stability [101] and hypoxia directly affects iron homeostasis in mammals (see Section 2.2. Regulation of Iron Absorption) and other invertebrates [114,115]. Although it has been established that tracheal cells sense hypoxia and induce terminal branch sprouting [102], the effect of sima and tango on iron homeostasis has not been directly investigated to date.

6. Differences in Iron Homeostasis between Mammals and Insects

Mammals and insects rely on different respiratory organs and differ on how they distribute the oxygen in the whole body: Mammals use lungs and the circulatory system to systemically distribute the oxygen whereas the insects use the tracheal system to distribute the oxygen [116]. In mammals, erythropoiesis occurs in the bone marrow and red blood cells are the carriers of oxygen from the lungs to tissues and of carbon dioxide from the tissues to the lungs. The iron requirement for hemoglobin production by far exceeds all other demands for iron [117]. In contrast, the insect tracheal system is a continuous tubular network that provides air directly to every organ and tissue throughout the body. Therefore, the insect circulatory system, which is open, is not primary used to transport oxygen and Drosophila hemoglobin is only expressed locally at the tip of the trachea [118,119,120]. Such a major difference in the organization of the respiratory and circulatory systems of insects and mammals is also reflected in the regulation of systemic iron homeostasis: the whole hemochromatosis-related pathway, including hepcidin, is lacking and also there is no EPO in Drosophila. Therefore, Drosophila is a poor model for some human conditions (i.e., hemochromatosis or some forms of anemia) but an excellent model to study tissue specific functions for iron, in processes that can be “masked” in the vertebrate models because of the physiologic priority of shuttling iron into erythropoiesis.
Another difference that should be kept in mind is the subcellular localization of ferritin, which in Drosophila and most other insects resides within the secretory system (ER and Golgi) and is secreted in the hemolymph in large quantities [2,19,52,121]. Although ferritin clearly serves an iron storage function in Drosophila [2,15,74] it remains to be shown whether it also serves as a transporter of iron between cells [52], as has been definitively shown for one of two tick ferritins [122]. Below, we briefly discuss the gaps in our current knowledge arising from the lack of conservation from key players of vertebrate iron homeostasis (Table 3).
Table 3. Mammalian iron metabolism proteins with no orthologs in Drosophila.
Table 3. Mammalian iron metabolism proteins with no orthologs in Drosophila.
Protein name (mammals)Key questions arising
FerroportinHow do insects export iron from cells?
HepcidinHow do insects signal peripheral iron sufficiency?
ErythropoietinNo erythropoiesis in insects; is there a diffusible signal for systemic hypoxia?
Transferrin ReceptorIs there a functional TsfR in flies? What is the function of Tsf1?
Is there a ferritin receptor and does ferritin mediate systemic iron transport?

6.1. Ferroportin

Ferroportin is the only known export protein for ferrous iron in mammalian cells [123]. There is no ferroportin homolog encoded in the Drosophila genome [52]. Therefore the question of how iron can exit an insect cell is of paramount importance, and a key unknown factor in the process of intestinal iron absorption. If an iron export protein exists in insects, it will be of interest to see if such a protein is also conserved in mammals. Ferritin secretion could be one mechanism of cellular iron export [52,122], although it would mean that only large quantities of iron could be released at any given time.

6.2. Hepcidin

The closest homologs of the iron-hormone hepcidin in Drosophila are the antimicrobial peptides of the defensin type [124,125]. These have been implicated in epithelial homeostasis in the Drosophila gut through their effects on intestinal microbiota [126]. They may also have an alternative function in systemic iron homeostasis, but this hypothesis requires experimental testing. In any case, it is highly probable that insects will signal their peripheral iron demands to the intestine, perhaps using circulating ferritin as a direct signal [2,70].

6.3. Erythropoietin

Unsurprisingly, there is no gene encoding for a protein with similarity to erythropoietin present in the fly genome. Nevertheless, it is worth asking if hypoxia releases any humoral factor in flies; in this sense the recent implication of estrogen-related receptor in the hypoxic response is intriguing [108]. The Malpighian tubules of Drosophila are its major excretory organ [127,128]. The Malpighian tubules clearly receive and respond to signals of stress or immune challenge [129], but it is not clear if they in turn secrete any hormones into the hemolymph, or if this is a function reserved for the fat bodies and other glands.

6.4. Transferrin Receptor

There has been a fair amount of discussion in reviews of insect iron metabolism over the fact that no Transferrin Receptor gene was found by homology searches in insect genomes [52,53,55]. Given the abundance of Tsf1 in the hemolymph, it is very likely that an uptake system for this protein exists. Either it has the same ancestral gene as mammalian TsfR, but has diverged so much in sequence that it cannot be recognized as such, or an independent Tsf uptake system exists in flies. Therefore, until functional experiments are performed it is better to leave the question of an insect TsfR open. The situation is somewhat similar with respect to putative ferritin receptors in mammals, where a few proteins have been suggested to participate in serum ferritin binding and internalization, but conclusive functional evidence is largely missing [130,131,132,133,134,135]. If hemolymph ferritin does indeed traffic iron from one cell to another, then an insect ferritin receptor is another important protein that awaits its discovery and functional analysis.

7. Functional Requirements of Iron in Drosophila

The functional relevance of iron in Drosophila biology transcends its key role in the generation of ATP as a cofactor in Krebs cycle and oxidative phosphorylation enzymes. As discussed above the specific requirement of iron binding to melanotransferrin for the formation of septate junctions in the formation of epithelia [67]. Below we discuss other developmental processes already known to depend on iron proteins. We also describe studies implicating iron in the immune and heat shock responses, the involvement of iron proteins in the maintenance of the circadian rhythm and finally the role of iron in neuronal degeneration, a field of study where many Drosophila models have been generated (Table 4) [136].
Table 4. Higher biological processes influenced by iron availability.
Table 4. Higher biological processes influenced by iron availability.
General processSpecific functionReferences
DevelopmentEpithelial junction formation[67]
Spermatogenesis[131]
Cell proliferation[135,136]
Ventral furrow formation [77]
Immune ResponseHemolymph ferritin and transferrin respond to infection[64,78]
Zygomycosis[137]
Wolbachia[138,139]
Sindbis viral entry[140]
Heat Shock ResponseUnknown (ferritin and transferrin are heat shock inducible)[141,142,143]
BehaviorTaste perception[63,69]
Circadian Rhythm[94]
Human Disease ModelsFriedreich’s Ataxia[9,144,145,146,147,148,149,150,151]
Alzheimer’s and Parkinson’s Disease[152,153,154,155]
Restless Legs Syndrome[156,157]
Neurodegeneration[158,159,160,161]

7.1. Iron Requirements for the Development of Drosophila melanogaster

From the genes studied so far, Fer1HCH, Fer2LCH and Tsf2 mutants are embryonic lethal [2,67], whereas Mvl and MCO3 mutants survive to adulthood [70]. Drosophila larvae bearing insertions in Heat shock protein cognate 20, a gene required for iron-sulfur cluster assembly, failed to grow during the 3rd instar larval stage and never initiated metamorphosis [90]. Drosophila mutant larvae for aminolevulinate synthase, encoding for the rate-limiting enzyme in heme biosynthesis, suffered massive water loss, possibly due to failed formation of a dityrosine-based cuticular barrier [162]. Mutants in the mitoferrin (mfrn) gene reached adulthood, but were male-sterile [163]. As mfrn is thought to transport iron into the matrix of mitochondria [164], it appears that mitochondrial iron is essential for spermatogenesis [163,165]. Whether the testis-specific mitochondrial ferritin (Fer3HCH [69]) or other iron genes have a role in this process has not yet been demonstrated; though a mutant recovered in a genetic screen for altered iron homeostasis [166] turned out to be male-sterile under iron deficient conditions (T.P. and F.M. unpublished observations). Iron has also been implicated in cell proliferation [167] and ferritin iron has been suggested to serve as a mitogen [168]; therefore one possibility is that the cell proliferation of spermatids requires an adequate supply of iron to occur. In any event, our present understanding of the role of iron in developmental processes remains rather limited, but there is sufficient evidence that disrupting iron homeostasis affects normal embryogenesis [2] and dietary iron chelation halts larval growth [69]. The recent use of the Synchrotron to visualize metals in tissues adds a further high-resolution in situ methodology for the study of the mutants already available [169].

7.2. Iron and the Immune Response

Iron sequestration is an important and evolutionarily conserved component of the innate immune response, because acquisition of iron is vital for pathogenic growth [170]. The importance of iron in Drosophila immunity can be demonstrated by the fact that regulation of iron proteins is increased in the presence of bacterial or fungal infections; specifically expression of Tsf1 mRNA [64] and protein [78] increases and a specific cleavage in Fer2LCH has also been detected [78]. Given that the mechanisms between the innate immune systems of both Drosophila and human are highly conserved [171], Drosophila may be used in trying to understand molecular aspects of poorly understood pathobiology in humans. For instance, it is known that individuals with elevated serum iron levels are at increased risk for zygomycosis, indicating a role of iron metabolism in the pathobiology of zygomycosis [172]. When Drosophila was injected with a standardized amount of Zygomycetes spores, rapid infection and death of wild type flies was observed, which could be partially blocked by iron chelation [137]. Similarly, Wolbachia, a natural symbiotic host of insects, directly affected ferritin levels in its hosts [138,139]. Beyond the “iron wars” between host and bacterial or fungal pathogens, DMT1 was recently identified as the cellular receptor for Sindbis virus, a prototypical member of the mosquito-borne alphaviruses [140]. As with the role of iron in development, it looks like there is much more left to discover over iron’s multiple functions in the provision of immunity.

7.3. Iron and the Heat-Shock Response

Three recent studies have shown significant induction of Tsf1 and ferritin proteins in response to growing the flies at high temperatures [141,142,143]. The precise function of these proteins in protection from heat stress remains unclear.

7.4. Iron Influences the Behavior of Drosophila melanogaster

The first evidence that mutations in a single gene could affect the behavior of an organism was obtained in studies of the circadian rhythms of Drosophila [173]. Rhythmic behavior is mediated by a group of about 150 “circadian” neurons in the central brain [174]. Heme has been previously implicated in the function of the circadian clock [175,176] and is a cofactor in relevant Drosophila nuclear receptors [177,178]. Our own investigations on whether heme biosynthetic genes were required in “circadian” neurons were inconclusive; instead we discovered that iron-sulfur cluster biosynthesis genes were implicated in the maintenance of circadian rhythms [94]. We also observed phenotypes with RNAi of Fer2LCH, but not Fer1HCH, and with two other genes: Tsf3 and CG1358, (FLVCR heme exporter homolog) [94]. Overexpression of ferritin in glial cells also led to an age-dependent decline in the ability to sustain circadian rhythms [158]. Support of the notion that iron may be directly involved in the maintenance of circadian rhythms came also from parallel studies in plants, where a retrograde signal from chloroplasts to nucleus signaling an iron deficiency was suggested to affect the period length of the clock [179,180,181,182]. More work is required to elucidate the mechanistic details of how iron may affect the circadian clock in flies.
Another behavior affected by iron relates to the perception of sweet taste [66,72]. As discussed earlier, Mvl has a key role in the preference for sugar shown by Drosophila flies (see Section 4.1). It appears that behavioral attraction to sugar may lie behind honeybee division of labor; addition of different metals in the hives of honeybees changed their foraging behavior [183,184].

7.5. Iron and Models of Human Disease

By far the most studied iron-related disease in a Drosophila model system is Friedreich’s ataxia [9,144,145,146,147,148,149,150,151], which in humans arises from reduced expression of the iron-sulfur cluster biosynthesis gene frataxin [185,186,187]. RNAi of frataxin in Drosophila resulted in adult flies with locomotion defects [146], likely explained by defective mitochondrial axonal transport and membrane potential [149] but also by the reduced activity of mitochondrial complexes [145]. RNAi of frataxin also increased sensitivity to oxidative stress [146,147] and resulted in accumulation of lipids and lipid peroxidation [9]. Iron has also been implicated in the Drosophila models of Restless Legs Syndrome [156,157], Parkinson’s [152,153] and Alzheimer’s [154,155] disease and in metal-induced neurodegeneration [159,160,161]. Finally, a number of dietary studies of compounds with iron chelating properties have been reported, but it is difficult to evaluate how much the effects seen in such studies are a direct consequence of a reduction in intestinal iron absorption [188,189,190]. For more information on “what can Drosophila teach us about iron-accumulation neurodegenerative disorders” the reader is referred to a recent review [136].

8. Conclusions

In contrast to other fields of biology, where work in Drosophila pioneered our present understanding, iron absorption is understood more extensively in mammals compared to Drosophila. This review compared and contrasted the two systems and pointed to future work required. The physiological significance of iron can be ascertained from studies of its impact on neurodegenerative disorders, development, immune response and behavior. Iron accumulation in different Drosophila species is conserved [191]. Drosophila shares with mammals the following proteins that function in iron absorption: Mvl (DMT1 homolog; iron import), ferritin (iron storage), IRP-1A (iron regulation), MCO1/MCO3 (ferroxidases), Tsf1 and Tsf2 (transferrin and melanotransferrin homologs, respectively). There are other conserved proteins which have not been investigated over a possible role in iron absorption or regulation, including CG1275/nemy (Dcytb homolog), CG30345 (HCP1 homolog), HO and sima/tango (HIFa/b homologs). The differences in physiology between mammals and Drosophila may explain why certain key players involved in iron regulation have no known orthologs, like hepcidin, and erythropoietin, whilst the lack of conservation of ferroportin and TsfR are perplexing, since cellular iron uptake and release are fundamental processes that would have been expected to be phylogenetically conserved. Thus, further research in this field is warranted.

Acknowledgments

This work was supported by CONACYT (project 179835 and graduate studentship to T.P.). The authors thank Christoph Metzendorf (Heidelberg University, Germany) and Mohamad Aslam (University of Cambridge, United Kingdom) and three anonymous reviewers for critical comments on the manuscript. The authors are grateful to James Waters (Princeton University, United States of America) for permission to use his photograph in the graphical representation that accompanies this paper on-line.

Conflict of Interest

The authors declare no conflict of interest.

References

  1. Law, J.H. Insects, oxygen, and iron. Biochem. Biophys. Res. Commun. 2002, 292, 1191–1195. [Google Scholar] [CrossRef]
  2. Missirlis, F.; Kosmidis, S.; Brody, T.; Mavrakis, M.; Holmberg, S.; Odenwald, W.F.; Skoulakis, E.M.; Rouault, T.A. Homeostatic mechanisms for iron storage revealed by genetic manipulations and live imaging of Drosophila ferritin. Genetics 2007, 177, 89–100. [Google Scholar] [CrossRef]
  3. Warburg, O. Iron, the oxygen-carrier of respiration-ferment. Science 1925, 61, 575–582. [Google Scholar]
  4. Brown, N.C.; Eliasson, R.; Reichard, P.; Thelander, L. Nonheme iron as a cofactor in ribonucleotide reductase from E. coli. Biochem. Biophys. Res. Commun. 1968, 30, 522–527. [Google Scholar] [CrossRef]
  5. Clark, D.V. Molecular and genetic analyses of Drosophila Prat, which encodes the first enzyme of de novo purine biosynthesis. Genetics 1994, 136, 547–557. [Google Scholar]
  6. Chavez, V.M.; Marques, G.; Delbecque, J.P.; Kobayashi, K.; Hollingsworth, M.; Burr, J.; Natzle, J.E.; O’Connor, M.B. The Drosophila disembodied gene controls late embryonic morphogenesis and codes for a cytochrome P450 enzyme that regulates embryonic ecdysone levels. Development 2000, 127, 4115–4126. [Google Scholar]
  7. Warren, J.T.; Petryk, A.; Marques, G.; Jarcho, M.; Parvy, J.P.; Dauphin-Villemant, C.; O’Connor, M.B.; Gilbert, L.I. Molecular and biochemical characterization of two P450 enzymes in the ecdysteroidogenic pathway of Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 2002, 99, 11043–11048. [Google Scholar] [CrossRef]
  8. Andersson, K.K.; Vassort, C.; Brennan, B.A.; Que, L., Jr.; Haavik, J.; Flatmark, T.; Gros, F.; Thibault, J. Purification and characterization of the blue-green rat phaeochromocytoma (PC12) tyrosine hydroxylase with a dopamine-Fe(III) complex. Reversal of the endogenous feedback inhibition by phosphorylation of serine-40. Biochem. J. 1992, 284, 687–695. [Google Scholar]
  9. Navarro, J.A.; Ohmann, E.; Sanchez, D.; Botella, J.A.; Liebisch, G.; Molto, M.D.; Ganfornina, M.D.; Schmitz, G.; Schneuwly, S. Altered lipid metabolism in a Drosophila model of Friedreich’s ataxia. Hum. Mol. Genet. 2010, 19, 2828–2840. [Google Scholar] [CrossRef]
  10. Lill, R.; Hoffmann, B.; Molik, S.; Pierik, A.J.; Rietzschel, N.; Stehling, O.; Uzarska, M.A.; Webert, H.; Wilbrecht, C.; Muhlenhoff, U. The role of mitochondria in cellular iron-sulfur protein biogenesis and iron metabolism. Biochim. Biophys. Acta 2012, 1823, 1491–1508. [Google Scholar]
  11. Pantopoulos, K.; Hentze, M.W. Rapid responses to oxidative stress mediated by iron regulatory protein. EMBO J. 1995, 14, 2917–2924. [Google Scholar]
  12. Jaakkola, P.; Mole, D.R.; Tian, Y.M.; Wilson, M.I.; Gielbert, J.; Gaskell, S.J.; von Kriegsheim, A.; Hebestreit, H.F.; Mukherji, M.; Schofield, C.J.; et al. Targeting of HIF-alpha to the von Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science 2001, 292, 468–472. [Google Scholar] [CrossRef]
  13. Missirlis, F.; Hu, J.; Kirby, K.; Hilliker, A.J.; Rouault, T.A.; Phillips, J.P. Compartment-specific protection of iron-sulfur proteins by superoxide dismutase. J. Biol. Chem. 2003, 278, 47365–47369. [Google Scholar]
  14. Sheftel, A.D.; Mason, A.B.; Ponka, P. The long history of iron in the Universe and in health and disease. Biochim. Biophys. Acta 2012, 1820, 167–187. [Google Scholar]
  15. Mehta, A.; Deshpande, A.; Bettedi, L.; Missirlis, F. Ferritin accumulation under iron scarcity in Drosophila iron cells. Biochimie 2009, 91, 1331–1334. [Google Scholar] [CrossRef]
  16. Knutson, M.D. Iron-sensing proteins that regulate hepcidin and enteric iron absorption. Annu. Rev. Nutr. 2010, 30, 149–171. [Google Scholar] [CrossRef]
  17. Poulson, D.F.; Bowen, V.T. Organization and function of the inorganic constituents of nuclei. Exp. Cell Res. 1952, 2, 161–180. [Google Scholar]
  18. Massie, H.R.; Aiello, V.R.; Williams, T.R. Iron accumulation during development and ageing of Drosophila. Mech. Ageing Dev. 1985, 29, 215–220. [Google Scholar] [CrossRef]
  19. Locke, M.; Leung, H. The induction and distribution of an insect ferritin—A new function for the endoplasmic reticulum. Tissue Cell 1984, 16, 739–766. [Google Scholar] [CrossRef]
  20. Sharp, P.A. Intestinal iron absorption: Regulation by dietary & systemic factors. International journal for vitamin and nutrition research. J. Int. Vitam. Nutr. 2010, 80, 231–242. [Google Scholar] [CrossRef]
  21. Fuqua, B.K.; Vulpe, C.D.; Anderson, G.J. Intestinal iron absorption. J. Trace Elem. Med. Biol. 2012, 26, 115–119. [Google Scholar] [CrossRef]
  22. Evstatiev, R.; Gasche, C. Iron sensing and signalling. Gut 2012, 61, 933–952. [Google Scholar] [CrossRef]
  23. Pantopoulos, K.; Porwal, S.K.; Tartakoff, A.; Devireddy, L. Mechanisms of mammalian iron homeostasis. Biochemistry 2012, 51, 5705–5724. [Google Scholar] [CrossRef]
  24. Gitlin, D.; Cruchaud, A. On the kinetics of iron absorption in mice. J. Clin. Investig. 1962, 41, 344–350. [Google Scholar] [CrossRef]
  25. Miret, S.; Simpson, R.J.; McKie, A.T. Physiology and molecular biology of dietary iron absorption. Annu. Rev. Nutr. 2003, 23, 283–301. [Google Scholar] [CrossRef]
  26. Gunshin, H.; Mackenzie, B.; Berger, U.V.; Gunshin, Y.; Romero, M.F.; Boron, W.F.; Nussberger, S.; Gollan, J.L.; Hediger, M.A. Cloning and characterization of a mammalian proton-coupled metal-ion transporter. Nature 1997, 388, 482–488. [Google Scholar] [CrossRef]
  27. McKie, A.T.; Barrow, D.; Latunde-Dada, G.O.; Rolfs, A.; Sager, G.; Mudaly, E.; Mudaly, M.; Richardson, C.; Barlow, D.; Bomford, A.; et al. An iron-regulated ferric reductase associated with the absorption of dietary iron. Science 2001, 291, 1755–1759. [Google Scholar] [CrossRef]
  28. Shayeghi, M.; Latunde-Dada, G.O.; Oakhill, J.S.; Laftah, A.H.; Takeuchi, K.; Halliday, N.; Khan, Y.; Warley, A.; McCann, F.E.; Hider, R.C.; et al. Identification of an intestinal heme transporter. Cell 2005, 122, 789–801. [Google Scholar] [CrossRef]
  29. Qiu, A.; Jansen, M.; Sakaris, A.; Min, S.H.; Chattopadhyay, S.; Tsai, E.; Sandoval, C.; Zhao, R.; Akabas, M.H.; Goldman, I.D. Identification of an intestinal folate transporter and the molecular basis for hereditary folate malabsorption. Cell 2006, 127, 917–928. [Google Scholar] [CrossRef]
  30. Raffin, S.B.; Woo, C.H.; Roost, K.T.; Price, D.C.; Schmid, R. Intestinal absorption of hemoglobin iron-heme cleavage by mucosal heme oxygenase. J. Clin. Investig. 1974, 54, 1344–1352. [Google Scholar] [CrossRef]
  31. Mottino, A.D.; Hoffman, T.; Jennes, L.; Vore, M. Expression and localization of multidrug resistant protein mrp2 in rat small intestine. J. Pharm. Exp. Ther. 2000, 293, 717–723. [Google Scholar]
  32. Jedlitschky, G.; Leier, I.; Buchholz, U.; Hummel-Eisenbeiss, J.; Burchell, B.; Keppler, D. ATP-dependent transport of bilirubin glucuronides by the multidrug resistance protein MRP1 and its hepatocyte canalicular isoform MRP2. Biochem. J. 1997, 327, 305–310. [Google Scholar]
  33. Shi, H.; Bencze, K.Z.; Stemmler, T.L.; Philpott, C.C. A cytosolic iron chaperone that delivers iron to ferritin. Science 2008, 320, 1207–1210. [Google Scholar] [CrossRef]
  34. Vanoaica, L.; Darshan, D.; Richman, L.; Schumann, K.; Kuhn, L.C. Intestinal ferritin H is required for an accurate control of iron absorption. Cell Metab. 2010, 12, 273–282. [Google Scholar] [CrossRef]
  35. 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]
  36. Vulpe, C.D.; Kuo, Y.M.; Murphy, T.L.; Cowley, L.; Askwith, C.; Libina, N.; Gitschier, J.; Anderson, G.J. Hephaestin, a ceruloplasmin homologue implicated in intestinal iron transport, is defective in the sla mouse. Nat. Genet. 1999, 21, 195–199. [Google Scholar] [CrossRef]
  37. Aisen, P. The transferrin receptor and the release of iron from transferrin. Adv. Exp. Med. Biol. 1994, 356, 31–40. [Google Scholar] [CrossRef]
  38. Fox, P.L. The copper-iron chronicles: The story of an intimate relationship. Biometals 2003, 16, 9–40. [Google Scholar] [CrossRef]
  39. Ganz, T.; Nemeth, E. Hepcidin and iron homeostasis. Biochim. Biophys. Acta 2012, 1823, 1434–1443. [Google Scholar]
  40. 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]
  41. Galy, B.; Ferring-Appel, D.; Kaden, S.; Grone, H.J.; Hentze, M.W. Iron regulatory proteins are essential for intestinal function and control key iron absorption molecules in the duodenum. Cell Metab. 2008, 7, 79–85. [Google Scholar] [CrossRef]
  42. Wang, J.; Pantopoulos, K. Regulation of cellular iron metabolism. Biochem. J. 2011, 434, 365–381. [Google Scholar] [CrossRef]
  43. Meyron-Holtz, E.G.; Ghosh, M.C.; Rouault, T.A. Mammalian tissue oxygen levels modulate iron-regulatory protein activities in vivo. Science 2004, 306, 2087–2090. [Google Scholar] [CrossRef]
  44. Mastrogiannaki, M.; Matak, P.; Keith, B.; Simon, M.C.; Vaulont, S.; Peyssonnaux, C. HIF-2alpha, but not HIF-1alpha, promotes iron absorption in mice. J. Clin. Investig. 2009, 119, 1159–1166. [Google Scholar] [CrossRef]
  45. Anderson, S.A.; Nizzi, C.P.; Chang, Y.I.; Deck, K.M.; Schmidt, P.J.; Galy, B.; Damnernsawad, A.; Broman, A.T.; Kendziorski, C.; Hentze, M.W.; et al. The IRP1-HIF-2alpha axis coordinates iron and oxygen sensing with erythropoiesis and iron absorption. Cell Metab. 2013, 17, 282–290. [Google Scholar] [CrossRef]
  46. Ghosh, M.C.; Zhang, D.L.; Jeong, S.Y.; Kovtunovych, G.; Ollivierre-Wilson, H.; Noguchi, A.; Tu, T.; Senecal, T.; Robinson, G.; Crooks, D.R.; et al. Deletion of iron regulatory protein 1 causes polycythemia and pulmonary hypertension in mice through translational derepression of HIF2alpha. Cell Metab. 2013, 17, 271–281. [Google Scholar] [CrossRef]
  47. Shah, Y.M.; Matsubara, T.; Ito, S.; Yim, S.H.; Gonzalez, F.J. Intestinal hypoxia-inducible transcription factors are essential for iron absorption following iron deficiency. Cell Metab. 2009, 9, 152–164. [Google Scholar] [CrossRef]
  48. Haase, V.H. Hypoxic regulation of erythropoiesis and iron metabolism. Am. J. Physiol. Renal Physiol. 2010, 299, F1–F13. [Google Scholar] [CrossRef]
  49. Kong, W.N.; Chang, Y.Z.; Wang, S.M.; Zhai, X.L.; Shang, J.X.; Li, L.X.; Duan, X.L. Effect of erythropoietin on hepcidin, DMT1 with IRE, and hephaestin gene expression in duodenum of rats. J. Gastroenterol. 2008, 43, 136–143. [Google Scholar] [CrossRef]
  50. Srai, S.K.; Chung, B.; Marks, J.; Pourvali, K.; Solanky, N.; Rapisarda, C.; Chaston, T.B.; Hanif, R.; Unwin, R.J.; Debnam, E.S.; Sharp, P.A. Erythropoietin regulates intestinal iron absorption in a rat model of chronic renal failure. Kidney Int. 2010, 78, 660–667. [Google Scholar] [CrossRef]
  51. Nichol, H.; Law, J.H. Iron economy in insects: Transport, metabolism, and storage. Annu. Rev. Entomol. 1992, 37, 195–215. [Google Scholar] [CrossRef]
  52. Nichol, H.; Law, J.H.; Winzerling, J.J. Iron metabolism in insects. Annu. Rev. Entomol. 2002, 47, 535–559. [Google Scholar] [CrossRef]
  53. Dunkov, B.; Georgieva, T. Insect iron binding proteins: Insights from the genomes. Insect Biochem. Mol. Biol. 2006, 36, 300–309. [Google Scholar] [CrossRef]
  54. Pham, D.Q.; Winzerling, J.J. Insect ferritins: Typical or atypical? Biochim. Biophys. Acta 2010, 1800, 824–833. [Google Scholar]
  55. Geiser, D.L.; Winzerling, J.J. Insect transferrins: Multifunctional proteins. Biochim. Biophys. Acta 2012, 1820, 437–451. [Google Scholar] [CrossRef]
  56. Nichol, H.; Locke, M. The localization of ferritin in insects. Tissue Cell 1990, 22, 767–777. [Google Scholar] [CrossRef]
  57. Shanbhag, S.; Tripathi, S. Epithelial ultrastructure and cellular mechanisms of acid and base transport in the Drosophila midgut. J. Exp. Biol. 2009, 212, 1731–1744. [Google Scholar] [CrossRef]
  58. Gutierrez, L.; Sabaratnam, N.; Aktar, R.; Bettedi, L.; Mandilaras, K.; Missirlis, F. Zinc accumulation in heterozygous mutants of fumble, the pantothenate kinase homologue of Drosophila. FEBS Lett. 2010, 584, 2942–2946. [Google Scholar] [CrossRef]
  59. Atanesyan, L.; Gunther, V.; Celniker, S.E.; Georgiev, O.; Schaffner, W. Characterization of MtnE, the fifth metallothionein member in Drosophila. J. Biol. Inorg. Chem. 2011, 16, 1047–1056. [Google Scholar] [CrossRef] [Green Version]
  60. Massie, H.R.; Aiello, V.R.; Williams, T.R. Inhibition of iron absorption prolongs the life span of Drosophila. Mech. Ageing Dev. 1993, 67, 227–237. [Google Scholar] [CrossRef]
  61. Rothenberger, S.; Mullner, E.W.; Kuhn, L.C. The mRNA-binding protein which controls ferritin and transferrin receptor expression is conserved during evolution. Nucleic Acids Res. 1990, 18, 1175–1179. [Google Scholar] [CrossRef]
  62. Charlesworth, A.; Georgieva, T.; Gospodov, I.; Law, J.H.; Dunkov, B.C.; Ralcheva, N.; Barillas-Mury, C.; Ralchev, K.; Kafatos, F.C. Isolation and properties of Drosophila melanogaster ferritin—Molecular cloning of a cDNA that encodes one subunit, and localization of the gene on the third chromosome. Eur. J. Biochem. 1997, 247, 470–475. [Google Scholar]
  63. Georgieva, T.; Dunkov, B.C.; Dimov, S.; Ralchev, K.; Law, J.H. Drosophila melanogaster ferritin: cDNA encoding a light chain homologue, temporal and tissue specific expression of both subunit types. Insect Biochem. Mol. Biol. 2002, 32, 295–302. [Google Scholar] [CrossRef]
  64. Yoshiga, T.; Georgieva, T.; Dunkov, B.C.; Harizanova, N.; Ralchev, K.; Law, J.H. Drosophila melanogaster transferrin. Cloning, deduced protein sequence, expression during the life cycle, gene localization and up-regulation on bacterial infection. Eur. J. Biochem. 1999, 260, 414–420. [Google Scholar]
  65. Muckenthaler, M.; Gunkel, N.; Frishman, D.; Cyrklaff, A.; Tomancak, P.; Hentze, M.W. Iron-regulatory protein-1 (IRP-1) is highly conserved in two invertebrate species—Characterization of IRP-1 homologues in Drosophila melanogaster and Caenorhabditis elegans. Eur. J. Biochem. 1998, 254, 230–237. [Google Scholar]
  66. Rodrigues, V.; Cheah, P.Y.; Ray, K.; Chia, W. Malvolio, the Drosophila homologue of mouse NRAMP-1 (Bcg), is expressed in macrophages and in the nervous system and is required for normal taste behaviour. EMBO J. 1995, 14, 3007–3020. [Google Scholar]
  67. Tiklova, K.; Senti, K.A.; Wang, S.; Graslund, A.; Samakovlis, C. Epithelial septate junction assembly relies on melanotransferrin iron binding and endocytosis in Drosophila. Nat. Cell Biol. 2010, 12, 1071–1077. [Google Scholar] [CrossRef]
  68. Adams, M.D.; Celniker, S.E.; Holt, R.A.; Evans, C.A.; Gocayne, J.D.; Amanatides, P.G.; Scherer, S.E.; Li, P.W.; Hoskins, R.A.; Galle, R.F.; et al. The genome sequence of Drosophila melanogaster. Science 2000, 287, 2185–2195. [Google Scholar] [CrossRef]
  69. Missirlis, F.; Holmberg, S.; Georgieva, T.; Dunkov, B.C.; Rouault, T.A.; Law, J.H. Characterization of mitochondrial ferritin in Drosophila. Proc. Natl. Acad. Sci. USA 2006, 103, 5893–5898. [Google Scholar]
  70. Bettedi, L.; Aslam, M.F.; Szular, J.; Mandilaras, K.; Missirlis, F. Iron depletion in the intestines of Malvolio mutant flies does not occur in the absence of a multicopper oxidase. J. Exp. Biol. 2011, 214, 971–978. [Google Scholar] [CrossRef]
  71. Lang, M.; Braun, C.L.; Kanost, M.R.; Gorman, M.J. Multicopper oxidase-1 is a ferroxidase essential for iron homeostasis in Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 2012, 109, 13337–13342. [Google Scholar] [CrossRef]
  72. Orgad, S.; Nelson, H.; Segal, D.; Nelson, N. Metal ions suppress the abnormal taste behavior of the Drosophila mutant malvolio. J. Exp. Biol. 1998, 201, 115–120. [Google Scholar]
  73. Folwell, J.L.; Barton, C.H.; Shepherd, D. Immunolocalisation of the D. melanogaster Nramp homologue Malvolio to gut and Malpighian tubules provides evidence that Malvolio and Nramp2 are orthologous. J. Exp. Biol. 2006, 209, 1988–1995. [Google Scholar] [CrossRef]
  74. Tang, X.; Zhou, B. Ferritin is the key to dietary iron absorption and tissue iron detoxification in Drosophila melanogaster. FASEB J. 2013, 27, 288–298. [Google Scholar] [CrossRef]
  75. Dunkov, B.C.; Georgieva, T. Organization of the ferritin genes in Drosophila melanogaster. DNA Cell Biol. 1999, 18, 937–944. [Google Scholar] [CrossRef]
  76. Hamburger, A.E.; West, A.P., Jr.; Hamburger, Z.A.; Hamburger, P.; Bjorkman, P.J. Crystal structure of a secreted insect ferritin reveals a symmetrical arrangement of heavy and light chains. J. Mol. Biol. 2005, 349, 558–569. [Google Scholar] [CrossRef]
  77. Puri, M.; Goyal, A.; Senutovich, N.; Dowd, S.R.; Minden, J.S. Building proteomic pathways using Drosophila ventral furrow formation as a model. Mol. Biosyst. 2008, 4, 1126–1135. [Google Scholar] [CrossRef]
  78. Levy, F.; Bulet, P.; Ehret-Sabatier, L. Proteomic analysis of the systemic immune response of Drosophila. Mol. Cell. Proteomics 2004, 3, 156–166. [Google Scholar]
  79. Kohler, S.A.; Henderson, B.R.; Kuhn, L.C. Succinate dehydrogenase b mRNA of Drosophila melanogaster has a functional iron-responsive element in its 5′-untranslated region. J. Biol. Chem. 1995, 270, 30781–30786. [Google Scholar]
  80. Gray, N.K.; Pantopoulos, K.; Dandekar, T.; Ackrell, B.A.; Hentze, M.W. Translational regulation of mammalian and Drosophila citric acid cycle enzymes via iron-responsive elements. Proc. Natl. Acad. Sci. USA 1996, 93, 4925–4930. [Google Scholar]
  81. Melefors, O. Translational regulation in vivo of the Drosophila melanogaster mRNA encoding succinate dehydrogenase iron protein via iron responsive elements. Biochem. Biophys. Res. Commun. 1996, 221, 437–441. [Google Scholar] [CrossRef]
  82. Lind, M.I.; Ekengren, S.; Melefors, O.; Soderhall, K. Drosophila ferritin mRNA: Alternative RNA splicing regulates the presence of the iron-responsive element. FEBS Lett. 1998, 436, 476–482. [Google Scholar] [CrossRef]
  83. Georgieva, T.; Dunkov, B.C.; Harizanova, N.; Ralchev, K.; Law, J.H. Iron availability dramatically alters the distribution of ferritin subunit messages in Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 1999, 96, 2716–2721. [Google Scholar] [CrossRef]
  84. Gu, H.F.; Lind, M.I.; Wieslander, L.; Landegren, U.; Soderhall, K.; Melefors, O. Using PRINS for gene mapping in polytene chromosomes. Chromosome Res. 1997, 5, 463–465. [Google Scholar] [CrossRef]
  85. Piccinelli, P.; Samuelsson, T. Evolution of the iron-responsive element. RNA 2007, 13, 952–966. [Google Scholar] [CrossRef]
  86. Lind, M.I.; Missirlis, F.; Melefors, O.; Uhrigshardt, H.; Kirby, K.; Phillips, J.P.; Soderhall, K.; Rouault, T.A. Of two cytosolic aconitases expressed in Drosophila, only one functions as an iron-regulatory protein. J. Biol. Chem. 2006, 281, 18707–18714. [Google Scholar] [CrossRef]
  87. Southon, A.; Farlow, A.; Norgate, M.; Burke, R.; Camakaris, J. Malvolio is a copper transporter in Drosophila melanogaster. J. Exp. Biol. 2008, 211, 709–716. [Google Scholar] [CrossRef]
  88. Dittmer, N.T.; Kanost, M.R. Insect multicopper oxidases: Diversity, properties, and physiological roles. Insect Biochem. Mol. Biol. 2010, 40, 179–188. [Google Scholar] [CrossRef]
  89. Suryo Rahmanto, Y.; Bal, S.; Loh, K.H.; Richardson, D.R. Melanotransferrin: Search for a function. Biochim. Biophys. Acta 2012, 1820, 237–243. [Google Scholar]
  90. Uhrigshardt, H.; Rouault, T.A.; Missirlis, F. Insertion mutants in Drosophila melanogaster Hsc20 halt larval growth and lead to reduced iron-sulfur cluster enzyme activities and impaired iron homeostasis. J. Biol. Inorg. Chem. 2013, 18, 441–449. [Google Scholar] [CrossRef]
  91. Kamyshev, N.G.; Iliadi, K.G.; Bragina, J.V.; Kamysheva, E.A.; Tokmatcheva, E.V.; Preat, T.; Savvateeva-Popova, E.V. Novel memory mutants in Drosophila: Behavioral characteristics of the mutant nemyP153. BMC Neurosci. 2002, 3, 9. [Google Scholar] [CrossRef]
  92. Iliadi, K.G.; Avivi, A.; Iliadi, N.N.; Knight, D.; Korol, A.B.; Nevo, E.; Taylor, P.; Moran, M.F.; Kamyshev, N.G.; Boulianne, G.L. Nemy encodes a cytochrome b561 that is required for Drosophila learning and memory. Proc. Natl. Acad. Sci. USA 2008, 105, 19986–19991. [Google Scholar] [CrossRef]
  93. Graveley, B.R.; Brooks, A.N.; Carlson, J.W.; Duff, M.O.; Landolin, J.M.; Yang, L.; Artieri, C.G.; van Baren, M.J.; Boley, N.; Booth, B.W.; et al. The developmental transcriptome of Drosophila melanogaster. Nature 2011, 471, 473–479. [Google Scholar] [CrossRef]
  94. Mandilaras, K.; Missirlis, F. Genes for iron metabolism influence circadian rhythms in Drosophila melanogaster. Metallomics 2012, 4, 928–936. [Google Scholar] [CrossRef]
  95. Zhang, X.; Sato, M.; Sasahara, M.; Migita, C.T.; Yoshida, T. Unique features of recombinant heme oxygenase of Drosophila melanogaster compared with those of other heme oxygenases studied. Eur. J. Biochem. 2004, 271, 1713–1724. [Google Scholar] [CrossRef]
  96. Cui, L.; Yoshioka, Y.; Suyari, O.; Kohno, Y.; Zhang, X.; Adachi, Y.; Ikehara, S.; Yoshida, T.; Yamaguchi, M.; Taketani, S. Relevant expression of Drosophila heme oxygenase is necessary for the normal development of insect tissues. Biochem. Biophys. Res. Commun. 2008, 377, 1156–1161. [Google Scholar] [CrossRef]
  97. Lavista-Llanos, S.; Centanin, L.; Irisarri, M.; Russo, D.M.; Gleadle, J.M.; Bocca, S.N.; Muzzopappa, M.; Ratcliffe, P.J.; Wappner, P. Control of the hypoxic response in Drosophila melanogaster by the basic helix-loop-helix PAS protein similar. Mol. Cell. Biol. 2002, 22, 6842–6853. [Google Scholar] [CrossRef]
  98. Romero, N.M.; Dekanty, A.; Wappner, P. Cellular and developmental adaptations to hypoxia: A Drosophila perspective. Methods Enzymol. 2007, 435, 123–144. [Google Scholar] [CrossRef]
  99. Harrison, J.F.; Haddad, G.G. Effects of oxygen on growth and size: Synthesis of molecular, organismal, and evolutionary studies with Drosophila melanogaster. Annu. Rev. Physiol. 2011, 73, 95–113. [Google Scholar] [CrossRef]
  100. Morton, D.B. Behavioral responses to hypoxia and hyperoxia in Drosophila larvae: Molecular and neuronal sensors. Fly 2011, 5, 119–125. [Google Scholar] [CrossRef]
  101. Gorr, T.A.; Tomita, T.; Wappner, P.; Bunn, H.F. Regulation of Drosophila hypoxia-inducible factor (HIF) activity in SL2 cells: Identification of a hypoxia-induced variant isoform of the HIFalpha homolog gene similar. J. Biol. Chem. 2004, 279, 36048–36058. [Google Scholar]
  102. Centanin, L.; Dekanty, A.; Romero, N.; Irisarri, M.; Gorr, T.A.; Wappner, P. Cell autonomy of HIF effects in Drosophila: Tracheal cells sense hypoxia and induce terminal branch sprouting. Dev. Cell 2008, 14, 547–558. [Google Scholar] [CrossRef] [Green Version]
  103. Choi, J.; Masaratana, P.; Latunde-Dada, G.O.; Arno, M.; Simpson, R.J.; McKie, A.T. Duodenal reductase activity and spleen iron stores are reduced and erythropoiesis is abnormal in Dcytb knockout mice exposed to hypoxic conditions. J. Nutr. 2012, 142, 1929–1934. [Google Scholar] [CrossRef]
  104. Zhang, D.L.; Su, D.; Berczi, A.; Vargas, A.; Asard, H. An ascorbate-reducible cytochrome b561 is localized in macrophage lysosomes. Biochim. Biophys. Acta 2006, 1760, 1903–1913. [Google Scholar]
  105. Sellami, A.; Wegener, C.; Veenstra, J.A. Functional significance of the copper transporter ATP7 in peptidergic neurons and endocrine cells in Drosophila melanogaster. FEBS Lett. 2012, 586, 3633–3638. [Google Scholar] [CrossRef]
  106. Wyman, S.; Simpson, R.J.; McKie, A.T.; Sharp, P.A. Dcytb (Cybrd1) functions as both a ferric and a cupric reductase in vitro. FEBS Lett. 2008, 582, 1901–1906. [Google Scholar] [CrossRef]
  107. Kidane, T.Z.; Farhad, R.; Lee, K.J.; Santos, A.; Russo, E.; Linder, M.C. Uptake of copper from plasma proteins in cells where expression of CTR1 has been modulated. Biometals 2012, 25, 697–709. [Google Scholar] [CrossRef]
  108. Li, Y.; Padmanabha, D.; Gentile, L.B.; Dumur, C.I.; Beckstead, R.B.; Baker, K.D. HIF- and non-HIF-regulated hypoxic responses require the estrogen-related receptor in Drosophila melanogaster. PLoS Genet. 2013, 9, e1003230. [Google Scholar] [CrossRef]
  109. Laftah, A.H.; Latunde-Dada, G.O.; Fakih, S.; Hider, R.C.; Simpson, R.J.; McKie, A.T. Haem and folate transport by proton-coupled folate transporter/haem carrier protein 1 (SLC46A1). Br. J. Nutr. 2009, 101, 1150–1156. [Google Scholar] [CrossRef]
  110. Dang, T.N.; Bishop, G.M.; Dringen, R.; Robinson, S.R. The putative heme transporter HCP1 is expressed in cultured astrocytes and contributes to the uptake of hemin. Glia 2010, 58, 55–65. [Google Scholar] [CrossRef]
  111. Le Blanc, S.; Garrick, M.D.; Arredondo, M. Heme carrier protein 1 transports heme and is involved in heme-Fe metabolism. Am. J. Physiol. Cell Physiol. 2012, 302, C1780–C1785. [Google Scholar] [CrossRef]
  112. Quigley, J.G.; Yang, Z.; Worthington, M.T.; Phillips, J.D.; Sabo, K.M.; Sabath, D.E.; Berg, C.L.; Sassa, S.; Wood, B.L.; Abkowitz, J.L. Identification of a human heme exporter that is essential for erythropoiesis. Cell 2004, 118, 757–766. [Google Scholar] [CrossRef]
  113. Khan, A.A.; Quigley, J.G. Control of intracellular heme levels: Heme transporters and heme oxygenases. Biochim. Biophys. Acta 2011, 1813, 668–682. [Google Scholar]
  114. Romney, S.J.; Newman, B.S.; Thacker, C.; Leibold, E.A. HIF-1 regulates iron homeostasis in Caenorhabditis elegans by activation and inhibition of genes involved in iron uptake and storage. PLoS Genet. 2011, 7, e1002394. [Google Scholar] [CrossRef]
  115. Ackerman, D.; Gems, D. Insulin/IGF-1 and hypoxia signaling act in concert to regulate iron homeostasis in Caenorhabditis elegans. PLoS Genet. 2012, 8, e1002498. [Google Scholar] [CrossRef]
  116. Beitel, G.J.; Krasnow, M.A. Genetic control of epithelial tube size in the Drosophila tracheal system. Development 2000, 127, 3271–3282. [Google Scholar]
  117. Hentze, M.W.; Muckenthaler, M.U.; Andrews, N.C. Balancing acts: Molecular control of mammalian iron metabolism. Cell 2004, 117, 285–297. [Google Scholar] [CrossRef]
  118. Hankeln, T.; Jaenicke, V.; Kiger, L.; Dewilde, S.; Ungerechts, G.; Schmidt, M.; Urban, J.; Marden, M.C.; Moens, L.; Burmester, T. Characterization of Drosophila hemoglobin. Evidence for hemoglobin-mediated respiration in insects. J. Biol. Chem. 2002, 277, 29012–29017. [Google Scholar] [CrossRef]
  119. Burmester, T.; Storf, J.; Hasenjager, A.; Klawitter, S.; Hankeln, T. The hemoglobin genes of Drosophila. FEBS J. 2006, 273, 468–480. [Google Scholar] [CrossRef]
  120. Gleixner, E.; Abriss, D.; Adryan, B.; Kraemer, M.; Gerlach, F.; Schuh, R.; Burmester, T.; Hankeln, T. Oxygen-induced changes in hemoglobin expression in Drosophila. FEBS J. 2008, 275, 5108–5116. [Google Scholar] [CrossRef]
  121. Karlsson, C.; Korayem, A.M.; Scherfer, C.; Loseva, O.; Dushay, M.S.; Theopold, U. Proteomic analysis of the Drosophila larval hemolymph clot. J. Biol. Chem. 2004, 279, 52033–52041. [Google Scholar]
  122. Hajdusek, O.; Sojka, D.; Kopacek, P.; Buresova, V.; Franta, Z.; Sauman, I.; Winzerling, J.; Grubhoffer, L. Knockdown of proteins involved in iron metabolism limits tick reproduction and development. Proc. Natl. Acad. Sci. USA 2009, 106, 1033–1038. [Google Scholar] [CrossRef]
  123. Ward, D.M.; Kaplan, J. Ferroportin-mediated iron transport: Expression and regulation. Biochim. Biophys. Acta 2012, 1823, 1426–1433. [Google Scholar]
  124. Bulet, P.; Hetru, C.; Dimarcq, J.L.; Hoffmann, D. Antimicrobial peptides in insects; structure and function. Dev. Comp. Immunol. 1999, 23, 329–344. [Google Scholar] [CrossRef]
  125. Verga Falzacappa, M.V.; Muckenthaler, M.U. Hepcidin: Iron-hormone and anti-microbial peptide. Gene 2005, 364, 37–44. [Google Scholar] [CrossRef]
  126. Charroux, B.; Royet, J. Elimination of plasmatocytes by targeted apoptosis reveals their role in multiple aspects of the Drosophila immune response. Proc. Natl. Acad. Sci. USA 2009, 106, 9797–9802. [Google Scholar] [CrossRef]
  127. Denholm, B.; Skaer, H. Bringing together components of the fly renal system. Curr. Opin. Genet. Dev. 2009, 19, 526–532. [Google Scholar] [CrossRef]
  128. Dow, J.A. Insights into the Malpighian tubule from functional genomics. J. Exp. Biol. 2009, 212, 435–445. [Google Scholar] [CrossRef]
  129. Davies, S.A.; Overend, G.; Sebastian, S.; Cundall, M.; Cabrero, P.; Dow, J.A.; Terhzaz, S. Immune and stress response ‘cross-talk’ in the Drosophila Malpighian tubule. J. Insect Physiol. 2012, 58, 488–497. [Google Scholar] [CrossRef]
  130. Todorich, B.; Zhang, X.; Slagle-Webb, B.; Seaman, W.E.; Connor, J.R. Tim-2 is the receptor for H-ferritin on oligodendrocytes. J. Neurochem. 2008, 107, 1495–1505. [Google Scholar] [CrossRef]
  131. Li, J.Y.; Paragas, N.; Ned, R.M.; Qiu, A.; Viltard, M.; Leete, T.; Drexler, I.R.; Chen, X.; Sanna-Cherchi, S.; Mohammed, F.; et al. Scara5 is a ferritin receptor mediating non-transferrin iron delivery. Dev. Cell 2009, 16, 35–46. [Google Scholar] [CrossRef]
  132. Todorich, B.; Zhang, X.; Connor, J.R. H-ferritin is the major source of iron for oligodendrocytes. Glia 2011, 59, 927–935. [Google Scholar] [CrossRef]
  133. Meyron-Holtz, E.G.; Moshe-Belizowski, S.; Cohen, L.A. A possible role for secreted ferritin in tissue iron distribution. J. Neural Transm. 2011, 118, 337–347. [Google Scholar] [CrossRef]
  134. Li, L.; Fang, C.J.; Ryan, J.C.; Niemi, E.C.; Lebron, J.A.; Bjorkman, P.J.; Arase, H.; Torti, F.M.; Torti, S.V.; Nakamura, M.C.; Seaman, W.E. Binding and uptake of H-ferritin are mediated by human transferrin receptor-1. Proc. Natl. Acad. Sci. USA 2010, 107, 3505–3510. [Google Scholar] [CrossRef]
  135. San Martin, C.D.; Garri, C.; Pizarro, F.; Walter, T.; Theil, E.C.; Nuñez, M.T. Caco-2 intestinal epithelial cells absorb soybean ferritin by mu2 (AP2)-dependent endocytosis. J. Nutr. 2008, 138, 659–666. [Google Scholar]
  136. Bekenstein, U.; Kadener, S. What can Drosophila teach us about iron-accumulation neurodegenerative disorders? J. Neural Transm. 2011, 118, 389–396. [Google Scholar] [CrossRef]
  137. Chamilos, G.; Lewis, R.E.; Hu, J.; Xiao, L.; Zal, T.; Gilliet, M.; Halder, G.; Kontoyiannis, D.P. Drosophila melanogaster as a model host to dissect the immunopathogenesis of zygomycosis. Proc. Natl. Acad. Sci. USA 2008, 105, 9367–9372. [Google Scholar]
  138. Kremer, N.; Voronin, D.; Charif, D.; Mavingui, P.; Mollereau, B.; Vavre, F. Wolbachia interferes with ferritin expression and iron metabolism in insects. PLoS Pathog. 2009, 5, e1000630. [Google Scholar] [CrossRef]
  139. Brownlie, J.C.; Cass, B.N.; Riegler, M.; Witsenburg, J.J.; Iturbe-Ormaetxe, I.; McGraw, E.A.; O’Neill, S.L. Evidence for metabolic provisioning by a common invertebrate endosymbiont, Wolbachia pipientis, during periods of nutritional stress. PLoS Pathog. 2009, 5, e1000368. [Google Scholar] [CrossRef]
  140. Rose, P.P.; Hanna, S.L.; Spiridigliozzi, A.; Wannissorn, N.; Beiting, D.P.; Ross, S.R.; Hardy, R.W.; Bambina, S.A.; Heise, M.T.; Cherry, S. Natural resistance-associated macrophage protein is a cellular receptor for sindbis virus in both insect and mammalian hosts. Cell Host Microbe 2011, 10, 97–104. [Google Scholar] [CrossRef]
  141. Pedersen, K.S.; Codrea, M.C.; Vermeulen, C.J.; Loeschcke, V.; Bendixen, E. Proteomic characterization of a temperature-sensitive conditional lethal in Drosophila melanogaster. Heredity 2010, 104, 125–134. [Google Scholar] [CrossRef]
  142. Robinson, R.A.; Kellie, J.F.; Kaufman, T.C.; Clemmer, D.E. Insights into aging through measurements of the Drosophila proteome as a function of temperature. Mech. Ageing Dev. 2010, 131, 584–590. [Google Scholar] [CrossRef]
  143. Colinet, H.; Overgaard, J.; Com, E.; Sorensen, J.G. Proteomic profiling of thermal acclimation in Drosophila melanogaster. Insect Biochem. Mol. Biol. 2013, 43, 352–365. [Google Scholar] [CrossRef]
  144. Canizares, J.; Blanca, J.M.; Navarro, J.A.; Monros, E.; Palau, F.; Molto, M.D. dfh is a Drosophila homolog of the Friedreich’s ataxia disease gene. Gene 2000, 256, 35–42. [Google Scholar] [CrossRef]
  145. Anderson, P.R.; Kirby, K.; Hilliker, A.J.; Phillips, J.P. RNAi-mediated suppression of the mitochondrial iron chaperone, frataxin, in Drosophila. Hum. Mol. Genet. 2005, 14, 3397–3405. [Google Scholar] [CrossRef]
  146. Llorens, J.V.; Navarro, J.A.; Martinez-Sebastian, M.J.; Baylies, M.K.; Schneuwly, S.; Botella, J.A.; Molto, M.D. Causative role of oxidative stress in a Drosophila model of Friedreich ataxia. FASEB J. 2007, 21, 333–344. [Google Scholar] [CrossRef]
  147. Anderson, P.R.; Kirby, K.; Orr, W.C.; Hilliker, A.J.; Phillips, J.P. Hydrogen peroxide scavenging rescues frataxin deficiency in a Drosophila model of Friedreich’s ataxia. Proc. Natl. Acad. Sci. USA 2008, 105, 611–616. [Google Scholar] [CrossRef]
  148. Runko, A.P.; Griswold, A.J.; Min, K.T. Overexpression of frataxin in the mitochondria increases resistance to oxidative stress and extends lifespan in Drosophila. FEBS Lett. 2008, 582, 715–719. [Google Scholar] [CrossRef]
  149. Shidara, Y.; Hollenbeck, P.J. Defects in mitochondrial axonal transport and membrane potential without increased reactive oxygen species production in a Drosophila model of Friedreich ataxia. J. Neurosci. 2010, 30, 11369–11378. [Google Scholar] [CrossRef]
  150. Navarro, J.A.; Llorens, J.V.; Soriano, S.; Botella, J.A.; Schneuwly, S.; Martinez-Sebastian, M.J.; Molto, M.D. Overexpression of human and fly frataxins in Drosophila provokes deleterious effects at biochemical, physiological and developmental levels. PLoS One 2011, 6, e21017. [Google Scholar]
  151. Soriano, S.; Llorens, J.V.; Blanco-Sobero, L.; Gutiérrez, L.; Calap-Quintana, P.; Morales, M.P.; Moltó, M.D.; Martínez-Sebastián, M.J. Deferiprone and idebenone rescue frataxin depletion phenotypes in a Drosophila model of Friedreich’s ataxia. Gene 2013, 521, 274–281. [Google Scholar]
  152. Bonilla-Ramirez, L.; Jimenez-Del-Rio, M.; Velez-Pardo, C. Low doses of paraquat and polyphenols prolong life span and locomotor activity in knock-down parkin Drosophila melanogaster exposed to oxidative stress stimuli: Implication in autosomal recessive juvenile Parkinsonism. Gene 2013, 521, 355–363. [Google Scholar]
  153. Esposito, G.; Vos, M.; Vilain, S.; Swerts, J.; de Sousa Valadas, J.; van Meensel; Schaap, O.; Verstreken, P. Aconitase causes iron toxicity in Drosophila pink1 mutants. PLoS Genet. 2013, 9, e1003478. [Google Scholar]
  154. Rival, T.; Page, R.M.; Chandraratna, D.S.; Sendall, T.J.; Ryder, E.; Liu, B.; Lewis, H.; Rosahl, T.; Hider, R.; Camargo, L.M.; et al. Fenton chemistry and oxidative stress mediate the toxicity of the beta-amyloid peptide in a Drosophila model of Alzheimer’s disease. Eur. J. Neurosci. 2009, 29, 1335–1347. [Google Scholar] [CrossRef]
  155. Liu, B.; Moloney, A.; Meehan, S.; Morris, K.; Thomas, S.E.; Serpell, L.C.; Hider, R.; Marciniak, S.J.; Lomas, D.A.; Crowther, D.C. Iron promotes the toxicity of amyloid beta peptide by impeding its ordered aggregation. J. Biol. Chem. 2011, 286, 4248–4256. [Google Scholar] [CrossRef]
  156. Freeman, A.; Pranski, E.; Miller, R.D.; Radmard, S.; Bernhard, D.; Jinnah, H.A.; Betarbet, R.; Rye, D.B.; Sanyal, S. Sleep fragmentation and motor restlessness in a Drosophila model of Restless Legs Syndrome. Curr. Biol. 2012, 22, 1142–1148. [Google Scholar] [CrossRef]
  157. Freeman, A.A.; Mandilaras, K.; Missirlis, F.; Sanyal, S. An emerging role for Cullin-3 mediated ubiquitination in sleep and circadian rhythm: Insights from Drosophila. Fly 2013, 7, 39–43. [Google Scholar] [CrossRef]
  158. Kosmidis, S.; Botella, J.A.; Mandilaras, K.; Schneuwly, S.; Skoulakis, E.M.; Rouault, T.A.; Missirlis, F. Ferritin overexpression in Drosophila glia leads to iron deposition in the optic lobes and late-onset behavioral defects. Neurobiol. Dis. 2011, 43, 213–219. [Google Scholar] [CrossRef]
  159. Bonilla-Ramirez, L.; Jimenez-Del-Rio, M.; Velez-Pardo, C. Acute and chronic metal exposure impairs locomotion activity in Drosophila melanogaster: A model to study Parkinsonism. Biometals 2011, 24, 1045–1057. [Google Scholar] [CrossRef]
  160. Wu, Z.; Du, Y.; Xue, H.; Wu, Y.; Zhou, B. Aluminum induces neurodegeneration and its toxicity arises from increased iron accumulation and reactive oxygen species (ROS) production. Neurobiol. Aging 2012, 33, 199.e1–199.e12. [Google Scholar] [CrossRef]
  161. Lozinsky, O.V.; Lushchak, O.V.; Storey, J.M.; Storey, K.B.; Lushchak, V.I. Sodium nitroprusside toxicity in Drosophila melanogaster: Delayed pupation, reduced adult emergence, and induced oxidative/nitrosative stress in eclosed flies. Arch. Insect Biochem. Physiol. 2012, 80, 166–185. [Google Scholar] [CrossRef]
  162. Shaik, K.S.; Meyer, F.; Vazquez, A.V.; Flotenmeyer, M.; Cerdan, M.E.; Moussian, B. Delta-aminolevulinate synthase is required for apical transcellular barrier formation in the skin of the Drosophila larva. Eur. J. Cell Biol. 2012, 91, 204–215. [Google Scholar] [CrossRef]
  163. Metzendorf, C.; Lind, M.I. Drosophila mitoferrin is essential for male fertility: Evidence for a role of mitochondrial iron metabolism during spermatogenesis. BMC Dev. Biol. 2010, 10, 68. [Google Scholar] [CrossRef]
  164. Metzendorf, C.; Wu, W.; Lind, M.I. Overexpression of Drosophila mitoferrin in l(2)mbn cells results in dysregulation of Fer1HCH expression. Biochem. J. 2009, 421, 463–471. [Google Scholar] [CrossRef]
  165. Hales, K.G. Iron testes: Sperm mitochondria as a context for dissecting iron metabolism. BMC Biol. 2010, 8, 79. [Google Scholar] [CrossRef]
  166. Mehta, A.; Deshpande, A.; Missirlis, F. Genetic screening for novel Drosophila mutants with discrepancies in iron metabolism. Biochem. Soc. Trans. 2008, 36, 1313–1316. [Google Scholar] [CrossRef]
  167. Metzendorf, C.; Lind, M.I. The role of iron in the proliferation of Drosophila l(2) mbn cells. Biochem. Biophys. Res. Commun. 2010, 400, 442–446. [Google Scholar] [CrossRef]
  168. Li, S. Identification of iron-loaded ferritin as an essential mitogen for cell proliferation and postembryonic development in Drosophila. Cell Res. 2010, 20, 1148–1157. [Google Scholar] [CrossRef]
  169. Lye, J.C.; Hwang, J.E.; Paterson, D.; de Jonge, M.D.; Howard, D.L.; Burke, R. Detection of genetically altered copper levels in Drosophila tissues by synchrotron X-ray fluorescence microscopy. PLoS One 2011, 6, e26867. [Google Scholar]
  170. Drakesmith, H.; Prentice, A.M. Hepcidin and the iron-infection axis. Science 2012, 338, 768–772. [Google Scholar] [CrossRef]
  171. Lemaitre, B.; Hoffmann, J. The host defense of Drosophila melanogaster. Annu. Rev. Immunol. 2007, 25, 697–743. [Google Scholar] [CrossRef]
  172. Kontoyiannis, D.P.; Lewis, R.E. Invasive zygomycosis: Update on pathogenesis, clinical manifestations, and management. Infect. Dis. Clin. North Am. 2006, 20, 581–607. [Google Scholar] [CrossRef]
  173. Konopka, R.J.; Benzer, S. Clock mutants of Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 1971, 68, 2112–2116. [Google Scholar] [CrossRef]
  174. Nitabach, M.N.; Taghert, P.H. Organization of the Drosophila circadian control circuit. Curr. Biol. 2008, 18, R84–R93. [Google Scholar] [CrossRef]
  175. Dioum, E.M.; Rutter, J.; Tuckerman, J.R.; Gonzalez, G.; Gilles-Gonzalez, M.A.; McKnight, S.L. NPAS2: A gas-responsive transcription factor. Science 2002, 298, 2385–2387. [Google Scholar] [CrossRef]
  176. Kaasik, K.; Lee, C.C. Reciprocal regulation of haem biosynthesis and the circadian clock in mammals. Nature 2004, 430, 467–471. [Google Scholar] [CrossRef]
  177. Reinking, J.; Lam, M.M.; Pardee, K.; Sampson, H.M.; Liu, S.; Yang, P.; Williams, S.; White, W.; Lajoie, G.; Edwards, A.; Krause, H.M. The Drosophila nuclear receptor e75 contains heme and is gas responsive. Cell 2005, 122, 195–207. [Google Scholar] [CrossRef]
  178. De Rosny, E.; de Groot, A.; Jullian-Binard, C.; Borel, F.; Suarez, C.; le Pape, L.; Fontecilla-Camps, J.C.; Jouve, H.M. DHR51, the Drosophila melanogaster homologue of the human photoreceptor cell-specific nuclear receptor, is a thiolate heme-binding protein. Biochemistry 2008, 47, 13252–13260. [Google Scholar] [CrossRef]
  179. Salome, P.A.; Oliva, M.; Weigel, D.; Kramer, U. Circadian clock adjustment to plant iron status depends on chloroplast and phytochrome function. EMBO J. 2013, 32, 511–523. [Google Scholar]
  180. Hong, S.; Kim, S.A.; Guerinot, M.L.; McClung, C.R. Reciprocal interaction of the circadian clock with the iron homeostasis network in Arabidopsis. Plant Physiol. 2013, 161, 893–903. [Google Scholar] [CrossRef]
  181. Chen, Y.Y.; Wang, Y.; Shin, L.J.; Wu, J.F.; Shanmugam, V.; Tsednee, M.; Lo, J.C.; Chen, C.C.; Wu, S.H.; Yeh, K.C. Iron is involved in the maintenance of circadian period length in Arabidopsis. Plant Physiol. 2013, 161, 1409–1420. [Google Scholar] [CrossRef]
  182. Wilson, G.T.; Connolly, E.L. Running a little late: Chloroplast Fe status and the circadian clock. EMBO J. 2013, 32, 490–492. [Google Scholar] [CrossRef]
  183. Ben-Shahar, Y.; Dudek, N.L.; Robinson, G.E. Phenotypic deconstruction reveals involvement of manganese transporter malvolio in honey bee division of labor. J. Exp. Biol. 2004, 207, 3281–3288. [Google Scholar] [CrossRef]
  184. Denison, R.; Raymond-Delpech, V. Insights into the molecular basis of social behaviour from studies on the honeybee, Apis mellifera. Invert. Neurosci. 2008, 8, 1–9. [Google Scholar] [CrossRef]
  185. Campuzano, V.; Montermini, L.; Molto, M.D.; Pianese, L.; Cossee, M.; Cavalcanti, F.; Monros, E.; Rodius, F.; Duclos, F.; Monticelli, A.; et al. Friedreich’s ataxia: Autosomal recessive disease caused by an intronic GAA triplet repeat expansion. Science 1996, 271, 1423–1427. [Google Scholar]
  186. Babcock, M.; de Silva, D.; Oaks, R.; Davis-Kaplan, S.; Jiralerspong, S.; Montermini, L.; Pandolfo, M.; Kaplan, J. Regulation of mitochondrial iron accumulation by Yfh1p, a putative homolog of frataxin. Science 1997, 276, 170–1712. [Google Scholar]
  187. Rötig, A.; de Lonlay, P.; Chretien, D.; Foury, F.; Koenig, M.; Sidi, D.; Munnich, A.; Rustin, P. Aconitase and mitochondrial iron-sulphur protein deficiency in Friedreich ataxia. Nat. Genet. 1997, 17, 215–217. [Google Scholar]
  188. Bahadorani, S.; Hilliker, A.J. Cocoa confers life span extension in Drosophila melanogaster. Nutr. Res. 2008, 28, 377–382. [Google Scholar] [CrossRef]
  189. Jimenez-Del-Rio, M.; Guzman-Martinez, C.; Velez-Pardo, C. The effects of polyphenols on survival and locomotor activity in Drosophila melanogaster exposed to iron and paraquat. Neurochem. Res. 2010, 35, 227–238. [Google Scholar] [CrossRef]
  190. Schriner, S.E.; Katoozi, N.S.; Pham, K.Q.; Gazarian, M.; Zarban, A.; Jafari, M. Extension of Drosophila lifespan by Rosa damascena associated with an increased sensitivity to heat. Biogerontology 2012, 13, 105–117. [Google Scholar] [CrossRef]
  191. Sadraie, M.; Missirlis, F. Evidence for evolutionary constraints in Drosophila metal biology. Biometals 2011, 24, 679–686. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Mandilaras, K.; Pathmanathan, T.; Missirlis, F. Iron Absorption in Drosophila melanogaster. Nutrients 2013, 5, 1622-1647. https://doi.org/10.3390/nu5051622

AMA Style

Mandilaras K, Pathmanathan T, Missirlis F. Iron Absorption in Drosophila melanogaster. Nutrients. 2013; 5(5):1622-1647. https://doi.org/10.3390/nu5051622

Chicago/Turabian Style

Mandilaras, Konstantinos, Tharse Pathmanathan, and Fanis Missirlis. 2013. "Iron Absorption in Drosophila melanogaster" Nutrients 5, no. 5: 1622-1647. https://doi.org/10.3390/nu5051622

Article Metrics

Back to TopTop