You are on page 1of 9

Molecular Genetics and Metabolism 128 (2019) 204–212

Contents lists available at ScienceDirect

Molecular Genetics and Metabolism


journal homepage: www.elsevier.com/locate/ymgme

Review article

Zebrafish as a model system to delineate the role of heme and iron T


metabolism during erythropoiesis

Jianbing Zhang, Iqbal Hamza
Department of Animal & Avian Sciences and Department of Cell Biology & Molecular Genetics, University of Maryland, College Park, MD 20742, USA

A R T I C LE I N FO A B S T R A C T

Keywords: Coordination of iron acquisition and heme synthesis is required for effective erythropoiesis. The small teleost
Zebrafish zebrafish (Danio rerio) is an ideal vertebrate animal model to replicate various aspects of human physiology and
Porphyria provides an efficient and cost-effective way to model human pathophysiology. Importantly, zebrafish ery-
Heme transport thropoiesis largely resembles mammalian erythropoiesis. Gene discovery by large-scale forward mutagenesis
Iron metabolism
screening has identified key components in heme and iron metabolism. Reverse genetic screens, using mor-
Erythropoiesis
pholino-knockdown and CRISPR/Cas9, coupled with the genetic tractability of the developing embryo have
further accelerated functional studies. Ultimately, the ex utero development of zebrafish embryos combined with
their transparency and developmental plasticity could provide a deeper understanding of the role of iron and
heme metabolism during early vertebrate embryonic development.

1. Introduction validated by morpholino-mediated knockdown or identified mutants


[4]. These conserved properties of zebrafish hematopoiesis make it a
The teleost fish, Danio rerio, also known as zebrafish, has gained good model for hematopoietic study and its application to mammalian
greater traction as a model organism for studying vertebrate develop- biology.
ment owing to some unique advantages in developmental biology. Primitive erythropoiesis in zebrafish occurs in the intermediate cell
Adult zebrafish are relatively small and one breeding pair can produce mass (ICM) of developing embryos, which is functionally equivalent to
100–200 progenies per spawning each week, which allows easy main- extra-embryonic yolk sac blood island region in mammals [5]. Primi-
tenance of animal strains with small space and ensures production of tive erythroid lineage arises from the ICM which is develops from the
numerous embryos for experiments (genetic screening work and posterior lateral mesoderm (PLM) region. Primitive erythropoiesis in
building transgenic lines). The external fertilization and rapid devel- zebrafish begins at the 4-somite stage with the expression of erythroid-
opment ex utero, combined with the property of transparency, allow specific transcription factor GATA1, a zinc-finger transcription factor
direct visualization and manipulation of developmental processes in necessary for early primitive erythroid progenitors differentiation [6].
early embryos, which are not possible in mice. The techniques for The circulating primitive erythroblasts further proliferate and differ-
transgenesis and gene manipulation are well-developed, rendering entiate into mature erythrocytes through a series of morphological al-
zebrafish as a genetically-tractable vertebrate animal model [1]. terations, with elongated nucleus around 4 dpf (days post fertilization),
and functionally survive to 10 dpf in peripheral blood stream [7].
2. Zebrafish erythropoiesis resembles those of higher vertebrates Mature erythrocytes display unique morphological characteristics, dis-
tinguishable from adult counterparts by less cytoplasm and remaining
Zebrafish has a similar hematopoietic program compared to higher nucleus.
vertebrates, spanning from developmental waves of hematopoiesis to Consistent with hematopoietic programs in mammals, definitive
conservation of producing comparable blood cell components. Two hematopoiesis in zebrafish marks its distinctive wave by generating all
successive waves of hematopoiesis, primitive and definitive, are also blood cell lineages throughout the lifespan. The initiation site of zeb-
found in the zebrafish [2,3]. Additionally, crucial regulators have been rafish definitive hematopoiesis has been identified by expression pat-
isolated as orthologs of many essential mammalian hematopoietic terns of mammalian definitive hematopoietic regulator orthologues, C-
regulators and functional conservation of these factors in zebrafish is MYB and RUNX1. These two transcription factors dictate the formation


Corresponding author: University of Maryland, College Park, 2413 ANSC, Bldg 142, College Park, MD 20742, USA.
E-mail address: hamza@umd.edu (I. Hamza).

https://doi.org/10.1016/j.ymgme.2018.12.007
Received 17 October 2018; Received in revised form 14 December 2018; Accepted 14 December 2018
Available online 24 December 2018
1096-7192/ © 2018 Elsevier Inc. All rights reserved.
J. Zhang, I. Hamza Molecular Genetics and Metabolism 128 (2019) 204–212

of definitive HSPCs (hematopoietic stem/progenitor cells) [8]. In zeb- protein (PCBP) [25]. PCBP1 and PCBP2 exhibit high affinity for ferritin
rafish, initial expression of runx1 and c-myb is located in the VDA in vitro and PCBP1 and its paralog PCBP2 co-immuoprecipitates with
(ventral wall of dorsal aorta) from 26 to 48 hpf (hours post fertiliza- ferritin in HEK293 cells. Co-expression of PCBP1 and PCBP2 together
tion), suggesting that definitive HSPCs are generated in the VDA. The with human ferritins in yeast causes iron deficiency response by in-
initiation of definitive hematopoiesis in the VDA happens as early as 36 creasing iron deposition into ferritin. While PCBP1 mediates delivery
hpf [9]. At around 4–5 dpf, hematopoietic cells are identifiable in and integration of iron into ferritin, nuclear receptor coactivator 4
pronephros/kidney and thereafter by 13 dpf. The fish kidney functions (NCOA4) promotes release of iron from ferritin by directing the ferritin
as a counterpart of the mammalian bone marrow (BM) to sustain de- complex to autophagosomes for degradation [26]. Binding of PCBP1 to
finitive hematopoiesis throughout its lifespan [10]. Posterior blood is- ferritin precedes NCOA4-ferritin interaction and coincides with globin
land (PBI), also called caudal hematopoietic tissue (CHT) and located synthesis during erythroid maturation. NCOA4-deficient cells exhibit
between the caudal artery and the caudal vein at the end of the tail, is massive accumulation of iron in ferritin with impaired hemoglobini-
recognized as a fetal hematopoietic organ in zebrafish and an inter- zation and enhanced erythroid cell death by ferritinophagy, since fer-
mediate hematopoietic site covering the intermediate period between ritin is an essential source of iron for heme production during terminal
the VDA and kidney as a counterpart to fetal liver in mammals [11,12]. erythroid differentiation.
Definitive erythrocytes are postulated to populate the circulation at Upon release from ferritin, iron is transported into the mitochondria
around 5 dpf as RBCs emerge at this stage in bloodless mutants with mediated by SLC25A37 (mitoferrin1, MFRN1), a protein belonging to
defects in producing primitive erythrocytes [13]. Transfusion experi- the family of mitochondrial solute carrier proteins. MFRN1 is expressed
ments reveal that primitive erythrocytes are the major circulating er- in the inner mitochondrial membrane and transports iron across mi-
ythroid components for the first 5 days and thereafter are gradually tochondrial membranes [27]. Mouse erythroblasts derived from Mfrn1-
replaced by presumptive definitive erythrocytes [7]. Like mammals, deficient embryonic stem cells show a complete block of iron in-
these definitive erythrocytes switch to express adult form of globins and corporation into heme. Defect in Mfrn1 results in profound hypo-
could be discriminated morphologically from primitive counterparts chromic anemia and erythroid maturation arrest owing to insufficient
[14,15]. mitochondrial iron uptake in zebrafish. Deletion of two yeast Mfrn
The kidney marrow (head kidney) is an adult hematopoietic organ homologs, Mrs3 and Mrs4 impairs incorporation of iron into PPIX and
in zebrafish which is functionally equivalent to BM in mammals. Based formation of Fe-S cluster assembly, collectively resulting in poor growth
on forward scatter (cell size) and side scatter (granularity), hemato- under low iron conditions. Deletion of Mfrn1 in mice is embryonic le-
poietic cells in the kidney marrow could be fractioned into four popu- thal and mice with targeted deletion of Mfrn1 in adult hematopoietic
lations: immature precursors of all lineages, lymphoid cells, mature tissues show severe anemia owing to deficits in erythroblast formation
erythroid cells, and myelo-monocytic cells (neutrophils, monocytes, [28].
macrophages, and eosinophils) [16]. Lethal irradiation followed by Large amounts of heme is synthesized during differentiation and
transplantation rescue in zebrafish reveals presence of definitive HSCs maturation of RBCs and therefore must be coupled with iron acquisi-
in the kidney marrow maintaining lifelong hematopoiesis program tion. The efficiency of converting iron to heme in maturing erythroid
[17]. cells is extremely high, resulting in heme-iron concentrations to be over
The spleen is postulated to be another hematopoietic tissue in the 40,000-fold greater than non-heme iron in mature RBCs [29,30]. One
adult zebrafish, although clear evidence is lacking to define its hema- mode of tackling this large requirement for heme would be upregu-
topoietic activity. Compared to the kidney marrow, zebrafish spleen is lating heme synthesis in the mitochondria and then mobilizing heme
not well-characterized. It was proposed that zebrafish spleen may out of the mitochondria for insertion into cytoplasmic globin to couple
function as a reservoir for RBCs where erythrocytes are stored and heme synthesis with increasing globin production. In contrast, it is also
destroyed [18]. Splenic macrophages can be found in the red pulp and essential to downregulate heme synthesis and decrease intracellular
contain phagosomes with erythrocytes and other cellular debris, in- iron content when globin production is low in the early stage of ery-
dicating the possibility of active erythrophagocytosis (EP) in the zeb- thropoietic development, since both free heme and iron are toxic to
rafish spleen [19,20]. erythroid cells by inducing oxidative stress. Failure to regulate heme
synthesis during erythropoiesis will cause either iron-deficient or iron-
3. Iron acquisition and heme biosynthesis during erythropoiesis overload anemia [31].
Although there are eight enzymatic steps for heme biosynthesis, the
The need for iron for heme synthesis during erythropoiesis man- rate-limiting step is the synthesis of ALA from glycine and succinyl-
dates an efficient pathway to uptake extracellular iron (Fig. 1). Ery- coenzyme A, catalyzed by ALAS [32] (Fig. 1). Two forms of ALAS exist,
throid cells rely on a high affinity system composed of transferrin (TF) ALAS1 and ALAS2. ALAS1 is ubiquitously expressed in all tissues, while
and transferrin receptor (TFR). One molecule of transferrin can bind to ALAS2 (or ALAS-E), is exclusively expressed in developing erythroid
two ferric iron atoms with an association coefficient of 10−20 M at cells [33,34]. ALAS2 is activated by the transcription factor GATA1, a
physiological pH [21]. Transferrin receptor 1 (TFR1, also known as master regulator for erythropoiesis. Chip-Seq analysis using G1E-ER4
CD71) tightly binds to TF, permitting developing erythroid cells to erythroid progenitor cell line derived from Gata1 mutant mice identi-
uptake iron efficiently from circulation. The complex of iron-bound TF fied two GATA-1 cis elements in the first and eighth introns of Alas2
and TFR1 is internalized by receptor-mediated endocytosis and the [35–37]. Disruption of these two GATA1-binding elements abrogates
ferric iron is then released from TF as the endosomes acidify. The re- expression of ALAS2 and subsequent inhibition of heme synthesis,
leased ferric iron is reduced to ferrous by STEAP3 (six-transmembrane which can be rescued by supplementing cells with high concentrations
epithelial antigen of prostate 3 reductase) [22]. The ferrous iron is of ALA, the product of ALAS [36,37]. Another regulation of ALAS2
transported out of the endosome by DMT1 (divalent metal transporter occurs at the post-transcriptional level, modulated by iron responsive
1, also known as NRAMP2 and SLC11A2) [23]. The apo-TF/TFR1 elements (IREs) in the 5′UTR. IREs interact with iron regulatory pro-
complex is then recycled back to the cell surface, where apo-TF dis- teins (IRPs), linking the regulation of heme biosynthesis in erythroid
sociates from the TFR1 and re-enters the circulation. The holo-TF/TFR1 cells to iron availability [38]. Under iron-deficient conditions, IRPs bind
and apo-TF/TFR1 cycle ensures optimal iron uptake from the circula- to IREs and inhibit Alas2 mRNA translation. Conversely, when in-
tion for hemoglobin production. tracellular iron levels increase, IRPs are either degraded or converted to
Since free iron is cytotoxic due to Fenton chemistry, iron is either an aconitase by an iron-sulfur [4Fe-4S] and Alas2 mRNA translation
stored or utilized upon transport into the cytosol [24]. Iron is stored in resumes to promote heme synthesis. The regulation of ALAS2 expres-
the cytosol by binding to ferritin with the aid of Poly r(C)-binding sion is coordinated with the cellular iron levels to tightly control

205
J. Zhang, I. Hamza Molecular Genetics and Metabolism 128 (2019) 204–212

Fig. 1. Heme synthesis and iron metabolism in ery-


throid cells.
Iron acquisition in erythroid cells is dependent on
endocytosis of Tf-bound Fe (Tf-Fe(III)) via the
transferrin receptor (TFR1). Iron (Fe(III)) is imported
into the cytoplasm by DMT1 after reduction by
STEAP3. Iron can be stored in ferritin (FTL) (FTL-
Iron) mediated by PCBP1 and can be released from
FTL promoted by NCOA4. MFRN1 is responsible for
bringing iron into the mitochondria for heme
synthesis. The red arrow represents overall heme
synthetic pathway. Heme biosynthesis initiates in the
mitochondrial matrix and is catalyzed by δ-amino-
levulinic acid synthase (ALAS2) to synthesize δ-
aminolevulinic acid (ALA) from glycine and suc-
cinyl-coenzyme A. ALA is subsequently transported
out of the mitochondria to the cytosol for the fol-
lowing four enzymatic reaction steps: ALA to por-
phobilinogen (PBG) catalyzed by ALA dehydratase
(ALAD); PBG to an unstable polymer hydro-
xymethylbilane by porphobilinogen deaminase
(PBGD); hydroxymethylbilane to uroporphyrinogen
III (UROgen III) by uroporphyrinogen synthase
(UROS); and UROgen III to coproporphyrinogen III
(CPgen III) by uroporphyrinogen decarboxylase
(UROD). CPgen III is then transported into the mi-
tochondria where coproporphyrinogen oxidase
(CPOX), a mitochondrial intermembrane space en-
zyme, catalyzes the formation of protoporphyrinogen IX (PPgen IX). The inner mitochondrial membrane enzyme protoporphyrinogen oxidase (PPOX) catalyzes the
formation of protoporphyrin IX (PPIX) from PPgen IX in the mitochondrial matrix. In the last step, ferrochelatase (FECH) catalyzes the insertion of ferrous iron (Fe2+)
into PPIX to form heme. Iron is also used for Fe-S cluster synthesis with involvement of GLRX5. IRP1 binding can inhibit translation of ALAS2 to prevent the
accumulation of toxic heme intermediates. Cellular iron efflux is mediated by FPN and requires iron oxidation on the extracellular side.

cellular heme content.


Another rate-limiting enzyme in the heme biosynthetic pathway is
FECH. Transcription of Fech spikes during terminal erythroid differ-
entiation and is controlled by the transcription factors SP1, NFE2 and
GATA1 [39]. Furthermore, the enzymatic activity of FECH is dependent
on the presence of [4Fe–4S] cluster, again linking iron levels to heme
synthesis [40]. Distinct erythroid-specific elements have been identified
in the promoter region of Fech, together with erythroid-specific alter-
native splicing in the 3′ noncoding region of Fech mRNA in the mouse
genome [41,42], underscoring the unique regulatory mechanisms for
heme synthesis in RBC maturation. Beside regulation of heme synthesis
in erythroid cells, globin synthesis is controlled by the heme/BACH1
axis, in which heme binds to BACH1, a transcription suppressor, to
relieve the depression for globin gene expression [43]. These regula-
tions collectively coordinate cellular iron levels, heme synthesis, and Fig. 2. The ins and outs of heme transport.
Mitochondrial isoform FLVCR1b transports heme into the cytosol. ABCB10 is
globin protein expression, in order to maintain heme and iron home-
reported to forms complex with FECH and MFRN1. It is not clear whether
ostasis in erythroid cells.
ABCB10 transport heme. The cell surface FLVCR1a and the ABC transporter
ABCG2 have been implicated in heme export. MPR5 is a heme exporter which
can transport heme from cytosol to the secretory pathway. HCP1 is a folate
4. Heme Transport and erythropoiesis
importer as well as a low-affinity heme importer. HRG-1 is a heme importer
that localizes to endosomal/lysosomal compartments, but also traffics to the
The hydrophobicity and cytotoxicity of free heme suggests the ex- plasma membrane.
istence of heme trafficking pathways [44] (Fig. 2). The final step of the
heme biosynthetic pathway occurs in the mitochondrial matrix (Fig. 1).
Thus heme must be exported out of the mitochondria for incorporation The Feline leukemia virus subgroup C receptor-related protein 1
into hemoproteins located in various subcellular compartments [45]. (FLVCR1) was identified as a heme exporter [48]. FLVCR1 belongs to
Studies have shown that ABCB10 may facilitate transport of heme out the family of MFS (major facilitator superfamily) proteins which
of the mitochondria in erythroid cells [46]. ABCB10 is a mitochondrial transport small solutes across membranes facilitated by a counter ion
ABC transporter located in the inner mitochondrial membrane. ABCB10 gradient. Overexpression of FLVCR1 significant reduces cellular heme
interacts with MFRN1 and FECH and stabilizes the complex [47]. Ex- content, suggesting that FLVCR1 is involved in heme export [48].
pression of ABCB10 is highly induced during erythroid differentiation FLVCR1-null mice are embryonic lethal with deficiencies in definitive
and ABCB10 overexpression strengthens hemoglobin synthesis in ery- erythropoiesis and suffer from craniofacial and limb deformities re-
throid cells. It has been shown that ABCB10-null mice display defective sembling those of patients with Diamond-Blackfan anemia (DBA).
erythropoiesis and lack of hemoglobinized RBCs, indicating that Flvcr1−/− mice develop a severe macrocytic anemia with proerythro-
ABCB10 is essential for erythropoiesis in vivo [46]. However, conclusive blast maturation arrest, suggesting that erythroid precursors may be
evidence for direct heme transport by ABCB10 is still lacking. exporting excess heme to avoid heme toxicity [49]. Two isoforms of

206
J. Zhang, I. Hamza Molecular Genetics and Metabolism 128 (2019) 204–212

FLVCR has been identified, FLVCR1a and FLVCR1b. While FLVCR1a heme importer in the intestine of C. elegans [65]. Four HRG1 homologs,
encodes a plasma membrane-localized heme transporter, FLVCR1b was CeHRG-1, CeHRG-4, CeHRG-5 and CeHRG-6 were found in the C. ele-
identified to be a mitochondrial isoform encoded from an alternate gans genome. Significant heme-induced inward currents were observed
transcription start-site, resulting in a shortened N-terminus containing a in Xenopus oocytes injected with Cehrg-1, Cehrg-4, and human HRG1
mitochondrial-targeting signal [50]. Thus, FLVCR1a and FLVCR1b mRNA, indicating heme-dependent transport across cell membranes
contributes to intercellular and intracellular heme transport respec- [65]. Human HRG1 (SLC48A1) mRNA is abundant in the brain, kidney,
tively. FLVCR1a is expressed in different hematopoietic cells and shows heart, skeletal muscle, in addition to cell lines derived from duodenum
weak expression in the fetal liver, pancreas and kidney [51]. FLVCR1a and bone marrow [65]. HRG1 localizes to acidic endosomal and lyso-
may export heme during EP, a process in which macrophage phago- somal organelles in HEK293 cells, and its binding affinity to heme in-
cytose senescent RBCs, as FLVCR1a has been showed to interact with creases as pH decreases. Human HRG1 interacts with the C subunit of
the extracellular heme-binding protein hemopexin and mediate heme the vacuolar proton ATPase (V-ATPase) pump and enhances endosomal
export that is at least 100-fold more efficient than in the absence of acidification [66]. These studies collectively suggest that HRG1 trans-
hemopexin [52]. FLVCR1a expression is increased during erythropoi- ports heme from the exoplasmic space or the lumen of acidic en-
esis and is at its greatest during intermediate stages of RBC maturation dosome–lysosomal compartments into the cytoplasm. Global tran-
when HMOX1 expression is low, implying that FLVCR1a helps to scriptomic expression profiling shows that Hrg1 mRNA is expressed
maintain stoichiometric amounts of heme and globin by exporting ex- during erythroblast maturation [67]. Why a heme importer would be
cess heme and preventing heme toxicity to RBCs [53]. Depletion of expressed during erythropoiesis since developing RBCs are capable of
FLVCR1b in HeLa cells results in accumulation of mitochondrial heme, de novo heme synthesis is puzzling.
indicating that FLVCR1b may play a role in heme export from the mi-
tochondria [50]. It is not known whether FLVCR1b resides on the inner 5. Zebrafish as a genetic model to study heme and iron
or outer mitochondrial membrane. Moreover, if mitochondrial heme metabolism
export by FLVCR1b is indispensable and indeed attenuated in Flvcr1−/

mice, it cannot explain why embryos from Flvcr1−/− null mice can The complete sequencing of the zebrafish genome makes it a pow-
survive until E14.5. Yeast does not appear to have an obvious FLVCR erful tool for gene discovery research as the zebrafish genome largely
homolog, yet is able to export heme from the mitochondria indicating resembles the human genome [68]. Large-scale forward genetic studies
that alternate mechanisms must exist for mitochondrial heme export. have been carried-out in some invertebrate model organisms, particu-
ABCG2, also known as breast cancer resistance protein (BCRP) has larly in the nematode and fruit fly. However, it is extremely expensive
been identified as a heme exporter in mammals [54]. ABCG2 is ex- to perform this approach in vertebrate animal model such as mice.
pressed in hematopoietic stem cells (HSCs) and erythroid progenitors. Zebrafish was the first vertebrate organism established for large-scale
Compared to high FLVCR1 expression during erythropoietic differ- forward genetic screening. Chemical mutagenesis is achieved by ex-
entiation, the expression levels of ABCG2 are particularly high at the posing adult male fish to N-ethyl-N-nitrosourea (ENU) for several days
early stages of hematopoiesis [55]. ABCG2 binds to heme directly which induces point mutations in the spermatogonia. These male fish
through an extracellular loop 3 with a porphyrin-binding domain [56]. are then mated with wildtype (WT) female fish to propagate the mu-
Ectopically expressed ABCG2 exports ZnMP, a heme analog, in K562 tation to F1 progenies [4,69]. Screening methods such as antibody
cells. However, direct evidence that ABCG2 exports heme is still lacking staining, whole mount in situ hybridization (WISH) and behavioral
[56]. Whether FLVCR1 and ABCG2 can function synergistically to ex- analysis can be employed to identify morphological or genetic pheno-
port heme during erythropoiesis is not clear. ABCG2-null human pa- types. By contrast, gene-specific knockdown mediated by morpholino-
tients are defined as Jr.(a–) blood group with a unique side population injection serves as an efficient tool for reverse genetic study [70].
of HSCs, but no apparent deficiencies in erythropoiesis [57,58]. Mutagenesis and targeted gene disruptions have contributed to un-
MRP-5 was identified as a heme exporter in Caenorhabditis elegans. covering the genes involved in heme and iron metabolism in zebrafish.
This roundworm is a unique model for uncovering heme trafficking Recent development of gene-editing tools such as TALENs and CRISPR/
pathways because it cannot synthesize heme but acquires environ- cas9 for targeted gene disruption complements transient morpholino
mental heme for sustenance. Thus, worms need to acquire dietary heme gene knockdowns that were typically used for reverse genetic studies
via the intestine and distribute heme from the intestine to other tissues [71,72].
[59,60]. C. elegans MRP-5 (CeMRP-5) localizes to the basolateral Several key genes which are involved in heme synthesis and meta-
membrane of intestinal cells and loss of MRP-5 results in accumulation bolism have been elucidated in zebrafish by both forward and reverse
of ZnMP in intestinal cells and growth retardation. Overexpression of genetic manipulation. In Table 1, we summarize the genes and corre-
CeMRP-5 and human MRP5 in yeast retards growth owing to heme sponding mutants identified in zebrafish together with related disease
depletion, which can be rescued by co-expressing C. elegans heme im- in humans. We have also compiled a list of mammalian heme-iron
porters. Interestingly, overexpression of CeMRP-5 and human MPR5 metabolism genes and performed Bulk-BLAST homology alignments to
causes heme levels to increase in the secretory pathway in both yeast identify potential orthologs in the zebrafish genome [73] (Supple-
and mammalian cells, as measured by a secretory pathway hemoprotein mental Table 1). With the advent of current gene-editing tools for tar-
reporter. However, the physiological role of MRP5 in vertebrates and its geted gene disruptions, generating new zebrafish mutants and alleles
involvement in erythropoiesis is unclear. will be a powerful approach to model human disorders of heme and
Heme carrier protein 1 (HCP1, SLC46A1) is a membrane protein iron metabolism in a facile vertebrate model.
expressed in enterocytes and was proposed to be an intestinal heme
transporter [61]. However, subsequent studies revealed that HCP1 is a 5.1. ALAS2
proton-coupled folate transporter (PCFT) rather than a heme trans-
porter [62]. Erythroblasts from HCP1 knockout mice showed deficiency Alas2 is the erythroid-specific enzyme for heme synthesis, while
in differentiation and high apoptosis rate resulting in severe macrocytic Alas1 is found in all other tissues. Zebrafish mutant sauternes (sau) was
normochromic anemia, which was ascribed to folate but not heme de- identified from ENU mutagenesis and positional cloning revealed that
ficiency [63]. Moreover, knockdown of HCP1 by shRNA in Caco-2 cells sau contains mutation in alas2 [74]. The sau mutants have a microcytic,
attenuated both heme and folate uptake but increased heme oxygenase hypochromic anemia, delayed erythroid maturation, and abnormal
expression, suggesting HCP1 could potentially function as a low affinity globin gene expression, suggesting that defects in heme synthesis can
heme importer [62,64]. affect globin protein production. Interestingly, sau mutants have
The Heme Responsive Gene -1 (HRG1, SLC48A1) was identified as a normal RBC numbers and show anemia around 33 hpf, possibly because

207
J. Zhang, I. Hamza Molecular Genetics and Metabolism 128 (2019) 204–212

Table 1
Summary of zebrafish mutants related to heme-iron metabolism.
Genes Mutant name or morpholino knockdown Potential human disease References

ALAS2 sauternes (sau) Congenital Sideroblastic Anemia (CSA) [74]


ALAD Alad−/− ALA dehydratase deficient porphyria (ADP) [76]
CPOX Cpox−/− hereditary coproporphyria (HCP) [76]
UROD yquem (yqe) Porphyria cutanea tarda (PCT)/hepato- erythropoietic porphyria (HEP) [77]
PPOX montalcino (mno) Human variegate porphyria [78]
FECH freixenet (frx)/dracula (drc) Erythropoietic protoporphyria [4] [78]
GRX5 shiraz (sir) Sideroblastic anemia [80] [81]
MFRN1 frascati (frs) Erythropoietic protoporphyria [27]
DMT-1 chardonnay (cdy) Hypochromic microcytic anemia [85] [86]
FPN1 weissherbst (wei) Iron-deficient anemia [87] [88]
TF gavi congenital hypotransferrinemia [90]
TFR1 chianti (cia) hypochromic anemia [92]
HRG1 hrg1a;hrg1b (CRISPR) unknown [73]
FLVCR Morpholino Knockdown Heme toxicity [96]

maternal heme may permit early RBC development or genetic com- homozygous mutants, revealing functional conservation. The zebrafish
pensation by Alas1. As mutations in alas2 cause congenital sideroblastic mno mutant could be very useful for further elucidating the patho-
anemia (CSA) in humans, sau represents the first animal model of this physiology of variegate porphyria and identifying chemical modifiers of
disease in zebrafish [74]. Most interestingly, sau zebrafish mutants are this disease.
viable even though it has only one-tenth of overall heme content
compared to WT zebrafish. Sau mutants can be fully rescued by sup- 5.5. FECH
plementing developing embryos with ALA, which further enhances its
value as a model to study the pathology and to evaluate potential cures Two different mutant alleles of fech have been identified in zebra-
for CSA in humans [75]. fish, freixenet (frx) and dracula (drc) [4,79]. Protoporphyrin IX accu-
mulates in fech mutant embryos owing to a deficiency in the activity of
5.2. ALAD and CPOX ferrochelatase, the terminal enzyme in the pathway for heme bio-
synthesis. The mutants show light-dependent hemolysis and liver dis-
Unlike ALAS, only one form of ALAD and CPOX was found in both eases, similar to that seen in humans with heredity erythropoietic
human and zebrafish. By using targeted TALEN and CRISPR/Cas9, protoporphyria (HEP) resulting from a disorder of ferrochelatase.
zebrafish alad−/− and cpox−/− were recently generated to model the
ALA-dehydratase-deficient porphyria (ADP) and hereditary copropor- 5.6. GRX5
phyria (HCP) [76]. alad−/− and cpox−/− mutants suffer from hypo-
chromic anemia, owing to deficiency in heme synthesis, with accu- Phenotypic analysis of shiraz (sir) mutant zebrafish revealed an in-
mulation of ALA and coproporphyrinogen-III. The abnormal timate connection between heme biosynthesis and [4Fe-4S] formation,
morphology of early RBCs was possibly due by accumulated globin connecting the two main uses for iron in the mitochondria, which were
without heme, representing sickle-cell anemia in humans with globin previously thought to be independent processes [80]. In sir mutants,
protein aggregations. hypochromic anemia, in the context of normal mitochondrial iron and
oxidative stress levels, was shown to be caused by a deletion in the
5.3. UROD glutaredoxin 5 (grx5) gene. GRX5 is required for the synthesis of Fe-S
clusters in the mitochondria [81]. The zebrafish protein also localizes to
Zebrafish mutant yqe contains a nonsense mutation in the gene the mitochondria and is capable of rescuing grx5-deficient yeast strain.
encoding UROD, which converts uroporphyrinogen to coproporphyr- Fe-S clusters are known to negatively regulate binding of IRP1 to IREs.
inogen. Homozygous mutation in urod leads to two forms of porphyrias, Decreased Fe-S cluster assembly in sir mutant leads to increased IRP1
porphyria cutanea tarda (PCT) and hepatoerythropoietic porphyria activity, which inhibits the expression of IRE-regulated target genes
(HEP) in human, similar to the phenotypes in zebrafish with photo- involved in heme biosynthesis, such as alas2. Indeed, alas2 expression is
sensitive porphyria syndrome [77]. Excessive amounts of uroporphyr- absent in sir mutants. Deletion of IREs in the alas2 mRNA rescued the
inogens and 7-carboxylate porphyrin accumulate in yqe embryos, re- anemic phenotype while overexpression of full-length alas2 mRNA did
presenting human HEP patients characterized by photosensitive skin not, suggesting that the function of GRX5 in regulating ALAS2 expres-
and excessive excretion of heme biosynthesis byproducts, uroporphyrin sion is through Fe-S clusters and IRP/IRE activity. An evolutionary
and 7-decarboxylate porphyrin in their urine. The zebrafish yqe mutant conserved role for GRX5 in regulating heme synthesis was confirmed in
phenocopies human patients facilitating studying of HEP pathogenesis human patients with GRX5 mutation [82]. Thus, shiraz mutants could
and development of new therapeutics. be a good animal model to screen for therapeutics related to defi-
ciencies in Fe-S cluster assembly.
5.4. PPOX
5.7. MFRN1
Zebrafish porphyria mutant montalcino (mno) contains mutation in
ppox, which catalyzes the oxidation of protoporphyrinogen, re- Two mitoferrin homologs are found in the zebrafish genome, Mfrn1
presenting human variegate porphyria [78]. Initially, at the onset of and Mfrn2. Positional cloning of frascati (frs) zebrafish mutants iden-
circulation, mno displays normal numbers of RBCs but are o-dianisidine tified a missense mutation in the mfrn1 (slc25A37) gene, an erythroid-
negative. A visible decrease in circulating erythrocytes can be observed specific form [27]. Frs mutants develop hypochromic anemia and show
after 36 hpf and the RBCs in the mutant embryos are fluorescent. The defects in erythroid maturation. Mouse Mfrn1 rescues the phenotypes in
mno mutant zebrafish could survive to approximately 25 dpf. Accord- zebrafish frs mutants. The same anemic phenotype was observed in a
ingly, human PPOX could partially rescue the hypochromia in mouse model with Mfrn1-knockout [28,83]. The MFRN2 (Slc25A28)

208
J. Zhang, I. Hamza Molecular Genetics and Metabolism 128 (2019) 204–212

paralog functions in mitochondrial iron import in non-erythroid tissues. erythroid differentiation were ascribed to mutations in tfr1a gene [92].
During early development, tfr1a transcripts are expressed specifically in
5.8. DMT1 erythrocytes. Importantly, cytoplasmic delivery of iron by microinjec-
tion at one-cell stage - but not intravenous iron injections rescue the
DMT1 is upregulated by dietary iron deficiency, and Belgrade rats hypochromia in cia mutants, indicating that tfr1a mutation prevents
that harbor a mutation in dmt1 suffer from iron deficiency anemia erythrocytes from taking-up and utilizing circulating iron. Intriguingly,
[23,84]. Zebrafish mutants chardonnay (cdy) carry a nonsense mutation a second tfr1 gene, tfr1b was identified together with tfr1a, a typical
in dmt1 with reduced hemoglobin levels and delayed erythrocyte ma- feature in zebrafish genome which has undergone whole genome du-
turation [85]. The Dmt1 protein is expressed in erythroid cells and the plication [93,94]. Whereas tfr1a is expressed in erythrocytes during
duodenum suggesting its role in erythropoiesis and intestinal iron ab- early development and cia mutants are anemic, tfr1b is ubiquitously
sorption. Cells overexpressing WT zebrafish dmt1 uptake nearly ten- expressed throughout embryogenesis and knockdown of tfr1b by mor-
times the amount of iron as non-transfected control cells, whereas the pholinos do not affect hemoglobinization. Tfr1b morphants have re-
cdy mutant protein is not functional. However, cdy mutants can survive tarded growth and develop brain necrosis, a phenotype that is similar to
to adulthood despite severe anemia, again suggesting an alternate path the neurologic defects observed in the mouse model [95], indicating
for iron absorption in zebrafish. In humans, mutations in DMT1 cause a that tfr1b may be involved in iron uptake in non-erythroid tissues. Tfr1a
phenotype of hypochromic microcytic anemia combined with iron (cia) and tfr1b deficient zebrafish embryos recapitulate the phenotype
overload, further supporting the possible existence of alternative me- of TFR1−/− mice [92].
chanisms for duodenal iron absorption that bypasses DMT1 [86].
5.12. FLVCR and HRG1
5.9. FPN1
Currently no zebrafish mutants have been reported with perturba-
The first identified iron exporter Fpn1 (Slc40A1) was found by po- tion in Flvcr. However, the function of zebrafish Flvcr is related to er-
sition-cloning of zebrafish weissherbst (weh) mutant [87]. weh mutant ythroid differentiation and maturation as determined by morpholino
embryos show hypochromic anemia with decreased hemoglobin levels, gene knockdowns [96]. Two splicing isoforms have been identified:
blocked erythroid maturation, and reduced numbers of erythrocytes. flvcr1a and flvcr1b. Flvcr1a is required for the expansion of committed
Erythroid cells of mutant embryos have significantly lower iron con- erythroid progenitors but cannot drive their terminal differentiation,
centrations compared to WT embryos, suggesting iron deficiency. Mi- while Flvcr1b contributes to the expansion phase and is required for
croinjection of iron-dextran rescues the anemic phenotype of weh mu- differentiation [96]. The coordinated expression of flvcr1a and flvcr1b
tants and continued injection of iron-dextran rescues the embryonic contributes to controlling the cytosolic heme pool required to sustain
lethality. These rescued fish, however, are only normal until 6 months regulation of erythroid progenitors and hemoglobin synthesis for he-
of age and eventually develop hypochromic anemia by 12 months, moglobinization during terminal maturation. Interestingly, treatment
suggesting that Fpn1 is also involved in adult red cell function, speci- with succinylacetone (SA), an inhibitor of heme synthesis, rescues the
fically iron-recycling in adult zebrafish. Compared with iron injected phenotype of flvcr1a morphants while heme supplementation restores
WT fish, the rescued mutants had increased iron staining in kidney hemoglobinzation of flvcr1b morphants, suggesting that the in-
macrophages, as well as increased staining in intestinal villi, suggesting tracellular heme pool during erythropoiesis is tightly regulated. As
that fpn1 mutation impairs iron export in these tissues. The iron-rescued Flvcr−/− mice are embryonic lethal, zebrafish flvcr mutants will be
weh mutants also have hepatic iron overload, with particularly high useful to dissect the functions of FLVCR in early development.
iron levels in the liver [88]. FPN1 also localizes to the yolk-syncytial Recently, zebrafish mutants for hrg1 (Slc48a1), a heme transporter
layer (YSL) during embryonic development, suggesting that FPN1 may with homologs in mammals, was generated by CRISPR/cas9 [73,97].
transport maternal iron from the yolk for embryogenesis. Both mice and Zebrafish genome contains two hrg1 paralogs - hrg1a (slc48a1b) and
humans have homologs of FPN1 with high similarities to zebrafish hrg1b (slc48a1a). Both are ubiquitously expressed throughout devel-
Fpn1. Mammalian FPN1 is robustly expressed in the placenta, duo- oping embryos and in the one-cell embryo. More importantly,
denum, and liver. At the protein level, human FPN1 is concentrated on hrg1a;hrg1b double mutants show accumulation of heme in the kidney
the basal surface of syncytiotrophoblasts in the placenta, an organ that macrophages due to defects in heme-iron recycling from damaged
is functionally like zebrafish YSL, indicating that human FPN1 plays a RBCs. RNAseq results show significant perturbation in heme-iron me-
role in maternal-fetal iron export. In mice, FPN1 is expressed on the tabolism and immune-related genes in the absence of hrg1. Altogether,
basolateral surface of enterocytes, suggesting a role in intestinal iron the fish studies affirm that Hrg1 is essential for recycling RBCs and that
transport [89]. this function is performed by the kidney marrow, which is the adult
hematopoietic organ in zebrafish [73].
5.10. TF
6. Conclusions
The zebrafish mutant gavi (gav) was shown to have mutations in
transferrin-a (Tf-a), which encodes the principal serum iron carrier Zebrafish has proven to be an indispensable vertebrate model to
[90]. Gav mutant embryos exhibit reduced tf-a expression and impaired study heme and iron metabolism with unique genetic advantages not
hemoglobin production with hypochromic anemia and embryonic afforded by mammalian models. The advantages of ex utero genetic
lethality by 14 dpf. In humans, phenotype of congenital hypo- manipulation and embryonic transparency allow convenient functional
transferrinemia caused by TF mutation is highly similar to those of gav validation of both zebrafish and human genes. Moreover, chemical-
mutants, including hypochromic anemia and embryonic death [91]. genetics drug screens in zebrafish provide a more comprehensive whole
The gav mutant is thus an ideal whole vertebrate anemia model for animal in-a-dish analyses for drug tests than typical cell-culture based
studying symptoms corresponding to human pathologies related to Tf. assays [98]. Given the high degree of conservation in the heme and iron
acquisition pathways in zebrafish and humans, it would be prudent to
5.11. TFR1 develop diseased models of humanized zebrafish for targeted high
throughput screens to identify more efficacious therapeutic drugs
Transferrin-bound iron is taken up into cells by binding to the against rare diseases.
transferrin receptor 1 (TFR1). Four different zebrafish chianti (cia) Supplementary data to this article can be found online at https://
mutants with varying degrees of hypochromic anemia and defective doi.org/10.1016/j.ymgme.2018.12.007.

209
J. Zhang, I. Hamza Molecular Genetics and Metabolism 128 (2019) 204–212

Author information [25] S. Leidgens, K.Z. Bullough, H. Shi, F. Li, M. Shakoury-Elizeh, T. Yabe,
P. Subramanian, E. Hsu, N. Natarajan, A. Nandal, T.L. Stemmler, C.C. Philpott, Each
member of the poly-r(C)-binding protein 1 (PCBP) family exhibits iron chaperone
I.H. is the President and Founder of Rakta Therapeutics Inc. (College activity toward Ferritin, J. Biol. Chem. 288 (2013) 17791–17802.
Park, MD), a company involved in the development of heme trans- [26] M.-s. Ryu, D. Zhang, O. Protchenko, M. Shakoury-Elizeh, C.C. Philpott, PCBP1 and
porter-related diagnostics. He declares no other competing financial NCOA4 regulate erythroid iron storage and heme biosynthesis, J. Clin. Investig.
(2017) 1–12.
interests. Correspondence and request for materials should be ad- [27] G.C. Shaw, J.J. Cope, L. Li, K. Corson, C. Hersey, G.E. Ackermann, B. Gwynn,
dressed to I.H. (hamza@umd.edu). A.J. Lambert, R.A. Wingert, D. Traver, N.S. Trede, B.A. Barut, Y. Zhou, E. Minet,
A. Donovan, A. Brownlie, R. Balzan, M.J. Weiss, L.L. Peters, J. Kaplan, L.I. Zon,
B.H. Paw, Mitoferrin is essential for erythroid iron assimilation, Nature 440 (2006)
Acknowledgements 96–100.
[28] M.-B.B. Troadec, D. Warner, J. Wallace, K. Thomas, G.J. Spangrude, J. Phillips,
IH is supported by grants from the US National Institutes of Health O. Khalimonchuk, B.H. Paw, D.M. Ward, J. Kaplan, Targeted deletion of the mouse
Mitoferrin1 gene: from anemia to protoporphyria, Blood 117 (2011) 5494–5502.
(DK074797, DK85035, AI67979, ES25661).
[29] D.R. Richardson, P. Ponka, D. Vyoral, Distribution of iron in reticulocytes after
inhibition of heme synthesis with succinylacetone: examination of the inter-
References mediates involved in iron metabolism, Blood 87 (1996) 3477–3488.
[30] A.-S. Zhang, A.D. Sheftel, P. Ponka, Intracellular kinetics of iron in reticulocytes:
evidence for endosome involvement in iron targeting to mitochondria, Blood 105
[1] G.J. Lieschke, P.D. Currie, Animal models of human disease: zebrafish swim into (2005) 368–375.
view, Nat. Rev. Genet. 8 (2007) 353–367. [31] M. Muñoz, J.A. García-Erce, Á.F. Remacha, Disorders of iron metabolism. Part II:
[2] A.J. Davidson, L.I. Zon, The ‘definitive’ (and ‘primitive’) guide to zebrafish hema- iron deficiency and iron overload, J. Clin. Pathol. 64 (2011) 287–296.
topoiesis, Oncogene 23 (2004) 7233. [32] G.A. Hunter, G.C. Ferreira, Molecular Enzymology of 5-Aminolevulinate Synthase,
[3] J.L. Galloway, L.I. Zon, 3 Ontogeny of Hematopoiesis: Examining the Emergence of the Gatekeeper of Heme Biosynthesis, Biochimica et Biophysica Acta - Proteins and
Hematopoietic Cells in the Vertebrate Embryo, in: B.T.-C.T.i.D. Biology (Ed.), Proteomics, (2011), pp. 1467–1473.
Academic Press, 2003, pp. 139–158. [33] B.K. May, S.C. Dogra, T.J. Sadlon, C.R. Bhasker, T.C. Cox, S.S. Bottomley, Molecular
[4] D.G. Ransom, P. Haffter, J. Odenthal, A. Brownlie, E. Vogelsang, R.N. Kelsh, Regulation of Heme Biosynthesis in Higher Vertebrates Progress in Nucleic Acid
M. Brand, F.J. van Eeden, M. Furutani-Seiki, M. Granato, M. Hammerschmidt, Research and Molecular Biology, 51 (1995), pp. 1–51.
C.P. Heisenberg, Y.J. Jiang, D.A. Kane, M.C. Mullins, C. Nüsslein-Volhard, [34] T.J. Sadlon, T. Dell'Oso, K.H. Surinya, B.K. May, T. Dell'Oso, K.H. Surinya, B.K. May,
Characterization of zebrafish mutants with defects in embryonic hematopoiesis, Regulation of erythroid 5-aminolevulinate synthase expression during erythropoi-
Development (Cambridge, England) 123 (1996) 311–319. esis, Int. J. Biochem. Cell Biol. 31 (1999) 1153–1167.
[5] Jill L.O. de Jong, Leonard I. Zon, Use of the Zebrafish System to Study Primitive and [35] M.J. Weiss, C. Yu, S.H. Orkin, Erythroid-cell-specific properties of transcription
Definitive Hematopoiesis Annual Review of Genetics, 39 (2005), pp. 481–501. factor GATA-1 revealed by phenotypic rescue of a gene-targeted cell line, Mol. Cell.
[6] S.E. Lyons, N.D. Lawson, L. Lei, P.E. Bennett, B.M. Weinstein, P.P. Liu, A nonsense Biol. 17 (1997) 1642–1651.
mutation in zebrafish gata1 causes the bloodless phenotype in vlad tepes, Proc. [36] N. Tanimura, E. Miller, K. Igarashi, D. Yang, J.N. Burstyn, C.N. Dewey,
Natl. Acad. Sci. 99 (2002) 5454–5459. E.H. Bresnick, Mechanism governing heme synthesis reveals a GATA factor/heme
[7] Q. Long, A. Meng, H. Wang, J.R. Jessen, M.J. Farrell, S. Lin, GATA-1 expression circuit that controls differentiation, EMBO Rep. 17 (2015) 1–17.
pattern can be recapitulated in living transgenic zebrafish using GFP reporter gene, [37] Y. Zhang, J. Zhang, W. An, Y. Wan, S. Ma, J. Yin, X. Li, J. Gao, W. Yuan, Y. Guo,
Development (Cambridge, England) 124 (1997) 4105–4111. J.D. Engel, L. Shi, T. Cheng, X. Zhu, Intron 1 GATA site enhances ALAS2 expression
[8] A.A.I.G. Cumano, Ontogeny of the Hematopoietic System Annual Review of indispensably during erythroid differentiation, Nucleic Acids Res. 45 (2017)
Immunology, 25 (2007), pp. 745–785. 657–671.
[9] C.E. Burns, T. Deblasio, Y. Zhou, J. Zhang, L. Zon, S.D. Nimer, Isolation and char- [38] N. Wilkinson, K. Pantopoulos, The IRP/IRE system in vivo: insights from mouse
acterization of runxa and runxb, zebrafish members of the runt family of tran- models, Front. Pharmacol. 5 (2014) 176.
scriptional regulators, Exp. Hematol. 30 (2002) 1381–1389. [39] A. Tugores, S.T. Magness, D.A. Brenner, A single promoter directs both house-
[10] C.E. Willett, A. Cortes, A. Zuasti, A.G. Zapata, Early hematopoiesis and developing keeping and erythroid preferential expression of the human ferrochelatase gene, J.
lymphoid organs in the zebrafish, Dev. Dyn. 214 (1999) 323–336. Biol. Chem. 269 (1994) 30789–30797.
[11] E. Murayama, K. Kissa, A. Zapata, E. Mordelet, V. Briolat, H.-F. Lin, R.I. Handin, [40] D.R. Crooks, M.C. Ghosh, R.G. Haller, W.-H. Tong, T.A. Rouault, Posttranslational
P. Herbomel, Tracing Hematopoietic Precursor Migration to Successive stability of the heme biosynthetic enzyme ferrochelatase is dependent on iron
Hematopoietic Organs during Zebrafish Development Immunity, 25 (2006), pp. availability and intact iron-sulfur cluster assembly machinery, Blood 115 (2010)
963–975. (860 LP – 869).
[12] H. Jin, J. Xu, Z. Wen, Migratory path of definitive hematopoietic stem/progenitor [41] S.T. Magness, A. Tugores, D.A. Brenner, Analysis of ferrochelatase expression
cells during zebrafish development, Blood 109 (2007) (5208 LP – 5214). during hematopoietic development of embryonic stem cells, Blood 95 (2000) (3568
[13] E.C. Liao, N.S. Trede, D. Ransom, A. Zapata, M. Kieran, L.I. Zon, Non-cell autono- LP – 3577).
mous requirement for the < em > bloodless < /em > gene in primitive hemato- [42] R.Y.Y. Chan, H.M. Schulman, P. Ponka, Expression of ferrochelatase mRNA in er-
poiesis of zebrafish, Development 129 (2002) (649 LP – 659). ythroid and non-erythroid cells, Biochem. J. 292 (1993) (343 LP – 349).
[14] F.-Y. Chan, J. Robinson, A. Brownlie, R.A. Shivdasani, A. Donovan, C. Brugnara, [43] T. Tahara, J. Sun, K. Nakanishi, M. Yamamoto, H. Mori, T. Saito, H. Fujita,
J. Kim, B.-C. Lau, H.E. Witkowska, L.I. Zon, Characterization of Adult α- and β- K. Igarashi, S. Taketani, Heme positively regulates the expression of β-globin at the
Globin Genes in the Zebrafish Blood, 89 (1997) (688 LP – 700). locus control region via the transcriptional factor bach1 in erythroid cells, J. Biol.
[15] A. Brownlie, C. Hersey, A.C. Oates, B.H. Paw, A.M. Falick, H.E. Witkowska, J. Flint, Chem. 279 (2004) 5480–5487.
D. Higgs, J. Jessen, N. Bahary, H. Zhu, S. Lin, L. Zon, Characterization of embryonic [44] S. Severance, I. Hamza, Trafficking of Heme and Porphyrins in Metazoa, (2009), pp.
globin genes of the zebrafish, Dev. Biol. 255 (2003) 48–61. 4596–4616.
[16] D. Traver, B.H. Paw, K.D. Poss, W.T. Penberthy, S. Lin, L.I. Zon, Transplantation and [45] I.J. Schultz, C. Chen, B.H. Paw, I. Hamza, Iron and porphyrin trafficking in heme
in vivo imaging of multilineage engraftment in zebrafish bloodless mutants, Nat. biogenesis, J. Biol. Chem. 285 (2010) 26753–26759.
Immunol. 4 (2003) 1238–1246. [46] M. Liesa, W. Qiu, O.S. Shirihai, Mitochondrial ABC Transporters Function: The Role
[17] D. Traver, A. Winzeler, H.M. Stern, E.A. Mayhall, D.M. Langenau, J.L. Kutok, of ABCB10 (ABC-me) as a Novel Player in Cellular Handling of Reactive Oxygen
A.T. Look, L.I. Zon, Effects of lethal irradiation in zebrafish and rescue by hema- Species Biochimica et Biophysica Acta (BBA) - Molecular Cell Research, 1823
topoietic cell transplantation, Blood 104 (2004) (1298 LP – 1305). (2012), pp. 1945–1957.
[18] D. Carradice, G.J. Lieschke, Zebrafish in hematology: sushi or science? Blood 111 [47] W. Chen, H.A. Dailey, B.H. Paw, Ferrochelatase forms an oligomeric complex with
(2008) 3331–3342. mitoferrin-1 and Abcb10 for erythroid heme biosynthesis, Blood 116 (2010) (628
[19] G.J. Lieschke, A.C. Oates, M.O. Crowhurst, A.C. Ward, J.E. Layton, Morphologic and LP – 630).
functional characterization of granulocytes and macrophages in embryonic and [48] J.G. Quigley, Z. Yang, M.T. Worthington, J.D. Phillips, K.M. Sabo, D.E. Sabath,
adult zebrafish, Blood 98 (2001) (3087 LP – 3096). C.L. Berg, S. Sassa, B.L. Wood, J.L. Abkowitz, Identification of a human heme ex-
[20] A.L. Menke, J.M. Spitsbergen, A.P.M. Wolterbeek, R.A. Woutersen, Normal porter that is essential for erythropoiesis, Cell 118 (2004) 757–766.
anatomy and histology of the adult zebrafish, Toxicol. Pathol. 39 (2011) 759–775. [49] S.B. Keel, R.T. Doty, Z. Yang, J.G. Quigley, J. Chen, S. Knoblaugh, P.D. Kingsley,
[21] P. Aisen, A. Leibman, J. Zweier, Stoichiometric and site characteristics of the I. De Domenico, M.B. Vaughn, J. Kaplan, J. Palis, J.L. Abkowitz, A heme export
binding of iron to human transferrin, J. Biol. Chem. 253 (1978) 1930–1937. protein is required for red blood cell differentiation and iron homeostasis, Science
[22] R.S. Ohgami, D.R. Campagna, E.L. Greer, B. Antiochos, A. McDonald, J. Chen, (New York, N.Y.) 319 (2008) 825–828.
J.J. Sharp, Y. Fujiwara, J.E. Barker, M.D. Fleming, Identification of a ferrireductase [50] D. Chiabrando, S. Marro, S. Mercurio, C. Giorgi, S. Petrillo, F. Vinchi, V. Fiorito,
required for efficient transferrin-dependent iron uptake in erythroid cells, Nat. S. Fagoonee, A. Camporeale, E. Turco, G.R. Merlo, L. Silengo, F. Altruda, P. Pinton,
Genet. 37 (2005) 1264–1269. E. Tolosano, The Mitochondrial Heme Exporter FLVCR1b Mediates Erythroid
[23] H. Gunshin, B. Mackenzie, U.V. Berger, Y. Gunshin, M.F. Romero, W.F. Boron, Differentiation, 122 (2012), pp. 4569–4579.
S. Nussberger, J.L. Gollan, M.A. Hediger, Cloning and characterization of a mam- [51] J.G. Quigley, C.C. Burns, M.M. Anderson, E.D. Lynch, K.M. Sabo, J. Overbaugh,
malian proton-coupled metal-ion transporter, Nature 388 (1997) 482–488. J.L. Abkowitz, Cloning of the cellular receptor for feline leukemia virus subgroup C
[24] C.C. Winterbourn, Toxicity of iron and hydrogen peroxide: the Fenton reaction, (FeLV-C), a retrovirus that induces red cell aplasia, Blood 95 (2000) (1093 LP –
Toxicol. Lett. 82-83 (1995) 969–974. 1099).

210
J. Zhang, I. Hamza Molecular Genetics and Metabolism 128 (2019) 204–212

[52] Z. Yang, J.D. Philips, R.T. Doty, P. Giraudi, J.D. Ostrow, C. Tiribelli, A. Smith, zebrafish: a recipe for functional genomics? Brief. Funct. Genomics Proteomics 1
J.L. Abkowitz, Kinetics and specificity of feline leukemia virus subgroup C receptor (2002) 239–256.
(FLVCR) export function and its dependence on hemopexin, J. Biol. Chem. 285 [71] P. Huang, A. Xiao, M. Zhou, Z. Zhu, S. Lin, B. Zhang, Heritable gene targeting in
(2010) 28874–28882. zebrafish using customized TALENs, Nat. Biotechnol. 29 (2011) 699–700.
[53] L.R. Alves, E.S. Costa, M.H.F. Sorgine, M.C.L. Nascimento-Silva, C. Teodosio, [72] L.-E. Jao, S.R. Wente, W. Chen, Efficient multiplex biallelic zebrafish genome
P. Brcena, H.C. Castro-Faria-Neto, P.T.P.C.T. Bozza, A. Orfao, P.L. Oliveira, editing using a CRISPR nuclease system, Proc. Natl. Acad. Sci. U. S. A. 110 (2013)
C.M. Maya-Monteiro, P. Bárcena, H.C. Castro-Faria-Neto, P.T.P.C.T. Bozza, 13904–13909.
A. Orfao, P.L. Oliveira, C.M. Maya-Monteiro, Heme-oxygenases during ery- [73] J. Zhang, I. Chambers, S. Yun, J. Phillips, M. Krause, I. Hamza, Hrg1 promotes
thropoiesis in K562 and human bone marrow cells, PLoS ONE 6 (2011) e21358. heme-iron recycling during hemolysis in the zebrafish kidney, PLoS Genet. 14
[54] P. Krishnamurthy, J.D. Schuetz, The ABC Transporter Abcg2/Bcrp: Role in Hypoxia (2018) e1007665.
Mediated Survival Biometals, 18 (2005), pp. 349–358. [74] A. Brownlie, A. Donovan, S.J. Pratt, B.H. Paw, A.C. Oates, C. Brugnara,
[55] K. Yamamoto, S. Suzu, Y. Yoshidomi, M. Hiyoshi, H. Harada, S. Okada, H.E. Witkowska, S. Sassa, L.I. Zon, Positional cloning of the zebrafish sauternes
Erythroblasts highly express the ABC transporter Bcrp1/ABCG2 but do not show the gene: a model for congenital sideroblastic anaemia, Nat. Genet. 20 (1998) 244–250.
side population (SP) phenotype, Immunol. Lett. 114 (2007) 52–58. [75] J.R. Kardon, Y.Y. Yien, N.C. Huston, D.S. Branco, G.J. Hildick-Smith, K.Y. Rhee,
[56] E. Desuzinges-Mandon, O. Arnaud, L. Martinez, F. Huché, A. Di Pietro, P. Falson, B.H. Paw, T.A. Baker, Mitochondrial ClpX activates a key enzyme for heme bio-
ABCG2 transports and transfers heme to albumin through its large extracellular synthesis and erythropoiesis, Cell 161 (2015) 858–867.
loop, J. Biol. Chem. 285 (2010) 33123–33133. [76] D. Sakamoto, H. Kudo, K. Inohaya, H. Yokoi, T. Narita, K. Naruse, H. Mitani,
[57] C. Saison, V. Helias, B.A. Ballif, T. Peyrard, H. Puy, T. Miyazaki, S. Perrot, K. Araki, A. Shima, Y. Ishikawa, Y. Imai, A. Kudo, A mutation in the gene for delta-
M. Vayssier-Taussat, M. Waldner, P.-Y. Le Pennec, J.-P. Cartron, L. Arnaud, Null aminolevulinic acid dehydratase (ALAD) causes hypochromic anemia in the me-
alleles of ABCG2 encoding the breast cancer resistance protein define the new blood daka, Oryzias latipes, Mech. Dev. 121 (2004) 747–752.
group system Junior, Nat. Genet. 44 (2012) 174–177. [77] H. Wang, Q. Long, S.D. Marty, S. Sassa, S. Lin, A zebrafish model for hepatoery-
[58] T. Zelinski, G. Coghlan, X.-Q. Liu, M.E. Reid, ABCG2 null alleles define the Jr(A−) thropoietic porphyria, Nat. Genet. 20 (1998) 239–243.
blood group phenotype, Nat. Genet. 44 (2012) 131–132. [78] K.A. Dooley, P.G. Fraenkel, N.B. Langer, B. Schmid, A.J. Davidson, G. Weber,
[59] A.U. Rao, L.K. Carta, E. Lesuisse, I. Hamza, Lack of heme synthesis in a free-living K. Chiang, H. Foott, C. Dwyer, R.A. Wingert, Y. Zhou, B.H. Paw, L.I. Zon,
eukaryote, Proc. Natl. Acad. Sci. U. S. A. 102 (2005) 4270–4275. Montalcino, a zebrafish model for variegate porphyria, Exp. Hematol. 36 (2008)
[60] T. Korolnek, J. Zhang, S. Beardsley, George L.L. Scheffer, I. Hamza, Control of 1132–1142.
metazoan heme homeostasis by a conserved multidrug resistance protein, Cell [79] S. Childs, B.M. Weinstein, M.A. Mohideen, S. Donohue, H. Bonkovsky,
Metab. 19 (2014) 1008–1019. M.C. Fishman, Zebrafish dracula encodes ferrochelatase and its mutation provides a
[61] M. Shayeghi, G.O. Latunde-Dada, J.S. Oakhill, A.H. Laftah, K. Takeuchi, model for erythropoietic protoporphyria, Curr. Biol. 10 (2000) 1001–1004.
N. Halliday, Y. Khan, A. Warley, F.E. McCann, R.C. Hider, D.M. Frazer, [80] R.A. Wingert, J.L. Galloway, B. Barut, H. Foott, P. Fraenkel, J.L. Axe, G.J. Weber,
G.J. Anderson, C.D. Vulpe, R.J. Simpson, A.T. McKie, Identification of an intestinal K. Dooley, A.J. Davidson, B. Schmid, B.H. Paw, G.C. Shaw, P. Kingsley, J. Palis,
heme transporter, Cell 122 (2005) 789–801. H. Schubert, O. Chen, J. Kaplan, L.I. Zon, Deficiency of glutaredoxin 5 reveals Fe-S
[62] S. Le Blanc, M.D. Garrick, M. Arredondo, Heme carrier protein 1 transports heme clusters are required for vertebrate haem synthesis, Nature 436 (2005) 1035–1039.
and is involved in heme-Fe metabolism, AJP 302 (2012) C1780–C1785. [81] M.T. Rodríguez-Manzaneque, J. Tamarit, G. Bellí, J. Ros, E. Herrero, Grx5 is a
[63] K.V. Salojin, R.M. Cabrera, W. Sun, W.C. Chang, C. Lin, L. Duncan, K.A. Platt, mitochondrial glutaredoxin required for the activity of iron/sulfur enzymes, Mol.
R. Read, P. Vogel, Q. Liu, R.H. Finnell, T. Oravecz, A mouse model of hereditary Biol. Cell 13 (2002) 1109–1121.
folate malabsorption: deletion of the < em > PCFT < /em > gene leads to systemic [82] C. Camaschella, A. Campanella, L. De Falco, L. Boschetto, R. Merlini, L. Silvestri,
folate deficiency, Blood 117 (2011) 4895–4904. S. Levi, A. Iolascon, The human counterpart of zebrafish shiraz shows sideroblastic-
[64] A.H. Laftah, G.O. Latunde-Dada, S. Fakih, R.C. Hider, R.J. Simpson, A.T. McKie, like microcytic anemia and iron overload, Blood 110 (2007) 1353–1358.
Haem and folate transport by proton-coupled folate transporter/haem carrier pro- [83] J. Chung, S.A. Anderson, B. Gwynn, K.M. Deck, M.J. Chen, N.B. Langer, G.C. Shaw,
tein 1 (SLC46A1), Br. J. Nutr. 101 (2008) 1150–1156. N.C. Huston, L.F. Boyer, S. Datta, P.N. Paradkar, L. Li, Z. Wei, A.J. Lambert, K. Sahr,
[65] A. Rajagopal, A.U. Rao, J. Amigo, M. Tian, S.K. Upadhyay, C. Hall, S. Uhm, J.G. Wittig, W. Chen, W. Lu, B. Galy, T.M. Schlaeger, M.W. Hentze, D.M. Ward,
M.K. Mathew, M.D. Fleming, B.H. Paw, M. Krause, I. Hamza, Haem homeostasis is J. Kaplan, R.S. Eisenstein, L.L. Peters, B.H. Paw, Iron regulatory protein-1 protects
regulated by the conserved and concerted functions of HRG-1 proteins, Nature 453 against mitoferrin-1-deficient porphyria, J. Biol. Chem. 289 (2014) 7835–7843.
(2008) 1127–1131. [84] M.D. Fleming, C.C. Trenor III, M.A. Su, D. Foernzler, D.R. Beier, W.E. Dietrich,
[66] K.M. O'Callaghan, V. Ayllon, J. O'Keeffe, Y. Wang, O.T. Cox, G. Loughran, N.C. Andrews, Microcytic anaemia mice have a mutation in Nramp2, a candidate
M. Forgac, R. O'Connor, Heme-binding protein HRG-1 is induced by insulin-like iron transporter gene, Nat. Genet. 16 (1997) 383–386.
growth factor I and associates with the vacuolar H(+)-ATPase to control endosomal [85] A. Donovan, A. Brownlie, M.O. Dorschner, Y. Zhou, S.J. Pratt, B.H. Paw,
pH and receptor trafficking, J. Biol. Chem. 285 (2010) 381–391. R.B. Phillips, C. Thisse, B. Thisse, L.I. Zon, The zebrafish mutant gene chardonnay
[67] X. An, V.P. Schulz, J. Li, K. Wu, J. Liu, F. Xue, J. Hu, N. Mohandas, P.G. Gallagher, (cdy) encodes divalent metal transporter 1 (DMT1), Blood 100 (2002) 4655–4659.
Global transcriptome analyses of human and murine terminal erythroid differ- [86] M.P. Mims, Y. Guan, D. Pospisilova, M. Priwitzerova, K. Indrak, P. Ponka,
entiation, Blood 123 (2014) (3466 LP – 3477). V. Divoky, J.T. Prchal, Identification of a human mutation of DMT1 in a patient
[68] K. Howe, M.D. Clark, C.F. Torroja, J. Torrance, C. Berthelot, M. Muffato, with microcytic anemia and iron overload, Blood 105 (2005) 1337–1342.
J.E. Collins, S. Humphray, K. McLaren, L. Matthews, S. McLaren, I. Sealy, [87] A. Donovan, A. Brownlie, Y. Zhou, J. Shepard, S.J. Pratt, J. Moynihan, B.H. Paw,
M. Caccamo, C. Churcher, C. Scott, J.C. Barrett, R. Koch, G.J. Rauch, S. White, A. Drejer, B. Barut, A. Zapata, T.C. Law, C. Brugnara, S.E. Lux, G.S. Pinkus,
W. Chow, B. Kilian, L.T. Quintais, J.A. Guerra-Assunção, Y. Zhou, Y. Gu, J. Yen, J.- J.L. Pinkus, P.D. Kingsley, J. Palis, M.D. Fleming, N.C. Andrews, L.I. Zon, Positional
H. Vogel, T. Eyre, S. Redmond, R. Banerjee, J. Chi, B. Fu, E. Langley, S.F. Maguire, cloning of zebrafish ferroportin1 identifies a conserved vertebrate iron exporter,
G.K. Laird, D. Lloyd, E. Kenyon, S. Donaldson, H. Sehra, J. Almeida-King, Nature 403 (2000) 776–781.
J. Loveland, S. Trevanion, M. Jones, M. Quail, D. Willey, A. Hunt, J. Burton, S. Sims, [88] P.G. Fraenkel, D. Traver, A. Donovan, D. Zahrieh, L.I. Zon, Ferroportin1 is required
K. McLay, B. Plumb, J. Davis, C. Clee, K. Oliver, R. Clark, C. Riddle, D. Elliot, for normal iron cycling in zebrafish, J. Clin. Investig. 115 (2005) 1532–1541.
D. Eliott, G. Threadgold, G. Harden, D. Ware, B. Mortimore, B. Mortimer, G. Kerry, [89] A. Donovan, C.A. Lima, J.L. Pinkus, G.S. Pinkus, L.I. Zon, S. Robine, N.C. Andrews,
P. Heath, B. Phillimore, A. Tracey, N. Corby, M. Dunn, C. Johnson, J. Wood, The iron exporter ferroportin/Slc40a1 is essential for iron homeostasis, Cell Metab.
S. Clark, S. Pelan, G. Griffiths, M. Smith, R. Glithero, P. Howden, N. Barker, 1 (2005) 191–200.
C. Stevens, J. Harley, K. Holt, G. Panagiotidis, J. Lovell, H. Beasley, C. Henderson, [90] P.G. Fraenkel, Y. Gibert, J.L. Holzheimer, V.J. Lattanzi, S.F. Burnett, K.A. Dooley,
D. Gordon, K. Auger, D. Wright, J. Collins, C. Raisen, L. Dyer, K. Leung, R.A. Wingert, L.I. Zon, Transferrin-a modulates hepcidin expression in zebrafish
L. Robertson, K. Ambridge, D. Leongamornlert, S. McGuire, R. Gilderthorp, embryos, Blood 113 (2009) 2843–2850.
C. Griffiths, D. Manthravadi, S. Nichol, G. Barker, S. Whitehead, M. Kay, J. Brown, [91] S. Goldwurm, C. Casati, N. Venturi, S. Strada, P. Santambrogio, S. Indraccolo,
C. Murnane, E. Gray, M. Humphries, N. Sycamore, D. Barker, D. Saunders, J. Wallis, P. Arosio, M. Cazzola, A. Piperno, G. Masera, A. Blondi, Biochemical and genetic
A. Babbage, S. Hammond, M. Mashreghi-Mohammadi, L. Barr, S. Martin, P. Wray, defects underlying human congenital hypotransferrinemia, Hematol. J. 1 (2000)
A. Ellington, N. Matthews, M. Ellwood, R. Woodmansey, G. Clark, J.D. Cooper, 390–398.
J. Cooper, A. Tromans, D. Grafham, C. Skuce, R. Pandian, R. Andrews, E. Harrison, [92] R.A. Wingert, A. Brownlie, J.L. Galloway, K. Dooley, P. Fraenkel, J.L. Axe,
A. Kimberley, J. Garnett, N. Fosker, R. Hall, P. Garner, D. Kelly, C. Bird, S. Palmer, A.J. Davidson, B. Barut, L. Noriega, X. Sheng, Y. Zhou, L.I. Zon, The chianti zeb-
I. Gehring, A. Berger, C.M. Dooley, Z. Ersan-Ürün, C. Eser, H. Geiger, M. Geisler, rafish mutant provides a model for erythroid-specific disruption of transferrin re-
L. Karotki, A. Kirn, J. Konantz, M. Konantz, M. Oberländer, S. Rudolph-Geiger, ceptor 1, Development (Cambridge, England) 131 (2004) 6225–6235.
M. Teucke, K. Osoegawa, B. Zhu, A. Rapp, S. Widaa, C. Langford, F. Yang, [93] J.H. Postlethwait, Y.-L.L. Yan, M.A. Gates, S. Horne, A. Amores, A. Brownlie,
N.P. Carter, J. Harrow, Z. Ning, J. Herrero, S.M.J. Searle, A. Enright, R. Geisler, A. Donovan, E.S. Egan, A. Force, Z. Gong, C. Goutel, A. Fritz, R. Kelsh, E. Knapik,
R.H.A. Plasterk, C. Lee, M. Westerfield, P.J. de Jong, L.I. Zon, J.H. Postlethwait, E. Liao, B. Paw, D. Ransom, A. Singer, M. Thomson, T.S. Abduljabbar, P. Yelick,
C. Nüsslein-Volhard, T.J.P. Hubbard, H. Roest Crollius, J. Rogers, D.L. Stemple, D. Beier, J.-S.S. Joly, D. Larhammar, F. Rosa, M. Westerfield, L.I. Zon, S.L. Johnson,
S. Begum, C. Lloyd, C. Lanz, G. Raddatz, S.C. Schuster, The zebrafish reference W.S. Talbot, Vertebrate genome evolution and the zebrafish gene map, Nat. Genet.
genome sequence and its relationship to the human genome, Nature 496 (2013) 18 (1998) 345–349.
498–503. [94] A. Amores, A. Force, Y.-L. Yan, L. Joly, C. Amemiya, A. Fritz, R.K. Ho, J. Langeland,
[69] W. Driever, L. Solnica-Krezel, A.F. Schier, S.C. Neuhauss, J. Malicki, D.L. Stemple, V. Prince, Y.-L. Wang, M. Westerfield, M. Ekker, J.H. Postlethwait, Zebrafish hox
D.Y. Stainier, F. Zwartkruis, S. Abdelilah, Z. Rangini, J. Belak, C. Boggs, A genetic clusters and vertebrate genome evolution, Science 282 (1998) (1711 LP – 1714).
screen for mutations affecting embryogenesis in zebrafish, Development 123 (1996) [95] J.E. Levy, O. Jin, Y. Fujiwara, F. Kuo, N.C. Andrews, Transferrin receptor is ne-
(37 LP – 46). cessary for development of erythrocytes and the nervous system, Nat. Genet. 21
[70] S. Sumanas, J.D. Larson, Morpholino phosphorodiamidate oligonucleotides in (1999) 396–399.

211
J. Zhang, I. Hamza Molecular Genetics and Metabolism 128 (2019) 204–212

[96] S. Mercurio, S. Petrillo, D. Chiabrando, Z.I. Bassi, D. Gays, A. Camporeale, K. Bishop, M.L. Calicchio, A. Lapierre, D.M. Ward, P. Liu, M.D. Fleming, I. Hamza,
A. Vacaru, B. Miniscalco, G. Valperga, L. Silengo, F. Altruda, M.H. Baron, HRG1 is essential for heme transport from the phagolysosome of macrophages
M.M. Santoro, E. Tolosano, The heme exporter Flvcr1 regulates expansion and during erythrophagocytosis, Cell Metab. 17 (2013) 261–270.
differentiation of committed erythroid progenitors by controlling intracellular [98] C.K. Kaufman, R.M. White, L. Zon, Chemical Genetic Screening in the Zebrafish
heme accumulation, Haematologica 100 (2015) 720–729. Embryo Nature Protocols, 4 (2009), pp. 1422–1432.
[97] C. White, X. Yuan, P.J. Schmidt, E. Bresciani, T.K. Samuel, D. Campagna, C. Hall,

212

You might also like