understanding of expansive erythropoiesis

219
The Pennsylvania State University The Graduate School Department of Biochemistry and Molecular Biology THE flexed-tail (f) MUTANT MOUSE: ADVANCING THE UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS A Thesis in Biochemistry, Microbiology and Molecular Biology by L. E. Lenox © 2005 L. E. Lenox Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy August 2005

Upload: others

Post on 13-Apr-2022

9 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

The Pennsylvania State University

The Graduate School

Department of Biochemistry and Molecular Biology

THE flexed-tail (f) MUTANT MOUSE: ADVANCING THE

UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

A Thesis in

Biochemistry, Microbiology and Molecular Biology

by

L. E. Lenox

© 2005 L. E. Lenox

Submitted in Partial Fulfillment of the Requirements

for the Degree of

Doctor of Philosophy

August 2005

Page 2: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

The thesis of L. E. Lenox was reviewed and approved* by the following:

Robert F. Paulson Associate Professor of Veterinary Science Thesis Advisor Chair of Committee Avery August Associate Professor of Immunology

Pamela H. Correll Associate Professor of Immunology

Ross C. Hardison Professor of Biochemistry and Molecular Biology

Andrew Henderson Associate Professor of Veterinary Science

Robert A. Schlegel Professor of Biochemistry and Molecular Biology Head of the Department of Biochemistry and Molecular Biology

*Signatures are on file in the Graduate School

Page 3: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

iii

ABSTRACT

The autosomal, recessive flexed-tail mutant (f/f) mouse has defects at times of

expansive erythropoiesis. The phenotype is evident both during development and in the

adult. f/f show a transient fetal/neonatal anemia that remits shortly after birth when the

main site of hematopoiesis has shifted from the fetal liver to the bone marrow. Adult

mice appear normal, but show a delay in their recovery to acute anemia. Analysis of the

flexed-tail (f) mutant has shown that the contribution from stress erythroid progenitors

resident in the spleen responding to hypoxia-induced BMP4/Madh5 dependent signals is

required for the rapid recovery to an acute anemia. The f mutation is a neomorphic

mutation of Madh5, with aberrantly spliced transcripts disrupting the normal BMP4

signaling pathway in the spleen following an erythropoietic challenge. Although this

splenic contribution to an acute anemia is critical for the rapid return to homeostasis, it is

not essential since flexed-tail mice are viable and both humans and mice can survive

without a spleen. To further understand the mechanisms of expansive erythropoiesis, we

have extended our analysis to splenectomized mice. These mice show altered kinetics of

recovery to a phenylhydrazine induced acute anemia with expansive erythropoiesis now

seen in the liver. Further, BMP4 is expressed in the liver and liver erythroid progenitors

exhibit properties similar to stress BFU-E in the spleen. This work has shown the

important role of the BMP4 pathway for regulating expansive erythropoiesis in extra-

medullary organs. It has also broadened our appreciation for splicing mutations, the

regulations within signaling pathways, and the contribution these pathways make to

signaling networks.

Page 4: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

iv

TABLE OF CONTENTS LIST OF FIGURES…......................................................................................................vi ACKNOWLEDGEMENTS..............................................................................................x Chapter 1. HEMATOPOIESIS, ERYTHROPOIESIS AND THE flexed-tail (f) MUTATION...............................................................................................1 Abstract……............................................................................................................1 Introduction to Hematopoiesis and the Development of the Hematopoietic System.................................................................................................................2 The development of the hematopoietic system……....................................3 Primitive and definitive hematopoiesis……................................................5 Changing sites of hematopoiesis throughout ontogeny…….......................6

The extra-embryonic location for the production/development of hematopoietic progenitor cells…............................................................7 The intra-embryonic location for the production of hematopoietic progenitor cells......................................................................................10 Migration of hematopoietic cells from microenvironments supporting their production to those specialized for their expansion.....................12

Introduction to Erythropoiesis...............................................................................15 Erythropoietin............................................................................................18 Regulation of erythropoietin......................................................................19 Erythropoietin signaling.............................................................................22 Hemoglobin................................................................................................24 Steady-state vs. Expansive Erythropoiesis.............................................................28 Microenvironments/Stroma Supporting Hematopoiesis........................................35 The flexed-tail (f) Mutant Mouse as a Means to Study Expansive Erythropoiesis...................................................................................................41 The embryonic defects of flexed-tail (f) mice…........................................41

The adult defects of flexed-tail (f) mice….................................................48 References..............................................................................................................55 Figures....................................................................................................................67 Chapter 2. MAPPING OF THE flexed-tail (f) LOCUS LEADS TO THE DISCOVERY THAT BMP4 AND Madh5 REGULATE THE ERYTHROID RESPONSE TO ACUTE ANEMIA.............................81 Forward..................................................................................................................81 Abstract..................................................................................................................83 Introduction............................................................................................................84 Methods..................................................................................................................87 Results....................................................................................................................89 Discussion..............................................................................................................99

Page 5: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

v

References............................................................................................................103

Figures..................................................................................................................106 Chapter 3. STRESS ERYTHROPOIESIS IN SPLENECTOMIZED MICE.........113 Abstract................................................................................................................113 Introduction..........................................................................................................115 Methods................................................................................................................118 Results..................................................................................................................120 Discussion............................................................................................................129 References............................................................................................................135 Figures..................................................................................................................137 Chapter 4. THE IMPACT OF flexed-tail...................................................................145 Abstract................................................................................................................145 The Impact of Our Analysis of the flexed-tail Locus on Our Appreciation for Splicing Mutations and Their Physiological Consequences...........................146

The Impact of Our Analysis of the flexed-tail Locus as It Relates to the TGF-β Family Signaling Pathway..................................................................149

TGF-β-BMP signaling.............................................................................149 Negative regulators of TGF-β/BMP signaling…....................................153 Interaction of TGF-β/BMP signaling with other signaling pathway.......155 Neomorphic properties of the flexed-tail truncated transcripts...............156

The Impact of Our Analysis of the flexed-tail Locus: A Reminder of the Subtleties in Place Behind Beautifully Orchestrated Physiological Mechanisms….................................................................................................163

Concluding Remarks............................................................................................166 References............................................................................................................167 Figures..................................................................................................................173 Appendix A. CLONING AND CHARACTERIZATION OF THE flexed-tail (f) LOCUS....................................................................................................183 References............................................................................................................192

Figures…..............................................................................................................193 Appendix B. SUPPLEMENTARY INFORMATION: GENOTYPING OF THE Sideroflexin (Sfxn) LOCUS...................................................................205

Page 6: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

vi

LIST OF FIGURES

Figure 1-1 The hematopoietic system 67

Figure 1-2 Dual stem cell model of hematopoietic stem cell ontogeny 68

Figure 1-3 Development and features of the ‘hemogenic’ regions during embryogenesis

69

Figure 1-4 Stages of erythroid development 70

Figure 1-5 Erythropoietin: regulation and signaling 71

Figure 1-6 Schematic of globin genes 72

Figure 1-7 CFU-E in the fetal liver of f/f and +/+ mice 73

Figure 1-8 Percentages of siderocytes in reticulocytes of f/f and f/+ control fetal and neonatal mice.

74

Figure 1-9 Diagram of hemoglobin biosynthesis 75

Figure 1-10 Chromatography of globin chains synthesized by day 18 +/+ (A) and f/f (B) fetal erythrocytes

76

Figure 1-11 Incorporation of 59Fe into heme from CFU-S of f/f and +/+ mice 77

Figure 1-12 Changes in spleen weight and reticulocyte count in flexed and wild-type recovering from phenylhydrazine induced acute anemia

78

Figure 1-13 Effects of phenylhydrazine treatment on enzymes of the hemoglobin biosynthetic pathway in spleen from flexed and wild-type mice

79

Figure 1-14 Transient endogenous spleen colonies (TE-CFU) in the spleens of wild-type and flexed-tail mice

80

Figure 2-1

Analysis of liver BFU-E expansion during the recovery from a PHZ induced hemolytic anemia

106

Figure 2-2 Genetic linkage map of the f locus and molecular analysis of Madh5 transcripts in f/f, f/+ and control mice

107

Page 7: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

vii

Figure 2-3 Analysis of BMP4 expression during the recovery from acute anemia

108

Figure 2-4 Analysis of the ability of BMP4 to induce the formation of stress BFU-E in spleen cells from untreated mice

109

Figure 2-5 Analysis of the recovery from acute anemia in f/Madh5- and +/Madh5- mice

110

Figure 2-6 Analysis of the effect of over-expression of the f mis-spliced Madh5 mRNAs on BMP4 signaling in W-20-17 osteoblast cells

111

Figure 2-7 Identification of the sub-population of progenitor cells from untreated spleen that respond to BMP4

112

Figure 3-1 Hematocrit values during the recovery to a phenylhydrazine (PHZ) induced hemolytic anemia

137

Figure 3-2 Analysis of BFU-E expansion of bone marrow BFU-E during the recovery from PHZ induced acute hemolytic anemia

138

Figure 3-3 H&E stained liver sections during recovery from a PHZ induced acute anemia

139

Figure 3-4 Analysis of liver BFU-E expansion during the recovery from a PHZ induced hemolytic anemia

140

Figure 3-5 Analysis of BMP4 expression in the liver during recovery from a PHZ induced acute anemia

141

Figure 3-6 H&E stained liver sections from flexed-tail (f/f) mice 142

Figure 3-7 Immunohistochemistry for SDF-1 in liver of wild-type mice during the recovery from a PHZ induced acute anemia

143

Figure 3-8 Immunohistochemistry for SDF-1 in liver of splenectomized mice during the recovery from a PHZ induced acute anemia

144

Figure 4-1 Calculation of 5’ splice site consensus sequence 173

Figure 4-2 Cascade of BMP signaling and levels of modulation 174

Figure 4-3 Signaling specificity in the TGF-β superfamily 175

Page 8: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

viii

Figure 4-4 Activation of the type I receptor kinase and recognition of R-Smads

176

Figure 4-5 Structure of the receptor activated R-Smads, common binding partner Smad4 and the inhibitory I-Smads

177

Figure 4-6 Map of full length Madh5 (Smad5) mRNA as well as the truncated transcripts found in f/f mice

178

Figure 4-7 RT-PCR from COS7 cells transfected with retroviral constructs containing full length and truncated Smad5 constructs

179

Figure 4-8 Transfected COS7 cells with constructs containing truncated transcripts

179

Figure 4-9 Predicted in-frame amino acid structure of truncated transcripts compared to wild-type Smad5

180

Figure 4-10 Gene expression of W-20-17 cells lines containing the f/f truncated transcripts

181

Figure 4-11 Current working model for the network regulating expansive erythropoiesis

182

Figure A-1 Physical map of Mus musculus Chr13 around region of Madh5 (Smad5)

193

Figure A-2 Strategy to find the flexed-tail (f) mutation 194

Figure A-3 Analysis for Madh5 (Smad5) on BAC clones 195

Figure A-4 H&E stained sections of the dorsal aorta from E10.5 f/f and f/+ littermates

196

Figure A-5 Northern blot for Madh5 (Smad5) mRNA levels in spleen of f/+ and f/f adult mice following an acute anemia

197

Figure A-6

Western blot for Madh5 (Smad5) protein levels in f/f, f/+ and +/+ mice

198

Page 9: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

ix

Figure A-7 Short-term rescue of f/f mutant spleen cells through retroviral transduction of full length Madh5 (Smad5) cDNA

199

Figure A-8 Long-term rescue of flexed-tail mutants: Madh5 transgenic f/f mice 200

Figure A-9 Genome Walking results of Madh5 genomic region from f/f, +/+ and E10 BAC

201

Figure A-10 Map of Madh5 subdivided for direct amplification, cloning and sequencing

202

Figure A-11

Secondary structure comparison between the difficult to amplify region of “Alaska” and the easily amplified comparable sized fragment of “Bombay” of the Madh5 gene

202

Figure A-12 Number of Ts in PolyT region (in Greece fragment) from flexed-tail and wild-type mice

203

Figure A-13 In vitro splicing assays to determine the functional consequences of variations in the PolyT tract in flexed-tail and wild-type mice

204

Figure B-1 Genotyping for the Sideroflexin (Sfxn) mutation 206

Figure B-2 Direct sequencing of the Sfxn1 exon 2 in f/f mice 207

Figure B-3 Analysis of Sfxn mutations in f/f mice by oligonucleotide ligation assay

208

Page 10: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

xACKNOWLEDGEMENTS

I would like to thank my family first and foremost for their love and support

throughout this process. I would not be where I am today without them. I would like to

thank my advisor, Dr. Robert Paulson, for giving me the opportunity to work in his lab.

He creates an environment conducive to scientific growth and development, built on

discussion, collaboration and curiosity. I am very thankful to the past and present

members of the Paulson lab, especially Anamaria Craici, Omid Harandi, Shailaja Hegde,

Andrew Lariviere, John Perry, Prashanth Porayette, Aparna Subramanian and Michele

Yon, who made it worth coming to work when the data didn’t. I appreciate the input

from and participation of my committee members which include Dr. Avery August, Dr.

Pamela Correll, Dr. Ross Hardison, and Dr. Andrew Henderson. Finally, I would like to

express my sincere gratitude to my friends and loved ones who have touched my life,

made things interesting, and kept me grounded on this journey. I look forward to using

what I have learned during this process to go on and make a difference in the world.

Page 11: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

Chapter 1

HEMATOPOIESIS, ERYTHROPOIESIS AND THE flexed-tail (f) MUTATION

Abstract

Hematopoiesis is the process of blood cell development. Erythropoiesis is the

branch of hematopoiesis that leads to the development of an erythrocyte. There are

distinct mechanisms of erythropoiesis and specialized microenvironments that support

these processes. Steady-state erythropoiesis occurs at a constant rate in the bone marrow

and relies on local sources of erythropoietin (Epo). Expansive erythropoiesis is a distinct

mechanism which occurs in the fetal liver during embryogenesis or in the adult spleen

following an erythropoietic stress. Expansive erythropoiesis leads to the rapid

expansion of erythroid progenitors and requires high levels of erythropoietin regulated by

hypoxia in the tissues. The flexed-tail (f) mutant mouse is a means to study expansive

erythropoiesis. These mutants have an embryonic/fetal anemia that remits after birth as

the main site of hematopoiesis has shifted to the bone marrow. The f/f reticulocytes

contain non-heme iron granules known as siderocytes. Adult f/f mice exhibit normal

steady state blood values but have a delay in their recovery to an acute anemia. This

return of normal hematocrit levels and the up regulation of heme biosynthetic enzymes

are delayed; however no siderocytic granules are present in the adult reticulocytes during

recovery. This mutant provides an important model for understanding the progenitors,

signals and microenvironments that regulate expansive erythropoiesis.

Page 12: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

2

Introduction to Hematopoiesis and the Development of the Hematopoietic

System

Hematopoiesis is the developmentally regulated and tissue specific process of

blood cell development. It is a highly conserved process in worms, fish, avian,

amphibian, and mammalian systems. Depicted as a hierarchal system cascading from a

pluripotent stem cell, this model has been firmly established through a series of

transplantation and cell fate experiments (Figure 1-1). Hematopoietic stem cells (HSC)

were first identified morphologically by a large nucleus with prominent nucleoli and a

deeply basophilic staining cytoplasms [1]. Currently the defining characteristics of the

hematopoietic stem cell are long-term, high level repopulation of all hematopoietic

lineages in the adult and the ability to self renew [2, 3]. In the murine system they can be

enriched for by their cell surface profile of kit hi, Sca-1 (Ly-6A/E)+ , with no lineage

specific surface expression, Lin- (Mac1-, B220-, CD3-, Gr1-) [3-5].

The HSC will commit to a particular lineage through multi-potent progenitors,

committed progenitors, and precursor cells ending with mature differentiated cells

specialized for a particular function [6]. There are erythrocytes for oxygen delivery,

myeloid cells (neutrophils and macrophages) to fight infection, mast and myeloid cells

which secrete histamines in an allergic response, megakaryocytes for blood clotting, and

B and T lymphocytes that are critical for adaptive immune function. The maturation

down a particular lineage involves the interaction with other cells, cytokines for

instructive or permissive signals, and cell intrinsic mechanisms [6]. Our understanding

of this developmental program has been complemented by work in embryonic stem (ES)

cells. ES cells are pluripotent cells from the inner cell mass of the mouse E3.5 blastocyst.

Page 13: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

3

When cultured in media lacking leukemia inhibitory factor (LIF), they can give rise to

cell aggregates termed embryoid bodies (EB). The EBs can be manipulated to

differentiate into the various ranges of embryonic tissues including epidermis, neuronal

and glial cells, muscle, endothelial and hematopoietic cells using specific cytokine

cocktails [7]. This system has added to our understanding of the origins of hematopoietic

potential, the relationship between lineages, and the “molecular definitions” of

progenitors and their progeny. For the most part, progenitor cells are still defined using

in vivo or in vitro colony assays, although advances have been made in distinguishing

them based on cell surface markers or gene expression profiles. As progenitors commit

to a lineage, they progressively gain morphological distinguishable characteristics, but

lose self-renewal capabilities. Mature hematopoietic cells have finite life spans ranging

from 2 days for neutrophils, to 30 days for erythrocytes, while the life of lymphoid cells

varies [4]. To maintain steady-state levels, the system must produce 1010 cells per day

to keep pace with normal losses [6]. It must also be able to adjust for large perturbations

in homeostasis such as blood loss or infection by invoking developmental pathways of

certain lineages, while leaving others unaffected.

The development of the hematopoietic system. Hematopoietic tissue is derived from

ventral mesoderm through the instructive signals of fibroblast growth factor (FGF),

transforming growth factor-β (TGF-β), bone morphogenic protein (BMP) and other

factors [1, 8]. BMP4 is known to be induced by and mediate the signals of Indian hedge

hog (Ihh), which is sufficient to reprogram anterior ectoderm and induce primitive

erythropoiesis in murine embryonic explants [9]. BMPs have been shown to be necessary

Page 14: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

4

following early specification of ventral mesoderm, by directly or indirectly targeting

critical hematopoietic transcription factors such as SCL, GATA-1, GATA-2, LMO-2, and

EKLF [10]. In fact, SCL is necessary and sufficient to specify hematopoietic mesoderm

[11]. SCL -/- mice are embryonic lethal at E9.5 due to a failure to develop embryonic

red cells [12, 13] and SCL -/- ES cells make no contributions to hematopoietic lineages in

chimeras [7]. Its been shown that the transcription factors SCL, LMO-2 and GATA-1

form a complex in vitro, specifying mesoderm to become blood, with over-expression of

the complex leading to embryos that are ventralized with blood throughout the dorsal

ventral axis [14].

A dual stem cell model has been proposed (Figure 1-2) whereby mesoderm

precursors migrate to both extra and intra-embryonic sites, with hematopoietic stem cell

activity arising independently at both locations [1]. It is clear from the anatomy of

hematopoietic development there is a close relationship between hematopoietic cells and

endothelial cells. In 1932, Murray et al. proposed the idea of a common ancestor to both

the hematopoietic and endothelial lineages termed the hemangioblast [15]. The initial

observation was made that both endothelial cells and hematopoietic cells emerge from a

cluster of identical cells from this mesodermal layer. The idea of this common ancestor

is supported by the close proximity of these cells as they develop, the similarity of their

cell surface markers [7], and gene expression profile with expression of CD34, Flk1,

TIE2, SCL, and GATA2 [3, 16]. Keller and colleagues have used the ES cell

differentiation system to identify a candidate for the hemangioblast. They found a

transient Flk1+ (VEGF receptor) population, termed the blast colony forming cell (Bl-

CFC) that in functional studies could give rise to both primitive and definitive

Page 15: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

5

hematopoietic precursors, as well as adhesive cells expressing an endothelial marker

(PECAM) [17, 18]. The important role of Flk1 expression for an

endothelial/hematopoietic ancestor is verified by in vivo evidence that Flk1-/- mice die

between E8.5-9.5 with no yolk sac blood islands, no organized blood vessels in the

embryo or yolk sac, with severely reduced hematopoietic progeny [19, 20]. This system

has also emphasized the importance of SCL for hematopoietic development since SCL -/-

ES cells can differentiate into Flk1+ mesoderm and down the endothelial lineage, but are

blocked in hematopoiesis at the hemangioblast stage [7].

Primitive and definitive hematopoiesis. The hematopoietic system occurs in two waves

that are defined by the morphology of the progeny, potential of the progenitor cells, and

the type of globins produced. Primitive hematopoiesis is a transient wave during

ontogeny that is replaced permanently by definitive hematopoiesis which is maintained

throughout the life of the animal. Primitive hematopoiesis shares characteristics with

lower vertebrates and amphibians such as large nucleated red blood cells. These

primitive cells produce only embryonic globins. The progeny have limited potential,

primarily producing primitive erythrocytes and macrophages in vivo, although other

lineages can be obtained using in vitro colony assays [4]. The onset of definitive

hematopoiesis occurs later in ontogeny and is characterized by enucleated erythrocytes,

fetal (in humans) and adult globin production, and progenitors with full potential, giving

rise to all blood cell lineages [4, 6]. The mechanisms that drive primitive erythropoiesis

are clearly distinct from those of definitive erythropoiesis. Definitive erythropoiesis is

much more dependent on Erythropoietin (Epo) than primitive, even though Epo increased

Page 16: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

6

the number of primitive erythroid cells and accumulation of globin transcripts in vitro [1].

Disruption of Epo, Epo receptor (EpoR) or Jak2, the downstream signaling kinases of the

receptor complex, almost completely abrogated definitive erythropoiesis in the fetal liver,

with only mild effects on primitive erythropoiesis [21-23]. Aside from the different

sensitivities to cytokines, there are distinct differences in transcriptional regulation of

primitive and definitive hematopoiesis. Although there are mutations shown to alter both

primitive and definitive hematopoiesis such as SCL/Tal1, LMO2 and Flk1, disruption of

the transcription factors Myb, AML-1 (Runx1), CBFβ, cause severe abnormalities in

definitive hematopoiesis, leaving yolk sac (primitive) hematopoiesis phenotypically

normal [4].

Changing sites of hematopoiesis throughout ontogeny. An interesting phenomenon of

the hematopoietic system is how hematopoietic locations change throughout ontogeny.

The properties of the system emphasize how microenvironments are specialized for

particular functions, establishing the role of a particular cell, the potential of that cell, and

whether that potential will be reached. In vivo, hematopoietic cells develop de novo in

both an extra-embryonic (yolk sac) [24] and intra-embryonic locations (P-Sp/AGM

region) (Figure 1-3) [25, 26]. These “hemogenic” areas share similar features since both

are composed of splanchnopleura (mesoderm against an endoderm layer), with no

hematopoietic potential found in somatopleura (mesoderm against ectoderm) [16, 27, 28].

The interactions of the mesodermal and endodermal layer as well as factors secreted by

these tissues are important for the development of both hematopoietic and endothelial

cells [29]. Although these hemogenic sites offer the microenvironment necessary for the

Page 17: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

7

formation of hematopoietic progenitors, only the yolk sac provides conditions supportive

of their development. The AGM environment, which develops from the para-aortic

splanchnopleura (P-Sp), is not optimal for their differentiation or expansion [4]. Once

circulation has been established (28-32 somite stage) and hematopoietic locations are no

longer isolated, cells will migrate to and colonize microenvironments specialized for

hematopoietic cell differentiation and expansion [3, 28]. This observation has added an

extra level of complexity to the debate in the field over the origins of stem cells

contributing to the overall workings of the adult hematopoietic system.

The extra-embryonic location for the production/development of hematopoietic

progenitor cells. The earliest blood cells emerge in the extra-embryonic mesoderm of

the avian and murine yolk sac (YS), and at analogous sites within the ventral blood island

(VBI) of Xenopus, or the neutral intermediate cell mass (ICM) associated with the tail in

bony fish [10]. In mice these blood islands form at mid/late primitive streak stage

(embryonic day 7.0, [E7.0]) [30], within 12 hrs of the start of mesoderm formation [1].

The blood islands start as a compact group of identical cells. External cells begin to

develop an endothelial morphology, while internal cells begin to lose their connections

and differentiate into primitive erythrocytes [16, 31]. These large, nucleated primitive

erythroid cells sustain the embryo until E12.5 when definitive enucleated cells enter

circulation from the fetal liver [31]. While primitive erythropoiesis constitutes essentially

all hematopoiesis in the early yolk sac (YS), small numbers of macrophage and

megakaryocyte progenitors and bi-potential (Ery and Mac at E7.5) can be found using in

vitro assays [32]. Before the onset of circulation high proliferative potential colony

Page 18: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

8

forming-cells (HPP-CFC), a quiescent multipotential cell of the adult bone marrow, are

found exclusively in the YS (E9.0) [31]. The HPP-CFC is the earliest multipotential

precursor that can be cultured in vitro without stromal support, capable of giving rise to

secondary HPP-CFC and other multipotential or unilineage progenitor cells in replating

assays. However, YS progenitors have limited potential as we see no lymphoid

progenitors present until after circulation has been established (E11-12) [25, 26, 33].

The YS does have definitive potential, which is consistent with what is seen in the

amphibian where the VBI does contribute to some extent to definitive hematopoiesis

[34]. At E8.25 definitive hematopoietic progenitors (large BFU-E producing adult

globins), and by E8.5 multipotential precursors (macrophages and granulocytes) are

detectible by in vitro colony assays, although at this time only embryonic red cells are

produced in vivo [30, 31]. Around the establishment of circulation (E8.5-9), myeloid

and B lymphoid progenitors arise with colonies composed of erythroid cells,

macrophages, granulocytes, mast cells, and megakaryocytes, identified by E9. Short term

repopulating cells, CFU-S can be detected in this region by E9.5 [35, 36]. Although

harboring progenitor cells with definitive multi-lineage potential, the YS

microenvironment does not appear sufficient to fully support the expansion and

differentiation of erythroid progenitors [37]. Further, precursors in the yolk sac do not

display extensive self-renewal since they are progressively replaced by a new population

of progenitors from an intra-embryonic location [16]. Using chick-chick and chick-quail

chimeric models, a number of investigators have failed to observe yolk sac contributions

to adult myelopoiesis or lymphopoiesis [28, 38, 39].

The YS does not contain a hematopoietic stem cell population capable of

Page 19: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

9

repopulating an irradiated adult until E11 when HSCs are already present in the blood

stream, AGM region, vitelline and umbilical arteries [33, 40-42]. Some have suggested

that this is due to the inability of the YS hematopoietic cells to home and engraft to the

adult microenvironment [28]. While others speculate it is due to the low level of MHI

class I on its surface, making it a prime target for destruction by NK cells [4]. Even

though no adult HSC is present in the YS prior to E11 the idea has emerged that there is

an “embryonic” hematopoietic stem cell present prior to this time. The embryonic stem

cells can be distinguished from the adult HSC by a slightly different surface expression

profile of Sca1lo, MAC1+, Kit+ and AA4.1antigen+ [2, 43, 44]. This embryonic HSC

has the potential to repopulate an adult, but it does not reach this potential until it has

been primed by an adequate microenvironment. Both the AGM stroma and the fetal liver

have been shown to induce the development of HSCs capable of repopulating adults.

AGM-S3 is a clonal endothelial cell line originally isolated from murine E10.5 AGM

region that can support primitive murine and human hematopoiesis in vitro. This line

provides factors that promote the generation of CFU-S (short-term repopulating) and

HSC (long-term repopulating) cells, and their ability to correctly home in an adult

environment. Prior to an established circulatory system, cells from E8-8.5 YS (and the P-

Sp which will become the AGM region, but at this stage has not generated true HSCs)

co-cultured on the AGM-S3 stromal line acquire long-term repopulating ability for

lethally irradiated adults [45]. The importance of the fetal liver environment for the

specification of definitive characteristics to embryonic cells has been demonstrated both

in vitro and in vivo. YS explants from 20-25 somite pairs (sp) reveal a transient wave of

erythroid cells that express embryonic and adult globins. Coculture with liver

Page 20: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

10

primordium led to a subsequent wave of YS erythroid cells that express adult globins

[37]. Yoder et al. demonstrated that the full potential of the embryonic HSC (YS HSC) is

dependent on the age of the recipient of the cells. They took advantage of the fact that

the fetal liver of newborn mice continues as a hematopoietic organ for several weeks after

birth. Pregnant dams (at E18 of gestation) were given a subcutaneous busulfan injection,

causing pups to be born in a sublethally myeoablated state. Injection of these conditioned

pups with E8-9 YS cells leads to the engraftment and contribution of these cells to adult

hematopoiesis. Transplanted cells gave rise to all blood cell lineages long term (>11

months). Bone marrow from these primary YS recipients reconstituted B and T

lymphocytes, granulocytes, and erythrocyte lineages in secondary lethally irradiated adult

mice. Thus embryonic HSCs when injected into a fetal/newborn environment are able to

acquire the necessary characteristics to repopulate the adult hematopoietic environment,

an attribute that is missed when directly injected into adults [46, 47].

The intra-embryonic location for the production of hematopoietic progenitor cells.

Although embryonic HSC can be “primed” for the adult environment, the chimeric

models have shown that it was an intra-embryonic location that contributes to adult

hematopoiesis [35, 38, 48]. In amphibians it is the dorsal lateral plate (DLP) that

contributes to adult definitive hematopoiesis [34], as is seen the corresponding region of

the murine system [49]. This intra-embryonic hemogenic tissue initiates as the visceral

para-aorta splanchnopleura (P-Sp) which will go on to form the dorsal aorta, gonads and

mesonephros (AGM region) following organogenesis [49]. Hematopoietic precursors are

found in the P-Sp/AGM region from E8.0-13, peaking about E10.5-11 [4]. At about the

Page 21: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

11

time definitive hematopoietic cells are present in YS, they are independently formed in

the P-Sp [1]. In vitro studies have demonstrated the presence of multipotent progenitors

(lymphoid and myeloid) in the P-Sp/AGM E8-10 embryos, which is the first time we see

progenitors with lymphoid potential [16, 49]. Prior to the establishment of the

circulatory system (which begins at the 8-10 somite pair stage: E8.5-9), multipotent

lymphomyeloid progenitors can be found in this region, although they are incapable of

reconstituting adult recipients [25, 26]. The P-Sp at this stage does appear to have

embryonic hematopoietic stem cell potential. Yoder and colleagues looked at

CD34+ckit+ cells from E9 YS and P-SP and found equivalent multilineage repopulating

ability into conditioned newborn recipient. Once circulation has been established, colony

forming cells capable of producing progeny of the erythroid, macrophage, granulocyte,

mast and megakaryocyte lineages were observed in the P-Sp/AGM region, as was the

case in the YS [50].

In addition to early embryonic HSC potential, it is in this intra-embryonic location

that “true” HSCs emerge which can be directly injected into lethally irradiated adult

recipients and contribute fully to adult hematopoiesis by re-establishing all blood lineages

in the long term [16, 25, 33, 35]. The development coincides with organogenesis which

leads to the formation of the AGM region (aorta gonad mesonephros region) from the

once P-Sp. At 37 somite pairs (E10.5-11.5), along the ventral wall of the dorsal aorta

(and other major vessels such as the vitelline and umbilical arteries) the first definitive

adult HSCs emerge as budding clusters from the endothelial wall [16, 33, 35, 40, 51].

Budding hematopoietic clusters from the ventral wall of the dorsal aorta is a phenomenon

seen not only in mammals, but also birds, zebrafish, and amphibian embryos [4]. The

Page 22: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

12

underlying mesenchyme secretes factors such as BMP4, which is critical for this process

[52, 53]. The transcription factor AML1/Cbfa2/Runx1 is expressed in the AGM (in

definitive hematopoietic progenitor cells and in endothelial cells from which these

hematopoietic cells are thought to emerge) [54] at the time the first HSC develop and

influences the temporal and spatial appearance of adult repopulating HSCs [55]. In fact,

Runx1 is required for the switch between primitive and definitive hematopoiesis. Runx1

-/- mice initiate yolk sac hematopoiesis, but die at E11-12.5 with only primitive nucleated

erythroblasts in the liver rudiment and a block in the establishment of definitive

hematopoietic progenitors with an absence of definitive erythroid, myeloid and

megakaryocytic cells [7]. de Bruijn et al. (2002) used GFP expressed from Sca-1 to

demonstrate that the first HSCs are localized within the endothelial cell layer lining the

wall of the dorsal aorta [2]. Around this time the fetal liver begins colonization (28-32

somite stage; E10-10.5) [1]. By E12, both the YS and FL now possess cells with HSC

activity, which have most likely traveled there from the AGM region through the

circulation [4].

Migration of hematopoietic cells from microenvironments supporting their

production to those specialized for their expansion. Once circulation has been

established, progenitors can migrate from environments designed for the production of

hematopoietic cells, to those specialized for their differentiation and expansion. By E12,

primitive hematopoiesis has sharply declined, and the fetal liver is the primary site of

definitive hematopoiesis of the developing embryo [28, 56]. The fetal liver does not

possess any de novo hematopoietic potential. When E9.5 fetal liver tissue was engrafted

Page 23: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

13

under the kidney capsule of an adult recipient, no hematopoietic elements were present,

even though the tissue survived. The introduction of hematopoietic cells into the

circulation of the FL tissue recipient resulted in the multi-lineage engraftment in the

implanted fetal tissue [57, 58]. The timing and origin of cells seeding the fetal liver is

controversial. The exact contribution of extra-embryonic (yolk sac) and intra-embryonic

(AGM) progenitors is not clear. Both have been shown to possess cells with definitive

hematopoietic activity in in vitro culture systems, but since circulation has been

established, pinpointing a cell’s origin is difficult. It is highly likely that cells from both

locations contribute to the overall pool of progenitors that home here. β1 and α4

integrins have been shown to be important for the homing of HSCs to the fetal liver

environment, while the CXCR4-SDF1 complex is required for the retention of cells in the

hematopoietic organ once migration has occurred [4]. The fetal liver microenvironment

supports the maturation of all lineages such as myeloid, mast, megakaryocytic and

lymphoid. In fact, it is fine tuned to support expansive erythropoiesis, as this is the

main requirement at this stage of development [56, 59].

Previous studies in humans have shown the unique ability of fetal burst forming

units (BFU-E) to expand in response to Epo alone, not requiring cytokines with burst

promoting activity (BPA) such as GM-CSF or IL-3 [60]. Kit ligand strongly synergized

with Epo to stimulate the growth of these BFU-E [61], and has been shown by others to

be essential for the function of CFU-E progenitors [62] . The critical role of Kit ligand or

its receptor in fetal liver erythropoiesis is evident by severe anemia associated with

mutations of the ligand or its receptor, affecting proliferation but not differentiation of

progenitor cells [63-65]. GM-CSF or IL-3 did not increase the clonogenicity of

Page 24: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

14

embryonic and fetal BFU-E unless Epo concentrations are suboptimal. This phenomenon

is distinct from BM progenitors, where only a small subset respond to Epo alone,

requiring other BPA to generate BFU-E [61]. It has also been shown that fetal

progenitors have a greater proliferative capacity in culture than their adult counterparts

[66, 67] and that fetal liver CFU-E were at least 5 times more sensitive to Epo than BM

CFU-E [67].

Although the fetal liver is the central point of hematopoiesis for the rapidly

growing embryo, by E18 the embryo is reaching full gestation (20 days) and achieving

steady state homeostatic conditions. From the fetal liver, cells will migrate to the adult

hematopoietic centers of the spleen (beginning E12) and bone marrow (by E15-16) [4].

In contrast to humans where the bone marrow provides sufficient expansive capabilities,

mice utilize both the spleen and bone marrow compartments for blood cell production for

the life of the adult [68].

Page 25: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

15

Introduction to Erythropoiesis

Erythropoiesis is the process of red blood cell development. The mature

erythrocyte is a small (8 µm diameter; 2 µm thickness) biconcave disk with a volume of

90 fL. It lacks a nucleus and mitochondria, with no ability to synthesize new proteins.

The role of the erythrocyte is highly specialized for oxygen transport and delivery using

the iron-containing molecule, hemoglobin. The average erythrocyte has a finite lifespan

of 120+/-20 days, with many trips through the microvasculature possible by its highly

resilient, 2 molecule thick membrane, consisting of tightly packed phospholipids. The

proper function and longevity of the erythrocyte is determined not only by the integrity of

its cell membrane, but also by its metabolism. Without a nucleus and mitochondria, the

cell has little ability to metabolize fatty acids or amino acids, and relies almost

exclusively on the break down of glucose for its energy requirements [69].

Historically, the progenitors of the erythroid lineage are defined by in vivo or in

vitro colony assays. The erythroid lineage is a branch of the myeloid lineage with the

earliest myeloid progenitor CFU-GEMM (colony forming unit-granulocyte, erythroid,

macrophage, and megakaryocyte) defined by in vitro colony assays or by an in vivo

colony assay for CFU-S (colony forming unit spleen) (refer to Figure 1-1). Currently,

these multipotential myeloid progenitors and their more lineage restricted progeny can be

enriched using cell surface markers. From both fetal liver and adult bone marrow,

multipotential common myeloid progenitors (CMP) can be isolated [IL7Rα-Lin-c-

Kit+Sca1-FcγRloCD34+], which give rise to either bi-potential megakaryocyte/erythrocyte

progenitors (MEP) [IL7Rα-Lin-c-Kit+Sca1-FcγRloCD34-] or granulocyte/macrophage

progenitors (GMP) [IL7Rα-Lin-c-Kit+Sca1-FcγRhiCD34+] as verified through in vitro

Page 26: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

16

colony assays [70, 71]. The earliest cell committed to the erythroid lineage is known as

the Burst Forming Unit Erythroid, or BFU-E (Figure 1-4). When plated in

methylcellulose media with a high concentration of erythropoietin (2 U/ml) and a factor

with burst promoting activity such as IL-3 [72], these cells produce a large macroscopic

cluster of 4 or more colonies, composed of primarily erythroid cells after 7-10 days in

culture. BFU-E are not actively proliferating (most are in the G0/G1 phase of cell cycle)

[73], but this depends on the strain of mice [74, 75]. In vivo, their proliferation appears

to be controlled by conditioning factors derived from lymphocytes and macrophages,

including stem cell factor (SCF), inerleukin-3 (IL-3) and granulocyte-macrophage

colony-stimulating factor (GM-CSF) [69]. Analysis of mice with null mutations in either

GM-CSF or IL-3 receptor indicates that these factors are not crucial for erythropoiesis.

However, mice deficient in SCF or its receptor (Kit) suffer from severe anemia,

suggesting a requirement of these signaling components for erythropoiesis [62]. The

early BFU-E have few erythropoietin receptors (EpoR), and thus respond minimally to

erythropoietin. As these cells mature, larger numbers of EpoR are expressed and late

BFU-E become more Epo responsive. The BFU-E will differentiate into a late stage

erythroid progenitor termed colony forming unit-erythroid, or CFU-E. These cells are

also defined by an in vitro colony assay, with a single progenitor generating single, small

colonies (8 or more erythroblasts) after 2 days in culture [76]. CFU-E are actively

proliferating, and respond strongly to Epo to stimulate growth and prevent apoptosis [77]

obtained with as little as 0.03 U/mL Epo in culture, which is 50-100 times less than that

to detect the more primitive BFU-E [78]. CFU-E will go through a series of programmed

maturation and differentiation steps leading to mature erythrocytes [69].

Page 27: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

17

Unlike the earlier progenitors, progeny of the CFU-E can be identified by their

morphology under a light microscope. Nucleated red cell precursors, or normoblasts, are

distinguished by their dense nuclear chromatin, lack of cytoplasmic granules, and later by

hemoglobin within the cell cytoplasm. The maturation of the normoblasts is divided into

three phases, early, intermediate and late, based on morphology of cells stained with

Wright’s stain. Along the erythroid developmental progression, there is a reduction in

cell size and mitochondria/RNA content as hemoglobin content increases until

enucleation, when cells are released into circulation. Early stage maturation cells

(pronormoblasts and basophilic normoblasts) are large (300-800fL), with nucleoli

generally not seen until the pronormoblast level. By the intermediate, polychromatic

stage, the nucleus is more compact and hemoglobin is present in the cytoplasm. In the

late maturation stage, eosinophilic or orthochromatic normoblasts have a dense, even

opaque nucleus, and pink cytoplasm from high hemoglobin content. Late stage

erythroblasts cease dividing and accumulate in the G0 phase. The nucleus will be

extruded, along with a degradation of organelle, to give a marrow reticulocyte, which is

still larger than a circulating mature erythrocyte and containing about two-thirds of its

eventual hemoglobin content. This cell gets its name because of the presence of residual

strands of RNA (reticulin). Under normal conditions, the cell is held in the marrow as the

remainder of hemoglobin synthesized and the cell gradually decreases its cell size, RNA

and mitochondria content. Finally, as the cell volume reaches that of a mature cell, it is

released into circulation (blood reticulocyte), with residual RNA for another 24 hrs

before becoming the mature erythrocyte. The total time for the development from the

CFU-E to mature erythrocyte takes about 7 days, but under anemic or hypoxic

Page 28: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

18

conditions, this can be shortened to as little as 5-6 days from the early release of marrow

reticulocytes into circulation, reducing the intermitotic interval or by skipping a mitotic

division [69, 77].

Erythropoietin. The key humoral regulator of erythropoiesis is the 34kDa glycoprotein

Erythropoietin (Epo) [79]. The primary role of this member of the cytokine superfamily

which includes thrombopoietin (Tpo), granulocyte-specific colony stimulating factor (G-

CSF) and IL-7 [80], is to promote the survival of sensitive progenitors by preventing

apoptosis [77, 81, 82], and to some extent to promote the proliferation and differentiation

of precursors [6, 83]. When added alone to cultures of purified erythroid progenitors,

Epo prevents apoptosis but does little to support cell proliferation (no great increase in

progenitor numbers), or induce differentiation (measured by globin synthesis or

enucleation). It works in concert with other factors such as SCF, required for optimal

progenitor cycling but itself is a poor mediator of cell survival, and insulin-like growth

factor (IGF-1), which is required for optimal erythroid differentiation (determined by

globin synthesis and enucleation) [84]. Epo and Epo receptor (EpoR) knock-out mice die

at E13-15, due to severe anemia from a lack of differentiation of their mature erythroid

progenitors (CFU-E) [21, 23, 85]. The similarity in the knock-out models gives strong

supporting evidence that there is a single receptor for Epo, and a single ligand for the

EpoR [86].

Although the lack of Epo or its receptor is embryonically lethal, they are not

essential for the development of erythroid progenitor cells. Erythroid colony forming

cells, early BFU-E and CFU-E progenitors, can be found in fetal liver cultures of Epo or

Page 29: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

19

EpoR -/- mice by including SCF or Tpo in the culture conditions [21, 23, 85]. Epo/EpoR

is crucial for the proliferation and survival of CFU-E and their terminal differentiation

[81, 84]. The period of Epo dependence is from just before the CFU-E stage through the

basophilic erythroblast stage with the beginning of hemoglobin synthesis [82]. This

window of Epo dependence is mirrored by the EpoR expression on erythroid progenitors,

from a low of 300 receptors/cell on late BFU-E to a high of 1100 receptors on CFU-E

and erythroblasts (refer to Figure 1-4). There are no Epo receptors on reticulocytes or

erythrocytes. EpoR has been shown to be expressed on megakaryocytes, endothelial

cells, heart, ovaries, placenta and brain, with a possible physiological role in cell survival

in these tissues as it been shown to be required for the survival of neuronal cells [87].

Within a population of erythroid progenitors, there exists heterogeneity with respect to

Epo sensitivities [84]. Cell divisions accompanying terminal differentiation are finely

controlled by cell cycle regulators, so when these progenitors are deprived of Epo, they

undergo apoptosis [73].

Regulation of erythropoietin. Developmental, tissue specific and environmental signals

all contribute to the precise regulation of Epo. Epo is expressed primarily by peritubular

interstitial cells in the cortex or the outer medulla of the kidney in adult mice, and also to

a lesser extent in the liver, particularly during fetal and neonatal development. Other

quantitatively less significant sites of production include brain, testis, lung, spleen,

placenta, bone marrow and ovary [79]. There are low basal levels of Epo in the serum,

with production greatly enhanced in response to hypoxia when more cells are recruited to

produce Epo, rather than individual cells increasing the amount of Epo they produce [88].

Page 30: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

20

The peritubular interstitial cells of the kidney detect a decrease in available oxygen

caused by decreased hemoglobin, decreased oxygen content or the higher oxygen

affinity of hemoglobin, all related to tissue oxygenation which is a function of the

number of circulating erythrocytes [69]. Epo production can be elicited in laboratory

animals by inducing hypoxemia, lowering the circulating blood mass by bleeding or PHZ

injection, or divalent metals like cobalt that mimic ferrous (Fe+2) oxide [89]. Regulation

of Epo by hypoxia and other stimuli occur at the transcriptional level [79, 90], and

requires protein synthesis since the translational inhibitor cyclohexamide can block the

hypoxic induction. There is a 50-100 fold increase in Epo mRNA in the human

hepatoma cells line Hep3B under hypoxic conditions [91]. The magnitude of Epo mRNA

induction is proportional to the degree of anemia, with maximum levels reached 4-8

hours after the hypoxic stimuli. There are three non-coding stretches of the Epo gene that

are highly conserved between humans and mouse: the promoter, the first intron and a

120 base pair (bp) region just 3’ to the polyadenylation site [79]. It is this 3’ enhancer

element that is critical for the regulation of Epo by hypoxia (Figure 1-5A) [92-94]. The

hypoxically inducible function of the Epo enhancer is dependent on three defined

regions. Located most 5’ in the enhancer is a response element for HIF-1 (CACGTGCT;

the consensus sequence being TACGTGCT ), which is the primary element that mediates

the transcriptional response to hypoxia [94]. 3’ to the HIF element is a CACA site. No

proteins are known to bind to this site, but mutations in this region abrogate hypoxia

inducible activity of the enhancer [79]. Unless the enhancer element is placed directly

upstream of the promoter, it requires a third site for hypoxically induced transcription

[93]. This is the DR-2 site, a direct repeat of two steroid hormone receptor half sites

Page 31: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

21

separated by 2 bp [92, 95]. The orphan nuclear receptor, HNF-4α has been shown to bind

specifically to this site [95]. This may contribute to the specificity of Epo gene

expression since HNF-4α has an expression profile limited to the renal cortex and liver,

as well as the intestine [79].

Binding of HIF-1 is the critical mediator of the hypoxic induction of Epo [79].

HIF-1 is a heterodimer composed of two subunits, both of which are basic helix-loop-

helix (bHLH) proteins in the PAS (PerAHR-ARNT-SIM) family of transcription factors

[96]. HIF-1β is the previously cloned aryl hydrocarbon receptor nuclear translocator

(ARNT) involved in the transcriptional regulation of genes in response to xenobiotics and

oxidant stress. It is expressed constitutively with mRNA and protein levels not

significantly affected by oxygen tension [79]. It is the HIF-1α subunit that confers the

hypoxic control to the heterodimer. HIF-1α is also expressed constitutively, but the

protein is rapidly degraded (half-life less than 5 minutes) under normoxic conditions [77,

97]. This subunit is only detectable in cells treated with hypoxia or stimuli that mimic

hypoxia (cobalt or iron chelators) [79]. Oxygen is an essential co-substrate for the Fe2-

dependent prolylhydroxylase that tags HIF-1α through proline hydroxylation on its

oxygen dependent degradation domain (ODD) for degradation by the ubiquitin-

proteasome pathway. Deletion of this domain leads to the stabilization of HIF-1α which

has constitutive DNA binding activity independent of oxygen tension [98]. Under

hypoxia, HIF-1α degradation is prevented, leading to its accumulation and the formation

of active HIF-1α-ARNT heterodimers [99]. A second hypoxia sensitive site on HIF-1α is

in its carboxyl-terminal transactivation domain (CAD). Under hypoxia, there is an

absence of arginine hydroxylation, so the transcriptional coactivators CBP and p300 can

Page 32: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

22

associate with HIF-1α, and enhance the binding of HIF-1α to hypoxia responsive

elements on target genes [77]. CBP/p300 has been shown to interact with HNF-4 [100].

Thus it appears the 3’ enhancer element of the Epo gene provides a scaffold for the

trimeric complex composed of the hypoxia regulated HIF-1α, the constitutively expressed

HNF-4, and the general transcriptional activator p300 to assemble and regulate Epo

transcription [79].

Erythropoietin signaling. Primarily through mechanisms of apoptotic suppression, Epo

acts to increase erythrocyte numbers and combat hypoxia by stimulating the early release

of maturing normoblasts from the marrow, increasing the amount of hemoglobin

synthesized per erythrocyte and stimulating the expansion of late BFU-E and all CFU-E

into mature red cells [84]. EpoR is a single chain receptor, with a highly related tertiary

structure to other family members including growth hormone receptor, prolactin and Tpo

(Figure 1-5B) [80]. It contains no kinase or other enzyme motif in its cytoplasmic

domain. Epo binding induces a conformational change in EpoR leading to

dimerization/oligomerisation of Epo receptors [101]. This allows the

transphosphorylation and activation of the two receptor associated JAK2 molecules

(Janus family receptor tyrosine kinases), which appears to be the initiator of EpoR signal

transduction [81]. JAK2 subsequently phosphorylates some or all of the eight tyrosine

residues in the intracellular domain of EpoR [101]. These phosphorylated tyrosines act

as docking sites and attract other intracellular proteins to bind the EpoR via their src

homology 2 (SH2) domains, leading to further phosphorylation of these proteins. The

signaling cascades propagated include the signal transducers and activator of

Page 33: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

23

transcription STAT5, phospholipid modifying enzymes (PI-3-kinase, PLC-γ, and SHIP),

regulators of Ras and MAP kinase signaling, tyrosine phosphatases (SHIP1 and SHIP2),

suppressors of cytokine signaling (CIS and SOCS3) and the Src family kinases [81].

The earliest detectible cellular action is the release of intracellular calcium ions, followed

in 1-2 hours by transcription of mRNA for several erythroid proteins, including the

globin chains [69]. An increase in Ca2+ influx is an early and necessary step in the

commitment to differentiation of murine erythroleukemia cells. The Ca2+ chelator EGTA

can inhibit Epo-induced murine erythroid colony growth. The increase in intracellular

Ca2+ may stimulate the expression the expression of proto-oncogenes or transcription

factor phosphorylation [101].

Following binding of Epo, the receptor:hormone complex is endocytosed into the

cell and degraded by the lysosome [102]. The amplitude and duration of EpoR signaling

is modulated by negative regulators. CIS1 (a member of the family cytokine-receptor

inhibitors) binds the Epo receptor to interfere with STAT5 binding to the receptor, and

may also accelerate the ubiquitination and degradation of activated EpoR. SOCS3, a

member of the family of suppressor of cytokine signaling, binds the kinase domain of

JAK-2, inhibiting its tyrosine kinase activity in vitro. SHP1 (SH2-containing protein

tyrosine phosphatase 1), dephosphorylates and inactivates JAK2 [81].

Epo signal transduction pathways have been shown to induce increased

expression and or activation of the anti-apoptotic Bcl2 family member Bcl-XL, along

with other transcription factors including GATA-1, NF-E2, SCL, NF-κb, EKLF and

Page 34: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

24

AP-1. Proto-oncogenes which are subsequently activated include c-myc, c-myb, c-fos and c-jun [101].

Hemoglobin. All of the cellular processes of the erythrocyte are finely tuned to create a

cell whose primary function is oxygen delivery. Hemoglobin is the molecule within an

erythrocyte that is directly responsible for oxygen transport and the main intracellular

protein constitutes about 33% of its contents. A reduction in hemoglobin synthesis,

which could result from deficiencies in iron supply or an impairment of prophyrin or

globin production, results in a microcytic, hypochromic anemia [69]. Each hemoglobin

molecule consists of two alpha-type globin chains and two beta-type globin chains, each

containing a heme molecule that can bind oxygen. Mice, along with humans, use

different hemoglobin species throughout ontogeny, a process referred to as hemoglobin

switching. The regulation of this process is primarily transcriptional [103]. In humans

there is an embryonic to fetal globin switch coinciding with the transition from yolk sac

to fetal liver hematopoiesis, and a switch from fetal to adult globins occurring near the

perinatal period when hematopoiesis shifts to the bone marrow. Mice, like most species

have a single switch from embryonic to adult globins once hematopoiesis has shifted

from primitive in the yolk sac to definitive in the fetal liver [104, 105]. The murine

alpha-globin gene on Chr 11, consists of three genes (ζ, α1 and α2) that are dependent on

the major regulatory element (αMRE) which appears as an erythroid specific DNaseI-

hypersensitive site 26 kb upstream of the ζ gene (Figure 1-6A) [105]. The ζ globin chain

is restricted to the embryonic or primitive erythroid lineage, with only transcripts from

the α1 or α2 expressed beyond E12.5 as determined by S1 nuclease protection analysis

Page 35: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

25

[106].

Both human and mouse beta-globin genes are regulated by a variety of

mechanisms including chromatin structure, promoter elements, cis and trans-regulatory

elements including enhancers, silencers and insulator elements and through the

interactions of transcription factors with these elements [103, 104, 107-113]. The murine

beta globin locus (Hbb) on Chr7 consists of four functional genes: the embryonic Hbb-y

and Hbb-bh1 and the adult Hbb-b1, and Hbb-b2 organized with the embryonic genes near

the 5’end and adult genes near the 3’ end of the locus (Figure 1-6B). Though not

contributing to functional globin proteins, the Hbb-bh0 gene is 3’ to Hbb-y and produces

minor embryonic mRNA, while the pseudogenes Hbb-bh2 and Hbb-bh3 lie downstream

of Hbb-bh1[114]. Inbred mouse strains possess one of three haplotypes at the β-globin

locus: Hbbd (diffuse), Hbbs (single) and Hbbp [115]. In strains of the diffuse type (Hbbd/

Hbbd) such as BALB/c, the two adult β chains βmaj and βmin, coded by the Hbb-b1 and

Hbb-b2 genes respectively, are made in unequal amounts and can be distinguished from

one another based on variations in their amino acid sequence (9 substitutions between

βdmin and βdmaj) [114, 116, 117]. Strains of the single haplotype (Hbbs/ Hbbs) such as

C57BL, contain two adult βglobin genes but make only one type of beta chain protein,

which is very similar to the βdmaj chain differing at just 3 amino acid positions [116,

117]. The Hbbp haplotype, such as in the AU/SsJ strain of mice, resembles Hbbd but

makes a variant βmin chain [117, 118].

A key cis-acting regulatory element of the beta globin gene is the Locus Control

Region (LCR). In mice it is comprised of six DNaseI hypersensitive sites (HS) located

from 5 to 28 kilobases upstream of the embryonic Hbb-y globin gene. The LCR is

Page 36: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

26

crucial for the high-level expression of the individual beta globin genes [107] with

distance from the LCR a contributing factor for controlling both the level and the timing

of expression [107, 119] . The LCR functions by establishing an open chromatin

structure of the beta globin genes, along with direct interactions between LCR elements

and the specific gene promoters through DNA looping [108], although knock-out studies

have shown the endogenous β-globin locus does not require the LCR to establish and

maintain an open chromatin conformation for developmental regulation of expression

[120]. Erythroid Kruppel-like factor (EKLF) is the key transcription factor required

specifically for the transcription of only the adult-type β-globin genes and binds to the

CCACCC element in the promoters of the mouse (and human) adult-type β-globin genes,

stabilizing the interaction between the LCR and the β-globin genes [121].

Although remarkably similar, there are slight variations between the human β-

globin genes and those of rodents. Without an embryonic-to-fetal switch as seen in the

human system, the mouse has no quantitative homologue to fetal hemoglobin (HbF)

[104]. Although no quantitatively homologous murine fetal hemoglobin exists, one of

the adult beta globins, βmin, qualitatively resembles that of the human HbF [122]. βmin

is developmentally controlled with an upregulation during gestation to roughly 45% of

total β-like globin gene expression [107] corresponding to the period in human gestation

when there is an upregulation of HbF [123]. Similarly, the amount of βmin drops

following birth to comprise only 20% the total amount of beta globin produced in the

adult. This percentage changes in adult mice under stimulation of erythropoiesis such as

phlebotomy, exposure to hydroxyurea, or injection of erythropoietin, with an increase in

βmin relative to βmaj [107]. There is interest in understanding how the HbF, and thus the

Page 37: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

27

ßmin globin, is regulated for devising new treatments for diseases such as sickle cell

anemia and β-thalassemia. Patients with elevated levels of HbF show milder forms of

these diseases [104, 124]. The current methods to artificially raise HbF levels are toxic,

so further understanding of the system may lead to alternative therapeutic targets and

better treatment options for patients.

Page 38: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

28

Steady State vs. Expansive Erythropoiesis

Steady-state erythropoiesis maintains circulating levels of erythrocytes that have

reached their finite life span. It relies on local sources of Epo and other factors to

replenish the pool of healthy erythrocytes. Steady state erythropoiesis occurs at a

constant rate, within the bone marrow microenvironment. Distinct from steady state is

stress or expansive erythropoiesis. Microenvironments supportive of expansive

erythropoiesis include the fetal liver and adult spleen, and to some extent the adult liver.

It relies on high levels of Epo and other factors, leading to the rapid production of

erythrocytes.

Hypoxia is the major stimulus to induce stress erythropoiesis and could be defined

in terms of hypoxemia, anemia, or increased hemoglobin-oxygen affinity at sea level

[89]. Erythropoietic expansion to extramedullary sites is noticed in response to

hypobaric hypoxia. In fact, the spleen appears to be the optimal microenvironment in

mice for erythroid production and maturation following hypoxia, whether from hypobaric

hypoxia, bleeding or hemolytic anemia [68, 125-128]. It is an accepted principle that

reduced oxygen supply or increased oxygen demand is accompanied by increased plasma

Epo (pEpo) titers. In the laboratory, Epo production can be induced by hypoxemia,

lowering circulating red cell mass by bleeding or phenylhydrazine injections, or cobalt

administration [89].

Once a hypoxic state is sensed in the body, expansive erythropoiesis is engaged,

yet the mechanisms of this process are only beginning to be understood. The erythroid

progenitors themselves are altered in response to hypoxia. Rich et al. (1982)

demonstrated that culturing bone marrow and fetal liver BFU-E or CFU-E in 5% oxygen

Page 39: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

29

(5% oxygen, 5% carbon dioxide, 90% nitrogen) resulted in an increase in colony number

and an increase in Epo sensitivity compared to cells grown in air (air supplemented with

5% carbon dioxide) [129]. Following hypoxia, hematopoiesis shifts to secondary

hematopoietic organs. Work has shown that after severe phlebotomy, the contribution of

the spleen to erythropoiesis increased from 10% in normal mice to over 40% in

phlebotomized mice [130]. Early experiments by Hara and Ogawa attributed the increase

in splenic contribution to the migration of cells from the bone marrow. They followed

erythropoietic precursors in the bone marrow, blood and spleen in adult mice with

phenylhydrazine (PHZ) induced acute anemia following a 3 day injection regime (60

mg/Kg injected on Day 0, 1, and 3). They saw an increase of CFU-E in the bone marrow

until Day 4 with the number of BFU-E declining until Day 10. Both CFU-E and BFU-E

increased until Day 4 in the spleen, and then declined to pretreatment levels. Only BFU-

E were recognized in murine blood, with the maximal increase of blood BFU-E on Day

2, two days prior to the peak in the spleen. To determine whether erythropoietic

stimulation resulted in changes of the progenitor’s proliferation state, they looked at the

proportion of precursors in the DNA synthesis phase. Cells in DNA synthesis phase

incorporate 3HTdR while their proliferative potential is destroyed by it. Using

methylcellulose clonal cell culture technique they looked at erythropoietic precursors in

the femur, spleen and blood of mice made anemic by bleeding, stimulated with Epo

injections or erythropoietically suppressed by the hypertransfusion with packed red blood

cells (polycythemia). Neither anemia nor polycythemia caused changes in the

proliferative stage of the precursors. This suggested to them that the changes in the

number of BFU-E seen in the spleen represent the migration of these early erythroid

Page 40: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

30

progenitors from their storage in the BM to the spleen for expansion and maturation,

rather than BFU-E proliferation [76]. However, work in our lab contradicts this model.

Our analysis shows that BMP4/Madh5 dependent signals, regulated by hypoxia, initiate

the differentiation and expansion of erythroid progenitors resident in the spleen. These

findings suggest a new model where stress erythroid progenitors, resident in the spleen

are poised to expand following an acute anemia [131].

Brandan et al. (1997) determined that the erythroid response to hypoxia was a

function of microenvironmental regulation rather than simple hormonal variance. Mice

were submitted to hypobaric hypoxia (HH) over an 18 day period. Plasma Epo titers rise

shortly after hypoxic exposure in both rodents and humans, but fall within 1-2 days

reaching normal yet above baseline levels after 3-4 days of sustained hypoxia. Cells

from bone marrow and spleen were evaluated over the time course for their proliferative

response to recombinant human erythropoietin (rHuEpo) analyzed by thymidine

incorporation assays as well as total nuclear cell counts and erythropoietic maturation

determined by 59Fe uptake. Bone marrow showed a very gradual yet sustained

proliferative response, with only a slight increase under stress which remained elevated

over the duration of the treatment. Spleen in contrast had a burst of maximal proliferative

response on day 6 of HH (26 fold increase), which returned to near control values after

day 10. Total nuclear cell counts increased in bone marrow and spleen, 1.5 and 5 times

respectively. While bone marrow total erythroid cell counts increased slightly and

remained elevated over the hypoxic period, splenic red cells rose abruptly to 30 times

over control on day 6 and fell sharply to near control levels by day 12. Splenic

contribution was approximately 60% of total production between days 6-8, with bone

Page 41: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

31

marrow making the major contribution (90%) by the end of the assay. The adaptive

response to hypoxia of erythroid cells exposed to the bone marrow or spleen

microenvironments is clearly different even with identical endogenous serum Epo titers

[125].

Aside from Epo, other factors are important for proper stress erythropoiesis.

There is a requirement for glucocorticoids, stem cell factor (SCF) and its receptor Kit,

(and Epo) for the in vitro expansion of immature erythroblasts to induce long-term

proliferation accompanied by differentiation arrest, with demonstrated roles for these

factors in stress erythropoiesis [126, 132-136]. It is known that stem cell factor (SCF)

and its receptor ckit are required for normal hematopoiesis in adult mouse. Mice with

mutations at the Steel locus (Sl) or dominant-white spotting (W) locus, encoding SCF and

the Kit receptor respectively [137-140], have a phenotype consisting of a macrocytic

anemia, decreased numbers of tissue mast cells and abnormalities in spermatocytes and

melanocytes [141]. In addition to these characteristics, the functional interaction of SCF

with Kit is also required for acute erythroid expansion following a hemolytic anemia.

Genetically anemic W/Wv mice, have defects in their recovery to a phenylhydrazine

induced acute anemia, with a delay in their splenic contribution (from the normal day2

until day 7 to 9) [136]. Mice treated with the monoclonal antibody ACK2, which

specifically recognizes the Kit receptor and blocks the hematopoietic growth-promoting

effects of SCF, had reduced femoral CFU-E, BFU-E and CFU-GM following a PHZ-

induced acute anemia, in fact less than half that found in phenylhydrazine treated

controls. Splenic hematopoiesis following an acute anemia was totally ablated in ACK2

antibody treated mice, which is in sharp contrast to control mice which saw a 25-50 fold

Page 42: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

32

increase in CFU-E and 6 to 10 fold increase in BFU-E or CFU-GM in the spleen.

Transplantation experiments using donor cells from phenylhydrazine injected mice

treated with ACK2 or untreated control showed a 75% decrease in the fraction of donor

derived BFU-E and CFU-GM in the bone marrow and spleen of recipient mice

transplanted with ACK2 exposed progenitors. This would suggest that a significant role

for the SCF/Kit interaction in stress erythropoiesis in the spleen is in the homing behavior

of hematopoietic progenitor cells mobilized following an acute anemia [126].

Glucocorticoids are released predominantly during stress to maintain homeostasis

[135, 142, 143]. In vitro, glucocorticoids have been shown to stimulate erythropoiesis

by enhancing the formation of murine erythroid colonies [144] and increasing the

proliferation of erythroid cells [145]. Bauer et al. (1999) has shown the requirement for

the glucocorticoid receptor (GR) for the expansion of immature erythroid cells during

stress erythropoiesis for the activation of erythroid progenitors. GRnull/null mice die at

birth, so they generated GRdim/dim mice which carry a targeted mutation in the

dimerization domain of the GR making it defective for cooperative DNA binding and

consequent transactivation of genes with glucocorticoid responsive elements (GREs), but

still able to transrepress. GRdim/dim mice have impaired regulation of GRE-dependent

genes but survive to adulthood. Erythroid cells from the fetal livers of GRnull/null and

GRdim/dim mice fail to undergo sustained proliferation in vitro like wild-type cells.

Furthermore, unlike wild-type mice which show nearly an eight-fold increase in the

numbers of CFU-E in their spleen following hemolytic anemia, GRdim/dim adult mice have

no increase in their number of CFU-E. Under the erythropoietic stress of sustained

hypoxia, GRdim/dim adult mice show a lack of rapid adaptation as determined by no

Page 43: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

33

observed increase in red blood cell counts, hemoglobin production, percent hematocrit, or

number of CFU-E from the spleen after being subjected to hypoxia (11% O2) for 2 Days.

The erythroid compartment of spleen cells responsible for GR-expansion following an

erythropoietic stress was restricted to the population displaying the unique combination

of early hematopoietic (CD34+ckit+) and the late state erythroid marker Ter119+ surface

antigens [135].

The involvement of a Ter119+ positive cell population in the spleen of

phenylhydrazine treated mice has been explored by others. Vannucchi et al (2000)

describe a bipotent (erythroid and megakaryocytic) early precursor cell which expresses

the erythroid Ter119 and megakaryocyte 4A5 surface markers. The Ter119+/4A5

population increased in both the bone marrow (from 1.3 to 3.8-4.7%) and spleens (from

<0.001 to 1.3-8.3%) of mice during recovery from a PHZ-induced acute anemia. This

population did not contain progenitor cells (CFU-E) and consisted of cells with the

morphology of blasts. PHZ-treated spleens expressed high levels of both erythroid (β-

globin and EpoR) and megakaryocytic (GpIIb, AchE and Mpl) genes. Due to the fact

that almost all (approx. 90%) of the colonies deriving from BFU-E and CFU-Mk (but

none of the colonies from CFU-GM and CFU-E) contained cells expressing erythroid and

megakaryocytic genes, and at least single erythroid BFU-E colonies transferred to

secondary TPO containing cultures that gave rise to cells with megakaryocyte

morphology, they concluded that normal BFU-E and most likely CFU-Mk are bipotent

for Erythroid/Megakaryocytic differentiation. This would concur with other

investigations which [146] would suggest that the E/Mk precursor is downstream to the

BFU-E level [147]. It makes physiological sense to maintain a population of cells

Page 44: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

34

capable of both erythroid and megakaryocytic differentiation to respond to erythropoietic

stress as it is often brought on by massive blood loss due to a ruptured vessel.

Page 45: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

35

Microenvironments/Stroma Supporting Hematopoiesis

Hematopoietic stem cells and their progenitor cells can proliferate and

differentiate in the microenvironments of the fetal liver, adult bone marrow, spleen and to

a certain extent the adult liver [148]. Although dispensable in vitro, in vivo the

macrophage have been considered to be a critical element of erythroid differentiation,

seen in the center of erythroblastic islands in the bone marrow and spleen making cell-to-

cell contact with the erythroblasts via very late antigen (VLA-4/vascular cell adhesion

molecule VCAM-1 interactions [130].

In fact, by selectively abrogating stromal macrophages from the splenic red pulp

with dichloromethylene diphosphonate encapsulated in multilamellar liposomes

(CL2MDP-liposome) in erythropoietically challenged mice, erythropoiesis was

suppressed at the level of CFU-E until 5 days after treatment when macrophages began to

appear in the red pulp [130]. However, it has not been demonstrated that macrophages

or macrophage cell lines support erythropoiesis in vitro. There are other components of

the microenvironments essential for supporting hematopoiesis and directing

differentiation. Stromal regulation of hematopoiesis has been proposed to occur by cell-

cell contact, extracellular matrices or through the secretion of hematopoietic growth

factors [149]. There is variation among hematopoiesis supporting microenvironments in

the types of cells comprising their stroma and the types of factors secreted by resident

cells, leading to varied affects on intrinsically distinct embryonic vs. adult precursors

from stroma stimuli through ontogeny [150].

Due to differences in the cellular composition of the hematopoietic organs, there

are differences in the type of hematopoiesis they are most suited to support. In the adult

Page 46: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

36

mouse, the bone marrow is the major hematopoietic organ where stem cells and their

progenitors develop and where myelopoiesis dominates [150]. In myeloid progenitor

CFU-S assay, granulocyte colonies predominate in the bone marrow, with other CFU-S

progeny preferentially expanded at other hematopoietic locations [59, 151].

Interestingly, when stem cells are injected into animals whose spleens contain plugs of

bone marrow, colonies in spleen were predominantly erythroid, while those forming in

the bone marrow plugs were predominantly granulocytic [151]. The heterogeneous

mixture of adherent cells making up the hematopoietic inductive environment (HIM) of

the bone marrow (supporting the self-renewal and commitment of hematopoietic stem

cells) contains a mixture of macrophages, fibroblast, endothelial cells, and preadipocytes

[150]. In fact, long-term maintenance of stem cell growth and differentiation in vitro is

dependent on these bone marrow derived adherent cells [152].

The spleen and fetal liver have been established as sites of hematopoiesis where

erythropoiesis prevails. In a CFU-S assay, erythroid colonies predominate in the spleen,

with erythroid spleen colonies found throughout the red pulp and granulocyte colonies

distributed in the white pulp [150]. The fetal liver is the main site of hematopoiesis of the

developing embryo, with a phase of rapid expansion of erythropoietic cell population

during 12 to 16 days of gestation. Stromal cell lines from both spleen and fetal liver have

been established demonstrating the erythropoietic inductive environment (EIM) created

in these organs. The MSS31 cell line from mouse newborn spleens selectively supports

the expansion, proliferation and differentiation of erythropoietic progenitor cells from

mouse fetal liver in a semisolid medium with the addition of Epo. After the addition of

fetal liver cells, erythroid progenitors adhered to the MSS31 cells, eventually maturing to

Page 47: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

37

hemoglobin producing cells that would detach and enucleate. This cell line exhibited

properties of endothelial cells having the ability to take up acetylated low-density

lipoprotein (Ac-LD) and form a capillary like structure in collagen matrices [151]. A

total of 5 lines were established from newborn spleens, which could support the

formation of large erythroid colonies from either mouse fetal liver cells or adult bone

marrow cells in the presence of Epo. They detected a new type of large colony composed

of up to 1000 benzidine-positive cells supported specifically on these spleen stromal

lines. This type of colony developing 4 to 6 days in culture was larger than a typical

CFU-E. It was also distinct from a typical BFU-E in its lower Epo dependency,

benzidine-staining, colony size, and timing of appearance [149]. Other cells lines known

to support hematopoietic stem cells such as the bone marrow derived preadipocyte-like

line PA6 and the mouse fibroblast cell line BALB3T3, do not posses the ability to

support the unique large erythroid colonies seen on the MSS cell lines [151]. Cell-to-cell

contact allowing short-range communication between the erythroid progenitor cells and

the MSS cells was required in that conditioned medium of MSS31 cells did not show any

effects on colony formation [149].

Mouse stromal lines established from E13 fetal livers have been established and

also shown to support the proliferation and differentiation of erythroid progenitors in a

semisolid medium in the presence of Epo [59]. Similar to that seen on the spleen stromal

lines, erythroid progenitors from mouse bone marrow of fetal liver were supported on

fetal liver stromal lines (FLS) in a semisolid medium in the presence of Epo after 4 days

in culture. The large colony forming cells from the adult bone marrow and fetal liver

require the same optimum Epo concentration (as low as 0.05 U/mL) enough for CFU-E,

Page 48: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

38

but not for BFU-E. These cells could not be separated from CFU-E in a density gradient,

which can fractionate BFU-E from CFU-E. Thus, the progenitors forming these large

colonies may be at a stage closer to CFU-E than BFU-E, with unique properties that

distinguish them even from the CFU-E. When in close contact with the FLS, erythroid

progenitors can divide rapidly (average generation time of 9.6 hours) with erythroid cells

producing hemoglobins capable of dividing more than 10 times on the FLS layer, which

can help explain the expansive erythropoiesis seen in the fetal liver during development

[59, 150].

Unlike the MSS cell lines with endothelial biology and morphology [151], the

FLS lines were reported to appear more epithelial-like [59]. Chargraoui et al. (2003) has

shown that the fetal liver contains a unique population of cells with endodermal and

mesodermal features and it is this cell population that comprises the hematopoietic

supportive environment. These cells in epithelial-to-mesenchymal transition (EMT) lines

express both mesenchymal and epithelial markers, with EMT cells seen transiently in the

fetal liver during its hematopoietic phase in gestation. As hepatocytes matured and the

hematopoietic capacity of the fetal liver gradually lost, EMTs began to decline, being

absent at the end of gestation and in the adult. By late gestation (E18) two populations

are apparent: epithelial cells which resemble mature hepatocytes and express CK-8, and a

smaller contribution of ASMA-positive mesenchymal cells similar to hepatic Ito

cells/microfibroblasts. The down-regulation of EMT and up-regulation of hepatocyte

differentiation in the fetal liver stromal cell line AFT024 is the direct affect of oncostatin

M (OSM). OSM-induced modifications in the fetal liver stromal cell phenotype are

associated with a decrease in hematopoietic supportive ability, with no effect of OSM on

Page 49: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

39

the bone marrow stromal line MS-5. OSM is expressed by CD45+ hematopoietic cells in

the fetal liver. During ontogeny, the CD45+ hematopoietic cells from the YS and AGM

migrate to the liver rudiment, finding a suitable environment of EMT cells. These cells

expand and differentiate leading to an increase in the amount of OSM, reaching levels

that would induce hepatocyte maturation of EMT cells. As hepatocyte maturation

proceeds and the EMT population decreasing, the liver eventually loses its ability to

support hematopoiesis [153].

In both cases, the EIM created by the MSS or FLS required cell-cell contact

and/or short-range communication between the erythroid progenitor cells and the stromal

layer for large erythroid colonies to form. These large colonies did not form when

progenitors were separated from the stromal layer by a diffusion chamber or a nucleopore

filter [59, 149]. Adhesion molecules and/or cell surface markers are a likely component

essential to the EIM environment. Development of erythroid cells on FLS is inhibited by

antibodies to very late activation antigen 4 (VLA-4 integrin) which is expressed on

erythroid cells [150]. All hematopoietic cells in the fetal liver express VLA-4, yet in

utero treatment of mice with an anti-VLA-4 monoclonal antibody specifically induced

anemia [154]. The stromal membrane associated protein-1 (SMAP-1) localizes to the red

pulp of the spleen where erythropoiesis dominates and expression during development

correlates with erythropoietic activity in the fetal liver. Antisense cDNA transfected into

MSS62 reduced expression of SMAP-1 and suppressed the large erythroid colony

formation, indicating the requirement of SMAP-1 in the EIM [150]. Mutations of the

receptor tyrosine kinase Kit coded by the murine dominant-white spotting locus (W) or

its ligand stem cell factor (SCF) [137, 138, 140] lead to deficiencies of germ cells,

Page 50: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

40

melanocytes, and hematopoiesis, including the erythroid lineage [141]. Not only is the

functional interaction between the EpoR and Kit for erythroid colony formation,

antibodies against Kit in the fetal liver stromal inhibited the proliferation of progenitor

cells [155].

Optimal erythropoietic expansion is not solely based on the EIM

microenvironment. Cell intrinsic qualities exist that prepare a progenitor to take full

advantage of an inductive environment. Such is the case with the unique partnership

created by the fetal liver progenitor cells and the microenvironment of the fetal liver

leading to the expansive erythropoiesis seen at this time. While both bone marrow and

fetal liver stromas effectively maintain CFU-GM, the stimulatory effect of the fetal liver

microenvironment in the long term maintenance of erythroid progenitors in culture is

specific for fetal BFU-E. Fetal liver stroma efficiently supports fetal BFU-E for 6-7

weeks in vitro, whereas bone marrow stroma was not able to maintain fetal BFU-E

beyond 4 weeks. When bone marrow cells were cultured on a fetal liver stromal layer,

the number of adult BFU-E declined precipitously [156]. The fact that there are

differences in the responsiveness of bone marrow or fetal erythroid precursors to the

same stromal cells suggests cell intrinsic differences between progenitors to a particular

environment.

Page 51: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

41

The flexed-tail (f) Mouse Mutant as a Means to Study Expansive Erythropoiesis

flexed-tail (f) is an autosomal recessive mouse mutation that arose spontaneously

and was initially identified in 1928 [157]. The hallmark features of this mutation are

defects in expansive erythropoiesis, a kinked tail which is caused by vertebral fusion, and

ventral white spotting caused by failure of melanoblasts to migrate from the neural crest

[158]. The latter two characteristics are not penetrant on all backgrounds. The primary

characteristics of this phenotype are apparent both during embryogenesis and in the adult

during the recovery from an acute anemia. In both cases the tissues of the animal are

under hypoxic stress.

The embryonic defects of flexed-tail (f/f) mice. The embryos show a transient yet

severe microcytic, hypochromic anemia that remits by two weeks after birth, with the

majority of fetal reticulocytes containing non-heme iron granules, which are referred to

as siderocytes [159, 160]. The number of red cells/mL in the blood of flexed-tail

embryos is approximately 80% that of normal mice, with these cells containing roughly

60-70% of the normal amount of hemoglobin, thus newborn mice contain about half the

normal amount of hemoglobin[161]. The characteristic siderocytes in f/f reticulocytes are

localized in the mitochondria which is similar to that seen in human sideroblastic anemia

[162].

Prior to E12 in the developing mouse, primitive erythroid cells are found in

circulation which originate from the blood islands of the yolk sac and are characterized

by large, nucleated cells, containing embryonic hemoglobins. Primitive erythropoiesis

Page 52: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

42

has not been fully characterized in the f/f mutants. Although their anemia is apparent by

E12 (lower red cell count, smaller body weight), this distinction breaks down at earlier

embryonic stages, with primitive erythrocytes being fully hemoglobinized [160]. The

severe f embryonic anemia is manifested at the shift between primitive and definitive

hematopoiesis. The locus affects not only the proliferation/differentiation of definitive

erythroid progenitors limiting the production and release of erythrocytes, but also the

maturation of the terminally differentiated cells leading to defective hemoglobin

synthesis and siderocytes in fetal reticulocytes [160, 161, 163]. Previous analysis had

focused on the affect of f/f on fetal liver erythropoiesis; however we have seen

differences between f/f and control littermates as early as the emergence of definitive

hematopoietic progenitors in the aorta-gonad-mesonephros (AGM) region (E10.5). f/+

control littermates have hematopoietic clusters budding from the ventral wall of the

dorsal aorta at E10.5, which are thought to be driven by signals from the surrounding

mesenchymal tissue [4]. This situation is similar to that seen in human embryos at a

corresponding stage of development [52]. These clusters are missing from the wall of the

f/f dorsal aorta at E10.5 and are not seen until E11.5. In addition, f/f embryos exhibit a

less dense mesenchymal layer which may affect the development of hematopoietic

clusters (see Appendix A; Figure A-4). Primitive erythropoiesis is completely replaced

by E12 as definitive hematopoietic progenitor cells from this intra-embryonic source

along with cells from extra-embryonic sources, such as the yolk sac, migrate to and

establish the fetal liver as the primary hematopoietic organ of the embryo. The primitive

to definitive progression is delayed in f/f mutants, in that fully hemoglobinized primitive

cells (nucleated erythroblasts) persist into E16, possibly to compensate for the deficiency

Page 53: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

43

of fully hemoglobinized definitive erythroblasts in circulation [160].

The anemia of the f/f mutant is especially pronounced between E13-16 [158],

which coincides with when the fetal liver is the primary hematopoietic organ of the

embryo. The f/f fetal liver is smaller in size and contains decreased numbers of erythroid

progenitors and identifiable erythroblasts [163-165]. Although the rate of increase in the

number of erythrocytes is as great in the f/f anemic fetuses compared to +/+, the defect

present in the fetal liver as early as E12 is retained with the same relative deficiency

through E16, suggesting the defect is in an erythroid progenitor migrating to or

proliferating in the fetal liver [158]. An early multi-potential myeloid progenitor capable

of committing to the erythroid lineage is the spleen colony forming unit or CFU-S. These

cells when injected into heavily irradiated recipients form macroscopic spleen colonies 9-

12 days after injection. There is no significant reduction in the ratio of CFU-S in the f/f

fetal livers, however the absolute number per liver is reduced (E12-15) [163, 164].

CFU-S consist primarily of erythroid cells. The CFU-S from f/f fetal livers produce

smaller colonies that are deficient in the erythroid cells as determined through

histological examination and the incorporation of 59Fe [163]. These results suggest that

erythroid differentiation in CFU-S is severely compromised.

Bateman et al. (1972) looked at the differentiation capacity of the colony forming

units (cfu) by measuring incorporation of 59Fe into the peripheral blood of recipients

transplanted with f/f fetal liver cells. He suggested the reduced number of recognizable

erythroblasts of f/f embryos is due to the low rate of proliferation of cfu and not a

reduction in the number of erythroblasts produced per cfu. More mature erythroid

committed progenitors are also affected by the f locus [163]. Both the absolute numbers

Page 54: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

44

and proportions of CFU-E in f/f fetal livers rise more slowly, with peak values reaching

only 50% that of normal CFU-E while also persisting longer in the fetal liver (Figure 1-7)

[164]. The number of CFU-E present in vivo has been shown to be regulated by

erythropoietin (Epo) [76, 166]. Control and f/f CFU-E have equivalent sensitivities to

Epo in in vitro colony assays [164]. Near the end of gestation (E18), reticulocytes which

still make up the majority of red blood cells of control animals (70%), are elevated in f/f

mutants (98%), as the hematopoietic system is still trying to control the fetal anemia

[162].

The effects of the f locus on fetal hemoglobin synthesis are just as severe as the

defects seen in the proliferation and differentiation of erythroid progenitors. As Cole et

al. suggests, the hemoglobin deficiencies may arise from the complex interaction between

disturbed proliferation of precursor cells and disturbed hemoglobin synthesis in

terminally differentiated cells [161]. Although comparable numbers of hemoglobin

deficient and siderocyte containing cells can be found between normal and flexed

littermates at the transition between primitive and definitive erythropoiesis at E12,

normal mice rapidly proceed by E14 to the production of fully hemoglobinized cells with

little or no siderocytic material (Figure 1-8). In sharp contrast, the hemoglobin

deficiency and persistence of siderocytes are prominent features of f/f mutant during

development [159, 160, 162]. Throughout gestation, the majority of fetal reticulocytes,

upwards of 90%, contain siderocytes. Ultra structure studies reveal they are localized

within the mitochondria which is similar to human sideroblastic anemia [162]. This

elevated percentage of cells containing siderocytes does not drop appreciably until birth,

with levels finally stabilized at roughly 3% within three weeks [167]. During

Page 55: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

45

development, the cells without siderocytes are just as hemoglobin deficient as those with

siderocytes [160]. This suggests the hemoglobin deficiency of flexed-tail mutants is not

completely tied to inefficiently utilizing iron, in that cells capable of maintaining

appropriate intracellular iron pools are still unable to synthesize enough hemoglobin.

Hemoglobin synthesis is a highly regulated process where the production of heme

is coordinated with the synthesis of α and β globin chains. f/f fetal reticulocytes have a

50% deficiency in beta globin chain synthesis which leads to an imbalance in α:β chain

ratio [162]. The deficiency in beta chain synthesis could be rescued by the addition of

heme, but not its precursor protoporphyrin (refer to Figure 1-9). This observation

suggests that there might be a defect in the conversion of protoporphyrin IX to the iron

containing molecule, heme, by heme synthetase. Iron uptake into f/f fetal reticulocytes is

normal while the utilization of absorbed iron for heme synthesis is reduced to less than

half the normal levels, with a similar reduction in the pool of iron available for heme

synthesis. The activity of heme synthetase in f/f reticulocyte homogenates (E17-18) was

similar to controls in vitro, when provided the precursor protoporphyrin, which had little

effect on the activity of normal reticulocyte homogenate. The activity of this enzyme

was actually elevated on a per liver and wet-weight basis when they compared fetal liver

homogenates between wild type and f/f mice at times the livers are composed of

comparable proportions of erythroid and non-erythroid cells and similar distributions of

erythroid types (f/+ E15; f/f E16) [161]. More importantly, addition of protoporphyrin

caused only a slight reduction in iron incorporation into heme from both genotypes,

indicating that it is unlikely heme synthetase activity in f/f fetal livers is restricted by

shortage of the precursor [168].

Page 56: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

46

Heme is part of an end product inhibition loop that can inhibit or repress

upstream heme biosynthetic enzymes such as ALA synthetase, so that heme and globin

synthesis remain coordinated [161]. The inhibition would be propagated through the

pathway, with decreased activity of one enzyme leading to decreased activity of

subsequent downstream enzymes. Coleman et al. (1969) looked at the activity of delta-

aminolevulinate dehydratase (ALD) which follows ALA synthetase in the heme

biosynthetic pathway. ALD converts delta-aminolevulinic acid to the heme precursor

porphobilinogen. Peak activity of ALD in fetal livers from control animals expressed as

activity/g of liver occurred on E14-15, and coincided with the probable period of

maximum heme biosynthesis. Mutation of the f gene had no effect on this profile. There

was concern that since ALD is expressed in the adult liver, which is not normally an

erythropoietic organ, non-hematopoietic levels of this enzymes could be obscuring any

possible changes in the activity seen in +/+ and f/f fetal livers. Therefore they evaluated

the activity of uroporphyrinogen synthetase, an enzyme more specific to hematopoietic

tissue and not found in appreciable quantities in the adult liver. No deficiency of

uroporphyrinogen synthetase was detectible in f/f fetal livers at E14-15, with only

moderate deficiency on E13. It should be noted that liver weight of flexed fetuses is only

50-60% that of normal fetuses, so the actual enzyme activity per liver is decreased in

flexed mice even through the amount per gram is normal [165].

While certain enzyme components of the heme biosynthetic pathway show

normal activity in f/f embryos, questions remain as to the cause of the hemoglobin

deficiencies and widespread siderocytes. While other groups concluded the major cause

of the anemia of f/f neonates to be decreased heme synthesis resulting from reduced

Page 57: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

47

formation of protoporphyrin or its precursors and aberrant regulation of globin chain

synthesis at the level of translation [161], Chui et al.(1977) suggest another possibility.

They attribute the excessive intracellular iron pools, and thus the siderocytes, to a defect

in the coordination regulation of hemoglobin components and a decreased utilization of

heme, rather than a problem in the heme biosynthetic pathway. Key to his argument is

the decreased production of globin chains with equivalent levels of free erythrocyte

protoporphyrin between control and f/f mice at E18 (Figure 1-10). By radiolabeling

newly synthesized globin chains and separating products by column chromatography,

they showed that f/f mutant reticulocytes exhibit a decrease in β-globin chain synthesis.

Fluorescence measurements from E18 +/+ or f/f fetal erythrocytes showed similar levels

of free protoporphyrin. Iron uptake into neonatal reticulocytes is normal; in fact one

would expect free erythrocyte levels to be markedly elevated if the mutant cells were iron

deficient. Not only is there equivalent levels of this precursor of heme, they go on to

report that there is an excess of free heme in mutant reticulocytes, evidenced by the fact

protein synthesis in f/f reticulocytes is more resistant than normal reticulocytes to the

inhibitory effects of three heme synthesis inhibitors (isoniazid [INH which acts at the step

involving pyridoxal phosphate], the iron chelator 2,2’-bipyridine and ethanol). One of

these inhibitors, INH, normally results in the inhibition of globin synthesis, with alpha

chain more susceptible than beta to the inhibitory effects due to the lack of intracellular

free heme. However, total globin synthesis in f/f cells was less vulnerable to the

inhibitory effects of INH than the control cells (106% vs. 26% of control). Further

evidence the defect of f is not a decrease in heme synthesis is that, hemin, the Fe+3

oxidation product of heme, which has been shown to be capable of stimulating globin

Page 58: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

48

synthesis in heme deficient reticulocytes by as much as 270% above control values, has

identical stimulatory effects on f/f globin synthesis (just 25% above basal values) as that

seen in normal fetal red cells. In fact, hemin preferentially stimulates alpha chain over

beta chain synthesis, but it fails to significantly alter alpha/beta ratios in f/f mutant cells

where alpha synthesis is already elevated compared to beta chain [162]. Many have

suggested the deficiency in hemoglobin synthesis is a consequence of disturbed

proliferation [161, 164] or the inability of the heme biosynthetic pathway to keep pace

with the rapid proliferation during fetal development [158]. Regardless of the root

cause, it is clear the reduced hemoglobin and siderocytic granules seen in f/f neonates are

stemming from an overall disruption in the coordinate regulation of hemoglobin

components. Arguably, the effect of the f mutation during embryonic development on

the proliferation/differentiation of hemoglobin producing cells and the production of

hemoglobin by those cells is complex. There is an overall disruption of the coordinate

regulation of hemoglobin components. The consequence of this disruption manifests as

the hemoglobin deficiencies and siderocytes seen in f/f mutant reticulocytes.

The adult defects of flexed-tail (f) mice. Thompson et al. (1966) determined that the f

locus was involved in the

“control of hematopoietic function that is manifested only under conditions of rapid growth, rather than simply a step specific to

fetal life.” [167]

This observation connects the two processes, the expansive erythropoiesis seen in the

developing embryo, and erythropoiesis following an acute anemia in adult. The severe

Page 59: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

49

anemia seen in the f neonate resolves within 2 weeks after birth. Adult mice appear

normal based on basic blood parameters. Their blood contains normal numbers of

erythrocytes and siderocytes are essentially absent [159, 160]. There are normal numbers

of erythroid progenitors such as BFU-E and CFU-E in f/f bone marrow and spleen [166].

The sensitivities of progenitors such as CFU-E to Epo and IL3 are similar to controls

[164, 166].

Differences between constituents of the hematopoietic system of the wild-type

and f/f adult mice begin to emerge as the mutant is dissected more closely. As was seen

from CFU-S from the f/f fetal liver, there are similar numbers of this multipotential

myeloid progenitor from f/f adult bone marrow, but they produce smaller spleen colonies

devoid of erythroid cells. There is a disturbance in the proliferation or differentiation of

these progenitors towards the erythroid lineage evidenced by the decreased ability of f/f

CFU-S to incorporate 59Fe into heme from the 6th to 10th day after transplantation (Figure

1-11). The defect manifests only under conditions of rapid proliferation, as there are

normal levels of iron incorporation by 30 days post transplantation once homeostasis has

been re-established. This is not the consequence of host cells replacing engrafted mutant

cells. Secondary CFU-S assays were performed by taking spleens engrafted with f/f or wt

bone marrow 20 or 30 days after initial transplantation, and reinjected into hosts to obtain

spleen colonies. f/f donor cells do persist and retain the characteristic iron incorporation

defect when subsequently transplanted into new hosts. The defective incorporation of

iron into heme is not a consequence of the f/f spleen microenvironment, but rather a cell

intrinsic deficiency since wild type donor cells into flexed recipients behave as if injected

into the wild type recipients [167].

Page 60: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

50

One may question whether this CFU-S defect is a delay in the appearance of cells

of the erythroid lineage from these progenitors, or a delay in the rate of hemoglobin

synthesis by individual erythroid progeny. Fowler et al. (1967) addressed this question

using radioautography to measure the uptake of 59Fe by individual cells versus overall

incorporation into heme. There were similar numbers of radioiron granules in nucleated

and non-nucleated cells from +/+ and f/f origins. However, overall there were fewer

59Fe-labeled cells per 100 nucleated cells in the f/f radioautographs compared to +/+

controls. The fact that the deficiency is in the number of cells with radioiron granules,

and not the amount of radioiron per cell demonstrates the f locus specifically affects the

proliferation of erythroid cells, leading to a delay in the production of hemoglobin-

synthesizing cells, rather than a decrease in the rate of hemoglobin synthesis by these

cells. The proliferation/differentiation defects in erythroid progeny from CFU-S do not

alter other hematopoietic lineages. Granulopoiesis from f/f CFU-S was unaffected, with

only a slightly longer lag period in the colony-forming cell growth curve, and no

significant differences in doubling times [169].

Adult f/f mutants do have defects in hematopoietic progenitors that can be

detected by specialized in vivo and in vitro colony assay. The consequences of these

differences are phenotypically apparent in the animal only during periods of

erythropoietic stress such as massive blood loss, irradiation, or phenylhydrazine (PHZ)

induced acute anemia. The hypoxic state caused by the destruction of cells capable of

transporting oxygen invokes expansive erythropoiesis to return blood levels back to

normal. f/f have a delay in this recovery to an acute anemia affecting kinetics rather than

magnitude of parameters ranging from reticulocyte counts in the blood to the up

Page 61: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

51

regulation of heme biosynthetic enzymes [165]. Following administration of a single

dose of PHZ (100mg/kg body weight), hematocrit, or packed red blood cell volumes,

drop from 50% to below 30% as a large portion of red blood cells are lysed by the

chemical. Wild-type and f/+ can return hematocrit values to the normal range (~50%),

within 7 days, while it takes f/f mutants 10-12 days to reach a normal blood profile [170].

For this recovery to occur, the hematopoietic system must shift from steady state

maintenance to stress mechanisms to rapidly return homeostasis. The bone marrow may

be a storage site for hematopoietic cells, but the core of the response to re-establish the

erythropoietic system occurs in the spleen [76, 165]. It is not surprising then that there is

no striking difference seen in the bone marrow of f/f and wild-type mice recovering an

acute anemia. Both show a rise in erythropoietic cells from 12%, to 20% in normal mice

by the fourth day after treatment, and 29% in f/f by the fifth day after treatment [165] (It

should be noted that in these experiments, a different PHZ regimen was used that

required 3 injections, dropping hematocrit to 25% 2 days post treatment). The delay in

the return of normal hematocrit levels of f/f mice can be attributed to the delay in the

expansion and maturation of erythroid progenitors in the spleen. An increase in spleen

weight closely parallels the rise in reticulocyte counts following PHZ treatment in both

flexed and normal genotypes. Maximum spleen weight is delayed in f/f compared to

controls, like the percentage of reticulocytes in the blood (Figure 1-12). These

reticulocytes do not contain siderocytes as seen in the embryo. These parameters, as well

as percentages of erythropoietic cells in the spleen determined morphologically by spleen

smears for nucleated hemoglobinized cells, rose more slowly and remained higher longer

in f/f mutants [165].

Page 62: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

52

The delay in erythropoietic cells in the spleen of f/f mice leads to the overall delay

in establishing normal blood parameters following an erythropoietic challenge. This

delay also extends to the enzymatic changes in the spleen which normally occur during

recovery period. Spleen cells actively producing hemoglobin increase from 2 to 60% of

the total number of cells following PHZ treatment in wild type mice, with an increase in

δ-aminolevulinate dehydratase (ALD) activity 24hrs post PHZ treatment. Similar

increases are seen in uroporphyrinogen synthetase, another enzyme in the heme

biosynthetic pathway. The increase in enzyme activity is not simply due to increase in the

size of the spleen, since activity levels are expressed in units per grams of tissue. The lag

seen in the proliferation of heme-synthesizing cells in f/f spleens leads to the lag in this

compensatory increase of both ALD and uroporphyrinogen synthetase by 2-3 days

(Figure 1-13). Although the activity of critical enzymes in the heme biosynthetic

pathway has been disturbed, there is no increase in siderocytes found in f/f adults

recovering from an acute anemia. These results demonstrate that the direct effect of f is

in the control of the differentiation or proliferation of the hematopoietic cells which

contain these enzymes [165].

Although f/f adult mice have normal numbers of BFU-E and CFU-E, they do

exhibit a defect in transient endogenous colony forming cells (TE-CFU). Unlike the in

vitro colony assays used for BFU-E and CFU-E, TE-CFU are defined in an in vivo assay,

requiring an erythropoietic stressor to engage them. Mice are given a sub lethal dose of

radiation (800 rads) that destroys many actively cycling cells and stimulated with 10 units

of Erythropoietin. These progenitors form macroscopic colonies on the spleen composed

primarily of maturing erythroblasts 4-6 days after stimulation. TE-CFU are more

Page 63: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

53

severely affected by the f locus than a mere change in the kinetics of their appearance or

functioning following an acute anemia. The nature of the TE-CFU makes it a likely

candidate for a progenitor whose role is highly specialized for rapid erythropoietic

repopulation which is required only at times of great erythropoietic need. These cells

have the ability to produce an excess of 105 progeny in approximately 5 days dependent

on the continual exposure to Epo [166]. Such an extreme condition parallels the hypoxic

state following a PHZ induced acute anemia, so it is not surprising a progenitor poised to

respond to acute erythroid stress is defective in f/f mutant mice.

In terms of development of an erythrocyte, TE-CFU are postulated to lie between

the multipotential myeloid progenitor (CFU-S) and the late stage erythroid committed

CFU-E based on the timing of their appearance, growth requirements and colony size.

The extent that these cells may overlap with the more immature erythroid progenitor, the

BFU-E has not been determined. Defining the exact relationship between the TE-CFU

and the other erythroid committed progenitors is complicated by the different types of

assays used to detect them. TE-CFU are virtually undetectable in f/f mutants along the

course of the experiment, while +/+ show upwards of 100 colonies per spleen at the peak

of their expression (Figure 1-14). Equivalent numbers of TE-CFU are reported between

f/f and +/+ at later times (9-12 days), but these seem inconsequential when compared to

the numbers seen in the initial wave of +/+ mice. TE-CFU are unable to be transplanted

when injected into irradiated hosts. To circumvent this attribute, chimeras were

established from +/+ or f/f bone marrow and used to test whether the lack of TE-CFU in

f/f mutants is a cell intrinsic defect or a microenvironmental consideration. f/f chimeras

lack the transient wave of colonies seen in the +/+ control chimeras demonstrating the

Page 64: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

54

cell intrinsic defect of the mutant progenitor cells. Clearly the f locus is having a

profound effect on the generation and/or maturation of this specific erythroid progenitor

[166].

Page 65: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

55

References

1. Yoder, M.C.a.P., J., Ventral (Yolk sac) Hematopoiesis in the Mouse, in

Hematopoiesis. A developmental approach, L.I. Zon, Editor. 2001, Oxford University Press, Inc.: New York. p. 180-191.

2. de Bruijn, M.F., X. Ma, C. Robin, K. Ottersbach, M.J. Sanchez, and E. Dzierzak, Hematopoietic stem cells localize to the endothelial cell layer in the midgestation mouse aorta. Immunity, 2002. 16(5): p. 673-83.

3. Orkin, S.H., Diversification of haematopoietic stem cells to specific lineages. Nat Rev Genet, 2000. 1(1): p. 57-64.

4. Godin, I. and A. Cumano, The hare and the tortoise: an embryonic haematopoietic race. Nat Rev Immunol, 2002. 2(8): p. 593-604.

5. Spangrude, G.J., S. Heimfeld, and I.L. Weissman, Purification and characterization of mouse hematopoietic stem cells. Science, 1988. 241(4861): p. 58-62.

6. Bagby Jr., G.C., Hematopoiesis, in The Molecular Basis of Blood Diseases, 2nd edition, G. Stamatoyannopoulos, et al., Editors. 1993, WB Saunders: Phialdelphia. p. 71-103.

7. Mikkola, H.K. and S.H. Orkin, The search for the hemangioblast. J Hematother Stem Cell Res, 2002. 11(1): p. 9-17.

8. Kessler, D.S. and D.A. Melton, Vertebrate embryonic induction: mesodermal and neural patterning. Science, 1994. 266(5185): p. 596-604.

9. Dyer, M., S. Farrington, D. Mohn, J. Munday, and M. Baron, Indian Hedgehog activates hematopoiesis and vasculogenesis and can respecify prospective neurectodermal cell fate in the mouse embryo. Development, 2001. 128: p. 1717-1730.

10. Schmerer, M. and T. Evans, Primitive erythropoiesis is regulated by Smad-dependent signaling in postgastrulation mesoderm. Blood, 2003. 102(9): p. 3196-205.

11. Mead, P.E., C.M. Kelley, P.S. Hahn, O. Piedad, and L.I. Zon, SCL specifies hematopoietic mesoderm in Xenopus embryos. Development, 1998. 125(14): p. 2611-20.

12. Robb, L., I. Lyons, R. Li, L. Hartley, F. Kontgen, R.P. Harvey, D. Metcalf, and C.G. Begley, Absence of yolk sac hematopoiesis from mice with a targeted disruption of the scl gene. Proc Natl Acad Sci U S A, 1995. 92(15): p. 7075-9.

13. Shivdasani, R.A., E.L. Mayer, and S.H. Orkin, Absence of blood formation in mice lacking the T-cell leukaemia oncoprotein tal-1/SCL. Nature, 1995. 373(6513): p. 432-4.

14. Mead, P.E., A.E. Deconinck, T.L. Huber, S.H. Orkin, and L.I. Zon, Primitive erythropoiesis in the Xenopus embryo: the synergistic role of LMO-2, SCL and GATA-binding proteins. Development, 2001. 128(12): p. 2301-8.

15. Murray, P.S.F., The development in vitro of the blood of the early chick embryo. Proc. Roy. Soc. London, 1932(111): p. 497-521.

Page 66: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

56

16. Garcia-Porrero, J.A., A. Manaia, J. Jimeno, L.L. Lasky, F. Dieterlen-Lievre, and I.E. Godin, Antigenic profiles of endothelial and hemopoietic lineages in murine intraembryonic hemogenic sites. Dev Comp Immunol, 1998. 22(3): p. 303-19.

17. Choi, K., M. Kennedy, A. Kazarov, J.C. Papadimitriou, and G. Keller, A common precursor for hematopoietic and endothelial cells. Development, 1998. 125(4): p. 725-32.

18. Kennedy, M., M. Firpo, K. Choi, C. Wall, S. Robertson, N. Kabrun, and G. Keller, A common precursor for primitive erythropoiesis and definitive haematopoiesis. Nature, 1997. 386(6624): p. 488-93.

19. Shalaby, F., J. Ho, W.L. Stanford, K.D. Fischer, A.C. Schuh, L. Schwartz, A. Bernstein, and J. Rossant, A requirement for Flk1 in primitive and definitive hematopoiesis and vasculogenesis. Cell, 1997. 89(6): p. 981-90.

20. Shalaby, F., J. Rossant, T.P. Yamaguchi, M. Gertsenstein, X.F. Wu, M.L. Breitman, and A.C. Schuh, Failure of blood-island formation and vasculogenesis in Flk-1-deficient mice. Nature, 1995. 376(6535): p. 62-6.

21. Lin, C.S., S.K. Lim, V. D'Agati, and F. Costantini, Differential effects of an erythropoietin receptor gene disruption on primitive and definitive erythropoiesis. Genes Dev, 1996. 10(2): p. 154-64.

22. Neubauer, H., A. Cumano, M. Muller, H. Wu, U. Huffstadt, and K. Pfeffer, Jak2 deficiency defines an essential developmental checkpoint in definitive hematopoiesis. Cell, 1998. 93(3): p. 397-409.

23. Wu, H., X. Liu, R. Jaenisch, and H.F. Lodish, Generation of committed erythroid BFU-E and CFU-E progenitors does not require erythropoietin or the erythropoietin receptor. Cell, 1995. 83(1): p. 59-67.

24. Silver, L. and J. Palis, Initiation of murine embryonic erythropoiesis: a spatial analysis. Blood, 1997. 89(4): p. 1154-64.

25. Cumano, A., F. Dieterlen-Lievre, and I. Godin, Lymphoid potential, probed before circulation in mouse, is restricted to caudal intraembryonic splanchnopleura. Cell, 1996. 86(6): p. 907-16.

26. Cumano, A., J.C. Ferraz, M. Klaine, J.P. Di Santo, and I. Godin, Intraembryonic, but not yolk sac hematopoietic precursors, isolated before circulation, provide long-term multilineage reconstitution. Immunity, 2001. 15(3): p. 477-85.

27. Pardanaud, L., D. Luton, M. Prigent, L.M. Bourcheix, M. Catala, and F. Dieterlen-Lievre, Two distinct endothelial lineages in ontogeny, one of them related to hemopoiesis. Development, 1996. 122(5): p. 1363-71.

28. Yoder, M.C., Introduction: spatial origin of murine hematopoietic stem cells. Blood, 2001. 98(1): p. 3-5.

29. Palis, J. and M.C. Yoder, Yolk-sac hematopoiesis: the first blood cells of mouse and man. Exp Hematol, 2001. 29(8): p. 927-36.

30. Palis, J., S. Robertson, M. Kennedy, C. Wall, and G. Keller, Development of erythroid and myeloid progenitors in the yolk sac and embryo proper of the mouse. Development, 1999. 126(22): p. 5073-84.

31. Palis, J., R.J. Chan, A. Koniski, R. Patel, M. Starr, and M.C. Yoder, Spatial and temporal emergence of high proliferative potential hematopoietic precursors

Page 67: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

57

during murine embryogenesis. Proc Natl Acad Sci U S A, 2001. 98(8): p. 4528-33.

32. Rich, I.N., The developmental biology of hemopoiesis: effect of growth factors on the colony formation by embryonic cells. Exp Hematol, 1992. 20(3): p. 368-70.

33. Muller, A.M., A. Medvinsky, J. Strouboulis, F. Grosveld, and E. Dzierzak, Development of hematopoietic stem cell activity in the mouse embryo. Immunity, 1994. 1(4): p. 291-301.

34. Chen, X.D. and J.B. Turpen, Intraembryonic origin of hepatic hematopoiesis in Xenopus laevis. J Immunol, 1995. 154(6): p. 2557-67.

35. Medvinsky, A. and E. Dzierzak, Definitive hematopoiesis is autonomously initiated by the AGM region. Cell, 1996. 86: p. 897-906.

36. Medvinsky, A.L., N.L. Samoylina, A.M. Muller, and E.A. Dzierzak, An early pre-liver intraembryonic source of CFU-S in the developing mouse. Nature, 1993. 364(6432): p. 64-7.

37. Cudennec, C.A., J.P. Thiery, and N.M. Le Douarin, In vitro induction of adult erythropoiesis in early mouse yolk sac. Proc Natl Acad Sci U S A, 1981. 78(4): p. 2412-6.

38. Dieterlen-Lievre, F., On the origin of haemopoietic stem cells in the avian embryo: an experimental approach. J Embryol Exp Morphol, 1975. 33(3): p. 607-19.

39. Martin, C., O. Lassila, T. Nurmi, J. Eskola, F. Dieterlen-Lievre, and P. Toivanen, Intraembryonic origin of lymphoid stem cells in the chicken: studies with sex chromosome and IgG allotype markers in histocompatible yolk sac-embryo chimaeras. Scand J Immunol, 1979. 10(4): p. 333-8.

40. de Bruijn, M.F., N.A. Speck, M.C. Peeters, and E. Dzierzak, Definitive hematopoietic stem cells first develop within the major arterial regions of the mouse embryo. Embo J, 2000. 19(11): p. 2465-74.

41. Huang, H. and R. Auerbach, Identification and characterization of hematopoietic stem cells from the yolk sac of the early mouse embryo. Proc Natl Acad Sci U S A, 1993. 90(21): p. 10110-4.

42. Lu, L.S., S.J. Wang, and R. Auerbach, In vitro and in vivo differentiation into B cells, T cells, and myeloid cells of primitive yolk sac hematopoietic precursor cells expanded > 100-fold by coculture with a clonal yolk sac endothelial cell line. Proc Natl Acad Sci U S A, 1996. 93(25): p. 14782-7.

43. Jordan, C.T., C.M. Astle, J. Zawadzki, K. Mackarehtschian, I.R. Lemischka, and D.E. Harrison, Long-term repopulating abilities of enriched fetal liver stem cells measured by competitive repopulation. Exp Hematol, 1995. 23(9): p. 1011-5.

44. Sanchez, M.J., A. Holmes, C. Miles, and E. Dzierzak, Characterization of the first definitive hematopoietic stem cells in the AGM and liver of the mouse embryo. Immunity, 1996. 5(6): p. 513-25.

45. Matsuoka, S., K. Tsuji, H. Hisakawa, M. Xu, Y. Ebihara, T. Ishii, D. Sugiyama, A. Manabe, R. Tanaka, Y. Ikeda, S. Asano, and T. Nakahata, Generation of definitive hematopoietic stem cells from murine early yolk sac and paraaortic splanchnopleures by aorta-gonad-mesonephros region-derived stromal cells. Blood, 2001. 98(1): p. 6-12.

Page 68: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

58

46. Yoder, M.C., J.G. Cumming, K. Hiatt, P. Mukherjee, and D.A. Williams, A novel method of myeloablation to enhance engraftment of adult bone marrow cells in newborn mice. Biol Blood Marrow Transplant, 1996. 2(2): p. 59-67.

47. Yoder, M.C., K. Hiatt, and P. Mukherjee, In vivo repopulating hematopoietic stem cells are present in the murine yolk sac at day 9.0 postcoitus. Proc Natl Acad Sci U S A, 1997. 94(13): p. 6776-80.

48. Lassila, O., J. Eskola, P. Toivanen, C. Martin, and F. Dieterlen-Lievre, The origin of lymphoid stem cells studied in chick yold sac-embryo chimaeras. Nature, 1978. 272(5651): p. 353-4.

49. Kumaravelu, P., L. Hook, A.M. Morrison, J. Ure, S. Zhao, S. Zuyev, J. Ansell, and A. Medvinsky, Quantitative developmental anatomy of definitive haematopoietic stem cells/long-term repopulating units (HSC/RUs): role of the aorta-gonad-mesonephros (AGM) region and the yolk sac in colonisation of the mouse embryonic liver. Development, 2002. 129(21): p. 4891-9.

50. Yoder, M.C., K. Hiatt, P. Dutt, P. Mukherjee, D.M. Bodine, and D. Orlic, Characterization of definitive lymphohematopoietic stem cells in the day 9 murine yolk sac. Immunity, 1997. 7(3): p. 335-44.

51. Wood, H.B., G. May, L. Healy, T. Enver, and G.M. Morriss-Kay, CD34 expression patterns during early mouse development are related to modes of blood vessel formation and reveal additional sites of hematopoiesis. Blood, 1997. 90(6): p. 2300-11.

52. Marshall, C., C. Kinnon, and A. Thrasher, Polarized expression of bone morphogenetic protein-4 in the human aorta-gonad-mesonephros region. Blood, 2000. 96: p. 1591-1593.

53. Marshall, C.J. and A.J. Thrasher, The embryonic origins of human haematopoiesis. Br J Haematol, 2001. 112(4): p. 838-50.

54. North, T., T. Gu, T. Stacy, Q. Wang, L. Howard, M. Binder, M. Marin-Padilla, and N. Speck, Cbfa2 is required for the formation of intra-aortic hematopoietic clusters. Development, 1999. 126: p. 2563-2575.

55. Cai, Z., M. de Bruijn, X. Ma, B. Dortland, T. Luteijn, R.J. Downing, and E. Dzierzak, Haploinsufficiency of AML1 affects the temporal and spatial generation of hematopoietic stem cells in the mouse embryo. Immunity, 2000. 13(4): p. 423-31.

56. Zon, L.I., Developmental biology of hematopoiesis. Blood, 1995. 86(8): p. 2876-91.

57. Houssaint, E., Differentiation of the mouse hepatic primordium. II. Extrinsic origin of the haemopoietic cell line. Cell Differ, 1981. 10(5): p. 243-52.

58. Johnson, G.R. and M.A. Moore, Role of stem cell migration in initiation of mouse foetal liver haemopoiesis. Nature, 1975. 258(5537): p. 726-8.

59. Ohneda, O., N. Yanai, and M. Obinata, Microenvironment created by stromal cells is essential for a rapid expansion of erythroid cells in mouse fetal liver. Development, 1990. 110(2): p. 379-84.

60. Peschle, C., A. Migliaccio, G. Migliaccio, R. Ciccariello, F. Letteri, G. Quattrini, G. Russo, and G. Mastroberardino, Identification and characterization of three classes of erythroid progenitors in human fetal liver. Blood, 1981. 58: p. 565-574.

Page 69: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

59

61. Muench, M. and R. Namikawa, Disparate regulation of human fetal erythropoiesis by the microenvironments of the liver and bone marrow. Blood Cells, Molecules and Diseases, 2001. 27: p. 377-390.

62. Wu, H., U. Klingmuller, A. Acurio, J.G. Hsiao, and H.F. Lodish, Functional interaction of erythropoietin and stem cell factor receptors is essential for erythroid colony formation. Proc Natl Acad Sci U S A, 1997. 94(5): p. 1806-10.

63. Bernex, F., P. De Sepulveda, C. Kress, C. Elbaz, C. Delouis, and J.J. Panthier, Spatial and temporal patterns of c-kit-expressing cells in WlacZ/+ and WlacZ/WlacZ mouse embryos. Development, 1996. 122(10): p. 3023-33.

64. Bernstein, A., L. Forrester, A.D. Reith, P. Dubreuil, and R. Rottapel, The murine W/c-kit and Steel loci and the control of hematopoiesis. Semin Hematol, 1991. 28(2): p. 138-42.

65. Ogawa, M., S. Nishikawa, K. Yoshinaga, S. Hayashi, T. Kunisada, J. Nakao, T. Kina, T. Sudo, H. Kodama, and S. Nishikawa, Expression and function of c-Kit in fetal hemopoietic progenitor cells: transition from the early c-Kit-independent to the late c-Kit-dependent wave of hemopoiesis in the murine embryo. Development, 1993. 117(3): p. 1089-98.

66. Lansdorp, P.M., W. Dragowska, and H. Mayani, Ontogeny-related changes in proliferative potential of human hematopoietic cells. J Exp Med, 1993. 178(3): p. 787-91.

67. Rich, I.N. and B. Kubanek, The ontogeny of erythropoiesis in the mouse detected by the erythroid colony-forming technique. II. Transition in erythropoietin sensitivity during development. J Embryol Exp Morphol, 1980. 58: p. 143-55.

68. Bozzini, C.E., M.E. Barrio Rendo, F.C. Devoto, and C.E. Epper, Studies on medullary and extramedullary erythropoiesis in the adult mouse. Am J Physiol, 1970. 219(3): p. 724-8.

69. Hillman, R.S.a.C.A.F., Red Cell Manual. 7 ed. 1996, Philadelphia: F.A. Davis Company.

70. Akashi, K., D. Traver, T. Miyamoto, and I.L. Weissman, A clonogenic common myeloid progenitor that gives rise to all myeloid lineages. Nature, 2000. 404(6774): p. 193-7.

71. Traver, D., T. Miyamoto, J. Christensen, J. Iwasaki-Arai, K. Akashi, and I.L. Weissman, Fetal liver myelopoiesis occurs through distinct, prospectively isolatable progenitor subsets. Blood, 2001. 98(3): p. 627-35.

72. Paulson, R.F., Personal communication, L.E. Lenox, Editor. 2005: State College, PA.

73. Koury, M.J., S.T. Sawyer, and S.J. Brandt, New insights into erythropoiesis. Curr Opin Hematol, 2002. 9(2): p. 93-100.

74. Melkun, E., M. Pilione, and R.F. Paulson, A naturally occurring point substitution in Cdc25A, and not Fv2/Stk, is associated with altered cell-cycle status of early erythroid progenitor cells. Blood, 2002. 100(10): p. 3804-11.

75. Suzuki, S. and A.A. Axelrad, Fv-2 locus controls the proportion of erythropoietic progenitor cells (BFU-E) synthesizing DNA in normal mice. Cell, 1980. 19(1): p. 225-36.

Page 70: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

60

76. Hara, H. and M. Ogawa, Erythropoietic precursors in mice under erythropoietic stimulation and suppression. Exp Hematol, 1977. 5(2): p. 141-8.

77. Israels, L.G., and E.D. Israels, Erythropoieisis: an overview, in Erythropoietins and erythropoiesis: molecular, cellular, preclinical, and clinical biology, G. Molineux, M. Foote, and S.G. Elliott, Editors. 2003, Birkhèauser Verlag: Basel; New York. p. 3-14.

78. Gregory, C.J. and A.C. Eaves, Three stages of erythropoietic progenitor cell differentiation distinguished by a number of physical and biologic properties. Blood, 1978. 51(3): p. 527-37.

79. Ebert, B.L. and H.F. Bunn, Regulation of the erythropoietin gene. Blood, 1999. 94(6): p. 1864-77.

80. Ihle, J.N., Signal Transduction in the Regulation of Hematopoeisis, in The Molecular Basis of Blood Disease, G. Stamatoyannopoulos, et al., Editors. 2001, WB Saunders, Co.: Philadelphia. p. 103-134.

81. Ghaffari, S., Jun-shen Huang, L., Zhang, J., and Lodish, H.F., Erythropoietin receptor signaling processes, in Erythropoietins and erythropoiesis: molecular, cellular, preclinical, and clinical biology, G. Molineux, M. Foote, and S.G. Elliott, Editors. 2003, Birkhèauser Verlag: Basel; New York. p. 65-85.

82. Koury, M., Kelley, L., and Bondurant, M., The Fate of Erythroid Progenitor Cells. Annals of the New York Academy of Sciences: Molecular, Cellular, and Developmental Biology of Erythropoietin and Erythropoiesis, 1994. 718: p. 259-267.

83. DeMartino, J., Martin, C., Mathey-Prevot, B., and D'Andrea, A., Erythropoietin Receptor Contains both Growth-promoting Activity and Differentaition-promoting Activity. Annals of the New York Academy of Sciences: Molecular, Cellular, and Developmental Biology of Erythropoietin and Erythropoiesis, 1994. 718: p. 213-221.

84. Kaushansky, K., Hematopoietic Growth Factors and Receptors, in The Molecular Basis of Blood Disease, G. Stamatoyannopoulos, et al., Editors. 2001, WB Saunders, Co.: Philadelphia. p. 25-79.

85. Kieran, M.W., A.C. Perkins, S.H. Orkin, and L.I. Zon, Thrombopoietin rescues in vitro erythroid colony formation from mouse embryos lacking the erythropoietin receptor. Proc Natl Acad Sci U S A, 1996. 93(17): p. 9126-31.

86. Molineux, G., Biology of erythropoietin, in Erythropoietins and erythropoiesis: molecular, cellular, preclinical, and clinical biology, G. Molineux, M. Foote, and S.G. Elliott, Editors. 2003, Birkhèauser Verlag: Basel; New York. p. 113-132.

87. Dame, C., Molecular biology of the erythropoietin receptor in hematopoietic and non-hematopoietic tissues, in Erythropoietins and erythropoiesis: molecular, cellular, preclinical, and clinical biology, G. Molineux, M. Foote, and S.G. Elliott, Editors. 2003, Birkhèauser Verlag: Basel; New York. p. 35-64.

88. Koury, S.T., M.J. Koury, M.C. Bondurant, J. Caro, and S.E. Graber, Quantitation of erythropoietin-producing cells in kidneys of mice by in situ hybridization: correlation with hematocrit, renal erythropoietin mRNA, and serum erythropoietin concentration. Blood, 1989. 74(2): p. 645-51.

Page 71: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

61

89. Bozzini, C.E., Alippi, R. M., Barcelo, A. C., Conti, M. I., Bozzini, C., Lezon, C. E., Olivera, M. I., The Biology of Stress Erythropoiesis and Erythropoietin Production. Annals of the New York Academy of Sciences: Molecular, Cellular, and Developmental Biology of Erythropoietin and Erythropoiesis, 1994. 718: p. 83-92.

90. Guillemin, K. and M. Krasnow, The hypoxic response: Huffing and HIFing. Cell, 1997. 89: p. 9-12.

91. Galson, D., Blanchard, K., Fandrey, J., Goldberg, M. and Bunn, H.F., Cis Elements That Regulate the Erythropoietin Gene. Annals of the New York Academy of Sciences: Molecular, Cellular, and Developmental Biology of Erythropoietin and Erythropoiesis, 1994. 718: p. 21-30.

92. Blanchard, K.L., A.M. Acquaviva, D.L. Galson, and H.F. Bunn, Hypoxic induction of the human erythropoietin gene: cooperation between the promoter and enhancer, each of which contains steroid receptor response elements. Mol Cell Biol, 1992. 12(12): p. 5373-85.

93. Pugh, C.W., B.L. Ebert, O. Ebrahim, and P.J. Ratcliffe, Characterisation of functional domains within the mouse erythropoietin 3' enhancer conveying oxygen-regulated responses in different cell lines. Biochim Biophys Acta, 1994. 1217(3): p. 297-306.

94. Semenza, G.L. and G.L. Wang, A nuclear factor induced by hypoxia via de novo protein synthesis binds to the human erythropoietin gene enhancer at a site required for transcriptional activation. Mol Cell Biol, 1992. 12(12): p. 5447-54.

95. Galson, D.L., T. Tsuchiya, D.S. Tendler, L.E. Huang, Y. Ren, T. Ogura, and H.F. Bunn, The orphan receptor hepatic nuclear factor 4 functions as a transcriptional activator for tissue-specific and hypoxia-specific erythropoietin gene expression and is antagonized by EAR3/COUP-TF1. Mol Cell Biol, 1995. 15(4): p. 2135-44.

96. Wang, G.L., B.H. Jiang, E.A. Rue, and G.L. Semenza, Hypoxia-inducible factor 1 is a basic-helix-loop-helix-PAS heterodimer regulated by cellular O2 tension. Proc Natl Acad Sci U S A, 1995. 92(12): p. 5510-4.

97. Semenza, G.L., HIF-1 and mechanisms of hypoxia sensing. Curr Opin Cell Biol, 2001. 13(2): p. 167-71.

98. Huang, L.E., J. Gu, M. Schau, and H.F. Bunn, Regulation of hypoxia-inducible factor 1alpha is mediated by an O2-dependent degradation domain via the ubiquitin-proteasome pathway. Proc Natl Acad Sci U S A, 1998. 95(14): p. 7987-92.

99. Semenza, G., HIF - 1 and human disease: one highly involved factor. Genes & Dev, 2000. 14: p. 1983-1991.

100. Yoshida, E., S. Aratani, H. Itou, M. Miyagishi, M. Takiguchi, T. Osumu, K. Murakami, and A. Fukamizu, Functional association between CBP and HNF4 in trans-activation. Biochem Biophys Res Commun, 1997. 241(3): p. 664-9.

101. Cheung, J.Y. and B.A. Miller, Molecular mechanisms of erythropoietin signaling. Nephron, 2001. 87(3): p. 215-22.

102. Sawyer, S., Introduction: The Erythropoietin Receptor and Signal Transduction. Annals of the New York Academy of Sciences: Molecular, Cellular, and

Page 72: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

62

Developmental Biology of Erythropoietin and Erythropoiesis, 1994. 718: p. 185-190.

103. Whitelaw, E., S.F. Tsai, P. Hogben, and S.H. Orkin, Regulated expression of globin chains and the erythroid transcription factor GATA-1 during erythropoiesis in the developing mouse. Mol Cell Biol, 1990. 10(12): p. 6596-606.

104. Stamatoyannopoulos, G., and Grosveld, F., Hemoglobin Switching, in The Molecular Basis of Blood Disease, G. Stamatoyannopoulos, et al., Editors. 2001, WB Saunders, Co.: Philadelphia. p. 135-182.

105. Gourdon, G., J.A. Sharpe, D.R. Higgs, and W.G. Wood, The mouse alpha-globin locus regulatory element. Blood, 1995. 86(2): p. 766-75.

106. Trimborn, T., J. Gribnau, F. Grosveld, and P. Fraser, Mechanisms of developmental control of transcription in the murine alpha- and beta-globin loci. Genes Dev, 1999. 13(1): p. 112-24.

107. Alami, R., M. Bender, Y. Feng, S. Fiering, B. Hug, T. Ley, M. Groudine, and E. Bouhassira, Deletions within the mouse beta-globin locus control region preferentially reduce beta(min) globin gene expression. Genomics, 2000. 63: p. 417-424.

108. Bulger, M. and M. Groudine, Looping versus linking: toward a model for long-distance gene activation. Genes Dev, 1999. 13(19): p. 2465-77.

109. Higgs, D.R., Do LCRs open chromatin domains? Cell, 1998. 95(3): p. 299-302. 110. Hug, B.A., R.L. Wesselschmidt, S. Fiering, M.A. Bender, E. Epner, M. Groudine,

and T.J. Ley, Analysis of mice containing a targeted deletion of beta-globin locus control region 5' hypersensitive site 3. Mol Cell Biol, 1996. 16(6): p. 2906-12.

111. Kioussis, D. and R. Festenstein, Locus control regions: overcoming heterochromatin-induced gene inactivation in mammals. Curr Opin Genet Dev, 1997. 7(5): p. 614-9.

112. Miller, J.L., Hemoglobin switching and modulation: genes, cells, and signals. Curr Opin Hematol, 2002. 9(2): p. 87-92.

113. Navas, P.A., K.R. Peterson, Q. Li, M. McArthur, and G. Stamatoyannopoulos, The 5'HS4 core element of the human beta-globin locus control region is required for high-level globin gene expression in definitive but not in primitive erythropoiesis. J Mol Biol, 2001. 312(1): p. 17-26.

114. Lyon, M.F., J.E. Barker, and R.A. Popp, Mouse globin gene nomenclature. J Hered, 1988. 79(2): p. 93-5.

115. Russell, E.S., and McFarland, E. C., Genetics of mouse hemoglobins. Annals of the New York Academy of Sciences, 1974. 241: p. 25-38.

116. Erhart, M.A., K.S. Simons, and S. Weaver, Evolution of the mouse beta-globin genes: a recent gene conversion in the Hbbs haplotype. Mol Biol Evol, 1985. 2(4): p. 304-20.

117. Weaver, S., M.B. Comer, C.L. Jahn, C.A. Hutchison, 3rd, and M.H. Edgell, The adult beta-globin genes of the "single" type mouse C57BL. Cell, 1981. 24(2): p. 403-11.

118. Jahn, C.L., C.A. Hutchison, 3rd, S.J. Phillips, S. Weaver, N.L. Haigwood, C.F. Voliva, and M.H. Edgell, DNA sequence organization of the beta-globin complex in the BALB/c mouse. Cell, 1980. 21(1): p. 159-68.

Page 73: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

63

119. Trimborn, T., J. Gribnau, F. Grosveld, and P. Fraser, Mechanisms of developmental control of transcription in the murine alpha and beta-globin loci. Genes and Development, 1999. 13: p. 112-124.

120. Bender, M.A., M. Bulger, J. Close, and M. Groudine, Beta-globin gene switching and DNase I sensitivity of the endogenous beta-globin locus in mice do not require the locus control region. Mol Cell, 2000. 5(2): p. 387-93.

121. Gillemans, N., R. Tewari, F. Lindeboom, R. Rottier, T. de Wit, M. Wijgerde, F. Grosveld, and S. Philipsen, Altered DNA-binding specificity mutants of EKLF and Sp1 show that EKLF is an activator of the beta-globin locus control region in vivo. Genes Dev, 1998. 12(18): p. 2863-73.

122. Whitney, J.B., 3rd, Differential control of the synthesis of two hemoglobin beta chains in normal mice. Cell, 1977. 12(4): p. 863-71.

123. Wawrzyniak, C. and R. Popp, Expression of the two adult β-globin genes in mouse yolk sac and fetal liver. Developmental Biology, 1987. 119: p. 299-301.

124. Oliveri, N. and D. Weatherall, The therapeutic reactivation of fetal hemoglobin. Human Molecular Genetics, 1998. 7: p. 1655-1658.

125. Brandan, N., M. Aguirre, R. Carmuega, M. Alvarez, and J. Juaristi, Proliferative and maturative behaviour patterns on murine bone marrow and spleen erythropoiesis along hypoxia. Acta Physiol Pharmacol Ther Latinoam, 1997. 47(2): p. 125-35.

126. Broudy, V., N. Lin, G. Priestley, K. Nocka, and N. Wolf, Interaction of stem cell factor and its receptor c-kit mediates lodgement and acute expansion of hematopoietic cells in the murine spleen. Blood, 1996. 88: p. 75-81.

127. Hara, H. and M. Ogawa, Erythropoietic precursors in mice with phenylhydrazine-induced anemia. American Journal of Hematology, 1976. 1: p. 453-458.

128. Ploemacher, R.E. and P.L. van Soest, Morphological investigation on phenylhydrazine-induced erythropoiesis in the adult mouse liver. Cell Tissue Res, 1977. 178(4): p. 435-61.

129. Rich, I. and B. Kubanek, The effect of reduced oxygen tension on colony formation of erythropoietic cells in vitro. British Journal of Haematology, 1982. 52: p. 579-588.

130. Sadahira, Y., T. Yasuda, T. Yoshino, T. Manabe, T. Takeishi, Y. Kobayashi, Y. Ebe, and M. Naito, Impaired splenic erythropoiesis in phlebotomized mice injected with CL2MDP-liposome: an experimental model for studying the role of stromal macrophages in erythropoiesis. J Leukoc Biol, 2000. 68(4): p. 464-70.

131. Lenox, L.E., J.M. Perry, and R.F. Paulson, BMP4 and Madh5 regulate the erythroid response to acute anemia. Blood, 2005. 105(7): p. 2741-8.

132. von Lindern, M., W. Zauner, G. Mellitzer, P. Steinlein, G. Fritsch, K. Huber, B. Lowenberg, and H. Beug, The glucocorticoid receptor cooperates with the erythropoietin receptor and c-Kit to enhance and sustain proliferation of erythroid progenitors in vitro. Blood, 1999. 94(2): p. 550-9.

133. Wessely, O., A. Bauer, C.T. Quang, E.M. Deiner, M. von Lindern, G. Mellitzer, P. Steinlein, J. Ghysdael, and H. Beug, A novel way to induce erythroid progenitor self renewal: cooperation of c-Kit with the erythropoietin receptor. Biol Chem, 1999. 380(2): p. 187-202.

Page 74: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

64

134. Wessely, O., E.M. Deiner, H. Beug, and M. von Lindern, The glucocorticoid receptor is a key regulator of the decision between self-renewal and differentiation in erythroid progenitors. Embo J, 1997. 16(2): p. 267-80.

135. Bauer, A., F. Tronche, O. Wessely, C. Kellendonk, H. Reichardt, P. Steinlein, G. Schutz, and H. Beug, The glucocorticoid receptor is required for stress erythropoiesis. Genes Dev, 1999. 13: p. 2996-3002.

136. Harrison, D. and E. Russell, The response of W/Wv and Sl/Sld amaemic mice to hematopoietic stimuli. Br. J. Haematol, 1972. 22: p. 155-167.

137. Geissler, E.N., M.A. Ryan, and D.E. Housman, The dominant-white spotting (W) locus of the mouse encodes the c-kit proto-oncogene. Cell, 1988. 55(1): p. 185-92.

138. Huang, E., K. Nocka, D.R. Beier, T.Y. Chu, J. Buck, H.W. Lahm, D. Wellner, P. Leder, and P. Besmer, The hematopoietic growth factor KL is encoded by the Sl locus and is the ligand of the c-kit receptor, the gene product of the W locus. Cell, 1990. 63(1): p. 225-33.

139. Zhu, H. and H. Bunn, Oxygen sensing and signaling: impact on the regulation of physiologically important genes. Respiration Physiology, 1999. 115: p. 239-247.

140. Zsebo, K.M., D.A. Williams, E.N. Geissler, V.C. Broudy, F.H. Martin, H.L. Atkins, R.Y. Hsu, N.C. Birkett, K.H. Okino, D.C. Murdock, and et al., Stem cell factor is encoded at the Sl locus of the mouse and is the ligand for the c-kit tyrosine kinase receptor. Cell, 1990. 63(1): p. 213-24.

141. Russell, E., Hereditary Anemias of the Mouse: A review for Geneticists. Advances in Genetics, 1979. 20: p. 357-459.

142. Axelrod, J. and T.D. Reisine, Stress hormones: their interaction and regulation. Science, 1984. 224(4648): p. 452-9.

143. Tronche, F., C. Kellendonk, H.M. Reichardt, and G. Schutz, Genetic dissection of glucocorticoid receptor function in mice. Curr Opin Genet Dev, 1998. 8(5): p. 532-8.

144. Golde, D.W., N. Bersch, and M.J. Cline, Potentiation of erythropoiesis in vitro by dexamethasone. J Clin Invest, 1976. 57(1): p. 57-62.

145. Udupa, K.B., H.M. Crabtree, and D.A. Lipschitz, In vitro culture of proerythroblasts: characterization of proliferative response to erythropoietin and steroids. Br J Haematol, 1986. 62(4): p. 705-14.

146. McLeod, D.L., M.M. Shreeve, and A.A. Axelrad, Chromosome marker evidence for the bipotentiality of BFU-E. Blood, 1980. 56(2): p. 318-22.

147. Vannucchi, A.M., F. Paoletti, S. Linari, C. Cellai, R. Caporale, P.R. Ferrini, M. Sanchez, G. Migliaccio, and A.R. Migliaccio, Identification and characterization of a bipotent (erythroid and megakaryocytic) cell precursor from the spleen of phenylhydrazine-treated mice. Blood, 2000. 95(8): p. 2559-68.

148. Ploemacher, R.E., P.L. van Soest, and O. Vos, Kinetics of erythropoiesis in the liver induced in adult mice by phenylhydrazine. Scand J Haematol, 1977. 19(5): p. 424-34.

149. Yanai, N., Y. Matsuya, and M. Obinata, Spleen stromal cell lines selectively support erythroid colony formation. Blood, 1989. 74(7): p. 2391-7.

Page 75: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

65

150. Obinata, M. and N. Yania, Cellular and molecular regulation of erythropoietic inductive microenvironment (EIM). Cell Structure and Function, 1999. 24: p. 171-179.

151. Yanai, N., T. Satoh, and M. Obinata, Endothelial cells create a hematopietic inductive microenvironment preferential to eythropoiesis in the mouse spleen. Cell Structure and Function, 1991. 16: p. 87-93.

152. Dexter, T.M., E. Spooncer, R. Schofield, B.I. Lord, and P. Simmons, Haemopoietic stem cells and the problem of self-renewal. Blood Cells, 1984. 10(2-3): p. 315-39.

153. Chagraoui, J., A. Lepage-Noll, A. Anjo, G. Uzan, and P. Charbord, Fetal liver stroma consists of cells in epithelial-to-mesenchymal transition. Blood, 2003. 101(8): p. 2973-82.

154. Hamamura, K., H. Matsuda, Y. Takeuchi, S. Habu, H. Yagita, and K. Okumura, A critical role of VLA-4 in erythropoiesis in vivo. Blood, 1996. 87(6): p. 2513-7.

155. Ohneda, O., N. Yanai, and M. Obinata, Combined action of c-kit and erythropoietin on erythroid progenitor cells. Development, 1992. 114(1): p. 245-52.

156. Slaper-Cortenbach, I., R. Ploemacher, and B. Lowenberg, Different stimulative effects of human bone marrow and fetal liver stromal cells on erythropoiesis in long-term culture. Blood, 1987. 69(1): p. 135-9.

157. Hunt, H. and D. Premar, Flexed-tail a mutation in the house mouse. Anat. Rec., 1928. 41: p. 117.

158. Russell, E.S., Margaret W. Thompson, and Eleanor C. McFarland, Analysis of effects of W and f genic substitutions on fetal mouse hematology. Genetics, 1968. 58: p. 259-270.

159. Gruneberg, H., The anaemia of the flexed-tail mouse (Mus musculus L.) I. Static and dynamic haematology. Journal of Genetics, 1942. 43: p. 45-68.

160. Gruneberg, H., The anaemia of the flexed-tail mice (Mus musculus L.) II. Siderocytes. Journal of Genetics, 1942. 44: p. 246-271.

161. Cole, R., J. Garlick, and R. Tarbutt, Disturbed haem and globin synthesis in reticulocytes of prenatal flexed-tail (f/f) anaemic mice. Genet. Res., Camb., 1974. 23: p. 125-135.

162. Chui, D., G. Sweeney, M. Patterson, and E. Russell, Hemoglobin synthesis in siderocytes of flexed-tail mutant (f/f) fetal mice. Blood, 1977. 50: p. 165-177.

163. Bateman, A. and R. Cole, Colony forming cells in the livers of prenatal flexed (f/f) anaemic mice. Cell Tissue Kinet., 1972. 5: p. 165-173.

164. Cole, R. and T. Regan, Haematopoietic Progenitor cells in the prenatal conegenitally anaemic "Flexed-tail" (f/f) mice. British Journal of Haematology, 1976. 33: p. 387-394.

165. Coleman, D., E. Russell, and E. Levin, Enzymatic studies of the hemopoietic defect in flexed mice. Genetics, 1969. 61: p. 631-642.

166. Gregory, C., E. McCulloch, and J. Till, The Cellular Basis for the defect in haematopoiesis in flexed-tail mice. III Restriction of the defect to erythropoietic progenitors capable of transient colony formation in vivo. British Journal of Haematology, 1975. 30: p. 401-410.

Page 76: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

66

167. Thompson, M., E. McCulloch, L. Siminovitch, and J. Till, The cellular basis for the defect in haematopoiesis in flexed-tail mice. I. the nature and persistence of the defect. Br J Haematol, 1966. 12: p. 152-160.

168. Cole, R., T. Regan, and R. Tarbutt, Haemoglobin synthesis in reticulocytes of prenatal f/f anaemic mice. B J Haematol, 1972. 23: p. 443-452.

169. Fowler, J., J. Till, E. McCulloch, and L. Siminovitch, The cellular basis for the defect in haematopoiesis in flexed-tail mice. II. The specificity of the defect for erythropoiesis. Br. J Haematol, 1967. 13: p. 256-264.

170. Lenox, L.E., hematocrit values during the recovery from a PHZ induced acute anemia in +/+, f/+ and f/f mice, Penn State University.

Page 77: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

67

Figu

re 1

-1.

The

hem

atop

oiet

ic sy

stem

.(Fr

om th

e Lu

nd S

tem

Cel

lCen

ter:

http

://w

ww

.mol

med

.lu.se

/hem

atop

oies

is_p

00.h

tm)

Page 78: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

68

Figure 1-2. Dual stem cell model of hematopoietic stem cell ontogeny.(From Palis and Yoder, 2001 [1]; hypothesis proposed by Godin et al. 1995. Proc. Natl. Acad. Sci. USA 92: 773-777). Mesoderm (M) cells spread laterally from the primitive streak to contribute to the formation of extra-embryonic and intra-embryonic tissue. Extra-embryonic yolk sac HSCs and intra-embryonic HSCs arise from mesoderm (M) during gastrulation. The aorta-gonad-mesonephros region (AGM) differentiates from paraaortic splanchnopleura(P-SP). Hematopoietic cells will seed the fetal liver once circulation has been established. Finally, cells will migrate to the bone marrow which remains the main site of hematopoiesis in the adult.

Page 79: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

69

Figu

re 1

-3.

Dev

elop

men

t and

feat

ures

of t

he ‘h

emog

enic

’ reg

ions

dur

ing

embr

yoge

nesi

s. (F

rom

Gar

cia-

Porr

ero

et a

l. 19

98 [1

6]).

Hem

ogen

icre

gion

s giv

e ris

e to

hem

atop

oiet

ic c

ells

from

the

extra

-em

bryo

nic

yolk

sac

(YS)

at e

mbr

yoni

c da

y 7.

5 (E

7.5)

and

intra

-em

bryo

nic

regi

on sp

lanc

hnop

leur

a (S

p) a

t the

pre

som

itic

stag

e (E

7.5)

. A

fter t

he 1

0-12

som

ite st

age,

in

tra-e

mbr

yoni

c he

mat

opoi

etic

pot

entia

l is r

estri

cted

to th

e pa

raao

rtic

spla

nchn

ople

ura

(P-S

p), w

hich

will

giv

e ris

e to

the

aorta

, gon

ads a

nd m

eson

ephr

os (A

GM

regi

on) f

ollo

win

g or

gano

gene

sis.

Dep

icte

d ar

e th

e po

tent

ials

of t

he h

emat

opoi

etic

ce

lls a

t the

var

ious

loca

tions

.

Page 80: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

(SCL/Tal1)(LMO2)

(GATA2)

(EKLF)(GATA1)

(KIT)

70

(SCL/Tal1)(LMO2)(GATA2)

(EKLF)(GATA1)

Figure 1-4. Stages of erythroid development. (From Koury, MJ et al. 2002 [73 ] and Israels, LG and Israels, ED 2003 [75 ]). Erythropoiesis from CFU-GEMM to erythrocyte; developmental stages, regulatory cytokines, cell-surface receptors, erythroid transcription factors and hemoglobin synthesis. Basic helix-loop-helix factor (SCL/TAL1); LIM-domain partner of TAL1 (LMO2); zinc finger factors that bind GATA sequences (GATA-1, GATA-2); erythroid Krüppel-like factor; stem cell factor receptor (KIT).

Page 81: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

71A)

B)

Figure 1-5. Erythropoietin: regulation and signaling. The 3’ enhancer element of the Epo gene is highly conserved between the murine and human systems. Shown in (A) is an expanded view of this enhancer in the human gene. Sites that are functionally critical for hypoxic induction are underscored in red. Binding of HIF-1, HNF-4, and p300 is illustrated (From Ebert, BL and Bunn, F 1999 [77]). (B) Model of the Epo receptor. Important residues or motifs are marked on the left side of the figure (cysteine: C; tyrosine: Y; arginine: R; serine: S; tryptophan: W), including their position. Important signaling pathways and negative regulatory loop are indicated on the right side of the model. P indicates the phosphorylation necessary for the activation of the pathway (From Dame, C 2003 [85]).

Page 82: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

72A)

bh2 bh3(chr7)

(βmaj) (βmin)

B)

Figure 1-6. Schematic of globin genes. (A) Murine α-globin loci. Genes are represented by solid boxes and vertical arrows represent DNase I hypersensitive sites (From Trimborn, T. et al. 1999 [104]) (B) Comparison of the human and mouse β-Globin clusters. The LCR is shown as a yellow box; numbers 1-6 represent the upstream hypersensitive sites. The horizontal green (embryos), blue (fetus) and red (adult) lines represent the developmental stages at which the various genes are expressed; the pseudogenes bh2 and bh3 are shown as gray lines; (βmaj and βmin) are the gene products from the b1 and b2 genes respectively in Hbbd (diffuse) mice (From Higgs, D. R. 1998 [107]).

Page 83: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

73

A B

Figure 1-7. CFU-E in the fetal liver of f/f and +/+ mice. (A) Numbers of CFU-E (able to form colonies of 16 or more cells) in livers of prenatal f/f ( ) and +/+ ( ) mice and numbers of colonies formed from +/+ ( ) and f/f ( ) livers without additional erythropoietin. (B) Proportions of CFU-E in livers of prenatal f/f and normal mice. (Cole and Regan 1976 [161]).

Page 84: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

74

Figure 1-8. Percentage of siderocytes in reticulocytes of f/f and f/+ control fetal and neonatal mice. The solid lines refers to anemics (f/f), the broken line to their normal siblings (f/+) (Gruneberg; II Siderocytes, 1942 [157]).

Page 85: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

75

Figure 1-9. Diagram of hemoglobin biosynthesis. (Taken from http://cls.umc.edu/COURSES/CLS312/hgbsynth.doc).

Coproporphyrinogen Oxidase(in mitochondria)

(in cytoplasm)

(in mitochondria)

Page 86: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

76

Figure 1-10. Chromatography of globin chains synthesized by day 18 +/+ (A) and f/f (B) fetal erythrocytes. Cells were labeled with 3H-leucine for 90 min. Non-isotopic adult hemolysate was added before globin chains were prepared and isolated. There were more +/+ than f/f fetal erythrocytes collected for this set of experiments, which accounted for the higher incorporation of 3H-leucine into globin chains in +/+ cells (Chui, et al. 1977 [159]).

Page 87: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

77

Figure 1-11. Incorporation of 59Fe into heme from CFU-S of f/f and +/+ mice.Incorporation of 59Fe into heme by cells derived from the spleens of irradiated (C3HXC57BL)F1 hosts which received transplants of 2X106 marrow cells (containing appox. 300 CFU), obtained from f/f ( ) or +/+ ( ) donors. 59 Fe incorporation is expressed as that percentage of the activity added to the incubation tube which was incorporated by 1.2X107 spleen cells in 45 minutes. The combined results of five experiments are given and the limits indicated are standard errors of the means (Thompson, M., et al. 1966. [164]).

Page 88: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

78

Figure 1-12. Changes in spleen weight and reticulocyte counts in flexed and wild-type mice recovering from phenylhydrazine induced acute anemia. Each point represents the average value obtained from at least 10 mice (Coleman, D.L. et al. 1969 [162]).

Page 89: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

A)+/+f/f

79

AL

D A

ctiv

ity (U

nits

/ g

Sple

en)

B)

Figure 1-13. Effects of phenylhydrazine treatment on enzymes of the hemoglobin biosynthetic pathway in spleen from flexed and wild-type mice. A) δ-aminolevulinate dehydratase (ALD) activity. Units of activity are µmoles of porphobilinogen produced per hour per g of spleen. Each point represents the average value obtained from 4 to 8 separate assays performed on individual spleens. B) Uroporphyrinogen synthetase activity. Each point represents the average value obtained from 4 to 6 separate assays. Each assay was run on the combined homogenates of spleen from 2 to 3 treated mice (Coleman, D.L. et al. 1969 [162]).

Page 90: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

80A)

B)

Figure 1-14. Transient endogenous spleen colonies (TE-CFU) in the spleens of wild-type and flexed-tail mice. A) +/+ or f/f mice (seven per group) were exposed to 800 rads of radiation followed by 10 units of Epo; spleens were examined 5 days later for TE-CFU. To test whether the marked differences seen between f/f and +/+ mice were cell autonomous, the experiments were repeated using chimeras prepared by repopulating B6C3F1 mice with bone marrow derived from either genotype. B) To examine the kinetics of endogenous colony formation in more detail, time course experiments were performed. Shown are the colony formation in the spleens of +/+ ( ) or f/f ( ) chimeras given 800 rads and 10 Units of Epo at time zero. Values graphed are the means ±SE for groups of four mice per point except on day 11 and 12 where only two or three mice remained (Gregory, C.J. et al.1975 [163]).

Page 91: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

81

Chapter 2

MAPPING OF THE flexed-tail (f) LOCUS LEADS TO THE DISCOVERY THAT BMP4 AND Madh5 REGULATE THE ERYTHROID RESPONSE TO ACUTE

ANEMIA

Forward

To clone the gene responsible for a mutation, first the locus must be

mapped to a precise interval on a particular chromosome. Candidate genes are selected,

and any mutations found within the candidate gene must be shown to co-segregate with

the mutant phenotype. It follows that biochemical evidence should be presented to verify

that the mutated gene is causing the phenotype. This could include variations of

transcripts or protein levels between wild-type and mutant animals or even recapitulating

the defective function and/or rescue of the defect in vitro. More conclusive than

biochemical evidence is functional evidence where a critical mechanism is shown to be

disrupted by a mutation in the candidate gene, which can be rescued by the wild-type

gene in vivo. Showing allelism with other known mutations is also a way to functionally

support the fact your candidate gene is the cause of the mutant phenotype.

Madh5 (Smad5) has been shown through in vivo and in vitro functional data to be

the flexed-tail locus. The complexity of the mutation and the genomic structure of

Smad5, as well as the nuances of the Smad5/BMP4 signaling pathway have surfaced

through various strategies used to prove f=Smad5. Appendix A contains information on

some of the experiments performed throughout the course of the flexed-tail project.

Page 92: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

82

Although not exhaustive, it describes the most relevant experiments devised to either

locate the mutation or provide evidence for Smad5 being the f locus. Due to the lack of

positive results, inconsistencies or ambiguities obtained, they were not contained in the

published Blood manuscript.

John Perry and I were co-first authors on the above mentioned Blood manuscript

(Lenox, L.E., J.M. Perry and R.F. Paulson, BMP4 and Madh5 regulate the erythroid response to acute

anemia. Blood, 2005. 105(7): p. 2741-8.) The following sections are directly from this

manuscript. The colony assays, including those from sorted cells, and the BMP4

expression data were the work of John M. Perry.

Page 93: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

83

Abstract

Acute anemia initiates a systemic response that results in the rapid mobilization

and differentiation of erythroid progenitors in the adult spleen. flexed-tail (f) mutant mice

exhibit normal steady state erythropoiesis, but are unable to rapidly respond to acute

erythropoietic stress. Here we show that f/f mutant mice have a mutation in Madh5. Our

analysis shows that BMP4/Madh5 dependent signaling, regulated by hypoxia, initiates

the differentiation and expansion of erythroid progenitors in the spleen. These findings

suggest a new model where stress erythroid progenitors, resident in the spleen are poised

to respond to changes in the microenvironment induced by acute anemia.

Page 94: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

84

Introduction

Erythropoiesis in the bone marrow is primarily homeostatic; however, the

situation is dramatically different in the adult spleen in response to acute erythropoietic

stress, where rapid, expansive erythropoiesis occurs. Previous work suggested a model

where acute anemia leads to tissue hypoxia, which induces erythropoietin (Epo)

expression in the kidney. Increased levels of serum Epo mobilize cells from the bone

marrow, which migrate to the spleen where they expand and differentiate [1, 2]. The

spleen contains a unique microenvironment that can support expansive erythropoiesis [3].

However the signals that regulate the increase in splenic erythropoiesis in response to

acute anemia are not clear. The expansive erythropoiesis observed in the adult spleen is

similar to fetal liver erythropoiesis during development [4]. In both cases rapid erythroid

development occurs.

Similar to the spleen, fetal liver stromal cells are capable of supporting the

expansion of erythroid progenitors [5]. Because of these common features it has been

suggested that splenic and fetal liver erythropoiesis may be mechanistically similar. This

link between the fetal liver and spleen is apparent in mice with a mutation at the flexed-

tail (f) locus. During fetal development, f/f mutant embryos exhibit a severe microcytic,

hypochromic anemia [6-8]. f/f fetal livers contain about 50% the normal number of

erythroid progenitors [9,10] and have a maturation defect, which results in the production

of large numbers of siderocytes or erythrocytes that contain non-heme iron granules

[8,11]. Despite these defects, the anemia of f/f mice resolves about two weeks after birth.

Adult f/f mice exhibit normal numbers of steady state erythroid progenitors [12].

However, they are unable to respond rapidly to acute erythropoietic stress. This defect is

Page 95: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

85

manifested by a delay in the expansion of erythroid progenitors in the spleen and a delay

in the appearance of reticulocytes following Phenylhydrazine (PHZ) induced acute

anemia [13]. Despite the delayed response, adult f/f mice do not exhibit the maturation

defect present during fetal liver erythropoiesis because no siderocytes are observed

during the recovery from acute anemia [14]. These observations demonstrate that the f

gene product plays a key role in regulating the expansion and maturation of erythroid

progenitors at times of great erythropoietic need.

Previous work has suggested that f/f mice have a mutation in sideroflexin 1

(sfxn1), a putative mitochondrial transporter, which is proposed to play a role in the

transport of molecules required for heme biosynthesis [15]. In this report, however, we

show that f/f mice have a mutation in the Madh5 gene, which directly affects the ability

of f/f mice to respond to acute anemia. Madh5 functions as a receptor activated Smad

downstream of the BMP2, 4 and 7 receptors [16,17]. Previous work has implicated

BMP’s and in particular BMP4 in the development of mesodermal cells that will give rise

to hematopoietic cells early in development [18]. Our work shows that in response to

acute anemia, BMP4 is rapidly induced in the spleen. BMP4 acts on an immature

progenitor cell causing it to differentiate into an Epo responsive stress erythroid

progenitor. Cell sorting experiments showed that BMP4 responsive cells exhibit the same

cell surface phenotype as the bone marrow derived Megakaryocyte-Erythroid progenitors

(MEPs) [19], however, only spleen MEPs respond to BMP4. These results demonstrate

that these spleen progenitors exhibit properties that are distinct from bone marrow

erythroid progenitors suggesting that they represent a population of “stress erythroid

Page 96: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

86

progenitors” resident in the spleen whose function is to rapidly generate erythrocytes at

times of great erythropoietic need.

Page 97: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

87

Methods

Mice. C57BL/6 and C57BL/6-f mice were obtained from Jackson Laboratory.

Madh5+/- mice were obtained from Dr. C. Deng [20]. All mice were approximately 6-8

weeks old, controls were age matched. Acute anemia was induced by injection of

Phenylhydrazine (Sigma, St. Louis, MO) at a concentration of 100 mg/Kg mouse in PBS.

Colony Assays for BFU-E. Splenocytes, bone marrow, and peripheral blood

cells were isolated from C57BL/6 control f/+ and f/f mice. 1x105/ml nucleated bone

marrow and peripheral blood cells and 2x106/ml nucleated splenocytes were plated in

methylcellulose media (StemCell Technologies, Vancouver, BC) containing 3U/ml Epo +

either 10 ng/ml IL3 (Sigma, St. Louis, MO) or 0.15-15 ng/ml BMP4 (R&D Systems,

Minneapolis, MN) where indicated. BFU-E were scored as described [21]. For the BMP4

pre-incubation experiment, splenocytes and bone marrow cells from C57BL/6 and f/f

mice were incubated for 24 hours in IMDM + 5% FCS + 15 ng/ml BMP4. Colony assays

were then performed as indicated above + 15 ng/ml BMP4 for each.

Characterization of the Epo sensitivity of the stress BFU-E. Colony assays

were performed as above on bone marrow and spleen cells in the presence of 0.1, 0.3, 1,

3, and 10 U/ml Epo as indicated. Additionally, bone marrow cells were supplemented

with 50 ng/ml SCF while splenocytes were supplemented with 15 ng/ml BMP4. Colonies

were scored as above.

Cloning of Madh5 mRNAs from f/f, f/+ and control mice. Total RNA was

isolated from single cell suspensions of spleen cells using the TRIzol reagent (Invitrogen,

Carlsbad, CA) according to manufacturer’s instructions. cDNA was generated and

Page 98: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

88

Madh5 cDNA was amplified using 5’-GGGGCCGAGCTGCTAAT-3’ and 5’-

CTATGAAACAGAAGAAATGGGG-3’ primers.

Analysis of BMP4 expression. Total RNA isolated from bone marrow or spleen

cells homogenized in TRIzol (Invitrogen, Carlsbad, CA) was reverse transcribed into

cDNA. PCR was performed using primers 5’-TGTGAGGAGTTTCCATCACG-3’ and

5’- TTATTCTTCTTCCTGGACCG-3’.

Staining of Spleen sections with anti-BMP4 antibodies. Spleens were harvested

on the indicated days post PHZ induced anemia, fixed in Zinc fixative and paraffin

embedded tissues sections were cut. The expression of BMP4 was analyzed as described

[22] using anti-BMP4 antibody (Novocastra Laboratories/Vector Laboratories,

Burlingame, CA). Slides were analyzed by confocal microscopy.

Cell staining and sorting. Bone marrow and spleen MEPs were sorted as

previously described [19] with the exception that FITC-conjugated anti-c-kit was used

(and APC-conjugated anti-c-kit and FITC-conjugated anti-CD34 eliminated) for spleen

sorts after determining that FITC conjugated anti-CD34 did not stain spleen cells. Cells

were washed twice and sorted using a Coulter Elite ESP flow cytometer. Cells were

plated in methylcellulose and scored as described in colony assays for BFU-E above.

Analysis of BMP4 signaling in W-20-17 osteoblast cells. The cDNAs coding

the f/f truncated transcripts, as well as full length Madh5 were cloned into the MSCVneo

retroviral construct. Recombinant virus was generated as previously described [23] and

used to infect W-20-17 cells (ATCC, Manassas, VA). Pools of neoR colonies were plated

and induction of alkaline phosphatase by BMP4 was measured as described [33].

Page 99: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

89

Results

flexed-tail mutant mice exhibit a delayed expansion of erythroid progenitors in the

spleen in response to acute anemia. In order to identify the origin of the defect, we

characterized the response of f/f, f/+ and control mice to acute anemia using a modified

PHZ induced hemolytic anemia protocol that rapidly induced severe anemia (hematocrit

30% treated v. 50% untreated) in 12 hours. Similar to previous findings, we observed that

early erythroid progenitors (BFU-E) in the bone marrow were not elevated in response to

anemia and in fact gradually declined over time (Figure 2-1A). The greatest difference

was observed in the spleen and peripheral blood. In the spleen, control mice exhibited an

expansion of BFU-E that peaked at 36 hours post anemia induction. Similarly, f/+ mice

showed the greatest expansion at 36 hours but unlike control mice we also observed a

significant expansion later at 4 and 6 days post anemia. In contrast, the expansion was

significantly delayed in the f/f mice where it peaked at 4 days post anemia induction

(Figure 2-1B). These results correlate with previous data showing that the f/f mice were

delayed in the expansion of erythroblasts in the spleen following anemia induction [13].

In the peripheral blood, however, we did not identify any BFU-E potentially migrating

from the bone marrow to the spleen at any of the time points in the f/f, f/+ or control mice

(data not shown). These results suggest a new model where progenitor cells resident in

the spleen mediate the response to acute anemia.

Splenic erythroid progenitors that expand in response to acute anemia exhibit

distinct properties. Bone marrow BFU-E colonies have a distinct morphology, develop

in 7 days in culture and require two signals to develop. The first signal is Epo and the

Page 100: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

90

second is a Burst-Promoting Activity (BPA), which in vivo is SCF, but in vitro IL-3 or

GM-CSF can substitute. We observed that the BFU-E colonies from spleen 36 hours post

anemia induction exhibited altered morphology. The colonies were larger, had more

small satellite colonies associated with the BFU-E. They also grew faster such that spleen

colonies grown for 5 days resembled bone marrow colonies that had grown for 7 days

and spleen BFU-E were routinely grown for 5 days out of convenience. In many ways

these erythroid progenitors resembled fetal liver erythroid progenitors, which are known

to exhibit a faster cell cycle than bone marrow BFU-E. Fetal liver BFU-E can also

develop in media containing only Epo, without an added BPA [24]. Given that f/f mice

also have a defect in fetal liver, we repeated the analysis of bone marrow, peripheral

blood and spleen BFU-E following induction of acute hemolytic anemia, however this

time cells were cultured in media containing only Epo. The bone marrow contained very

few cells that could form BFU-E colonies in Epo only media (Figure 2-1C). In the control

and f/+ spleens however, the expansion of BFU-E at 36 hours was completely

recapitulated when the cells were plated in Epo only media (Figure 2-1D). In fact more

splenic BFU-E colonies developed in this media. Similar to the initial observations, the f/f

mice exhibited a delay in the expansion of BFU-E with the maximum number of colonies

observed at 4 days post anemia induction. The number of spleen BFU-E observed under

the Epo only culture conditions responded in a linear manner when increasing numbers of

cells were plated, which suggests that the spleen cells are not producing a BPA (data not

shown). Once again we did not identify any BFU-E in the peripheral blood at any of the

time points in the f/f, f/+ or control mice indicating that these BFU-E are resident in the

spleen.

Page 101: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

91

In addition to shorter cell cycling times, fetal liver erythroid progenitors also

exhibit an increased sensitivity to Epo [24]. We tested the Epo-sensitivity of splenic

erythroid progenitors. We observed that these progenitor cells were actually less sensitive

to Epo than bone marrow cells (Figure 2-1E). Decreased Epo sensitivity is an ideal

property of a stress progenitor, because differentiation of these progenitors would be

dependent on the high serum Epo concentrations only present during the response to

acute anemia. Taken together, these results suggest that the spleen contains a distinct

population of erythroid progenitors that are poised to respond to acute erythroid stress.

These progenitors, which we will refer to as “stress BFU-E”, form large burst colonies in

5 days, require only Epo at relatively high levels to develop and are resident in the spleen.

f/f mice have mutation in the Madh5 gene. The flexed-tail locus is located on mouse

chromosome 13 [25]. We generated a panel of 408 F2 progeny using a F1(C57BL/6-f/f X

BALB/c) intercross. F2 progeny were scored at birth for anemia by hematocrit and for

the presence of siderocytes by staining blood smears for iron deposits. We constructed a

high resolution genetic linkage map of the f locus and initially localized the gene 0.6 cM

distal to the microsatellite marker D13MIT13. Further analysis of markers showed that

the f locus co-segregated with the marker D13Mit208 (Figure 2-2A). Our linkage

mapping results differ from the recent work from Fleming et al. [15]. They mapped the f

locus to a more proximal position on chromosome 13 and identified a mutation in the

sideroflexin 1 (sfxn1) gene, which they proposed caused the f/f mutant phenotype. Since

there is only a single allele of the f mutation, all f/f must carry the same mutation [26].

Like Fleming et al., our colony of C57BL/6J-f mice was derived from C57BL/6J-f mice

Page 102: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

92

obtained from the Jackson Laboratory. However, when we scored our f/f progeny for the

presence of the mutation in exon 2 of Sfxn1, we identified f/f progeny that exhibited

severe anemia and siderocytes at birth, but were heterozygous for the insertion mutation

(Appendix B). These results suggest that in our colony, the f mutation has been separated

from the mutation in Sfxn1 by recombination and thus, mutation of Sfxn1 cannot be the

cause of the f/f mutant phenotype.

In order to identify candidate genes for the f locus, we initially took advantage of

the fact that one of our flanking markers, D13Mit13, is located in the IL-9 gene [27]. This

region of mouse chromosome 13 is homologous to human chromosome 5q31.

Comparison of the human and mouse gene maps in region immediately surrounding IL-9

revealed MADH5 was located in this region. The possibility that Madh5 was encoded by

the f locus was supported by the recent mouse genome sequence release that showed that

D13MIT208 is located in the Madh5 gene. Previous work in Xenopus and mice has

demonstrated that BMP4/Madh5 dependent signals play a key role in the development of

erythroid cells [18,28]. Madh5 is highly expressed in the fetal liver during development

[29] and we have observed Madh5 expression in the spleen of mice recovering from PHZ

induced acute anemia (Data not shown).

To determine whether Madh5 is mutated in f/f mice, we cloned the entire coding

region of Madh5 by RT-PCR from spleen RNA isolated from C57BL/6J-+/+, C57BL/6J-

f/+ and C57BL/6J-f/f mice. Only the expected product was observed in wildtype mice,

however in both f/+ and f/f mice an additional band was observed (Figure 2-2B). The

majority of the mRNA in f/+ mice is the wildtype fragment, while in f/f mice the majority

of the mRNA is a truncated mRNA. The level of wildtype message in f/f mutant mice

Page 103: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

93

varies, but most express <30% wildtype mRNA. Analysis of the sequence of the

truncated mRNA showed that it is a mixture of two mis-spliced mRNAs. The consistent

feature of these mutant mRNAs was the deletion of exon 2, which contains the AUG

initiator codon. In one of the mis-spliced mRNAs we also observed deletion of exon 4

and insertion of a 15 nucleotide sequence at the splice junction between exons 6 and 7. In

the other mis-spliced mRNA, we observed an aberrant splice into the middle of exon 3

(Figure 2-2). There were no other alterations in the coding sequence of Madh5 observed

in f/f mice. Southern blot analysis of the Madh5 genomic region did not identify any

deletions or rearrangements in the Madh5 locus suggesting that the alterations in the

Madh5 mRNA in f/f mice are due to defects in mRNA splicing (data not shown). We

have sequenced the entire Madh5 transcription unit from f/f and control mice and have

not identified a consistent mutation, which suggests that the defect may lie in the

promoter or 3’ to the Madh5 gene. We are currently investigating this possibility.

BMP4 expression is induced in the spleen just prior to the expansion of “stress BFU-

E”. The identification of aberrantly spliced Madh5 mRNA in f/f mice suggests a role for

the BMP2, 4 and 7 signaling pathways in the response to acute anemia [16, 30]. We

investigated the expression of BMP2, 4 and 7 in the spleen during the response to acute

anemia by RT-PCR. BMP2 is not expressed in the spleen at any time point during the

recovery from acute anemia, while BMP7 is expressed at low levels at all times tested

(data not shown). BMP4 is not expressed in the spleen of untreated mice, however

expression is initiated at 12 hours, peaks at 24 hours with lower levels at 36 and 48 hours

post anemia induction. We also observe low levels at 6 and 8 days post anemia induction.

Page 104: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

94

The highest expression observed at 24 hours post anemia induction is just prior to the

expansion of “Stress BFU-E” in the spleen (Figure 2-3A). Staining of spleen sections

with anti-BMP4 antibodies showed that BMP4 protein was not present in untreated mice,

however high-levels of BMP4 was observed throughout the red pulp of the spleen at 24

and 36 hours post anemia induction, but expression is severely decreased by 48 hours and

essentially off by 96 hours post anemia induction (Figure 2-3B). BMP4 was excluded

from the white pulp.

The expression of BMP4 is tightly regulated during the response to acute anemia.

During the analysis of the BMP4 expression by RT-PCR, we also tested the expression in

f/f mice. Surprisingly, BMP4 expression is expanded in the mutant mice and untreated

mice at all the time points during the response to acute anemia exhibit BMP4 expression

(Figure 2-3A). The expression of BMP4 in untreated f/f correlates with the observation of

stress BFU-E in these mice (See Figure 2-1D). Despite the constitutive mRNA

expression, BMP4 protein expression was not observed at all time points suggesting that

BMP4 expression is regulated post-transcriptionally. These observed differences in

BMP4 expression in f/f mice suggest that the regulation of BMP4 may require a Madh5-

dependent signal to inhibit the expression of BMP4 in the spleens of untreated mice.

The increase in serum Epo concentration that occurs during the response to acute

anemia is regulated at the transcriptional level by the hypoxia inducible transcription

factor complex, HIF-1 [31]. Given that Epo expression is regulated by hypoxia we tested

whether BMP4 expression is also hypoxia inducible. MSS31 spleen stromal cells [32],

which support erythroid progenitor cell expansion in vitro, were grown in normoxic (20%

O2) and hypoxic (6% O2) conditions. At low oxygen concentration, the expression of

Page 105: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

95

BMP4 was significantly increased (Figure 2-3C). Analysis of the BMP4 gene revealed the

presence of a putative HIF element in the 3’UTR, which is conserved in the human, rat

and mouse BMP4 genes.

BMP4 causes the differentiation of an immature progenitor cell into an Epo

responsive stress BFU-E. Given that BMP4 was induced in the spleen just prior to the

expansion of stress BFU-E, we next tested whether culturing spleen cells from untreated

mice in Epo and BMP4 could induce the expansion of stress BFU-E. Spleen cells from

untreated f/f and control mice were plated in methylcellulose media containing 3U/ml

Epo and various concentrations of BMP4. Control spleen cells responded in a dose

dependent manner to BMP4 with a 6.1 fold increase in the number of stress BFU-E seen

at a 15 ng/ml BMP4 dose (Figure 2-4A). f/f spleen cells failed to respond to BMP4 except

at the highest concentration, which is consistent with their defect in Madh5. BMP4 had

very little effect on the number of BFU-E in the bone marrow suggesting that there are

very few BMP4 responsive erythroid progenitor cells in the bone marrow (Figure 2-4A).

We did not detect any BMP4 responsive cells in the peripheral blood in untreated or in

mice treated with PHZ to induce anemia (data not shown).

BMP4 can induce the expansion of stress BFU-E, but the mechanism by which

BMP4 affects these cells is not clear. One might imagine two possible roles for BMP4.

First BMP4 could synergize with Epo to promote the differentiation of stress BFU-E,

much like SCF synergizes with Epo to increase the number and size of bone marrow

BFU-E [33]. Alternatively, BMP4 may act on an earlier cell inducing it to differentiate

into an Epo responsive stress BFU-E. This possibility is similar to the situation in

Page 106: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

96

Xenopus embryos where BMP4 can induce mesodermal cells to become erythroid

progenitors [18]. To test these possibilities, we pre-incubated spleen cells from untreated

mice with and without BMP4 for 24 hours, washed the cells and then plated them in

methylcellulose media containing either Epo alone or Epo and BMP4. The cells pre-

incubated with BMP4 and plated in Epo alone gave rise to as many stress BFU-E

colonies as the cells plated in Epo and BMP4 (Figure 2-4B). These results suggest that a

short exposure to BMP4 promotes the differentiation of an immature BMP4 responsive

(BMP4R) cell into an Epo responsive stress BFU-E.

f is a gain of function allele of Madh5. Our analysis shows that the f mutation maps to

the Madh5 locus, f/f mice express misspliced mRNAs, and spleen cells from f/f fail to

respond to BMP4. Although all of these data are consistent with f being a mutation in

Madh5, we crossed f mice with Madh5+/- mice [20] to generate f/Madh5- mice to test

whether f was allelic to Madh5. Figure 2-5A shows the expansion of Stress BFU-E in

f/Madh5- and +/Madh5- during the recovery from acute anemia. Both genotypes exhibit

an altered recovery when compared to control (compare Figure 2-5A with Figure 2-1D).

The peak expansion of stress BFU-E in f/Madh5- is delayed until 48 hours, while

+/Madh5- exhibit an increase in Stress BFU-Es at 36 hours, they continue to expand at 48

hours. Analysis of BMP4 mRNA expression in these mice showed that both f/Madh5-

and +/Madh5- expressed BMP4 at all time points during the recovery (Figure 2-5B).

These data are similar to what is observed in f/f and f/+ mice suggesting that a Madh5

dependent signal is required for the regulation of BMP4 in the spleen. These results show

that the phenotype of f/Madh5- is more severe than f/+ and +/Madh5-, which

Page 107: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

97

demonstrates that f is allelic to a targeted mutation in Madh5. Classical genetic analysis,

however, would predict that if f was a hypomorphic or loss of function mutation in

Madh5 then f/Madh5- should have a more severe phenotype than f/f. We observed the

opposite in that the f/f phenotype was the most severe. These data suggest that the f

mutation in Madh5 is a neomorphic or gain of function mutation in Madh5. To test this

possibility, we expressed each of the two mis-spliced Madh5 mRNAs identified in f/f

mice in a cell line that responds to BMP4 and analyzed whether these mRNAs could

affect BMP4 dependent responses. W-20-17 is a mouse osteoblast cell line that

differentiates in response to BMP4 [34]. The mutant transcripts as well as controls were

cloned in to the MSCV-neo retroviral vector and W-20-17 cells were infected with

recombinant retroviruses. Although the mis-spliced messages lack the initiator ATG of

wildtype Madh5, both messages contain in-frame ATGs that could be used to initiate the

translation of truncated forms of the Madh5 protein (Figure 2-6A). Treatment of W-20-17

cells with BMP4 induces the osteoblast differentiation program as measured by an

increase in Alkaline Phosphatase (AP) activity. W-20-17 cells that express either f mutant

message have a profound defect in BMP4 dependent induction of AP activity with

mutant message #1 exhibiting the most severe defect (Figure 2-6B). These results show

that the mis-spliced Madh5 mRNAs present in f/f mice dominantly suppress BMP4

dependent signals. Furthermore, W-20-17 cells do not express endogenous Madh5 (Data

not shown), but rather rely on Madh1 and Madh8 to transmit BMP4 signals, which

suggests that the mis-spliced mRNAs also inhibit signaling through Madh1 and Madh8.

These results explain why f/Madh5- mice have a less severe phenotype than f/f because

the f/Madh5- mice express lower levels of the mis-spliced mRNAs.

Page 108: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

98

Spleen Megakaryocyte-Erythroid progenitors (MEPs) are the BMP4 responsive

cells. In order to further characterize the BMP4R cell, we fractionated spleen cells and

assayed for response to BMP4. Initially we sorted cells based on their expression of

lineage restricted markers. BMP4R cells were observed in the Lin- population (Figure 2-

7A). Previous work had demonstrated that mice with a mutation in the Kit receptor,

Dominant white spotting (W) mice, failed to respond rapidly to acute anemia, which

suggested that the BMP4R cell may also be Kit+ [35]. Analysis of Lin- Kit+ spleen cells

confirmed this hypothesis as BMP4R cells were not detected in Lin-Kit- spleen cells

(Figure 2-7B). In the bone marrow, erythroid progenitors are derived from the Common

Myeloid Progenitor (CMP) and the Megakaryocyte-Erythroid Progenitor (MEP) [19].

CMPs and MEPs are Kit+, but differ in their expression of CD34. We analyzed the

expression of CD34 in Lin- Kit+ spleen cells and observed that they were CD34-,

suggesting that the spleen does not contain CMPs (data not shown). Isolation of MEPs

from spleen showed that they responded to BMP4 (Figure 2-7C). Interestingly, MEPs

from isolated from bone marrow failed to respond to BMP4. These results suggest that

the unique microenvironment of the spleen alters the properties of MEPs rendering them

responsive to BMP4 and further underscores our assertion that distinct erythroid

progenitors are present in the spleen poised to respond to acute erythroid stress.

Page 109: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

99

Discussion

Acute anemia induces a robust erythroid response that occurs in the spleen. Our

data demonstrate that in response to PHZ induced anemia, dedicated stress progenitors

resident in the spleen respond to a BMP4 dependent signal, which leads to the rapid

expansion of erythroid progenitors. These data differ from the previous model which

suggested that high serum Epo levels induced by tissue hypoxia mobilized erythroid

progenitors from the bone marrow to spleen where they expanded and differentiated [1,2].

This early model was based on the observation that BFU-E were present in the peripheral

blood during the recovery from acute anemia [1]. However, we did not observe migration

of erythroid progenitors in our studies and all of our data suggests that stress BFU-E are

resident in the spleen. One reason for the differences in our data could be that earlier

experiments cultured BFU-E for 10 days rather than the 5-7 days used in our studies. The

longer culture conditions may have allowed more immature cells to develop. In addition,

we used a modified protocol to induce anemia, which utilized a single injection of a high

dose of PHZ. This protocol results in a synchronous and reproducible expansion of

erythroid progenitors. The earlier experiments used multiple injections of a low dose of

PHZ. We have found that multiple low doses of PHZ did not allow us to look at early

events during the recovery and did not produce reliable results (L. Lenox and RF Paulson

unpublished observations 1999).

The earlier model suggested that bone marrow progenitors develop in the spleen

during the recovery. Our data however demonstrates that specialized BMP4R cells are

resident in the spleen. BMP4 induces these cells to differentiate leading to the expansion

of stress BFU-E in the spleen but not in the bone marrow. Stress BFU-E exhibit

Page 110: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

100

properties that are distinct from bone marrow BFU-E in that they form large burst

colonies in 5 days rather than 7 days and they require only Epo without any added BPA

for BFU-E formation. Fractionation of spleen cells revealed that BMP4R cells are

contained within the MEP population. However, only spleen MEPs are able to respond to

BMP4. This observation suggests that the spleen microenvironment may provide a

specialized signal that enables spleen MEPs to respond to BMP4 [3]. This situation would

be similar to chondrogenesis, where pre-somitic mesoderm cells require BMP signals to

differentiate into chondrocytes, but are unable to respond to BMP unless they first

encounter a Sonic hedgehog signal [36].

The notion that f/f mice have defect in a spleen progenitor was first suggested

almost 30 years ago. Gregory et al. showed that f/f mice had normal numbers of BFU-E

and CFU-E, but were defective in transient endogenous colony forming units (TE-CFU)

[12]. This population of cells was defined by an in vivo assay that identifies progenitor

cells that form endogenous spleen colonies following sublethal irradiation and

stimulation of erythropoiesis by Epo injection or bleeding [37]. BMP4R cells exhibit many

of the properties expected in a putative TE-CFU. They are resident in the spleen, rapidly

expand at times of great erythropoietic need and require high levels of Epo for

differentiation. In addition to TE-CFU, previous work has identified other stress erythroid

progenitors in the spleen. Mice with mutations in the glucocorticoid receptor are slow to

respond to acute anemia and it has been suggested that a CD34+ Kit+ TER119+

population of cells fails to expand in these mutant mice [38]. Another group identified a

4A5+ TER119+ bipotential megakaryocyte-erythroid progenitor that expands in the

spleen following PHZ induced anemia [39]. However, both of these progenitors express

Page 111: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

101

the lineage-restricted marker, TER119, which is not present on BMP4R cells or BFU-E in

general [40]. The relationship between these progenitors and the BMP4R cells and the

stress BFU-E is not understood, but the most likely possibility is that they are derived

from BMP4R cells.

Our analysis of f/f mice has identified a mutation in Madh5, which causes aberrant

splicing. The mis-spliced Madh5 mRNAs exert a dominant negative effect on BMP4

signaling which suggests that f represents a gain of function allele of Madh5. These mis-

spliced mRNAs inhibit BMP4 dependent signals in W-20-17 osteoblast cells, which do

not express endogenous Madh5, suggesting that the f mutant mRNAs inhibit BMP4

signaling mediated by Madh1 and/or Madh8. This possibility is consistent with our

observation that f/f exhibits a more severe phenotype than f/Madh5- mice. In this case, f/f

mice express higher levels of mis-spliced mRNAs and thus would have a more

significant impairment of BMP4 dependent signaling than f/Madh5- mice.

Previous work from Fleming et al. suggested that f was a mutation in the putative

mitochondrial transporter Sfxn1 [15]. We show that the mutation in Sfxn1has been

separated from the f locus by recombination in f/f mice in our colony. All of our

observations characterizing the BMP4 dependent expansion of stress BFU-E in the spleen

during the recovery from acute anemia and the defective response in f/f mice support the

idea that the f locus encodes Madh5. In addition, other phenotypes associated with f/f

mice, tail flexures and white belly spots, can easily be explained by defects in the BMP4

signaling pathway. The tail flexures are caused by defects in chondrogenesis that result in

vertebrae fusion [41]. BMP4 plays a key role in regulating chondrocyte development [36].

Furthermore, white belly spots are caused by defects in the migration of neural crest

Page 112: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

102

derived melanocytes [42]. Inhibition of BMP4 signaling in chick embryos impairs the

ability of neural crest cells to migrate [43]. Thus the combination of the observed defects

in BMP4 dependent signals in f/f mice with the genetic interactions of f and Madh5-

alleles demonstrates that the f locus encodes Madh5.

Page 113: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

103

References

1. Hara H, Ogawa M. Erythropoietic precursors in mice with phenylhydrazine induced anemia. Am J Hematol. 1976; 1:453-458.

2. Broudy V, Lin N, Priestley G, Nocka K, Wolf N. Interaction of stem cell factor and its receptor c-kit mediates lodgement and acute expansion of hematopoietic cells in the murine spleen. Blood. 1996; 88:75-81.

3. Obinata M, Yania N. Cellular and molecular regulation of erythropoietic inductive microenvironment (EIM). Cell Struct Funct. 1999; 24:171-179.

4. Palis J, Segel G. Developmental biology of erythropoiesis. Blood Rev.1998; 12:106-114.

5. Ohneda O, Yanai N, Obinata M. Microenvironment created by stromal cells is essential for a rapid expansion of erythroid cells in the mouse fetal liver. Development. 1990; 110:379-384.

6. Mixter R, Hunt H. Anemia in the flexed tailed mouse, Mus musculus. Genetics. 1933; 18:367-387.

7. Gruneberg H. The anaemia of the flexed-tail mouse (Mus musculus L.) I. Static and dynamic haematology. J Genetics. 1942; 43:45-68.

8. Gruneberg H. The anaemia of the flexed-tail mice (Mus musculus L.) II. Siderocytes. J Genetics. 1942; 44:246-271.

9. Bateman A, Cole R. Colony forming cells in the livers of prenatal flexed (f/f) anaemic mice. Cell Tissue Kinet. 1972; 5:165-173.

10. Cole R, Regan T. Haematopoietic Progenitor cells in the prenatal congenitally anaemic "Flexed-tail" (f/f) mice. Br J Haematol. 1976; 33:387-394.

11. Chui D, Sweeney G, Patterson M, Russell E. Hemoglobin synthesis in siderocytes of flexed-tail mutant (f/f) fetal mice. Blood. 1977; 50:165-177.

12. Gregory C, McCulloch E, Till J. The Cellular Basis for the defect in haematopoiesis in flexed-tail mice. III Restriction of the defect to erythropoietic progenitors capable of transient colony formation in vivo. Br J Haematol. 1975; 30:401-410.

13. Coleman D, Russell E, Levin E. Enzymatic studies of the hemopoietic defect in flexed mice. Genetics. 1969; 61:631-642.

14. Cole R, Regan T, Tarbutt R. Haemoglobin synthesis in reticulocytes of prenatal f/f anaemic mice. B J Haematol. 1972; 23:443-452.

15. Fleming M, Campagna D, Haslett J, Trenor C, Andrews N. A mutation in a mitochondrial transmembrane protein is responsible for the pleiotropic hematological and skeletal phenotype of flexed-tail (f/f) mice. Genes Dev. 2001; 15:652-657.

16. Massague J. How cells read TGF-β signals. Nat Rev Mol Cell Biol. 2000; 1:169-178.

17. Massague J, Chen YG. Controlling TGF-β signaling. Genes Dev. 2000; 14:505-520.

18. Huber T, Zhou Y, Mead P, Zon L. Cooperative effects of growth factors involved in the induction of hematopoietic mesoderm. Blood. 1998; 92:4128-4137.

Page 114: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

104

19. Akashi K, Traver D, Miyamoto T, Weissman I. A clonogenic common myeloid progenitor that gives rise to all myeloid lineages. Nature. 2000; 404:193-197.

20. Yang X, Castilla L, Xu X, et al.. Angiogenesis defects and mesenchymal apoptosis in mice lacking SMAD5. Development. 1999; 126:1571-1580.

21. Finkelstein L, Ney P, Liu Q, Paulson R, Correll P. Sf-Stk kinase activity and the Grb2 binding site are required for Epo-independent growth of primary erythroblasts infected with Friend virus. Oncogene. 2002; 21:3562-3570.

22. Marshall C, Kinnon C, Thrasher A. Polarized expression of bone morphogenetic protein-4 in the human aorta-gonad-mesonephros region. Blood. 2000; 96:1591-1593.

23. Melkun E, Pilione M, Paulson R. A naturally occurring point substitution in Cdc25A, and not Fv2/Stk, is associated with altered cell cycle status of early erythroid progenitor cells. Blood. 2002; 100:3804-3811.

24. Peschle C, Migliaccio A, Migliaccio G, et al.. Identification and characterization of three classes of erythroid progenitors in human fetal liver. Blood. 1981; 58:565-574.

25. Lyon M, Rastan S, Brown S. Genetic Variants and Strains of the Laboratory Mouse (ed 3rd). Oxford: Oxford University Press; 1996.

26. Hunt H, Premar D. Flexed-tail a mutation in the house mouse. Anat. Rec. 1928; 41:117. 27. Bult CJ, Blake JA, Richardson JE, et al.. The Mouse Genome Database (MGD): integrating biology with the genome. Nucleic Acids Res. 2004; 32 Database issue:D476-481.

28. Nakayama N, Lee J, Chiu L. Vascular endothelial growth factor synergistically enhances bone morphogenetic protein-4 dependent lymphohematopoietic cell generation from embryonic stem cells in vitro. Blood. 2000; 95:2275-2284.

29. Flanders K, Kim E, Roberts A. Immunohistochemical expression of Smads 1-6 in the 15-day gestation mouse embryo: signaling by BMPs and TGF-betas. Dev Dyn. 2001; 220:141-154.

30. Hogan B. Bone morphogenetic proteins: multifunctional regulators of vertebrate development. Genes Dev. 1996; 10:1580-1594.

31. Ebert B, Bunn H. Regulation of the Erythropoietin Gene. Blood. 1999; 94:1864-1877.

32. Yanai N, Satoh T, Obinata M. Endothelial cells create a hematopoietic inductive microenvironment preferential to eythropoiesis in the mouse spleen. Cell Struct Funct. 1991; 16:87-93.

33. McNiece I, Langley K, Zsebo K. Recombinant human stem cell factor synergises with GM-CSF, G-CSF, IL-3 and Epo to stimulate human progenitor cells of the myeloid and erythroid lineages. Exp. Hematol. 1991; 19:226-231.

34. Binnerts ME, Wen X, Cante-Barrett K, et al.. Human Crossveinless-2 is a novel inhibitor of bone morphogenetic proteins. Biochem Biophys Res Commun. 2004; 315:272-280.

35. Harrison D, Russell E. The response of W/Wv and Sl/Sld amaemic mice to hematopoietic stimuli. Br. J. Haematol. 1972; 22:155-167.

Page 115: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

105

36. Murtaugh L, Chyung J, Lassar A. Sonic Hedgehog promotes somatic chondrogenesis by altering the cellular response to BMP signaling. Genes Dev. 1999; 13:225-237.

37. Gregory C, McCulloch E, Till J. Transient erythropoietic spleen colonies: effects of erythropoietin in normal and genetically anemic W/Wv mice. J Cell Physiol. 1975; 86:1-8.

38. Bauer A, Tronche F, Wessely O, et al.. The glucocorticoid receptor is required for stress erythropoiesis. Genes Dev. 1999; 13:2996-3002.

39. Vannucchi A, Paoletti F, Linari S, et al.. Identification and characterization of a bipotent (erythroid and megakaryocytic) cell precursor from the spleen of phenylhydrazine-treated mice. Blood. 2000; 95:2559-2568.

40. Kina T, Ikuta K, Takayama E, et al.. The monoclonal antibody TER-119 recognizes a molecule associated with glycophorin A and specifically marks the late stages of murine erythroid lineage. Br J Haematol. 2000; 109:280-288.

41. Kamenoff R. Effects of the flexed-tail gene on the development of the house mouse. J Morphol. 1935; 58:117-155.

42. Christiansen J, Coles E, Wilkinson D. Molecular control of neural crest formation, migration and differentiation. Curr Opin Cell Biol. 2000; 12:719-724.

43. Sela-Donenfeld D, Kalcheim C. Regulation of the onset of neural crest migration by coordinated activity of BMP4 and Noggin in the dorsal tube. Development. 1999; 126:4749-4762.

Page 116: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

0

50

100

150

0 12 24 36 48 4 6 8

BFU

-E

DControl

f/ff/+

Hours DaysPost PHZ Injection

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+

Hours DaysPost PHZ Injection

BFU

-E

A

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+

Hours DaysPost PHZ Injection

BFU

-E

C

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+

Hours DaysPost PHZ Injection

BFU

-E

B

Bone marrow Epo+IL-3

Bone marrow Epo

SpleenEpo+IL-3

SpleenEpo

0

25

50

75

100

0.1 0.3 1 3 10

Epo concentration U/ml

BFU

-E (%

Max

imal

) Bone MarrowSpleen

E.

0

50

100

150

0 12 24 36 48 4 6 8

BFU

-E

DControl

f/ff/+

Hours DaysPost PHZ Injection

0

50

100

150

0 12 24 36 48 4 6 8

BFU

-E

DControl

f/ff/+Control

f/ff/ff/+

Hours DaysPost PHZ Injection

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+

Hours DaysPost PHZ Injection

BFU

-E

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+

Hours DaysPost PHZ Injection

BFU

-E

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+

Hours DaysPost PHZ Injection

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+Control

f/ff/ff/+

Hours DaysPost PHZ Injection

BFU

-E

A

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+

Hours DaysPost PHZ Injection

BFU

-E

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+

Hours DaysPost PHZ Injection

BFU

-E

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+

Hours DaysPost PHZ Injection

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+Control

f/ff/ff/+

Hours DaysPost PHZ Injection

BFU

-E

C

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+

Hours DaysPost PHZ Injection

BFU

-E

C

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+

Hours DaysPost PHZ Injection

BFU

-E

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+

Hours DaysPost PHZ Injection

BFU

-E

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+

Hours DaysPost PHZ Injection

0

50

100

150

0 12 24 36 48 4 6 8

Control

f/ff/+Control

f/ff/ff/+

Hours DaysPost PHZ Injection

BFU

-E

B

Bone marrow Epo+IL-3

Bone marrow Epo

SpleenEpo+IL-3

SpleenEpo

0

25

50

75

100

0.1 0.3 1 3 10

Epo concentration U/ml

BFU

-E (%

Max

imal

) Bone MarrowSpleen

E.

0

25

50

75

100

0.1 0.3 1 3 10

Epo concentration U/ml

BFU

-E (%

Max

imal

) Bone MarrowSpleen

0

25

50

75

100

0.1 0.3 1 3 10

Epo concentration U/ml

BFU

-E (%

Max

imal

) Bone MarrowSpleen

E.

106

Figure 2-1. Analysis of BFU-E expansion during the recovery from PHZ induced acute hemolytic anemia. (A) Bone marrow and (B) spleen BFU-E from C57BL/6-f/f, C57BL/6-f/+ and C57BL/6-+/+ control mice during the recovery from PHZ induced acute anemia. Cells were plated in methylcellulose media containing Epo (3 U/ml) and IL-3 (10 ng/ml). (C) Bone marrow and (D) spleen BFU-E from C57BL/6-f/f, C57BL/6-f/+ and C57BL/6-+/+ control mice during the recovery from PHZ induced acute anemia. Cells were plated in methylcellulose media containing only Epo (3 U/ml). (E) Sensitivity of bone marrow and spleen BFU-E to Epo. A total of 5x105 bone marrow or spleen cells were plated in methylcellulose media containing the indicated concentrations of Epo plus 50 ng/ml SCF (bone marrow, ) or 15 ng/ml BMP4 (spleen, ). The asterisk indicates P= 0.02. Error bars represent standard deviation.

Page 117: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

D13Mit248

D13Mit65

flexed -tail

D13Mit279

D13Mit208

D13Mit21

D13Mit41

D13Mit186

Sfxn1

D13Mit250D13Mit13

D13Mit122

C57

BL/

6-+/

+

C57

BL/

6-f/

+

C57

BL/

6-f/

f

*

Wildtype

flexed-tail

1 ∆3 5 6 74

StopATG

1 2 4 5 6 73

cctgttcatttccag

1 3 5 6 7

**

A.

B.

D13Mit248

D13Mit65

flexed -tail

D13Mit279

D13Mit208

D13Mit21

D13Mit41

D13Mit186

Sfxn1

D13Mit250D13Mit13

D13Mit122

C57

BL/

6-+/

+

C57

BL/

6-f/

+

C57

BL/

6-f/

f

*

Wildtype

flexed-tail

1 ∆3 5 6 741 ∆3 5 6 74

StopATG

1 2 4 5 6 731 2 4 5 6 73

cctgttcatttc

1 3 5 6 7

cctgttcattt

1 3 5 6 71 3 5 6 7

**

A.

B.

107

Figure 2-2. Genetic linkage map of the f locus and molecular analysis of Madh5transcripts in f/f, f/+ and control mice. (A) Linkage map of the f locus on mouse chromosome 13. The position of the markers on chromosome 13 according to NCBIm33 mouse genome assembly is indicated. The position of Madh5 and Sfxn1 are shown. (B) The coding region of the Madh5 was cloned by RT-PCR of spleen RNA from the indicated mice. The arrowheads indicate the position of the wildtype and mutant mRNAs. The * indicates a non-specific background band. The exon structure of the wildtype and f/f mRNAs is indicated at the right. The f/f mouseshown here is an example of a mutant mouse that expresses very little wildtypeMadh5 mRNA.

Page 118: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

+/+

Number of Cycles

20

25

30

Post-PHZ Injection

f/f

BMP4HPRT

Hours Days0 12 24 36 48 4 6 8

Hours Days0 12 24 36 48 4 6 8

20

25

30

Number of Cycles

BMP4HPRT

+/+

Number of Cycles

20

25

30

Post-PHZ Injection

f/f

BMP4HPRTBMP4HPRT

Hours Days0 12 24 36 48 4 6 8

Hours Days0 12 24 36 48 4 6 8

Hours Days0 12 24 36 48 4 6 8

Hours Days0 12 24 36 48 4 6 8

20

25

30

Number of Cycles

BMP4HPRTBMP4HPRT

+/+

f/f

0 Hour 24 Hour 36 Hour 48 Hour 96 Hour

+/+

f/f

0 Hour 24 Hour 36 Hour 48 Hour 96 Hour

20% 6%[O2]BMP4

HPRT

CTTGACCTTATTTATGACTTTACGTGCAAA

CTTGACCTTATTTATGACTTTACGTGCAAA||||||||||||||||||||||||||||||CTTGACCTTATTTATGACTTTACGTGCAAA

TGAATG

||||||||||||||||||||||||||||||Mouse

Human

HNF4 site HIF SiteRat

20% 6%[O2]BMP4

HPRT

CTTGACCTTATTTATGACTTTACGTGCAAA

CTTGACCTTATTTATGACTTTACGTGCAAA||||||||||||||||||||||||||||||CTTGACCTTATTTATGACTTTACGTGCAAA

TGAATG

||||||||||||||||||||||||||||||Mouse

Human

HNF4 site HIF SiteRat

CTTGACCTTATTTATGACTTTACGTGCAAA

CTTGACCTTATTTATGACTTTACGTGCAAA||||||||||||||||||||||||||||||CTTGACCTTATTTATGACTTTACGTGCAAA

TGAATG

||||||||||||||||||||||||||||||Mouse

Human

HNF4 site HIF SiteRat

B)

C)

A)108

Figure 2-3. Analysis of BMP4 expression during the recovery from acute anemia. (A)RT-PCR analysis of BMP4 expression in C57BL/6-+/+ control (Top) and C57BL/6-f/f (bottom) mice. Arrows indicate the positions of the BMP4 specific band and the HPRT control band and the number of PCR cycles is indicated at the right. (B) Spleen sections from C57BL/6-+/+ (Top) and C57BL/6-f/f (bottom) mice stained with anti-BMP4 antibodies at the indicated times following PHZ induced acute anemia. (C) RT-PCR analysis of BMP4 expression in MSS31 spleen stromal cells grown at normoxic (20%) and hypoxic (6%) conditions (Top). The position of the putative hypoxia inducible element in the BMP4 gene is indicated and an alignment of this sequence from mouse, human and rat is presented (Bottom).

Page 119: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

109

Spleen

Bone Marrow

0

50

Epo Epo+BMP4

Stre

ss B

FU-E

+ BMP4- BMP4

Pre-incubation

Fold

incr

ease

0

2

4

6

1

0

2

4

6

0 0.15 1.5 15BMP4 (ng/m

Fold

Incr

ease

1

Controlf/f

A.

B.

Spleen

Bone Marrow

0

50

Epo Epo+BMP4

Stre

ss B

FU-E

+ BMP4- BMP4

Pre-incubation

Fold

incr

ease

0

2

4

6

1

0

2

4

6

0 0.15 1.5 15BMP4 (ng/ml)

Fold

Incr

ease

1

Controlf/f

Controlf/f

A.

B.

Figure 2-4. Analysis of the ability of BMP4 to induce the formation of stress BFU-E in spleen cells from untreated mice. (A) Bone marrow (Top) and spleen (Bottom) cells from untreated f/f and control mice were plated in methylcellulose media containing Epo (3 U/ml) and the indicated concentration of BMP4. (B) Spleen cells from C57BL/6-+/+ mice were pre-incubated with BMP4 (15 ng/ml) for 24 hours, washed and then plated in the methylcellulose media containing either Epo (3 U/ml) alone or Epo + BMP4 (15 ng/ml).

Page 120: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

110

0

100

200

300

400

0 12 24 36 48 4 6 8

Post-PHZ Injection

BFU

-E

f/Madh5-+/Madh5-

Hours Days

f/Madh5-

+/Madh5-BMP4

HPRT

0 12 24 36 48 4 6 8

Hours DaysPost PHZ treatment

B.

A.

0

100

200

300

400

0 12 24 36 48 4 6 8

Post-PHZ Injection

BFU

-E

f/Madh5-+/Madh5-

Hours Days

f/ -

+/ -BMP4

HPRT f/ -

+/ -BMP4

HPRT

+/ -BMP4

HPRT

0 12 24 36 48 4 6 8

Hours DaysPost PHZ treatment

B.

A.

Figure 2-5. Analysis of the recovery from acute anemia in f/Madh5- and +/Madh5-mice. (A) Analysis of the expansion of stress BFU-E in the spleen of C57BL/6-f/Madh5- and C57BL/6-+/Madh5- mice. Spleen cells were plated in methylcellulose media containing 3 U/ml Epo. (B) Expression of BMP4 in the spleen of C57BL/6-f/Madh5- and C57BL/6-+/Madh5- mice. The BMP4 and HPRT control bands are indicated by the arrows. These results are from 25 cycles of PCR.

Page 121: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

111

-BMP4+BMP4

0123456789

Control Wildtype Mutant #1 Mutant #2

Alk

alin

e Ph

osph

atas

eA

ctiv

ity

Wildtype

Mutant #1

Mutant #2

ATG

*

[*****]

B.

A.

-BMP4+BMP4

0123456789

Control Wildtype Mutant #1 Mutant #2

Alk

alin

e Ph

osph

atas

eA

ctiv

ity

-BMP4+BMP4

0123456789

Control Wildtype Mutant #1 Mutant #2

Alk

alin

e Ph

osph

atas

eA

ctiv

ity

Wildtype

Mutant #1

Mutant #2

ATG

*

[*****]

Wildtype

Mutant #1

Mutant #2

ATG

*

[*****]

B.

A.

Figure 2-6. Analysis of the effect of over-expression of the f mis-spliced Madh5 mRNAs on BMP4 signaling in W-20-17 osteoblast cells. (A) Schematic representation of the f mis-spliced Madh5 mRNAs and control Madh5 mRNA. The position of the endogenous ATG is indicated in the wildtype mRNA. * indicate the positions of putative in-frame ATGs in the mis-spliced mRNAs. (B) Induction of alkaline phosphatase activity by BMP4 in control W-20-17 cells and W-20-17 cells expressing wildtype or mis-splice Madh5 mRNAs. Alkaline phosphatase activity was normalized to protein concentration and expressed in relative units.

Page 122: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

112

0

1

2

3

4

5

6

Unfractionated Lin+ Lin-

BFU

-E p

er 1

05ce

llsN.D.

EpoEpo + BMP4

020406080

100120140160

BFU

-E p

er 1

05ce

lls

N.D.N.D.

Lin- Kit+ Lin- Kit-

EpoEpo + BMP4

0

10

20

Bone Marrow Spleen

BFU

-E p

er 3

x104

cells Epo

Epo + BMP4

A.

B.

C.

0

1

2

3

4

5

6

Unfractionated Lin+ Lin-

BFU

-E p

er 1

05ce

llsN.D.

EpoEpo + BMP4

0

1

2

3

4

5

6

Unfractionated Lin+ Lin-

BFU

-E p

er 1

05ce

llsN.D.

EpoEpo + BMP4EpoEpo + BMP4

020406080

100120140160

BFU

-E p

er 1

05ce

lls

N.D.N.D.

Lin- Kit+ Lin- Kit-

EpoEpo + BMP4

020406080

100120140160

BFU

-E p

er 1

05ce

lls

N.D.N.D.

Lin- Kit+ Lin- Kit-

EpoEpo + BMP4EpoEpo + BMP4

0

10

20

Bone Marrow Spleen

BFU

-E p

er 3

x104

cells Epo

Epo + BMP4

0

10

20

Bone Marrow Spleen

BFU

-E p

er 3

x104

cells Epo

Epo + BMP4EpoEpo + BMP4

A.

B.

C.

Figure 2-7. Identification of the sub-population of progenitor cells from untreated spleen that responds to BMP4. (A) Unfractionated, Lin+ and Lin –cells from untreated wildtype spleen were plated in 3 U/ml Epo + 15 ng/ml BMP4 and the induction of stress BFU-E was analyzed. (B) Kit+ Lin- and Kit- Lin- cells from untreated wildtype spleen were plated in 3 U/ml Epo + 15ng/ml BMP4 and the induction of stress BFU-E was analyzed. (C) MEPs (Lin- Sca1- IL-7Rα- Kit+ CD34- FcγRlow) isolated from bone marrow and spleen of wildtype mice were plated in 3 U/ml Epo + 15 ng/ml BMP4 and the induction of stress BFU-E was analyzed. N.D.None Detected.

Page 123: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

113

Chapter 3

STRESS ERYTHROPOIESIS IN SPLENECTOMIZED MICE

Abstract

The spleen has been shown to be the main site for reestablishing homeostasis

following an erythropoietic stress. This is due to the fact it contains highly specialized

erythroid progenitors which respond to BMP4 [1], as well as the erythropoietic inductive

environment created (EIM) created by the stroma of this organ [2]. Mice defective in this

expansion of erythroid progenitors, such as the flexed-tail (f/f) mutants, which exhibit a

disruption of BMP4/Madh5 signaling due to a splicing defect in Madh5, show a delay in

the recovery to an acute anemia. Considering the critical role of the spleen and its

progenitors in response to an erythropoietic stress, we wanted to characterize the stress

erythroid response in splenectomized mice and determine if their defect was even more

severe than the delay seen in flexed-tail mice. Splenectomized mice show different

kinetics in their recovery to a phenylhydrazine induced acute anemia. Although there is

no increase in the bone marrow contribution to erythropoiesis over that seen in wild-type

controls, the liver of splenectomized mice shows evidence of erythropoietic activity.

BMP4 is expressed in the liver and confers properties to erythroid progenitors in the liver

of splenectomized mice similar to those seen in progenitors expanding in the spleen of

wild-type mice. Since there are no erythroid progenitors in the liver prior to an acute

anemia, we wanted to determine what signals may be important for progenitor cells to

Page 124: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

114

traffic here from other locations. A potential candidate is the chemokine stromal cell-

derived factor-1 (SDF-1 or CXCL12). SDF-1 is a factor whose expression has been

shown to increase in ischemic tissue and is involved in the mobilization of progenitor

cells. SDF-1 is expressed in the liver of wild-type and splenectomized mice and could be

playing a role in the recruitment of erythroid progenitor cells to the liver following a

phenylhydrazine induced acute anemia. The differences between splenectomized and

control mice were not significant, so the role of SDF-1 in homing of progenitor cells to

the liver will require functional analysis of SDF-1/CXCR4 signaling. Interestingly, even

though f/f mice are defective in the stress erythroid response in the spleen, they show no

morphological evidence for erythropoietic activity in the liver when it is observed in the

liver of splenectomized mice. Considering that BMP4 is expressed at sites participating

in stress erythropoiesis (be it the spleen or the liver) and it has a role in the expansion of

stress erythroid progenitors at these locations, this might suggest that splenectomized f/f

mice would have a further delay in erythropoiesis in the liver following an erythropoietic

challenge.

Page 125: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

115

Introduction

The spleen is the primary organ for re-establishing homeostasis following an

erythropoietic challenge such as irradiation, hypoxia or an acute anemia [3-6]. Work in

our lab has characterized the role of the spleen in stress erythropoiesis and has shown the

spleen contains an endogenous erythroid progenitor which differentiates under the control

of BMP4. BMP4 expression, regulated by hypoxia, increases in the spleen at 24 hours

post PHZ induced acute anemia. This BMP4 signal drives the expansion and

differentiation of erythroid progenitors resident in the spleen [7]. BMP4 induces the

differentiation of a highly specialized “stress” BFU-E. These “stress” BFU-E exhibit

properties distinct from steady state bone marrow progenitors. They grow faster in

culture, forming a BFU-E colony in 5 days rather than 7 and develop in the absence of a

burst promoting factor, requiring only Epo to form BFU-E. This process of expansive

erythropoiesis is supported by the unique erythropoietic inductive microenvironment

(EIM) established in the spleen. The ability of the spleen EIM to induce the expansion of

erythroid progenitors has been demonstrated using MSS31 cell line from mouse newborn

spleens [2]. After the addition of fetal liver cells and erythropoietin, erythroid

progenitors adhered to the MSS31 cells, eventually maturing to hemoglobin producing

cells that would detach and enucleate. Cell-to-cell contact allowing short-range

communication between the erythroid progenitor cells and the MSS cells is required in

that conditioned medium of MSS31 cells did not show any effects on colony formation.

Although the EIM of the spleen is known to be a composite of secreted factors,

progenitor cells, stromal cells, and adhesion molecules, the properties that distinguishes

this EIM from other microenvironments has not been fully characterized.

Page 126: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

116

The expansive erythropoiesis in the spleen is altered in mice with a mutation at

the flexed-tail locus (f) [8, 9]. f/f are defective in the BMP4/Smad5 signaling pathway,

which disrupts the expansion of splenic erythroid progenitors and leads to a delay in the

recovery to acute anemia [7, 10]. Since a defect in the BMP signaling pathway in the

spleen manifests as a delay in re-establishing homeostasis, we wanted to characterize the

response in the complete absence of a spleen. Both flexed-tail and splenectomized mice

are viable, but it is reasonable to hypothesize that the lack of the specialized splenic

microenvironment by splenectomy would impart an even more severe phenotype during

the immediate recovery to an acute stress than that seen in the flexed-tail mutant.

. In this report we show that splenectomy does affect the ability of mice to

respond to an erythropoietic stress. Splenectomized mice show delayed kinetics of

hematocrit recovery following a PHZ induced acute anemia. Although the bone marrow

does not greatly increase its contribution to the recovery, the delay in recovery is not

severe since erythropoiesis is initiated in the liver. Unlike the spleen, the liver does not

harbor any erythroid specific progenitors. However, the liver expresses BMP4 and the

progenitor cells participating in the liver resemble splenic stress BFU-E which proliferate

with Epo alone. Even though f/f mice are defective in their splenic response, they do not

have a response in their liver. Since erythroid progenitors are not present in the liver

prior to the anemia, progenitors must migrate to the liver for their expansion. The signal

directing the homing of progenitors to the liver may be the chemokine stromal-derived

factor-1 (SDF-1 or CXCL12). SDF-1 has been shown to be a key regulator of stem cell

migration and is involved in the recruitment of CXCR4+ progenitor cells to ischemic

tissue for tissue regeneration. SDF-1 is expressed in the liver during the recovery from

Page 127: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

117

acute anemia and may be involved in the recruitment of erythroid progenitor cells to the

liver in splenectomized mice. The similarities that have emerged between expansive

erythropoiesis in the spleen and that seen in the liver of splenectomized mice helps define

the signals that are important for initiating expansive erythropoiesis in extra-medullary

organs in general, the nature of the progenitors participating and how progenitors may

locate environments which will provide the most productive site for their expansion.

.

Page 128: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

118

Methods

Mice. C57BL/6 and C57BL/6 splenectomized mice were obtained from Jackson

Laboratory. Spleens were removed at weaning. All mice were 6-8 weeks old and

controls were age matched. Splenectomized mice were not used until at least 3 weeks

after splenectomy. Acute anemia was induced by a single injection of Phenylhydrazine

(Sigma, St. Louis , MO) at a concentration of 100 mg/Kg mouse in PBS.

Hematocrit Analysis. Percent hematocrit was determined by the packed red

blood cell volume over total volume of blood obtained using a capillary tube ocular

bleed.

Colony Assays for BFU-E. Single cell suspensions from bone marrow or liver

were obtained at indicated points post PHZ induced acute anemia in IMDM + 5% heat

inactivated FCS. For liver suspensions, 22 mg/50 mL collagenase Type I (Worthington,

Lakewood, NJ) was included in the media. Erythrocytes were lysed with cold 0.16M

ammonium chloride (Sigma, St. Louis, MO). In addition, liver suspensions were passed

through a 70 µM nylon cell strainer (BD Falcon, Bedford, MA) along with separation of

debris and hepatocytes from mononuclear cells using a Nycoprep 1.077 gradient (AXIS-

SHIELD PoC AS, Oslo, Norway). 1x105 /mL bone marrow or 1.4x105/mL liver cells

were plated in methylcellulose media (StemCell Technologies, Vancouver, BC)

containing 3U/mL Epo ± either 10ng/mL IL3 (Sigma, St. Louis, MO) or 15ng/mL BMP4

(R&D Systems, Minneapolis, MN) where indicated. BFU-E were scored as described

[11].

H&E staining of liver sections. Livers were harvested at the indicated times post

PHZ induced anemia, fixed in 4% paraformaldehyde, and paraffin embedded tissue

Page 129: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

119

sections cut and stained using standard hematoxylin and eosin staining procedures

(Electron Microscope Facility, Penn State University, University Park, PA).

Staining of liver sections with BMP4 and SDF-1 antibodies. Livers were

harvested at the indicated times post PHZ induced anemia, fixed in 4% paraformaldehyde

and paraffin embedded tissue sections cut. Sections were de-paraffinized using

Histoclear (National Diagnostics, Atlanta, GA) and rehydrated through ethanol gradients

before blocking endogenous peroxidase activity for 15 minutes in 3% H2O2 in PBS.

After 3x 2 min washes in PBS, slides were blocked with Protein Blocking Agent

(Immunotech, Marseille, France) in a humid chamber for 40-60 min. Expression of

BMP4 was analyzed using anti-BMP4 (Novocastra Laboratories/Vector Laboratories,

Burlingame, CA) (1:20) in PBS + 1% BSA (Fisher Scientific, Hampton, NJ) and

expression of SDF-1 using anti-mouse CXCR12 β subunit (eBiosciences, San Diego,

CA) (1:200) in PBS + 1% BSA. After 3x 5min washes with PBS, HRP-conjugated

secondary antibodies were diluted in PBS +1 % BSA (1:50) for goat anti-mouse, or

(1:100) for goat anti-rabbit, and incubated for 40 min. After washing in PBS, HRP

activity was detected using HRP detection kit (Pharmingen, San Diego, CA) according to

manufacturer’s instructions. Cells were counterstained with hematoxylin where

indicated.

Page 130: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

120

Results

The kinetics of recovery from acute anemia is delayed in splenectomized mice.

Analysis of the flexed-tail (f) mutant mice has shown that the contribution from stress

erythroid progenitors resident in the spleen, which respond to hypoxia-induced

BMP4/Madh5 dependent signals, is critical for quickly re-establishing homeostatic

hematocrit levels following an erythropoietic challenge [7]. Although this mechanism is

required for the rapid recovery from an acute anemia, it is clearly dispensable for bone

marrow steady state erythropoiesis since flexed-tail mice are viable. This point is further

demonstrated by the observation that both mice and humans can survive without a spleen.

To determine if splenectomized mice have a defect in their recovery to an acute anemia,

we followed the hematocrit of splenectomized mice after a single dose injection of PHZ,

dropping hematocrit from 50% to below 40% within 12 hrs. This allows us to evaluate

the severity of splenectomy during the initial phase of expansive erythropoiesis and

characterize the compensatory mechanism of stress erythropoiesis during the critical

window of splenic participation. Surprisingly, splenectomized mice show no delay in

their recovery of normal hematocrit values following an acute anemia (Figure3-1). Both

splenectomized and control mice re-establish 50% hematocrit values by eight days post-

treatment. Although splenectomized mice recover in the same time as control mice, they

show different kinetics of recovery. Control mice stabilize and show a rise in hematocrit

by three days after treatment, which correlates with the maturing stress BFU-E in the

spleen which peaks at 36 post treatment [1]. Splenectomized mice, lacking a source of

Page 131: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

121

the stress BFU-E from the spleen, show a longer lag period with their hematocrit not

stabilized or increasing until six days post treatment.

Bone marrow erythropoiesis does not increase in splenectomized mice during the

recovery from acute anemia. Since the defect in the recovery to an acute anemia is not

that drastic in splenectomized mice, we wanted to determine how splenectomized mice

compensate for the loss of the erythropoietic inductive environment (EIM) of the spleen.

An obvious location for compensation would be the bone marrow. The bone marrow is

the location of homeostatic erythropoiesis and bone marrow BFU-E require two signals

to differentiate, Epo plus a burst promoting factor, such as IL3. The absence of a spleen

may augment the production of erythroid progenitors in the bone marrow. To test this

possibility we used in vitro colony assay to evaluate the number of early erythroid

progenitors (BFU-E) over the recovery period. We saw only a small elevation in the

number of BFU-E in the bone marrow of splenectomized mice both at 12 hours and 48

hours (less than two fold) (Figure 3-2). It seems that these slight differences in steady

state BM BFU-E, although contributing to the recovery, did not appear great enough to

explain the near normal recovery time of the splenectomized mice. This is consistent

with that reported by Bozzini et al. [3] where splenectomized mice with extended

exposure (9 day; 23,000ft, 19hr/day) to hypoxia or injected with erythropoietin showed

no increase in the rate of erythropoiesis in their bone marrow, though they were capable

of increasing their total circulating red cell volume.

Perhaps the differences seen in the bone marrow were not detected under the

conditions used for classical BFU-E. It is known that the spleen harbors erythroid

Page 132: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

122

progenitors poised for expansive erythropoiesis with unique properties that distinguish

them from classical bone marrow BFU-Es. Endogenous splenic erythroid progenitors

proliferate more quickly than classical BM BFU-E (5 days vs. 7-10 days in culture

respectively), to produce mature erythrocytes more rapidly. Unlike progenitor cells from

the BM which show little response to BMP4, endogenous splenic erythroid progenitors

differentiate following exposure to BMP4, to a “stress BFU-E.” These stress BFU-E

expand in the presence of high levels of Epo, independent of a factor with burst-

promoting activity [1]. To determine if the bone marrow of splenectomized mice harbor

progenitors with properties more like the splenic stress BFU-E poised for expansive

erythropoiesis, we did in vitro colony assay for cells that respond to BMP4 + Epo, or Epo

alone. There were no statistical differences seen in the number of BMP4 responsive

erythroid progenitors or in erythroid progenitor that responded to Epo alone from the

bone marrow of control and splenectomized mice (data not shown). These data

demonstrate that the bulk of the compensation for the lack of the EIM of the spleen or its

endogenous progenitors is not seen from the bone marrow. Even under this extreme

condition, the bone marrow does not alter its microenvironment or spectrum of

progenitor cells to support expansive erythropoiesis.

The liver becomes the site of extramedullary erythropoiesis in splenectomized mice.

Since the bone marrow is not making any dramatic contribution to the timely recovery of

splenectomized mice, another hematopoietic organ must be involved. Erythropoiesis is

known to be supported in non-splenic extra-medullary sites including the liver, axillary

lymph nodes and thymus following erythropoietic stimulation. It is known that the liver

Page 133: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

123

and lymph nodes in splenectomized mice show a considerable increase in red cell

production through morphological evidence and the incorporation of 59Fe by these tissues

following extended exposure to a hypoxic atmosphere (20-25 days) or injected with

9units Epo daily for 14 days [3]. We looked at H&E stained liver sections from control

and splenectomized mice over the recovery period to the acute anemia for evidence of

hematopoietic activity (Figure 3-3). A few, tiny hematopoietic clusters (staining dark

purple against the pink hepatocytes) can be seen in control mice by day 4 post treatment,

but these clusters are small and sparse, with no great increase in their numbers over the

recovery period. In sharp contrast to the control mice, the liver of splenectomized mice

show the emergence of these clusters at day 4, with prominent, wide spread, large-sized

clusters by day 6, continuing at least through day 8 post treatment. Closer inspection of

these clusters (Figure 3-3B) reveals the presence of red staining mature erythrocytes at

the center, surrounded by more immature erythropoietic progenitors. The data

demonstrate that without a spleen present, robust expansive erythropoiesis occurs in the

liver.

BFU-E expand in the liver during the recovery from an acute anemia but are not

resident in the liver. Before treatment, no appreciable numbers of erythroid progenitors

are present in the livers of control or splenectomized mice as determined by in vitro

colony assays (Figure 3-4). This is consistent with what others have reported in both the

murine [6] and human systems [12]). However, from the H&E stained sections it is clear

that the liver becomes a critical contributing microenvironment for erythroid progenitors

following the acute anemia in splenectomized mice. We analyzed BFU-E in the liver by

Page 134: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

124

in vitro colony assays. During recovery from PHZ induced anemia, BFU-E expanded in

both splenectomized and control mice, however the increase was much greater in

splenectomized mice. Between untreated and 3 days post PHZ treatment, livers of

splenectomized mice exhibited a 65-fold expansion of BFU-E compared, which is more

than 3 times greater than the 20-fold increase seen in wild-type livers (Figure 3-4). The

fact that no erythroid specific progenitors are found in untreated mice suggests that the

microenvironment of the liver is not fully suited to replace that of the spleen, in that it

does not harbor erythroid progenitors poised to respond to an erythropoietic stress.

Liver expresses BMP4 and liver erythroid progenitors exhibit properties similar to

spleen stress BFU-E. BMP4 is the key signal that initiates stress erythropoiesis in the

spleen. It is known that BMP4 is induced in the spleen 24 hours following induction of

an acute anemia. The spleen stromal line MSS31 grown under hypoxic conditions also

upregulates the expression of BMP4 [7]. In order to test whether the liver also

upregulates BMP4, we stained liver sections from splenectomized and control mice for

BMP4. This analysis showed that BMP4 is constitutively expressed in the liver in both

control and splenectomized mice over the recovery period (Figure 3-5). Not all cells in

the cross section stain positively for BMP4 expression with staining often concentrated

around the bile ducts. However, there appears to be no correlation between the

hematopoietic clusters, and localization of BMP4 expression.

Although BMP4 expression does not appear to be regulated by hypoxia in the

liver as it is in the spleen, the fact that it is expressed in the liver during the recovery to

acute anemia reveals a potential role of BMP4 signaling for the expansion of progenitors

Page 135: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

125

in the liver following erythropoietic challenge. These observations suggest that the liver

progenitors may respond in a similar manner to that seen with endogenous progenitors in

the spleen. In the spleen there are two types of progenitors. An immature progenitor,

the BMP4 responsive cell (BMP4R), which responds to the upregulation of BMP4 in the

spleen following an acute anemia and differentiates into an Epo responsive stress

erythroid progenitor (stress BFU-E). Splenic erythroid progenitors only need transient

exposure to BMP4 to render them Epo responsive. If BMP4R cells similar to those

observed in the spleen migrate to the liver in splenectomized mice, they would be

converted to stress BFU-E, which form BFU-E colonies in the presence of Epo alone. To

determine if this was true of erythroid progenitors in the liver, in vitro colony assays were

repeated under the condition of Epo + BMP4 or Epo alone. The majority of BFU-E

colonies seen at day 3 post anemia in splenectomized mice are actually BFU-E that are

responding to Epo alone (Figure 3-4B). Including BMP4 in the media did not increase

the number of BFU-E, which is consistent with the idea that the endogenous BMP4

expression in the liver is sufficient to drive the differentiation of any potential BMP4R

cells (data not shown). These progenitors forming colonies in the liver, encountering a

BMP4 signal already present, are now capable of rapid proliferation independent of a

burst promoting factor, which is a characteristic more similar to splenic stress BFU-E

than steady-state BM erythroid progenitors.

No evidence of erythropoietic activity in the liver of flexed-tail mice during the

period it is observed in splenectomized mice. When the spleen is absent, there is

morphological evidence for erythropoiesis in the liver starting at day 4 post PHZ-induced

Page 136: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

126

acute anemia, with prominent clusters observed at day 6. Although the spleen is present

in flexed-tail mice, the mechanisms of expansive erythropoiesis in the spleen are

defective. We looked at H&E liver from flexed-tail mice at day 4 and 6 post PHZ-

induced acute anemia to determine if flexed-tail mice increase erythropoiesis in the liver

to compensate for their defective splenic mechanisms. We saw no prominent

hematopoietic clusters in the liver of flexed-tail mice as we had seen at day 4 and 6 in

splenectomized mice (Figure 3-6). The liver expresses BMP4 during the recovery and

confers properties to erythroid progenitors in the liver of splenectomized mice similar to

those seen in progenitors expanding in the spleen of wild-type mice. flexed-tail mice

show a delay in the expansion of the stress erythroid progenitors in their spleen, due to

the inability to respond to BMP4 signals regulating their expansion. It is unclear if the

lack of erythropoiesis in the liver of flexed-tail mice at these time points is due to the

preference for spleen microenvironment for expansive erythropoiesis, even if it is

defective, or if the mechanism in the liver would also be delayed considering the role

BMP4 has in the expansion of stress erythroid progenitors in the liver.

SDF-1 is expressed in the livers of wild-type and splenectomized mice. Despite the

fact that the liver is capable of supporting expansive erythropoiesis, it is clearly a second

choice to the spleen. Although the liver is expressing BMP4 at the same time with the

spleen following an acute anemia, no significant erythropoiesis is occurring in the liver

when a spleen is present. This suggests that there must be a signal for cells to specifically

home to the liver and initiate erythroid differentiation once it is determined there is no

splenic contribution to the recovery. One possible candidate for this homing signal is the

Page 137: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

127

chemokine stromal cell-derived factor-1 (SDF-1 or CXCL12). SDF-1 through interaction

with its receptor CXCR4, has been established as a key regulator of murine stem cell

migration, homing, and anchorage of repopulating cells to the bone marrow, as well as

release of maturing cells into the blood stream [13]. SDF-1 has been shown to be

expressed by liver bile duct epithelium with expression increased in the liver by stressors

such as irradiation or inflammation. This upregulation leads to the SDF1 dependent

recruitment of CD34+/CXCR4+ hematopoietic progenitors to the liver for repair [14].

Ceradini et al. has shown SDF-1 expression can be regulated by other physiological

stressors such as ischemia. The hypoxia inducible factor HIF-1 is up-regulated in

endothelial cells in ischemic tissue in direct proportion to reduced oxygen tension. HIF-1

levels mediated by the hypoxic gradient induced SDF-1 expression which increased the

adhesion, migration and homing of circulating CXCR4+ progenitors cells to the

ischemic tissue for tissue regeneration [15]. An acute anemia creates hypoxic conditions

in the tissues, which could lead to increased expression of SDF-1 in the liver of

splenectomized mice, and the homing of progenitor cells to this organ for expansion. To

address this possible mechanism of progenitor cell recruitment, immunohistochemical

analysis of liver sections for SDF-1 expression from control and splenectomized mice

was performed over the recovery period (Figure 3-7 and Figure 3-8). In both

splenectomized and wild-type mice, SDF-1 can be seen in hepatocytes just below the

endothelial layer of hepatic vessels, as well as the bile duct epithelia. In untreated mice,

there is much more overall staining in splenectomized mice compared to wild-type,

especially in hepatocytes surrounding vessels. By 12 hrs post PHZ, both splenectomized

and wild-type mice show SDF-1 expression throughout the liver, with more over-all

Page 138: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

128

staining apparent in the splenectomized mice. The most dramatic increase in SDF-1 is

seen at 48 hours, as SDF-1 expression in both is more punctuated as individual

hepatocytes show high level expression compared to more moderate amounts in

neighboring cells. At 4 and 6 days post anemia, SDF-1 expression evens out throughout

the liver, although individual high-expressing cells can still be seen below the endothelial

lining of the vessel walls. In splenectomized mice at day 8 there is a slight elevation in

over-all SDF-1 staining compared to wild-type controls, but levels have decreased in both

and almost returned to that seen prior to the anemia. Because SDF-1 is expressed in the

liver during the recovery from an acute anemia, it may have a role in the recruitment of

erythroid progenitors, but the results are not conclusive at this time.

Page 139: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

129

Discussion

The splenic mechanisms of expansive erythropoiesis during the immediate

response to an acute anemia are critical for the rapid return to steady state blood

parameters, but are not essential for survival as splenectomized mice are viable. Our

results have shown splenectomized mice display different kinetics of recovery to an acute

anemia. Consistent with previous reports, erythropoiesis does not increase substantially

in the bone marrow during the immediate response to acute anemia of splenectomized

mice, but the liver now becomes an important organ for re-establishing homeostasis. The

liver [6], lymph nodes [3], thymus [16] have all been shown to participate in

erythropoietic activity under various conditions including bone marrow suppression from

radiation exposure [12, 17], malaria infection [18], phenylhydrazine induced hemolytic

anemia [19] and splenectomy [3]. The liver is of particular interest since it was the

primary site of hematopoiesis during development supporting the expansion of erythroid

progenitors. Although once supportive of expansive erythropoiesis, the liver

microenvironment loses this ability as cells differentiate from stroma of epithelial-to-

mesenchymal transition (EMT) to mature hepatocytes, directed by the increase in

oncostatin M (OSM) from the influx of hematopoietic progenitors from the yolk sac and

AGM [20]. Though no longer specialized for erythropoietic activity, the hematopoietic

potential in the liver is re-established under stress conditions sharing similarities with the

erythropoietic supportive environment of the spleen.

Evidence of the erythropoietic capacity of the adult liver has been determined

morphologically by the existence of erythroblastic islands forming in the liver following

Page 140: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

130

an acute anemia using an extended phenylhydrazine regimen (over a 4 day period) [6].

Bozzini et al. took splenectomized mice exposed to hypoxia or injected with

erythropoietin, and observed the inability of the splenectomized mice to fully respond to

these erythropoietic stimuli, although noted an increase in red cell production in the liver

(and lymph nodes) [3]. These data verifies the ability of the adult liver to support

erythropoiesis, but does not define the parameters necessary for establishing the suitable

erythropoietic microenvironment, the nature of the participating progenitors, or its role in

the immediate recovery to an acute anemia in the absence of a spleen. We have shown

that parallels can be drawn between expansive erythropoiesis seen in the spleen and

erythropoiesis initiated in the liver. BMP4 is expressed in both the spleen and the liver

following an acute anemia. In the spleen BMP4 is tightly regulated and its up-regulation

is coordinated with the onset of the acute anemia. In the liver there is no great fluctuation

in BMP4 levels but rather it is ubiquitously expressed. Regardless, a large proportion of

progenitors that migrate to this BMP4 rich environment, experience a BMP4 signal and

are now capable of responding to Epo alone, independent of a factor with burst

promoting activity. We have not observed the migration of BFU-E that can expand in the

presence of Epo alone in the blood. The characteristic of stress BFU-E to respond to Epo

alone sets expansive erythroid progenitors, whether endogenous to the spleen or those

seen in the liver, apart from their steady-state bone marrow counterparts. Although the

majority of the progenitors participating in the liver exhibit stress BFU-E properties,

some steady state BFU-E are present because inclusion of IL-3 increases the number of

BFU-E in the liver.

One difference between progenitors participating in the spleen, and those in the

Page 141: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

131

liver of splenectomized mice is their origin. BMP4 responsive cells reside in the spleen

of mice under homeostatic conditions. We know that the population of BMP4-responsive

cells in the spleen participating in expansive erythropoiesis show limited ability to self-

renew and must be replenished following an erythropoietic challenge. A full response of

this population can not be re-detected after the initial 36 hr burst until 21 days after the

onset of the anemia (Perry JP, unpublished results). Through transplantation experiments

it appears that the source of replenishing the splenic BMP4 responsive progenitors is the

bone marrow. Signals within the splenic microenvironment render the transplanted bone

marrow cells capable of responding to BMP4, since the un-transplanted bone marrow

progenitors show no appreciable response to BMP4 signaling before homing to the spleen

microenvironment (JP Perry in preparation). Considering that this population of

progenitors with BMP4 responsive potential originates in the bone marrow, the extreme

conditions of lacking the spleen may alter conditions enough to allow those normally sent

to the spleen, to be retained in the bone marrow or sent to the liver. Neither of these

situations is consistent with what we have seen in our colony assays from the bone

marrow or liver. There are no great increases in BMP4 responsive cells, progenitors that

can respond to Epo alone, or steady-state BFU-E (requiring a burst promoting factor) in

the bone marrow or liver prior to the onset of the anemia. Although capable of

supporting erythropoiesis, the liver microenvironment is not sufficient to replace the

splenic microenvironment to maintain an expansive erythroid progenitor population.

Although the liver lacks erythroid progenitors it does contain both CFU-S and

hematopoietic stem cells, both capable of differentiating down the erythroid pathway [17,

21]. The fact that no BFU-E are present within the liver prior to the onset of the anemia

Page 142: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

132

would suggest a migration from the bone marrow, or a differentiation of these multi-

potential progenitors. Most likely it is a combination of both sources. The single dose of

PHZ we use to induce a synchronous response to the acute anemia, rather than multiple

doses initiating multiple waves, is important for characterizing the immediate response to

acute anemia. However, it makes establishing a connection between what others have

reported for the migration of progenitors following an erythropoietic challenge [19, 22]

and the origins and timing of BFU-E we see in the liver, ambiguous.

Whether endogenous multipotential cells are differentiating or progenitors are

migrating from the bone marrow, we know that no appreciable amount of erythropoiesis

occurs in the liver if a spleen is present, even if the splenic mechanisms are defective.

H&E stained liver sections from flexed-tail mice at Day 4 and Day 6 of recovery from

PHZ acute anemia are comparable to wild-type. We have seen that exposure to BMP4

alters the growth factors requirements of erythroid progenitors under erythropoietic

stress, but how do progenitor cells migrate to the appropriate microenvironment that can

support their expansion? The observation that CXCR4-positive endothelial progenitor

cells are specifically recruited to sites of injury by hypoxic gradients via HIF-1-induced

expression of SDF-1 and the observation that SDF-1 is important for the mobilization of

hematopoietic progenitor cells, made SDF-1 up-regulation in the liver of splenectomized

mice an attractive model for the recruitment of erythroid progenitors during the recovery

from an acute anemia. SDF-1 is expressed in the liver following an acute anemia in both

wild-type and splenectomized mice. Though overall expression appears slightly higher in

splenectomized mice throughout the recovery from acute anemia, there are no dramatic

differences seen between mice with or without a spleen. So although SDF-1 may be a

Page 143: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

133

factor important for the recruitment of progenitor cells to the liver, there may be other

mechanisms at work to either help establish erythropoiesis in the liver of splenectomized

mice, or extinguish the response when a spleen is present. The intrinsic properties of the

progenitors themselves may be important for the regulation of their own trafficking and

will need to be addressed in future work.

However, the mechanisms affecting differences in erythropoietic activity in the

liver following an acute anemia may be less a function of microenvironmental factors or

characteristics intrinsic to the progenitors themselves and more a function of the

physiology of the organism. The liver is unique among organs in that it receives the

majority of its blood supply (approx. 75%) from the hepatic portal vein which carries

venous blood that is largely depleted of oxygen. The blood entering the liver by the

hepatic portal vein comes from the digestive tract and the major abdominal organs such

as the spleen and pancreas [23]. The majority of progenitors mobilized from the bone

marrow would reach the spleen first. Finding a suitable environment for their expansion,

most cells could lodge in the spleen, with few progenitors continuing on through

circulation to eventually reach the liver. When the spleen is absent, progenitors do not

find a suitable environment until they reach the liver and thus the large numbers of

erythroid progenitors seen in the liver of splenectomized mice following an acute anemia.

SDF-1, known to be regulated in a HIF-1 dependent manner may show expression in the

liver of both wild-type and splenectomized mice due to the decreased oxygenation of the

blood of this organ. In fact, the level of ischemia in the tissue, or the duration of the

hypoxia may be an important component to this whole process. The wild-type mice with

a burst in BFU-E in the spleen at 36 hrs were able to stabilize their hematocrits between

Page 144: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

134

2-3 days post anemia before the liver is showing any morphological evidence of

erythropoietic activity [1]. This might suggest that a mechanism which is shutting down

the response in the spleen, is also important for shutting down any response that might

have initiated in the liver. Even if the mobilization of progenitors from the bone marrow

is a process set into motion at the point the anemic/hypoxic threshold is crossed, the early

contribution from the spleen alleviates the hypoxia to a point where the environment in

the liver no longer contains all the necessary signals to support expansion of these

progenitors.

Page 145: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

135

References

1. Lenox, L.E., J.M. Perry, and R.F. Paulson, BMP4 and Madh5 regulate the erythroid response to acute anemia. Blood, 2005. 105(7): p. 2741-8.

2. Obinata, M. and N. Yania, Cellular and molecular regulation of erythropoietic inductive microenvironment (EIM). Cell Structure and Function, 1999. 24: p. 171-179.

3. Bozzini, C.E., M.E. Barrio Rendo, F.C. Devoto, and C.E. Epper, Studies on medullary and extramedullary erythropoiesis in the adult mouse. Am J Physiol, 1970. 219(3): p. 724-8.

4. Hara, H. and M. Ogawa, Erthropoietic precursors in mice with phenylhydrazine-induced anemia. Am J Hematol, 1976. 1(4): p. 453-8.

5. Broudy, V.C., N.L. Lin, G.V. Priestley, K. Nocka, and N.S. Wolf, Interaction of stem cell factor and its receptor c-kit mediates lodgment and acute expansion of hematopoietic cells in the murine spleen. Blood, 1996. 88(1): p. 75-81.

6. Ploemacher, R.E. and P.L. van Soest, Morphological investigation on phenylhydrazine-induced erythropoiesis in the adult mouse liver. Cell Tissue Res, 1977. 178(4): p. 435-61.

7. Lenox, L.E., J.M. Perry, and R.F. Paulson, BMP4 and Madh5 regulate the erythroid response to acute anemia. Blood, 2004.

8. Hunt, H. and D. Premar, Flexed-tail a mutation in the house mouse. Anat. Rec., 1928. 41: p. 117.

9. Mixter, R. and H. Hunt, Anemia in the flexed tailed mouse, Mus musculus. Genetics, 1933. 18: p. 367-387.

10. Coleman, D., E. Russell, and E. Levin, Enzymatic studies of the hemopoietic defect in flexed mice. Genetics, 1969. 61: p. 631-642.

11. Finkelstein, L.D., P.A. Ney, Q.P. Liu, R.F. Paulson, and P.H. Correll, Sf-Stk kinase activity and the Grb2 binding site are required for Epo-independent growth of primary erythroblasts infected with Friend virus. Oncogene, 2002. 21(22): p. 3562-70.

12. Golden-Mason, L. and C. O'Farrelly, Having it all? Stem cells, haematopoiesis and lymphopoiesis in adult human liver. Immunol Cell Biol, 2002. 80(1): p. 45-51.

13. Lapidot, T. and I. Petit, Current understanding of stem cell mobilization: the roles of chemokines, proteolytic enzymes, adhesion molecules, cytokines, and stromal cells. Exp Hematol, 2002. 30(9): p. 973-81.

14. Kollet, O., S. Shivtiel, Y.Q. Chen, J. Suriawinata, S.N. Thung, M.D. Dabeva, J. Kahn, A. Spiegel, A. Dar, S. Samira, P. Goichberg, A. Kalinkovich, F. Arenzana-Seisdedos, A. Nagler, I. Hardan, M. Revel, D.A. Shafritz, and T. Lapidot, HGF, SDF-1, and MMP-9 are involved in stress-induced human CD34+ stem cell recruitment to the liver. J Clin Invest, 2003. 112(2): p. 160-9.

15. Ceradini, D.J., A.R. Kulkarni, M.J. Callaghan, O.M. Tepper, N. Bastidas, M.E. Kleinman, J.M. Capla, R.D. Galiano, J.P. Levine, and G.C. Gurtner, Progenitor cell trafficking is regulated by hypoxic gradients through HIF-1 induction of SDF-1. Nat Med, 2004. 10(8): p. 858-64.

Page 146: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

136

16. Albert, S., P. Wolf, and I. Pryjma, Evidence of Erytheropoiesis in the Thymus of Mice. J Reticuloendothel Soc, 1965. 85: p. 30-9.

17. Taniguchi, H., T. Toyoshima, K. Fukao, and H. Nakauchi, Presence of hematopoietic stem cells in the adult liver. Nat Med, 1996. 2(2): p. 198-203.

18. Halder, R.C., T. Abe, M.K. Mannoor, S.R. Morshed, A. Ariyasinghe, H. Watanabe, H. Kawamura, H. Sekikawa, H. Hamada, Y. Nishiyama, H. Ishikawa, K. Toba, and T. Abo, Onset of hepatic erythropoiesis after malarial infection in mice. Parasitol Int, 2003. 52(4): p. 259-68.

19. Ploemacher, R.E., P.L. van Soest, and O. Vos, Kinetics of erythropoiesis in the liver induced in adult mice by phenylhydrazine. Scand J Haematol, 1977. 19(5): p. 424-34.

20. Chagraoui, J., A. Lepage-Noll, A. Anjo, G. Uzan, and P. Charbord, Fetal liver stroma consists of cells in epithelial-to-mesenchymal transition. Blood, 2003. 101(8): p. 2973-82.

21. Taniguchi, H., T. Toyoshima, K. Fukao, and H. Nakauchi, Evidence for the presence of hematopoietic stem cells in the adult liver. Transplant Proc, 1995. 27(1): p. 196-9.

22. Hara, H. and M. Ogawa, Erythropoietic precursors in mice with phenylhydrazine-induced anemia. American Journal of Hematology, 1976. 1: p. 453-458.

23. Ross, M.H., Kaye, G. I. and Pawlina, W., Histology. 4th ed. 2003, Baltimore: Lippincott Williams and Wilkins. 875.

Page 147: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

137

0.0%

10.0%

20.0%

30.0%

40.0%

50.0%

60.0%

70.0%

0 12 24 36 48 3 4 5 6 8

Time post PHZ treatment

(%) H

emat

ocrit

c57 averagesplenectomized average

****

*

Hours Days

0.0%

10.0%

20.0%

30.0%

40.0%

50.0%

60.0%

70.0%

0 12 24 36 48 3 4 5 6 8

Time post PHZ treatment

(%) H

emat

ocrit

c57 averagesplenectomized average

****

*

Hours Days

Figure 3-1. Hematocrit values during the recovery to a phenylhydrzaine(PHZ) induced hemolytic anemia. Percent hematocrit (packed red blood cell volume) were plotted for C57BL/6-+/+ ( ) or C57BL/6-splenectomized ( ) mice. * indicated p<0.05, ** indicates p<0.01.

ss

Page 148: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

138

BFU

-E p

er 1

X10^

5 B

M c

ells *

***

**

Time post PHZ treatment

05

101520253035404550

0 12 24 36 48 4 6 8

C57 control

C57 splenectomized

Hours Days

BFU

-E p

er 1

X10^

5 B

M c

ells *

***

**

Time post PHZ treatment

05

101520253035404550

0 12 24 36 48 4 6 8

C57 control

C57 splenectomized

Hours Days

Figure 3-2. Analysis of BFU-E expansion of bone marrow BFU-E during the recovery from PHZ induced acute hemolytic anemia. Bone marrow BFU-E from C57BL/6-+/+ (black bar) and C57BL/6-splenectomized (white bar). 1X105 bone marrow cells were plated in methylcellulose media containing Epo (3 U/mL) and IL-3 (10 ng/mL). *indicates p<0.05, **indicates p<0.01.

Page 149: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

139

Figu

re 3

-3.

H&

E st

aine

d liv

er se

ctio

ns d

urin

g re

cove

ry fr

om a

PH

Z in

duce

d ac

ute

anem

ia.

(A) 4

00X

mag

nific

atio

n of

H&

E st

aine

d liv

er se

ctio

ns fr

om C

57B

L/6-

+/+

cont

rol a

nd sp

lene

ctom

ized

mic

e. (

B) 1

0X fu

rther

mag

nific

atio

nof

the

hem

atop

oiet

ic c

lust

ers i

n th

e re

gion

of t

he y

ello

w b

ox sh

own

in p

art (

A) [

tota

l of 4

000X

mag

nific

atio

n].

Con

trol

2 D

ays

4 D

ays

6 D

ays

8 D

ays

Sple

nect

omiz

ed

Unt

reat

ed

Con

trol

2 D

ays

4 D

ays

6 D

ays

8 D

ays

Sple

nect

omiz

ed

Unt

reat

ed

Con

trol

2 D

ays

4 D

ays

6 D

ays

8 D

ays

Sple

nect

omiz

ed

Unt

reat

edA

)

B)

Page 150: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

140

B) Epo only

A) Epo + IL3

0

20

40

60

80

100

120

140

160

Live

r BFU

-E:

1.4X

10^5

cel

ls/w

ell

O 2 3 4 5

Days post PHZ treatment

Live

r BFU

-E:

1.4X

10^5

cel

ls/w

ell

0

20

40

60

80

100

120

140

160

O 2 3 4 5

Days post PHZ treatment

control splenectomized

control splenectomized

B) Epo only

A) Epo + IL3

0

20

40

60

80

100

120

140

160

Live

r BFU

-E:

1.4X

10^5

cel

ls/w

ell

O 2 3 4 5

Days post PHZ treatment

Live

r BFU

-E:

1.4X

10^5

cel

ls/w

ell

0

20

40

60

80

100

120

140

160

O 2 3 4 5

Days post PHZ treatment

0

20

40

60

80

100

120

140

160

O 2 3 4 5

Days post PHZ treatment

control splenectomized

control splenectomized

Figure 3-4. Analysis of liver BFU-E expansion during the recovery from a PHZ induced hemolytic anemia. 1.4X105 liver cells from C57BL/6-+/+ control (black bar) or C57BL/6-splenectomized mice (white bar) were harvested at various points during the recovery to a PHZ induced acute anemia and plated in methylcellulose media containing (A) Epo (3 U/mL) + IL-3 (10 ng/mL) or (B) Epo (3 U/mL) alone. * indicates p<0.05, **indicates p<0.01.

Page 151: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

141

Figu

re 3

-5.

Ana

lysi

s of B

MP4

exp

ress

ion

in th

e liv

er d

urin

g re

cove

ry fr

om a

PH

Z in

duce

d ac

ute

anem

ia.

(A) 4

00X

m

agni

ficat

ion

of li

ver s

ectio

ns fr

om C

57B

L/6-

+/+

cont

rol a

nd C

57B

L/6-

sple

nect

omiz

ed m

ice

stai

ned

with

ant

i-BM

P4

antib

odie

s at t

he in

dica

ted

times

follo

win

g a

PHZ

indu

ced

acut

e an

emia

and

cou

nter

-sta

ined

with

hem

atox

ylin

. (B

) Liv

er

sect

ions

show

n in

par

t (A

) sta

ined

with

out t

he B

MP4

ant

ibod

y bu

tonl

y w

ith th

e se

cond

ary-

HR

P co

njug

ated

ant

ibod

y an

d co

unte

rsta

ined

as a

con

trol.

*in

dica

tes n

o co

unte

r-st

ain.

Con

trol

Sple

nect

omiz

ed

B)

Con

trol

Sple

nect

omiz

ed

B)

*

2 D

ays

4 D

ays

6 D

ays

8 D

ays

Unt

reat

ed

Con

trol

Sple

nect

omiz

ed

A)

2 D

ays

4 D

ays

6 D

ays

8 D

ays

Unt

reat

ed

Con

trol

Sple

nect

omiz

ed

A)

*

Page 152: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

142

Day 6 Day 4

f/f mouse #1

100 µ m 100 µ m

f/f mouse #2

Figure 3-6. H&E stained liver sections from flexed-tail (f/f) mice. Liver sections from two different f/f mice recovering from a PHZ induced acute anemia. At these time points (Day 4 and Day 6 post PHZ) splenectomized mice have numerous, large hematopoietic clusters present in their livers (seeFigure 3-3), which are not present in the f/f livers.

Page 153: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

143

0hr 12hr 24hr 36hr

200µm200µm

SDF-1

Control

48hr 4Day 6Day 8Day

200µm200µm

SDF-1

Control

Figure 3-7. Immunohistochemistry for SDF-1 in the liver of wild-type mice during the recovery from a PHZ induced acute anemia. Liver sections taken at various time points from wild-type mice recovering from a PHZ induced acute anemia. These sections were stained using an SDF-1 (CXCL12 β subunit) antibody, or HRP-conjugated secondary only (Control). No counter-stain was used.

Page 154: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

144

0hr 12hr 24hr 36hr

200µm200µm

SDF-1

Control

y y y

200µm200µm

48hr 4Day 6Day 8Day

SDF-1

Control

Iso Cont

Figure 3-8. Immunohistochemistry for SDF-1 in the liver of splenectomized mice during the recovery from a PHZ induced acute anemia. Liver sections taken at various time points from splenectomized mice recovering from a PHZinduced acute anemia. These sections were stained using an SDF-1 (CXCL12 βsubunit) antibody, or HRP-conjugated secondary only (Control). No counter-stain was used. Also shown is rabbit IgG serum isotype control (Iso Cont).

Page 155: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

145

Chapter 4

THE IMPACT OF flexed-tail

Abstract

The analysis of the flexed-tail locus broadens our appreciation for splicing

mutations and their physiological consequences, as well as the complexities and controls

in place within signaling networks. The flexed-tail mutation in Madh5 may be added to

the growing list of splicing-affecting genomic variants (SpaGV), which are often over-

looked as “silent” genomic variants. There is evidence that the truncated transcripts

from flexed-tail mice are producing truncated proteins. Analysis of the composition of

these truncated proteins begins to address the neomorphic properties observed as they

disrupt the TGF-β/BMP signaling pathways. We are reminded of the subtleties in place

behind precisely orchestrated signaling networks as these aberrant molecules are

controlled from propagating widespread detrimental effects.

Page 156: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

146

The Impact of Our Analysis of the flexed-tail Locus on Our Appreciation for

Splicing Mutations and Their Physiological Consequences

We have shown through allelism and functional assays that the cause of the

flexed-tail phenotype is a splicing mutation in Madh5 (Smad5). Although disappointing,

the lack of the actual mutation causing the splicing abnormalities may not be as

surprising as originally perceived. Considering splicing as straight-forward as two

consecutive transesterification reactions joining exons while removing introns

distinguished by minimal motifs (GU and AG dinucleotides at exon-intron and intron-

exon junctions, a polypyrimidine tract and a branch point A) is simplistic at best. Recent

work has demonstrated the relevance of splicing enhancers and silencers regulating

splicing. These elements can appear as nonsense mutations contained within exons,

portions of introns, or sequences completely outside the gene, located from tens to

thousands of base pairs from the splice sites they affect. A recent review by Pagani and

Baralle (2004), emphasizes the need for diligence beyond sequence comparison when

evaluating for the molecular basis of disease, for far too often over-looked “silent”

genomic variants (GV) appear harmless because they do not affect amino acid coding

sequence or splice junction nucleotides. The effect of these single nucleotide

substitutions or small insertions and deletions, often in SNPs and simple sequence

repeats, must in their opinion be evaluated in splicing functional assays to determine

which are deleterious and which are really benign. These splicing-affecting genomic

variants (SpaGV) lead to splicing abnormalities which include inducing exon skipping,

activation of cryptic splice sites or alterations in the balance of alternatively spliced

Page 157: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

147

isoforms, all could potentially lead to disease states [1]. Considering the daunting task

for the splicing machinery to correctly identify exons, only 145 nucleotides [2] on

average, separated by sometimes kilobases of sequence contaminated with pseudogenes

and cryptic splice sites, it is not surprising that weakly defined splice junctions whose

surrounding nucleotides fall outside the most highly conserved consensus [3] [as is the

case of the 5’splice junction of intron 1 of Madh5: score of 66.5 out of 100; see Figure 4-

1 for description of calculation], would be highly susceptible to perturbations caused by

seemingly benign genetic variations.

Aside from exon splicing enhancers (ESE), which are often bound by the splicing

activating serine arginine (SR) protein, or the exon splicing silencers (ESS), which bind

heterogeneous nuclear ribonucleoproteins particles (hnRNP), Pagani and Baralle shed

light onto more “hidden pathways to disease.” Their analysis examines splicing-affecting

genomic variants (SpaGV) that stem from variants affecting RNA secondary structure,

modifiers of SpaGVs, and the transcription-splicing connection. Variants in promoter or

transcriptional enhancer elements often affect the splicing process due to the coordinate

regulation of transcription and splicing in vivo. Our sequence analysis of Madh5 did not

include the promoter, or the entire 3’ UTR (our reference cDNA sequence GI# 6678773

has been replaced by GI#42734445, which has roughly 2000 more base pairs at the 3’

end of the final exon) which could contain sequences important for transcription

regulation. Mindful of the potential for subtle variations leading to splicing anomalies,

we had been focused on the analysis of a potentially variant polyT tract in intron 4 of the

flexed-tail mutant (see Appendix A). Whether the flexed-tail mutation in Madh5 can be

added to the growing list of SpaGV in genes that contribute to pathologies including

Page 158: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

148

ataxia telangiectasia [4], nerofibromin 1 [5], and cystic fibrosis transmembrane regulator,

[6-8] will be a matter of time and diligence.

As genomes are more thoroughly analyzed beyond the mere content of genes,

more will be learned about mutations that can affect splicing. Those genome variations

which lie outside the coding region of the gene contribute to the complexity of the

molecular mechanism of splicing and their analysis will augment our limited knowledge

of the splicing process. From this it may become more evident if and how a mutation that

may exist in a locus such as sideroflexin could be modifying the flexed-tail splicing

defect in Smad5 and contributing to the variations seen among mice. Questions such as

these will be less complex when we are able to analyze the progeny from our current

breeding of f/+; sideroflexin +/+ crosses that will separate the loci.

Page 159: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

149

The Impact of Our Analysis of the flexed-tail Locus as It Relates to the TGF-β

Family Signaling Pathway

We have shown that the flexed-tail Madh5 splicing mutation leads to mice having

defective BMP4 signaling. The defect not only disrupts the balance of correctly spliced

transcripts, but also creates aberrantly spliced truncated transcripts that possess

neomorphic properties. A general understanding of the intricacies of the TGF-β

signaling pathway suggests that these mutant forms could overcome the built-in

safeguards of this signaling system.

TGF-β/BMP signaling. The TGF-β superfamily of secreted signaling molecules is

comprised of 35 structurally related pleiotropic cytokines in vertebrates including the

transforming growth factors (TGF-βs), bone morphogenic proteins (BMPs), activins,

nodals and related proteins [9]. They are important molecules initiating a diverse range

of cellular processes from differentiation, proliferation, migration and apoptosis with key

roles in development and carcinogenesis. TGF-β responses can be cell type specific and

are dependent on both the concentration of TGF-β signaling components and the activity

of other signaling pathways which can synergize or antagonize the TGF-β pathway.

Although this pathway appears simple, combinatorial interactions in the heteromeric

receptor and Smad complexes, receptor-interacting and Smad-interacting proteins, and

cooperation with sequence-specific transcription factors allow substantial versatility and

diversification of TGF-β family responses [10].

Page 160: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

150

The extra-cellular TGF-β family members signal by stimulating the formation of

specific heteromeric complexes of type I and type II serine/threonine kinases membrane

receptors which exist in homodimers at the cell surface in the absence of ligand (Figure

4-2) [10]. There are five known mammalian type II receptors which bind the TGF-β

ligands and phosphorylate and activate the type I receptors, of which there are seven

mammalian members. The combination of type II and type I receptors in this tetrameric

complex allows differential ligand binding or differential signaling in response to the

same ligand [11, 12]. For example the type II receptor Act RII can combine with the type

I receptor ActRIB (ALK4) and mediate activin signaling, but its interaction with the

BMP-RIA (ALK3) allows BMP binding and signaling (Figure 4-3). Thus, a ligand can

induce different signaling pathways depending on the composition of the receptor

complex. TβRII interacts not only with the type I receptor TβRI (ALK5) which activates

Smad2 and Smad3, but also with ALK1 to activate Smad1 and Smad5 [10].

The activated type I receptors, which are responsible for downstream signaling

specificity [12], phosphorylate and thus stimulate intracellular signaling molecules

known as Smads. Smads are highly conserved molecules first identified as the products

of the Drosophila Mad and C. elegans Sma genes, which lie downstream of the BMP-

analogous ligand-receptor systems in these organisms [12-14]. Functionally, Smads fall

into three subfamilies: Receptor regulated Smads (R-Smads: Smad1, Smad2, Smad3,

Smad5, Smad8,), the common mediator (Co-Smads: Smad4), and the inhibitory Smads

(I-Smads: Smad6 and Smad7). R-Smads are further subdivided into two classes: Smad2

and -3 which transduce activin/TGF-β signals, and Smad1, -5, and -8 which preferentially

transduce BMP signals. Access of the R-Smads to the type I receptor is facilitated by

Page 161: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

151

auxiliary proteins such as Smad anchor for receptor activation (SARA) [9], however such

an accessory protein has not been conclusively identified in the BMP pathway [12]. R-

Smads and Smad4 are expressed in most if not all cell types [10], while the induction of

Smad6 and Smad7 is highly regulated by extracellular signals [12, 15-17].

Certain structural elements in the receptor, such as the 9 amino acid L45 loop [18]

and within the MH2 domain of Smad molecule itself such as the L3 loop [19] along with

alpha helix 1 [20], are important for dictating the specificity of recognition of the type I

receptor and the Smad protein (Figure 4-4). It is noted that these regions provide

minimal components for specificity in vitro and larger regions may be important for the

correct assembly of receptor complexes, accessory adaptor molecules and Smad protein

complexes in vivo [20].

The R-Smads and Co-Smads share two conserved domains: N-term Mad

homology-1 domain (MHI) and the C-term Mad-homology -2 domain (MH2), separated

by the more variable linker region (Figure 4-5). I-Smads show only weak homology with

the N-term MHI domain, although their MH2 domain shows high homology to the MH2

domains of R-Smads and Co-Smad. The MHI domain is important for DNA binding and

interaction with transcription factors. It contains a lysine-rich nuclear localization

sequence that is conserved in all R-Smads and been shown in Smad1 and Smad3 to

confer nuclear import [21]. Although Smad2 also has this lysine-rich sequence in its

MHI domain, upon phosphorylation by a type I receptor it is released from the anchoring

SARA and translocates to the nucleus by a cytosolic-factor-independent import activity

that requires a region in the MH2 domain [22]. While Smads alone can bind to specific

DNA sequences, their binding affinity is considered to be too weak to serve as effective

Page 162: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

152

and highly specific DNA binding proteins in vivo [23, 24]. The linker region of Smad1/5

contains MAPK phosphorylation sites, PSXP motifs, and their phosphorylation has been

shown to inhibit activated Smads from nuclear accumulation. This ERK-mediated

inhibition of nuclear accumulation occurs without interfering with the formation of

Smad1-Smad4 complex [25]. The MH2 domain is the motif which interacts with the

receptor (R-Smads, Smad6/7), mediates protein-protein interactions with Smads,

transcriptional coactivators and corepressors. The MH2 domain of both R-Smads and Co-

Smad activate transcription primarily through the ability of this domain to recruit the

general transcriptional coactivators p300 and CBP [10, 15, 24, 26].

Both the MHI and MH2 have been shown to be involved in protein-protein

interactions with a continually growing list of transcription factors. These transcription

factors can act as specificity determinants because they can be cell-type specific and be

regulated by other signaling pathways [9]. The plethora of interacting proteins provides a

mechanistic explanation for the documented crosstalk between the Smad pathway and

many other signaling networks, which range from the Ras/MAPK pathway to the Wnt/β-

catenin and nuclear hormone signaling cascades [15, 16].

Inactive cytoplasmic Smads are intrinsically auto-inhibited through an

intramolecular interaction between the MH1 and MH2 domains [12, 27]. In vitro, N-

terminal deletion mutants of Smad1 have been used to mimic activated Smad1 [28].

Upon ligand binding, the type II receptor phosphorylates the type I receptor, which in

turn propagates the signal by phosphorylating R-Smads at the SSXS motif at the C-term

end of their MH2 domain. Once activated, Smad molecules go through a conformational

change which relieves the inhibitory folding of the MHI domain on the MH2 domain.

Page 163: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

153

This open conformation allows the R-Smad to form heteromeric complexes with the Co-

Smad, Smad4, with oligomerisation achieved by the interaction of the phosphorylated C-

term tail of R-Smads with the L3 loop of another Smad [12, 29]. This R-Smad/Co-Smad

complex accumulates in the nucleus, and controls gene expression in a cell-type specific

and ligand dose-dependent manner through interactions with transcription factors,

coactivators and corepressors [12, 30].

Negative regulators of TGF-β/ BMP signaling. Regulation of the TGF-β family

pathway is achieved at various levels, including extra-cellularly, at the membrane site and

intra-cellularly (refer to Figure 4-2). Many of these factors are TGF-β/BMP-inducible

and inhibit the BMP pathway, thus establishing negative feedback loops. The active

local concentrations of TGF-β ligands have been shown to be important for rendering a

particular biological effect, especially during development. The active local

concentration and consequently the extent of action of these morphogens are controlled in

part by the influence of extracellular modulators. In vertebrates, these extracellular

modulators include noggin, chordin, chordin–like, follistatin, FSRP, the DAN/Cerberus

protein family and sclerostin [31]. Noggin for example has been shown to bind to and

antagonize BMP-2, -4, and -7 signals by interfering with the ability of the BMP to bind

cognate cell-surface receptors [32].

At the membrane, BMP and activin membrane bound inhibitor (BAMBI) is a

molecule which shows sequences similarity to TGF-β receptors but lacks the intracellular

kinase domain. As a naturally occurring dominant-negative pseudoreceptor whose

expression appears to be induced by BMPs, BAMBI blocks TGF-β, activin, and BMP

Page 164: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

154

signaling by interacting with receptors of the TGF-β family and preventing the formation

of functional receptor complexes [24, 31, 33, 34].

Within the cytoplasm I-Smads and Smad ubiquitination regulatory factor (Smurf)

are some of the more well characterized modulators. Induction of Smad6 and Smad7 is

highly regulated by extracellular signals including epidermal growth factor (EGF), TGF-

β1, Activin and BMP-7, suggesting the existence of auto-inhibitory feedback

mechanisms [12, 15-17]. In fact, the promoter region of the mouse Smad6 contains a

BMP responsive element, which directly binds a Smad4/5 complex [35]. Inhibitory

Smads, through their MH2 domains [10], can bind stably to the intracellular domain of

activated BMP type I rectors and prevent activation of the R-Smads by the receptor [36-

38]. They have been shown to recruit a complex of GADD34 and the catalytic subunit of

protein phosphatase 1 to the TGF-β type I receptor to de-phosphorylate and inactivate the

receptor. In addition Smad6 has been shown to compete with Smad4 for binding to

receptor-activated Smad-1 yielding apparently inactive Smad1-Smad6 complexes [24,

39].

In addition to competing for receptors and/or Smad4 binding, or by recruiting

phosphatase complexes to the activated receptors to attenuate signaling, I-Smads recruit

Smad ubiquitination regulatory factors (Smurf1 and Smurf2) to components of the

pathway. Smurf1 and Smurf2 are members of the HECT (homologous to E6-associated

protein C terminus) class of E3 ubiquitin ligases. Smurf1 interacts with Smad1 and -5

through the PPXY motif (PY motif) located in the linker region of Smad1 and targets

these molecules for ubiquitination, leading to proteasomal degradation [24, 40, 41].

Page 165: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

155

Smurf1 also binds to BMP type I receptors via I-Smads and induces ubiquitination and

degradation of these receptors [28, 31].

Within the nucleus multiple factors have been shown to act as transcriptional

corepressors of BMP-Smad signaling. As an example, the cellular counterpart of the

retroviral oncogene (v-ski) known as c-ski and the related SnoN protein physically

interact with the MH2 domain of Smad2, -3, and -4 to directly repress their ability to

activate TGF-β target genes [24]. Recently, Ski was also shown to interact with the MH2

domain of Smad1 and Smad5 in a BMP signaling-dependent manner [42].

Interaction of TGF-β/BMP signaling pathway with other signaling pathways. Cross-

talk between the TGF-β/BMP pathway and other signaling pathways is an area field that

has been recently emerging. These mechanisms include the Wnt/β-Catenin,

Ca2+/Calmodulin signaling, Wnt/Ca2+ signaling, Erk-MAPK pathway, and the JAK-

STAT pathway [24]. Both the Erk-MAPK and the JAK-STAT pathway have roles in

Erythropoietin signaling and erythroid development and may prove important for our

discussion, although the analysis of their role in TGFβ signaling has been mainly in vitro.

Activation of receptor tyrosine kinases (RTKs) by ligands such as epidermal growth

factor (EGF) or hepatocyte growth factor (HGF) subsequently activates the Erk

(extracellular signal regulated kinases) subfamily of mitogen-activated protein kinases

(MAPK), which in turn phosphorylate serine residues in consensus PXSP motifs located

in the linker region of Smad1. As a consequence, Smad1 is inhibited from accumulating

in the nucleus. This balance between phosphorylation and activation by BMPs and the

opposing regulation by the RTKs phosphorylation, determines the levels of Smad1

Page 166: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

156

activity in the nucleus, and so the role of Smad1 in the control of cell fate [25].

Leukemia inhibitory factor (LIF), which acts through the gp130 receptor and STAT3, can

act synergistically with BMP2 in inducing astrocyte differentiation in culture. This is

shown to be due to the formation of Smad1 and STAT3 complex bridged by the general

transcriptional coactivator p300 [43]. Furthermore, BMP signals may be transduced by

MAP kinases in addition to Smads [24]. Considering the diverse role of the TGF-β family

members and the limited number of signaling components, the complexity and interaction

among various signaling pathways is not surprising, but much remains to be learned

about the intricacies of these events.

Neomorphic properties of the flexed-tail truncated transcripts. By stably transducing

the two truncated transcripts of Smad5 from flexed-tail into a BMP responsive bone

marrow stromal cell line (W-20-17) [44, 45], it has been established that it is not so much

the reduced levels of the full length transcript that is leading to the flexed-tail phenotype,

but more so the presence of the truncated versions (refer to Chapter 2 and Figure 2-6, pg.

110) [46]. The presence of the truncated transcripts consistently affected the basal level

of alkaline phosphatase activity in these cells, as well as rendering the lines containing

them unable to respond to BMP4 signals. BMP normally causes the cells to differentiate

with an increase seen in alkaline phosphatase activity. If the line showed any response

to BMP4, it was never to the level reached by the W-20-17 line control (empty MSCV

vector), suggesting that any response they may show may never reach physiologically

significant thresholds. Transcript #1 appeared to be the more detrimental transcript as

Page 167: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

157

this line grew more slowly than the lines containing the other constructs and consistently

showed the least response to BMP4.

So why would an osteoblast cell line which normally differentiates in response to

BMP signals be unresponsive with the presence of the flexed-tail truncated versions of

Smad5 transcript? This is consistent with the behavior of the BMP4 responsive erythroid

cells contained in the spleen and we have evidence suggesting that these transcripts

produce truncated proteins. Analysis of the transcripts using NetStart1.0 [47, 48]

predicted downstream, high potential translational start sites in-frame with the legitimate

translational start site missing from the deletion of exon 2 in the truncated transcripts

(Figure 4-6). If translation is initiated at these downstream ATGs (at 1190bp in #1; at

1370, 1388, 1412 or 1445 in #2; reference sequence is GI #6678773) , the truncated

transcripts would produce truncated Smad5 proteins of 33.05kDa and 26.8kDa from

transcript #1 and #2 respectively. To determine if this was the case, COS7 cells were

transiently transfected with the MSCV constructs containing full length Smad5 (Neo22),

truncated transcript #1 (lel-139) and transcript #2 (lel-118). Figure 4-7 verifies the

presence of the mRNA from these constructs. Lysates from these COS7 cells were

immunoprecipitated using a column bound (Seize column, Pierce) C-term Smad5 specific

anti-body (D-20, Santa Cruz). Figure 4-8 shows a band at the predicted size for a protein

made from transcript #1.

How could such truncated proteins interrupt the response to BMP signals and

disrupt the auto-regulatory loop of the signaling pathway? Closer inspection of the

composition of the protein may provide some insights. Figure 4-9 shows the comparison

Page 168: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

158

of the truncated Smad5 proteins compared to full length versions. The most notable

difference is the absence of the MH1 domain. Without an MHI domain, there is no

portion of the molecule to collapse upon and control the MH2 or active portion of the

molecule. The C-term domain has an effector function that becomes apparent when the

isolated domain is tested in transcriptional assays [49] and mesoderm induction assays

[50] [25]. In fact, groups use deletion mutant of the closely related Smad1 which lacks

the N-term Mad-homology domain to mimic activated Smad1 [28]. The fact that the

MHI domain which normally binds to the MH2 in the inactive state to prevent aberrant

activation is absent suggests that these truncated proteins may be functioning by

misregulated, overactive signaling.

Possibly more detrimental than the loss of a regulatory motif in the molecule, is

the features the truncated protein would still contain. Through the presence of the MH2

domain, the proteins would possess the necessary elements of the Smad molecule to bind

the receptor, associate with Smad4 and transactivate transcription. Alpha helix 1 and the

L3 loop of the MH2 domain are present in both truncated proteins. The critically

residues of the L3 loop (amino acids [AAs] 425 and 428 have not been disrupted by the

inserted 5 amino acids of transcript #1 (see Figure 4-9). Both transcripts contain the C-

terminal phosphorylation sites required to activate the molecule by the type I receptors.

Considering the composition of these molecules, and the mechanisms by which other

inhibitors such as BAMBI and the inhibitory Smads function, one way these proteins

could be interfering in correct Smad pathway signaling is by competing with full length

Smad5 for necessary components. The unregulated MH2 domain could bind to type I

Page 169: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

159

receptors, preventing full length Smad5 access. These truncated proteins, activated by

the receptor can interact with the Co-Smad, again causing competitive inhibition of full

length Smad5. Whether these aberrant complexes could be shuttled to the nucleus is

unclear since the N-term NLS is not present. However, if they did make it to the nucleus,

they are likely to be able to activate/repress transcription through functional

transcriptional complexes and the active MH2 domain. We’ve seen evidence for a

disruption of the negative feedback loop in flexed-tail mice by the constitutive

transcription of BMP4 in the spleens of these animals. Although both BAMBI and

Smad6 have been shown by RT-PCR to be affected during recovery in wild-type mice

(increase in BAMBI; decrease in Smad6), again, these aberrant transcripts may

themselves be competing with full length Smad5 and preventing correct regulation of the

pathway.

As signaling pathways must be tightly controlled, safeguards are in place to

prevent a single disruption from causing catastrophic consequences. The BMP signaling

pathway has established safeguards to even these aberrant active proteins through over-

lapping functions of other Smad members, tight regulation of the duration and intensity

of the signaling and the requirement for the combinatory network of other signaling

pathways. In the case of the predicted truncated proteins, evidence for their regulation is

revealed by their structure. The portion of the linker region from transcript #1 does

contain the Erk-MAPK phosphorylation sites (AAs 188, 196, 205 and 213) shown to

regulate nuclear accumulation of activated Smad1 (comparison of GI#6678773 amino

acid sequence with [25]). It is possibly, and even likely these truncated proteins would be

Page 170: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

160

highly unstable due to the action of Smurf1. The inhibitory activity of Smurf1 is not

necessarily correlated with its ability to bind to Smad1/5 directly. Smurf1 associated

with Smad1/5 indirectly through I-Smads (Smad6) can induce their ubiquitination and

degradation. Although Smurf1 did not ubiquitinate a Smad1 mutant with a deletion of it

PY motif (∆PY), it could do so in the presence of Smad6. Smurf1 induced ubiquitination

of Smad1(∆PY) more strongly than it did that of full length Smad1, suggesting that the

Smurf1-Smad6 complex targets activated R-Smads more efficiently than non-activated

R-Smads [28].

Interestingly, the paper describing the W-20-17 features [45] noted that lower

basal levels of alkaline phosphatase activity, as we had seen with the presence of the

truncated transcripts, was observed when they included TGF-β in the culture media. The

original W-20 lines, although starting at statistically lower basal levels, were still able to

differentiate and increase levels of alkaline phosphatase activity following exposure to

BMPs. Others have used C2C12 mesenchymal cells to identify genes differentially

regulated by BMP2 and TGF-β [51]. The following genes were included in a list shown

in C2C12 cells to have varied expression in the presence of BMP2 (B), BMP2/TGF-β

(BT), or TGF-β (T) in the culture conditions, with their relative expression levels denoted

in brackets: Hey1 (Hairy/enhancer-of-split related with YRPW motif 1) [B>BT>T] ,

Hes1 [down-regulated by BMP2 and up-regulated by TGF-β; TGF-β inhibited BMP2

down-regulation], (hairy/enhancer of split1), Car3 (carbonic anhydrase 3) [B>BT>T],

Tnc (Tenascin C) [B<BT<T]. Using RT-PCR from BMP4 stimulated W-20-17 cell lines

containing the stably transduced transcripts, we evaluated the levels of these genes. The

most drastic difference between control and transcript containing lines was seen in the

Page 171: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

161

Car3 gene (Figure 4-10). There was almost no Car3 expression from the line containing

transcript#1, compared to MSCV vector control which showed equivalent amounts of

Car3 and the housekeeping gene HPRT. This decrease in the amount of Car3 expression

in the presence of truncated transcript #1 suggests that these cells may be altering their

gene expression profile as if there is TGF-β present. This would be an interesting

situation since Smad5 can act downstream of both BMPs and potentially TGFβs [12, 52,

53]. These results are preliminary and more stringent evaluation would be necessary to

draw such conclusions, but the study of the truncated versions may provide insights into

how Smad5 is specified to propagate a TGF-β vs. BMP signal.

A location for this distinction between TGF-β vs. BMP signal could be occurring

at the level of the receptor complex. A point of convergence of the pathways occurs at

the type I receptor, ALK1 (refer to Figure 4-3). The maintenance and specificity of the

system requires that each member of the type I receptor family be able to discriminate

among different groups of Smad proteins [20]. It is repeatedly stated that the type I

receptors Activin receptor like kinase (ALK) 1, 2, 3 (BMPR-IA) and 6 (BMPR-IB)

phosphorylate Smad1, -5, and -8, whereas ALK 4, 5, and 7 phosphorylate 2 and 3.

However, recent studies in endothelial cells have shown that TGF-β can bind to and

transducer signals through ALK1 and ALK5 [54, 55]. ALK5 is widely expressed, but

ALK1 is predominantly expressed in endothelial cells at specific sites of interactions

between epithelial and mesenchymal cells [56]. The TGFβ-ALK5 signaling has an

opposite effects from the TGFβ-ALK1 pathways on endothelial cell behavior. ALK5

inhibits endothelial cell migration and proliferation, whereas ALK1 stimulates both

Page 172: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

162

processes [9, 57, 58]. The possibility that receptor combinations acting as a branch point

of TGF-β family signal modulation could be occurring in other cell types and involve

other type I receptors such as the closely related ALK2 is an area for future research [9].

The type of receptors expressed by a particular cell may help dictate how the cell

interprets signals from multiple TGF-β family members, and through the use of one

Smad molecule over another, how those interpretations are relayed to the nucleus to

regulate gene expression.

Page 173: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

163

The Impact of Our Analysis of the flexed-tail Locus: A Reminder of the Subtleties in

Place Behind Beautifully Orchestrated Physiological Mechanisms

Apart from regulations in place within a signaling pathway, the requirement for the

composite of signals from other pathways adds another dimension to the functions of cell

systems. The converging pathways often affect feedback loops to help fine tune the

cell’s instructions. As much as negative feedbacks may be important to quench

responses, so too are positive feedback loops to achieve threshold levels to propagate

cellular responses. In flexed-tail mice, BMP4 expression and signaling is disrupted,

along with the apparent disruption of a negative feedback loop regulating BMP4

expression, as BMP4 mRNA is always present. When we compare the signals

converging to drive stress erythropoiesis, such as hypoxia and BMP4, we can see

parallels to networks of signaling in other related systems. There is evidence for TGF-β2

signaling combining with hypoxic signals causing autocrine regulation of the upstream

TGF-β2 molecule in endothelial cells. Exposure of human umbilical vein endothelia

cells (HUVECs) to hypoxia (1% O2) increases gene expression and bioactivation of TGF-

β2 and induces the phosphorylation and nuclear transport of Smad2 and Smad3 and their

association with DNA [59]. Furthermore, hypoxia and TGF-β cooperate in the induction

of the promoter activity of vascular endothelial growth factor (VEGF), which is a major

stimulus in the promotion of angiogenesis. Optimal HIF-1α dependent induction of

VEGF promoter was obtained in the presence of Smad3 and co-immunoprecipitation

experiments revealed that HIF-1α physically associates with Smad3, through the MH1

and MH2 domains [60]. Long term hypoxic conditions result in the late up-regulation of

Page 174: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

164

TGF-β1, again suggesting the induction of an autocrine loop by hypoxia [60, 61]. The

cross-talk between Smad3 and HIF-1α signaling pathways appears to be an important

mechanism by which endothelial cells, relatives of hematopoietic cells from the

hemangioblast, respond to hypoxic stress [59]. The working model that is currently

under investigation in our lab for the network created by the BMP, hypoxia and the stem

cell factor (SCF) pathways to regulate erythroid development and stress erythropoiesis is

shown in Figure 4-11. How the cell integrates the directions from the various molecular

players will broaden our understanding of erythropoiesis and other cellular processes.

Stepping back from the molecular dissection of signaling molecules in BMP

signaling, and the signaling networks created within a cell from instructions from its

microenvironment, the next level of complexity involves the interaction of cells with new

microenvironments. This is a regular occurrence for hematopoietic progenitor cells that

must traffic to specialized microenvironments, especially when stress conditions arise. In

the case of stress erythropoiesis in splenectomized mice, we have seen that there are

signals common to different microenvironments, such as BMP4 expression in both the

spleen and liver. The conservation of important signaling pathways across different

microenvironments allows for the flexibility of progenitors to function even in

suboptimal locations. For this reason, we could postulate that splenectomized flexed-tail

mutants might show an even greater delay in their recovery kinetics than that seen in

splenectomized wild-type mice. How the parts of cell intrinsic mechanisms and

environmental cues combine to establish the whole process of proper cell trafficking for

Page 175: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

165

expansion or differentiation will be important not only for understanding stress

erythropoiesis but also for other branches of hematopoiesis and development.

Page 176: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

166

Concluding Remarks

The molecular mechanisms that regulate physiological responses within an organism are

orchestrated with such complexity that viewing them from afar would appear

undiscernibly chaotic. Through closer inspection of the parts, be it the organ, progenitor

cell, signaling molecule or transcription factor, understanding where they fit into the

whole life process becomes more obtainable. The functional parts begin to coalesce, to

provide durability, flexibility and adaptability for the organism against the stresses of its

environment. To realize that a spontaneous mutation, manifesting as the slightly kinked

tail on a mouse in a barn of an observant farmer, could years later not only introduce us

to new signals in the regulation of erythropoiesis, but also lead us to the intricacies of

gene regulation and signaling networks, is nothing less than inspiring. Its lesson

emphasizes the need to be mindful of the subtleties in the life processes, as they may be

having a bigger impact on the system as a whole than initially appreciated.

Page 177: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

167

References

1. Pagani, F. and F.E. Baralle, Genomic variants in exons and introns: identifying the splicing spoilers. Nat Rev Genet, 2004. 5(5): p. 389-96.

2. Lander, E.S., L.M. Linton, B. Birren, C. Nusbaum, M.C. Zody, J. Baldwin, K. Devon, K. Dewar, M. Doyle, W. FitzHugh, R. Funke, D. Gage, K. Harris, A. Heaford, J. Howland, L. Kann, J. Lehoczky, R. LeVine, P. McEwan, K. McKernan, J. Meldrim, J.P. Mesirov, C. Miranda, W. Morris, J. Naylor, C. Raymond, M. Rosetti, R. Santos, A. Sheridan, C. Sougnez, N. Stange-Thomann, N. Stojanovic, A. Subramanian, D. Wyman, J. Rogers, J. Sulston, R. Ainscough, S. Beck, D. Bentley, J. Burton, C. Clee, N. Carter, A. Coulson, R. Deadman, P. Deloukas, A. Dunham, I. Dunham, R. Durbin, L. French, D. Grafham, S. Gregory, T. Hubbard, S. Humphray, A. Hunt, M. Jones, C. Lloyd, A. McMurray, L. Matthews, S. Mercer, S. Milne, J.C. Mullikin, A. Mungall, R. Plumb, M. Ross, R. Shownkeen, S. Sims, R.H. Waterston, R.K. Wilson, L.W. Hillier, J.D. McPherson, M.A. Marra, E.R. Mardis, L.A. Fulton, A.T. Chinwalla, K.H. Pepin, W.R. Gish, S.L. Chissoe, M.C. Wendl, K.D. Delehaunty, T.L. Miner, A. Delehaunty, J.B. Kramer, L.L. Cook, R.S. Fulton, D.L. Johnson, P.J. Minx, S.W. Clifton, T. Hawkins, E. Branscomb, P. Predki, P. Richardson, S. Wenning, T. Slezak, N. Doggett, J.F. Cheng, A. Olsen, S. Lucas, C. Elkin, E. Uberbacher, M. Frazier, R.A. Gibbs, D.M. Muzny, S.E. Scherer, J.B. Bouck, E.J. Sodergren, K.C. Worley, C.M. Rives, J.H. Gorrell, M.L. Metzker, S.L. Naylor, R.S. Kucherlapati, D.L. Nelson, G.M. Weinstock, Y. Sakaki, A. Fujiyama, M. Hattori, T. Yada, A. Toyoda, T. Itoh, C. Kawagoe, H. Watanabe, Y. Totoki, T. Taylor, J. Weissenbach, R. Heilig, W. Saurin, F. Artiguenave, P. Brottier, T. Bruls, E. Pelletier, C. Robert, P. Wincker, D.R. Smith, L. Doucette-Stamm, M. Rubenfield, K. Weinstock, H.M. Lee, J. Dubois, A. Rosenthal, M. Platzer, G. Nyakatura, S. Taudien, A. Rump, H. Yang, J. Yu, J. Wang, G. Huang, J. Gu, L. Hood, L. Rowen, A. Madan, S. Qin, R.W. Davis, N.A. Federspiel, A.P. Abola, M.J. Proctor, R.M. Myers, J. Schmutz, M. Dickson, J. Grimwood, D.R. Cox, M.V. Olson, R. Kaul, C. Raymond, N. Shimizu, K. Kawasaki, S. Minoshima, G.A. Evans, M. Athanasiou, R. Schultz, B.A. Roe, F. Chen, H. Pan, J. Ramser, H. Lehrach, R. Reinhardt, W.R. McCombie, M. de la Bastide, N. Dedhia, H. Blocker, K. Hornischer, G. Nordsiek, R. Agarwala, L. Aravind, J.A. Bailey, A. Bateman, S. Batzoglou, E. Birney, P. Bork, D.G. Brown, C.B. Burge, L. Cerutti, H.C. Chen, D. Church, M. Clamp, R.R. Copley, T. Doerks, S.R. Eddy, E.E. Eichler, T.S. Furey, J. Galagan, J.G. Gilbert, C. Harmon, Y. Hayashizaki, D. Haussler, H. Hermjakob, K. Hokamp, W. Jang, L.S. Johnson, T.A. Jones, S. Kasif, A. Kaspryzk, S. Kennedy, W.J. Kent, P. Kitts, E.V. Koonin, I. Korf, D. Kulp, D. Lancet, T.M. Lowe, A. McLysaght, T. Mikkelsen, J.V. Moran, N. Mulder, V.J. Pollara, C.P. Ponting, G. Schuler, J. Schultz, G. Slater, A.F. Smit, E. Stupka, J. Szustakowski, D. Thierry-Mieg, J. Thierry-Mieg, L. Wagner, J. Wallis, R. Wheeler, A. Williams, Y.I. Wolf, K.H. Wolfe, S.P. Yang, R.F. Yeh, F. Collins, M.S. Guyer, J. Peterson, A. Felsenfeld, K.A. Wetterstrand, A. Patrinos, M.J.

Page 178: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

168

Morgan, P. de Jong, J.J. Catanese, K. Osoegawa, H. Shizuya, S. Choi and Y.J. Chen, Initial sequencing and analysis of the human genome. Nature, 2001. 409(6822): p. 860-921.

3. Shapiro, M.B. and P. Senapathy, RNA splice junctions of different classes of eukaryotes: sequence statistics and functional implications in gene expression. Nucleic Acids Res, 1987. 15(17): p. 7155-74.

4. Teraoka, S.N., M. Telatar, S. Becker-Catania, T. Liang, S. Onengut, A. Tolun, L. Chessa, O. Sanal, E. Bernatowska, R.A. Gatti, and P. Concannon, Splicing defects in the ataxia-telangiectasia gene, ATM: underlying mutations and consequences. Am J Hum Genet, 1999. 64(6): p. 1617-31.

5. Ars, E., E. Serra, J. Garcia, H. Kruyer, A. Gaona, C. Lazaro, and X. Estivill, Mutations affecting mRNA splicing are the most common molecular defects in patients with neurofibromatosis type 1. Hum Mol Genet, 2000. 9(2): p. 237-47.

6. Cartegni, L., S.L. Chew, and A.R. Krainer, Listening to silence and understanding nonsense: exonic mutations that affect splicing. Nat Rev Genet, 2002. 3(4): p. 285-98.

7. Chu, C.S., B.C. Trapnell, S. Curristin, G.R. Cutting, and R.G. Crystal, Genetic basis of variable exon 9 skipping in cystic fibrosis transmembrane conductance regulator mRNA. Nat Genet, 1993. 3(2): p. 151-6.

8. Pagani, F., C. Stuani, M. Tzetis, E. Kanavakis, A. Efthymiadou, S. Doudounakis, T. Casals, and F.E. Baralle, New type of disease causing mutations: the example of the composite exonic regulatory elements of splicing in CFTR exon 12. Hum Mol Genet, 2003. 12(10): p. 1111-20.

9. ten Dijke, P. and C.S. Hill, New insights into TGF-beta-Smad signalling. Trends Biochem Sci, 2004. 29(5): p. 265-73.

10. Derynck, R. and Y.E. Zhang, Smad-dependent and Smad-independent pathways in TGF-beta family signalling. Nature, 2003. 425(6958): p. 577-84.

11. Derynck, R. and X.H. Feng, TGF-beta receptor signaling. Biochim Biophys Acta, 1997. 1333(2): p. F105-50.

12. Moustakas, A., S. Souchelnytskyi, and C.H. Heldin, Smad regulation in TGF-beta signal transduction. J Cell Sci, 2001. 114(Pt 24): p. 4359-69.

13. Patterson, G.I. and R.W. Padgett, TGF beta-related pathways. Roles in Caenorhabditis elegans development. Trends Genet, 2000. 16(1): p. 27-33.

14. Whitman, M., Smads and early developmental signaling by the TGFbeta superfamily. Genes Dev, 1998. 12(16): p. 2445-62.

15. Itoh, S., F. Itoh, M.J. Goumans, and P. Ten Dijke, Signaling of transforming growth factor-beta family members through Smad proteins. Eur J Biochem, 2000. 267(24): p. 6954-67.

16. Massague, J., How cells read TGF-β signals. Nature Reviews Molecular Cell Biology, 2000. 1: p. 169-178.

17. Nakao, A., M. Afrakhte, A. Moren, T. Nakayama, J.L. Christian, R. Heuchel, S. Itoh, M. Kawabata, N.E. Heldin, C.H. Heldin, and P. ten Dijke, Identification of Smad7, a TGFbeta-inducible antagonist of TGF-beta signalling. Nature, 1997. 389(6651): p. 631-5.

Page 179: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

169

18. Persson, U., H. Izumi, S. Souchelnytskyi, S. Itoh, S. Grimsby, U. Engstrom, C.H. Heldin, K. Funa, and P. ten Dijke, The L45 loop in type I receptors for TGF-beta family members is a critical determinant in specifying Smad isoform activation. FEBS Lett, 1998. 434(1-2): p. 83-7.

19. Lo, R.S., Y.G. Chen, Y. Shi, N.P. Pavletich, and J. Massague, The L3 loop: a structural motif determining specific interactions between SMAD proteins and TGF-beta receptors. Embo J, 1998. 17(4): p. 996-1005.

20. Chen, Y.G. and J. Massague, Smad1 recognition and activation by the ALK1 group of transforming growth factor-beta family receptors. J Biol Chem, 1999. 274(6): p. 3672-7.

21. Xiao, Z., N. Watson, C. Rodriguez, and H.F. Lodish, Nucleocytoplasmic shuttling of Smad1 conferred by its nuclear localization and nuclear export signals. J Biol Chem, 2001. 276(42): p. 39404-10.

22. Xu, L., Y.G. Chen, and J. Massague, The nuclear import function of Smad2 is masked by SARA and unmasked by TGFbeta-dependent phosphorylation. Nat Cell Biol, 2000. 2(8): p. 559-62.

23. Shi, Y., Y.F. Wang, L. Jayaraman, H. Yang, J. Massague, and N.P. Pavletich, Crystal structure of a Smad MH1 domain bound to DNA: insights on DNA binding in TGF-beta signaling. Cell, 1998. 94(5): p. 585-94.

24. von Bubnoff, A. and K.W. Cho, Intracellular BMP signaling regulation in vertebrates: pathway or network? Dev Biol, 2001. 239(1): p. 1-14.

25. Kretzschmar, M., J. Doody, and J. Massague, Opposing BMP and EGF signalling pathways converge on the TGF-beta family mediator Smad1. Nature, 1997. 389(6651): p. 618-22.

26. Itoh, S., J. Ericsson, J. Nishikawa, C.H. Heldin, and P. ten Dijke, The transcriptional co-activator P/CAF potentiates TGF-beta/Smad signaling. Nucleic Acids Res, 2000. 28(21): p. 4291-8.

27. Hata, A., R.S. Lo, D. Wotton, G. Lagna, and J. Massague, Mutations increasing autoinhibition inactivate tumour suppressors Smad2 and Smad4. Nature, 1997. 388(6637): p. 82-7.

28. Murakami, G., T. Watabe, K. Takaoka, K. Miyazono, and T. Imamura, Cooperative inhibition of bone morphogenetic protein signaling by Smurf1 and inhibitory Smads. Mol Biol Cell, 2003. 14(7): p. 2809-17.

29. Correia, J.J., B.M. Chacko, S.S. Lam, and K. Lin, Sedimentation studies reveal a direct role of phosphorylation in Smad3:Smad4 homo- and hetero-trimerization. Biochemistry, 2001. 40(5): p. 1473-82.

30. Shi, Y. and J. Massague, Mechanisms of TGF-beta signaling from cell membrane to the nucleus. Cell, 2003. 113(6): p. 685-700.

31. Balemans, W. and W. Van Hul, Extracellular regulation of BMP signaling in vertebrates: a cocktail of modulators. Dev Biol, 2002. 250(2): p. 231-50.

32. Zimmerman, L.B., J.M. De Jesus-Escobar, and R.M. Harland, The Spemann organizer signal noggin binds and inactivates bone morphogenetic protein 4. Cell, 1996. 86(4): p. 599-606.

Page 180: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

170

33. Grotewold, L., M. Plum, R. Dildrop, T. Peters, and U. Ruther, Bambi is coexpressed with Bmp-4 during mouse embryogenesis. Mech Dev, 2001. 100(2): p. 327-30.

34. Onichtchouk, D., Y.G. Chen, R. Dosch, V. Gawantka, H. Delius, J. Massague, and C. Niehrs, Silencing of TGF-beta signalling by the pseudoreceptor BAMBI. Nature, 1999. 401(6752): p. 480-5.

35. Ishida, W., T. Hamamoto, K. Kusanagi, K. Yagi, M. Kawabata, K. Takehara, T.K. Sampath, M. Kato, and K. Miyazono, Smad6 is a Smad1/5-induced smad inhibitor. Characterization of bone morphogenetic protein-responsive element in the mouse Smad6 promoter. J Biol Chem, 2000. 275(9): p. 6075-9.

36. Imamura, T., M. Takase, A. Nishihara, E. Oeda, J. Hanai, M. Kawabata, and K. Miyazono, Smad6 inhibits signalling by the TGF-beta superfamily. Nature, 1997. 389(6651): p. 622-6.

37. Inoue, H., T. Imamura, Y. Ishidou, M. Takase, Y. Udagawa, Y. Oka, K. Tsuneizumi, T. Tabata, K. Miyazono, and M. Kawabata, Interplay of signal mediators of decapentaplegic (Dpp): molecular characterization of mothers against dpp, Medea, and daughters against dpp. Mol Biol Cell, 1998. 9(8): p. 2145-56.

38. Souchelnytskyi, S., T. Nakayama, A. Nakao, A. Moren, C.H. Heldin, J.L. Christian, and P. ten Dijke, Physical and functional interaction of murine and Xenopus Smad7 with bone morphogenetic protein receptors and transforming growth factor-beta receptors. J Biol Chem, 1998. 273(39): p. 25364-70.

39. Hata, A., G. Lagna, J. Massague, and A. Hemmati-Brivanlou, Smad6 inhibits BMP/Smad1 signaling by specifically competing with the Smad4 tumor suppressor. Genes Dev, 1998. 12(2): p. 186-97.

40. Zhu, H., P. Kavsak, S. Abdollah, J.L. Wrana, and G.H. Thomsen, A SMAD ubiquitin ligase targets the BMP pathway and affects embryonic pattern formation. Nature, 1999. 400(6745): p. 687-93.

41. Rotin, D., WW (WWP) domains: from structure to function. Curr Top Microbiol Immunol, 1998. 228: p. 115-33.

42. Wang, W., F.V. Mariani, R.M. Harland, and K. Luo, Ski represses bone morphogenic protein signaling in Xenopus and mammalian cells. Proc Natl Acad Sci U S A, 2000. 97(26): p. 14394-9.

43. Nakashima, K., M. Yanagisawa, H. Arakawa, N. Kimura, T. Hisatsune, M. Kawabata, K. Miyazono, and T. Taga, Synergistic signaling in fetal brain by STAT3-Smad1 complex bridged by p300. Science, 1999. 284(5413): p. 479-82.

44. Binnerts, M.E., X. Wen, K. Cante-Barrett, J. Bright, H.T. Chen, V. Asundi, P. Sattari, T. Tang, B. Boyle, W. Funk, and F. Rupp, Human Crossveinless-2 is a novel inhibitor of bone morphogenetic proteins. Biochem Biophys Res Commun, 2004. 315(2): p. 272-80.

45. Thies, R.S., M. Bauduy, B.A. Ashton, L. Kurtzberg, J.M. Wozney, and V. Rosen, Recombinant human bone morphogenetic protein-2 induces osteoblastic differentiation in W-20-17 stromal cells. Endocrinology, 1992. 130(3): p. 1318-24.

Page 181: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

171

46. Lenox, L.E., J.M. Perry, and R.F. Paulson, BMP4 and Madh5 regulate the erythroid response to acute anemia. Blood, 2005. 105(7): p. 2741-8.

47. Pedersen, A.G. and H. Nielsen, Neural network prediction of translation initiation sites in eukaryotes: perspectives for EST and genome analysis. Proc Int Conf Intell Syst Mol Biol, 1997. 5: p. 226-33.

48. Pedersen, A.G.a.H.N. NetStart 1.0 Prediction Server. [cited; Available from: http://www.cbs.dtu.dk/services/NetStart/.

49. Liu, F., A. Hata, J.C. Baker, J. Doody, J. Carcamo, R.M. Harland, and J. Massague, A human Mad protein acting as a BMP-regulated transcriptional activator. Nature, 1996. 381(6583): p. 620-3.

50. Baker, J.C. and R.M. Harland, A novel mesoderm inducer, Madr2, functions in the activin signal transduction pathway. Genes Dev, 1996. 10(15): p. 1880-9.

51. de Jong, D.S., W.T. Steegenga, J.M. Hendriks, E.J. van Zoelen, W. Olijve, and K.J. Dechering, Regulation of Notch signaling genes during BMP2-induced differentiation of osteoblast precursor cells. Biochem Biophys Res Commun, 2004. 320(1): p. 100-7.

52. Fuchs, O., O. Simakova, P. Klener, J. Cmejlova, J. Zivny, J. Zavadil, and T. Stopka, Inhibition of Smad5 in human hematopoietic progenitors blocks erythroid differentiation induced by BMP4. Blood Cells Mol Dis, 2002. 28(2): p. 221-33.

53. Bruno, E., S.K. Horrigan, D. Van Den Berg, E. Rozler, P.R. Fitting, S.T. Moss, C. Westbrook, and R. Hoffman, The Smad5 gene is involved in the intracellular signaling pathways that mediate the inhibitory effects of transforming growth factor-beta on human hematopoiesis. Blood, 1998. 91(6): p. 1917-23.

54. Goumans, M.J., G. Valdimarsdottir, S. Itoh, F. Lebrin, J. Larsson, C. Mummery, S. Karlsson, and P. ten Dijke, Activin receptor-like kinase (ALK)1 is an antagonistic mediator of lateral TGFbeta/ALK5 signaling. Mol Cell, 2003. 12(4): p. 817-28.

55. Oh, S.P., T. Seki, K.A. Goss, T. Imamura, Y. Yi, P.K. Donahoe, L. Li, K. Miyazono, P. ten Dijke, S. Kim, and E. Li, Activin receptor-like kinase 1 modulates transforming growth factor-beta 1 signaling in the regulation of angiogenesis. Proc Natl Acad Sci U S A, 2000. 97(6): p. 2626-31.

56. Roelen, B.A., M.A. van Rooijen, and C.L. Mummery, Expression of ALK-1, a type 1 serine/threonine kinase receptor, coincides with sites of vasculogenesis and angiogenesis in early mouse development. Dev Dyn, 1997. 209(4): p. 418-30.

57. Matsuyama, S., M. Iwadate, M. Kondo, M. Saitoh, A. Hanyu, K. Shimizu, H. Aburatani, H.K. Mishima, T. Imamura, K. Miyazono, and K. Miyazawa, SB-431542 and Gleevec inhibit transforming growth factor-beta-induced proliferation of human osteosarcoma cells. Cancer Res, 2003. 63(22): p. 7791-8.

58. Ward, S.M., J.S. Desgrosellier, X. Zhuang, J.V. Barnett, and J.B. Galper, Transforming growth factor beta (TGFbeta) signaling via differential activation of activin receptor-like kinases 2 and 5 during cardiac development. Role in regulating parasympathetic responsiveness. J Biol Chem, 2002. 277(51): p. 50183-9.

Page 182: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

172

59. Zhang, H., H.O. Akman, E.L. Smith, J. Zhao, J.E. Murphy-Ullrich, and O.A. Batuman, Cellular response to hypoxia involves signaling via Smad proteins. Blood, 2003. 101(6): p. 2253-60.

60. Sanchez-Elsner, T., L.M. Botella, B. Velasco, A. Corbi, L. Attisano, and C. Bernabeu, Synergistic cooperation between hypoxia and transforming growth factor-beta pathways on human vascular endothelial growth factor gene expression. J Biol Chem, 2001. 276(42): p. 38527-35.

61. Falanga, V., S.W. Qian, D. Danielpour, M.H. Katz, A.B. Roberts, and M.B. Sporn, Hypoxia upregulates the synthesis of TGF-beta 1 by human dermal fibroblasts. J Invest Dermatol, 1991. 97(4): p. 634-7.

Page 183: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

A) 173

Score = 100 (t-mint)/(maxt-mint) (Pos)

-3-2-1+1+2+3+4+5+6

•t is the total of percentages for the 8 nucleotides occurring in the subsequence being scored

•mint and maxt are the minimum and maximum possible totals (the sum of the lowest and highest percentages in each of the eight positions)

B)

Figure 4-1. Calculation of 5’ splice site consensus sequence. (A) Table of the tabulated nucleotide frequencies (percent at which a nucleotide exists at a location within the splice region when compared to the total number of sequences evaluated: compiled from a total of 393 rodent sequences from GeneBank) for the 5’ splice site locations for rodents (Shapiro, M.B. and P. Senapathy, 1987 [3]). Consensus nucleotides are shown in the right-hand column in the green box; nucleotide position labeled in blue under position (Pos) column; orange boxes denote percents used for calculation in example shown in (B). (B) Example for the 5’ splice site calculation for the first exon-intron junction of murine Madh5. Capital letters denote the end of exon 1, while the lower case letters denote the first nucleotides of intron 1. (C) Map of the calculated 5’ splice site scores for murine Madh5.

1680 bps1680 bps1680 bps1680 bps

IAPIAPIAPIAP

= exon unaltered in f/f truncated transcript

= exon altered in f/f truncated transcript = intron

= site of 14bp insertion= exon unaltered in f/f truncated transcript

= exon altered in f/f truncated transcript = intron

= site of 14bp insertion

EXAMPLE: For the murine Madh5 sequence TGT_gtgagg

t = 11+14+7+100+100+37+73+82+19 = 443mint = 58maxt = 637Score = 100(443-58)/(637-58) = 66.5

C)66.5 99.5

92.686.2 80.7

96.9

insertion

Page 184: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

174

Figure 4-2. Cascade of BMP signaling and levels of modulation. (From Balemans, W and Van Hul, W. 2002 [31]). BMP dimers bind to serine/threonine kinase receptors type I and II. Upon ligand binding, type II receptorstransphosphorylate type I receptors. The latter phosphorylate members of the Smad family of transcription factors. These Smads are subsequently translocated to the nucleus, where they activate transcription of target genes. (1) I-Smads and Smurfs regulate intracellular signaling by preventing further Smad signaling and consequent activation of gene transcription, (2) Pseudoreceptor BAMBI modulates BMP signaling at the membrane site by binding to BMP type II receptors, and (3) Extracellular antagonists modulate binding of BMP dimersto the BMP type I and type II receptors.

Page 185: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

175

Ligands Activins TGF-βs BMPsGDFsMIS

Type II ActRII/IIB TβRII BMPRIIReceptors ActRII/IIB

MISRII

Type I ALK4 ALK5 ALK1 ALK3 Receptors ALK6

ALK2

Activin/TGF-βSmad pathway

BMPSmad pathway

*

Figure 4-3. Signaling specificity in the TGF-β superfamily. (From Moustakas et al., 2001[12]) Classification of mammalian Smad signaling cascade into activin-TGF-β (red) and BMP (blue) pathway. Representative examples of mammalian ligands, type II receptors, type I receptors, R-Smads, Co-Smads, and I-Smads are depicted in pathways linked by arrows or signs of inhibition. Bifurcation of the TGF- β pathway at the level of type I receptors towards both TGF- and BMP Smads is marked by an asterisk (*). Nomenclature of proteins: growth and differentiation factors (GDFs), Mullerian inhibiting substances (MIS), activin type II and type IIB receptor (ActRII/IIB), TGF-β type II receptor (TβRII), BMP type II receptor (BMPRII), MIS type II receptor (MISRII), activin receptor-like kinases 1-6 (ALK1-ALK6).

R-Smads Smad2 Smad1Smad3 Smad5

Smad8

Co-Smads Smad4

I-Smads Smad7 Smad6

Page 186: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

176A)

L3 loop

pS-X-pS

SXS

GS domainL45 loop

B)

Figure 4-4. Activation of the type I receptor kinase and recognition of R-Smad. (From Shi, Y. and J. Massague, 2003 [30] and Chen, Y.G. and J. Massague, 1999 [20]) (A) In the basal state, type I receptor (TβRI) remains unphosphorylated. This conformation can be recognized by the type II receptor (TβRII) for phosphorylation in the GS domain. After phosphorylation, the TβRI uses its GS domain and the L45 loop to interact with the basic pocket and L3 loop of an R-Smad, resulting in its phosphorylation in the C-terminal SXS motif. The pThr/pSer-X-pSer motif on the R-Smad and the type I receptor is shown as green spheres [30]. (B) α-helix 1 and L3 sequences of R-Smads. Structural elements (arrows, β-strands; round boxes, α-helices) correspond to the Smad4 MH2 domain. Subtype-specific residues are boxed. Numbering of the last residue in each sequence corresponds to the Smad species not in parentheses. Not shown in (A) is α-helix 1. This structure has also been shown to be important for R-Smad recognition and activation by the type I receptors ALK1/2. Through comparison to the crystal structure of Smad4, the subtype specific residues of α-helix 1 are exposed to solvent in the vicinity of the L3 loop [20].

Page 187: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

177

Generic Smad Structure

I-Smad

Smad4

R-Smad(Smad2)

Figure 4-5. Structure of the receptor activated R-Smads, common binding partner Smad4 and the inhibitory I-Smads. (From ten Dijke, P. and C.S. Hill, 2004 [9]) The MH1 (red) and MH2 (blue) domains are conserved among Smads. Two regions that are conserved among R-Smads but not other Smads are indicated by pale pink boxes. Non-conserved regions (including the linker) are shown in yellow. Smad2 contains two inserts in its MH1 domain (L1 and exon3) that are not found in other R-Smads. Motifs shown include: phosphorylated C-terminal SxS motif (pSXpS); (NES) nuclear export signal; (NLS) nuclear localization signal; (PY), the PPxYmotif that mediates binding to Smurf1 and Smurf2.

Page 188: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

178

ATG

mutant transcript#1

mutant transcript#2

*[ ]*****

wild type

ATGATG

mutant transcript#1

mutant transcript#2

**[ ]*****[ ]*****

wild type

ATG

mutant transcript#1

mutant transcript#2

*[ ]*****

wild type

ATGATG

mutant transcript#1

mutant transcript#2

**[ ]*****[ ]*****

wild type

Figure 4-6. Map of full length Madh5 (Smad5) as well as the truncated transcripts found in f/f mice. Colored boxes represent exons. Dotted line denotes skipped exon(s). * indicates location of in-frame ATGs with (translation initiation sites predicted by NetStart with scores > 0.5; http://www.cbs.dtu.dk/services/NetStart/) [47] downstream of the known Madh5 translational start site (ATG). Actual start codon located in exon 2 has a score of 0.691; in frame ATG of mutant transcript #1 score of 0.585; in frame ATGs of mutant transcript #2 are 0.563, 0.588, 0.524, 0.588 and 0.608 respectively.

Page 189: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

Figure 4-8. Transfected COS7 cells with constructs containing truncated transcripts. 48 hr transfection of COS7 cells with truncated transcripts, the empty MSCV-neo vector, and full length Smad5 transcript. Lysates were first immunoprecipitated with Seize-columns (Peirce) binding the N-term Smad1/5/8 antibody N-18 (Santa Cruz), which helped to eliminate non-specific binding and which would explain why no appreciable amounts of full length Smad5 are seen in the control lanes. Subsequently, the lysates were immunoprecipitated with Seize column (Peirce) bound C-term Smad5 antibody (D-20 from Santa Cruz), which would bind the predicted truncated proteins from the f/f truncated transcripts. The lysates were run on a 4-20% gradient gel, transferred, and the blot was probed with D-20 Smad5 (1:1000) and an anti-goat secondary (Santa Cruz) (1:8000). The predicted sizes for the transcripts are: lel-118 (transcript #2-F) 26.8 kDa; lel-139 (transcript #1-F) 33.1 kDa. (MSCV) is the empty MSCV-neo construct lysates; (1a-1c) elutants from the immunoprecipitation column binding protein from lysates of the truncated transcript #1 in MSCV-neo in Forward orientation; (2a-2c) elutants from the immunoprecipitation column binding protein from lysates of the truncated transcript #2 in MSCV-neo in the Forward orientation; (FL) is the full-length Smad5 in MSCV-neo (clone Neo22) lysates; (*) denotes a protein running at the predicted size for transcript #1.

83.6

39.631.4

17.4

kDa1a 1b 1c 2a 2b 2cMSCV FL

Figure 4-7. RT-PCR from COS7 cells transfected with retroviral constructs containing full length and truncated Smad5 constructs. 5µg total RNA from transfected COS7 cells (48 hr) was reverse transcribed according to SuperScriptII(Invitrogen) protocol using 250 ng random hexamers. cDNA was amplified using Smad5-A-5F and Smad5-A-3R primers. (FL) Full length Smad5 (Neo22 construct); (2) truncated transcript #2 in forward orientation (lel-118); (1) truncated transcript #1 in forward orientation (lel-139).

1000bp

1500bp

FL 2 1

2000bp

179

*

Page 190: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

180

Receptor interaction and phosphorylationHomo- and hetero-oligomerization

Nuclear import/exportBinding to DNA cofactors

Binding to co-activators and co-repressors

DNA-bindingNuclear localization

Binding to DNA cofactors

1 133

PPXY (PY) motif recognized by WW domains of Smurf1 [41, 40]

Corresponding Smad5 Erk Kinase phosphorylation sites located at amino acids 188, 196, 205, 213 [25]

Figure 4-9. Predicted in-frame amino acid structure of truncated transcripts compared to wild-type Smad5. Reference for domains Entrez Gene ID #17129. (http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?CMD=search&DB=gene). Full length Smad5 codes for a 465 amino acid (aa) protein of 52 kDa. Truncated transcript #1 codes for a predicted 33.3 kDa protein combining aa’s 167-217 to 265-465, with 5 aa’s inserted between 418-419 according to those constructs used in the transduced W-20-17 cell assays. Truncated transcript #2’s longest open reading frame is predicted to be 26.8 kDa and consists of aa’s 227-465.

SXS motif of R-Smads which is phosphorylated by type I receptors to activate the molecule [30]

Alpha helix-1: conserved structure shown in Smad1 to be important in addition to L3 loop for the recognition of type I receptors [20]

MH2

265

MH1 MH2

223 465320-328 416-434

WT

MH2

PVHFQ167 217 265

227

Transcript #1

Transcript #2

L3 loop: structural motif determining specific interactions between Smad proteins an type I receptors, as well as for homo-trimerization and hetero-oligomerization (aa’s425,428) [19,30]

Page 191: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

181

HPRT (338bps)

Car3 (291bps)

200bps

400bps

MSCV Smd5 #1R #1F#2R#2F

Figure 4-10. Gene expression of W-20-17 cell lines containing the f/f truncated transcripts. RT-PCR for HPRT and Car3 from W-20-17 cells lines stimulated with BMP4 (15 ng/mL) for 24 hrs. Lanes labeled as follows: empty MSCV-neo (MSCV); full length Smad5 clone Neo22 (Smd5); transcript #1-reverse orientation (#1R); transcript #2-forward orientation (#2F); transcript #2-reverse orientation (#2F); transcript #1-forward orientation (#1F).

Page 192: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

182

BMP4 Hypoxia SCF

BMPR1α

Smad5 PERK

Kit receptor

MAPK

Rsk

ATF4

•SCL•Lmo2

•GATA1•Other factors

HIF1

eIF2α

ER stress

?

?

?

BMP4 Hypoxia SCF

BMPR1α

Smad5 PERK

Kit receptor

MAPK

Rsk

ATF4

•SCL•Lmo2

•GATA1•Other factors

HIF1

eIF2α

ER stress

?

?

?

Figure 4-11. Current working model for the network regulating expansive erythropoiesis (R.F. Paulson lab). The model diagrams the possible connection of the BMP, hypoxia and stem cell factor (SCF) signaling pathways converging to regulate erythroid development and stress erythropoiesis, specifically in regulating the expansion and differentiation of stress erythroid progenitors in the spleen. (Slide provided by Omid F. Harandi).

Page 193: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

183

Appendix A

CLONING AND CHARACTERIZATION of the flexed-tail (f) LOCUS

The pursuit of a candidate gene for flexed-tail. The flexed-tail locus was located on

mouse chromosome 13 [1]. A panel of 408 F2 progeny from a F1(C57BL/6-f/f X

BALB/c) intercross were scored at birth for anemia by hematocrit and for the presence of

siderocytes by staining blood smears for iron deposits. A high resolution genetic linkage

map of the f locus was generated and initially localized the gene 0.6 cM distal to the

microsatellite marker D13MIT13. Further analysis by positional cloning of markers using

a TI31 Radiation Hybrid panel showed that the f locus co-segregated with the marker

D13Mit208. From here YAC and BAC libraries were screened for clones that contained

the flanking markers of flexed-tail (D13Mit13 and D13/Mit250) determined from the

genetic linkage map. Three YAC clones E10 (sized at approx. 110 Kb), E14 (sized at

115 Kb) and B24 (sized at 125 Kb) were initially determined by a PCR screen to contain

the flanking markers. At the time, no extensive data base existed that contained ordered

sequence of the mouse genome. With the development of the murine genomic resources,

it has become clear that the initial PCR screens were flawed. Although correct sized

bands were generated by the screens, sequence comparison from the ends of the BACs as

well as the generation of a comprehensive map of the mouse genome confirms that they

could not in fact contain both flanking markers. These markers lie too far apart to be

contained on the BACs of the determined sizes (Figure A-1).

Page 194: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

184

Proceeding with the information available at the time, the strategy to clone the

flexed-tail locus consisted of using the BAC clones and various methods to generate

sequence information between the flanking markers, or to identify transcriptional units

contained within this region (Figure A-2). The GPS-1 Genome Priming System

(Clonetech) is a Tn7 transposon insertion system which will randomly insert a primer

binding site throughout the BAC clone. Using a library of these insertions, the sequence

of the BAC can be compiled and candidate genes revealed. An alternative system was

the Exon Trapping System (Invitrogen). This technique utilized a specialized splicing

vector which contained cloned genomic segments (from the BAC clones). When this

vector is transfected into COS7 cells, splicing will occur and exons isolated, cloned and

sequenced.

Exon information generated from both methods was further utilized to try and

generate candidate genes for flexed-tail. Exons were used as probes to screen a cDNA

library generated from E16 fetal liver to try and determine the full length transcript which

had contained the particular exon. 5’ and 3’ RACE (Rapid Amplification of cDNA ends)

helped generate larger transcripts containing a particular fragment, again to try and

isolate a candidate gene for flexed-tail. No solid candidate genes were obtained by these

methods.

The Human and Mouse Genome Projects, as well as the data bases set up to

support these endeavors proved to be most useful recourse for finding candidate genes for

the flexed-tail locus. The comparison of the Human and Mouse Genomes on the NCBI

website (http://www.ncbi.nlm.nih.gov/Homology/) showed that the D13Mit13, which

contained in the IL-9 gene, was in close proximity to Madh5 (Smad5) in both the human

Page 195: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

185

and mouse genomes. Smad5 was determined by both PCR and Southern blot to be

contained on our BACs (E10 and E14), as well as being expressed in the adult spleen

following a phenylhydrazine induced acute anemia (D3 post PHZ using a 2-injection

regimen) (Figure A-3). Smad5 is a downstream signaling molecule for the TGFβ

superfamily of morphogens, acting primarily down stream of BMPs. BMP4 is

specifically expressed along the ventral wall of the dorsal aorta (part of the AGM region),

at the onset of definitive hematopoietic cells in this region [2]. Early H&E sections of f/f

and f/+ littermates suggest defects in this region at this time (Figure A-4). The embryonic

defects of the flexed-tail mutants including a more detailed analysis of the fetal liver and

AGM are the pursuit of Prashanth Porayette, a fellow graduate student in the Paulson lab.

Smad5 is localized to a region in the human genome with known homology to the

region of mouse Chr 13 where flexed-tail has been isolated, and has been shown through

PCR and Southern analysis to be contained on our BAC clones. The BMP signaling

pathway is important for erythroid development, and Smad5 has been shown to be

expressed in the adult spleen following an acute anemia. For these reasons, Smad5 was

more extensively pursued as the candidate for the flexed-tail locus. A summary of the

results is contained in our paper published in Blood (2005) [3].

Northern blot analysis. It was discovered that flexed-tail mutants were preferentially

producing a truncated Madh5 (Smad5) mRNA with significantly less full length

transcript detected in PCR. To get a more quantitative analysis of these differences

Northern blots from D=4 PHZ treated adult spleens were performed. Although initially

Page 196: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

186

showing a roughly 2-fold decrease in the amount of full length Smad5 in f/f compared to

f/+ controls, these results could not be consistently repeated (Figure A-5). RNase

Protection assays fared no better at showing a definitive comparison of the amount of full

length Smad5 transcript in flexed-tail mice.

Western blot analysis. With the indication that there were differences in the amount of

correctly spliced transcripts from f/f mice as compared to +/+ or f/+, it follows that there

may be altered Smad5 protein levels in the mutants. Western blotting was attempted

using various protein sources (whole cell lysates: spleen; nuclear extracts; PHZ treated

spleen; retrovirally transduced spleen COS7, 293T, W-20-17 cells), using antibodies

against Smad5 (N-term (1/5/8); C-term (5); (phospho)-1/5/8; (phospho)-Smad5), and

using various procedures (whole cell lysates or immunoprecipitation: anti- Smad5;

Smad4; Smad1/5/8; Pierce Column to eliminate heavy chain using (1/5/8 or 5)). The

most promising results were generated using 24 hr post PHZ adult spleen lysates

immunoprecipitated with (anti-Smad1/5/8) and probed with anti-Smad5 (D20) (Figure A-

6). This result was not consistently seen. Difficulty in duplicating the results could be

attributed to the inherent variation in Smad5 amongst f/f mice, low levels of Smad5

protein in primary cells, stability of the protein, or specificity of the antibodies.

Short-term rescue of flexed-tail—retroviral transduction of f/f spleen cells. Spleen

erythroid progenitors from f/f mutants do not respond to BMP4 in erythroid colony

assays. To rescue these progenitors, f/f progenitors were retrovirally transduced with full

length Smad5 cDNA. At the time it was hypothesized that providing a wild-type copy

Page 197: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

187

would produce enough full length transcript to render them capable of a BMP4 response.

Retroviral rescues were attempted using various conditions for the transduction (varied

time of transduction, combination of various factors including SCF, IL3, Epo, BMP4) as

well as various conditions in the plating media (no SCF, no IL3, but +/- Epo, +/- BMP4),

and even sorted Lin- vs. whole spleen cell suspensions. The conditions that showed the

best “rescue” of f/f as determined by the number of hemoglobinized colonies used Epo,

IL3, SCF in the transduction, and incubation with virus for 5 hours (Figure A-7A). The

morphology of the colonies was not consistent with pure BFU-E and showed a large

proportion of mixed morphology colonies. There was also no response to BMP4. The

condition that led to the expected number and morphology of colonies (BMP4, SCF and

IL3 in the transduction, incubated for 5 hours), showed little difference between any of

phenotypes with none showing a response to BMP4 (Figure A-7B). We know now that

the mere exposure of the spleen cells to BMP4 is enough to drive their differentiation to

an Epo responsive cell, which no longer retains BMP4 its responsiveness. Including

BMP4 during the transduction probably just differentiated the cells beyond what would

respond in the in vitro culture assays. Altering the cytokine cocktails during the

transduction was having unanticipated effects on our spleen progenitor cells, and thus no

conclusive results could be obtained using these methods.

Long-term rescue--Smad5 transgenic f/f mice. Before the observation of the

neomorphic properties of the flexed-tail truncated transcripts, it was hypothesized that a

wild-type copy of the Madh5 gene would rescue the recessive mutation. Using our E10

BAC we generated transgenic mice containing what we thought to be the full length

Page 198: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

188

Smad5. Only 2 founders (out of 9 litters) were generated, with only #1439 a productive

breeder. These were crossed to our f/f mutants. F1s that were BAC+ were crossed again

to f/f. f/f BAC+ and f/f BAC- were compared for their hematocrit recovery following a

PHZ induced acute anemia. Initial results were promising, but by increasing the number

of mice for more statistically relevant results, we found no statistical difference in the

hematocrit values between f/f BAC+ and f/f BAC- over the recovery period (Figure A-8).

This suggests the flexed-tail phenotype was not rescued with our BAC. This could be

due to the founder not expressing the BAC, or to the fact the entire Smad5 was not

contained on E10 (3’-most end not on BAC: see 10/29/01), and thus Smad5 from the

BAC was not correctly regulated. Even if given a wild-type version of Smad5, the

flexed-tail mice are still producing the aberrant versions which have been shown to

disrupt BMP signaling.

Cloning of the flexed-tail mutation. The splicing defect in f/f mutants was not due to

any mutations in the coding region. For this reason, we tried alternative methods to

sequence the genomic sequence of Smad5 between flexed-tail and wild-type mice. One

technique, genome walking, enabled me to compile roughly 10 Kb of genomic sequence

from f/f (not including that compiled from our C57 controls, or off the BAC clones). A

map of this data is included in (Figure A-9). The biggest problem with this technique

was specificity. Artifacts and repeat regions, including a defective retroviral element

(intercisternal A particle, (IAP)) between exon 1 and exon 2, made analyzing the

sequence messy and complex at best. Since the sequence generated is genome wide,

pseudogenes also complicate the analysis.

Page 199: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

189

To circumvent these problems, the entire Smad5 gene (from exon 1 to 7), or

roughly 39 Kb was directly amplified in shorter fragments (3-6 kb stretches), cloned and

sequenced (Laurie Lenox, John Perry, Prashanth Porayette and Michelle Yon). The

shorter fragments with their respective names are shown in (Figure A-10). The biggest

difficulties with direct sequencing of the Smad5 gene was getting fragments amplified for

cloning. A particularly difficult stretch was the 5’ most end of the gene containing exon

1. This region is highly GC rich (approx. 70%), with tight secondary structure, even at

the temperatures optimal for polymerase extension (72 degrees). A comparison of the

secondary structure using mFold (http://mfold.burnet.edu.au) between equal sized

fragments from the easily amplified Bombay fragment (43% GC rich), and the nearly

impossible 5’ end of Alaska is shown in (Figure A-11). The 5’ end of Alaska was

eventually sequence by adjusting PCR conditions, primer design (short 200-500bp

stretches), and through the use of EpiCenter Fail Safe polymerase, allowing efficient

extension at 88 degrees, rather than the standard 72 degrees. Sequences were compared

between flexed-tail and C57 wild-type clones, as well as comparisons to the Ensembl

genome database (Ensembl Genome Browser: http://www.ensembl.org/). Verification

was established between at least two clones, from two independent mice. The only

potential mutation that held up to this scrutiny was a variation in a poly T stretch located

in the Greece fragment (intron 4). Sequences originally cloned from flexed-tail showed

14Ts whereas the stretch from the C57 (as well as the database) contained 16Ts.

Mutations in poly-pyrimidine stretches have been shown to aberrantly effect splicing [4,

5]. The most notable is in the cystic fibrosis transmembrane conductance regulator

(CFTR) gene. In the CFTR gene there are variants in poly-T tract in intron 8 (5T, 7T,

Page 200: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

190

9T), with each allele leading to the generation of correctly and aberrantly spliced CFTR

transcripts lacking exon 9. There is an inverse correlation between the level of aberrantly

spliced transcript and the length of the poly(T) tract. The more correctly spliced

transcript, the less severe the disease [6, 7]. The PolyT stretch of Smad5 was much

longer than that seen in CFTR. A number of techniques were used to verify the number

of T’s from f/f and C57 mice.

Determining the number of T’s in the Greece fragment polyT stretch. Due to the

length of the polyT stretch in question (14 vs. 16 Ts), all techniques requiring an

amplification step have proven problematic mostly likely due to polymerase slipping.

Directly cloning and sequencing the PolyT stretch using high fidelity polymerases would

suggest a distribution of Ts between flexed-tail and wild-type mice, centering around 15-

16Ts (Figure A-12). Alternative techniques have been attempted including radioactive

PCR (to decrease the number of cycles required and thus decreasing artifacts), primer

extension, Oligo-nucleotide ligation assays (3 different primer sets), screening of a

genomic library, as well as the Transgenomic Mutation Detection Kit. In all cases, it was

a balancing act between signal strength and specificity. Since none of these techniques

could yield definitive results, we tried to show functionally the consequences of

differences in the PolyT region. Currently Shailaja Hedge in our lab is performing

recombination cloning. This is a cloning technique that does not rely on PCR

amplification, but rather on a recombination competent E. coli strain. Utilizing this

technique will hopefully reduce the artifacts associated with amplification and determine

the true number of Ts of the polyT stretch between f/f and +/+ mice.

Page 201: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

191

Functional consequences of the PolyT tract-- In vitro splicing assay. We utilized the

Exon Trapping vector (pSPL3), and cloned different genomic fragments (Greece,

Florence-Greece, fragment of Florence-Greece) between flexed-tail or wild-type mice,

verified the number of Ts in the polyT stretch, then used the constructs in a splicing assay

to determine whether there were specific differences in the spliced products dependent on

the size of the polyT stretch. Primer combinations to amplify the spliced products were

positioned within the splicing vector itself, or within exons to amplify more specific

products (Figure A-13). Predicted splice products (Exon3:4:5) as well as those seen in

vivo in the flexed-tail (aberrant 3:4:5) were obtained. Although promising, overall

analysis of the products using these techniques was difficult. There were many spliced

products produced, from both flexed-tail and wild-type, often varying by only one or two

bps at the splice junctions, including numerous exon:vector or vector:vector

combinations. The combinations were so infinite and inconsistent that no conclusive

results as to how the variations in the polyT stretch were specifically affecting splicing

could be determined.

Page 202: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

192

References

1. Lyon, M., S. Rastan, and S. Brown, Genetic Variants and Strains of the Laboratory Mouse. 3rd ed. 1996, Oxford: Oxford University Press.

2. Marshall, C., C. Kinnon, and A. Thrasher, Polarized expression of bone morphogenetic protein-4 in the human aorta-gonad-mesonephros region. Blood, 2000. 96: p. 1591-1593.

3. Lenox, L.E., J.M. Perry, and R.F. Paulson, BMP4 and Madh5 regulate the erythroid response to acute anemia. Blood, 2005. 105(7): p. 2741-8.

4. Giannini, G., E. Ristori, F. Cerignoli, C. Rinaldi, M. Zani, A. Viel, L. Ottini, M. Crescenzi, S. Martinotti, M. Bignami, L. Frati, I. Screpanti, and A. Gulino, Human MRE11 is inactivated in mismatch repair-deficient cancers. EMBO Rep, 2002. 3(3): p. 248-54.

5. Nissim-Rafinia, M., O. Chiba-Falek, G. Sharon, A. Boss, and B. Kerem, Cellular and viral splicing factors can modify the splicing pattern of CFTR transcripts carrying splicing mutations. Hum Mol Genet, 2000. 9(12): p. 1771-8.

6. Chu, C.S., B.C. Trapnell, S. Curristin, G.R. Cutting, and R.G. Crystal, Genetic basis of variable exon 9 skipping in cystic fibrosis transmembrane conductance regulator mRNA. Nat Genet, 1993. 3(2): p. 151-6.

7. Pagani, F., C. Stuani, M. Tzetis, E. Kanavakis, A. Efthymiadou, S. Doudounakis, T. Casals, and F.E. Baralle, New type of disease causing mutations: the example of the composite exonic regulatory elements of splicing in CFTR exon 12. Hum Mol Genet, 2003. 12(10): p. 1111-20.

Page 203: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

193

D13Mit13D13Mit250Smad5

D13Mit208D13Mit13D13Mit250Smad5

D13Mit208

Figure A-1. Physical Map of Mus musculus Chr13 around region of Madh5(Smad5). http://www.ensembl.org/Mus_musculus/

Page 204: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

194

GPSTM-1 Exon Trap

3 BAC’s

cDNA libraryfrom E16 fetal liver

Probe library with “trapped”

exons

Poly A+ RNA (from mouse fetal livers E16)

Marathon cDNA Adaptor

AP1 primerGene

specificF R AP1 primer

ds cDNA

5’-3’-x-

-x-3’

-x-3’

-5’

-5’

5’ and 3’ RACE

GPSTM-1 Exon Trap

3 BAC’s

cDNA libraryfrom E16 fetal liver

Probe library with “trapped”

exons

Poly A+ RNA (from mouse fetal livers E16)

Marathon cDNA Adaptor

AP1 primerGene

specificF R AP1 primer

ds cDNA

5’-3’-x-

-x-3’

-x-3’

-5’

-5’

5’ and 3’ RACE

Generate Sequence information between Mit13 and Mit250 and/oridentify transcription units

Generate Sequence information between Mit13 and Mit250 and/oridentify transcription units

Figure A-2. Strategy to find the flexed-tail (f) mutation. Sequence information from 3 BAC clones containing the flanking markers of flexed-tail (f) was generated using the GPS-1 (Clontech) Genome priming system while the Exon trapping system (Invitrogen) pulled out exons contained within the BAC clones. 5’ and 3’ Rapid Amplification of cDNA Ends (RACE) was used to determine the full length transcripts pulled from the BAC. Exons were also used to probed a cDNA library generated from E16 fetal liver RNA to determine genes that are expressed during the stage in development flexed-tail mice are anemic which are also contained on the BAC clones.

Page 205: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

195

Figure A-3. Analysis for Madh5 (Smad5) on BAC clones. (A) PCR showing Madh5 on the E10 BAC as well as RT-PCR showing expression in the spleen of an adult f/f on day 3 following an acute anemia. (B) The BAC clones E10 and E14, were digested with a restriction enzyme to produce 15 Kb fragments. This DNA was used in a Southern blot to determine if Madh5 was contained within the BAC clones which contained the flanking markers of flexed-tail. Shown are the blots probed using a Madh5 probe (Smad5), as well as probes to the ends of the vector sequence of the BAC (T7 and SP6); (M) denotes lane for size markers.

T7 SP6Mouse Genomic DNA

~15kb

pBeloBAC

T7

E10

E14

Smad5

E10

E14M

25Kb

SP6E

10 E14

SP6E

10 E14

SP6E

10 E14

T7 SP6Mouse Genomic DNA

~15kb

pBeloBAC

T7

E10

E14

T7

E10

E14

Smad5

E10

E14M

25Kb

Smad5

E10

E14M

25Kb

SP6E

10 E14

SP6E

10 E14

SP6E

10 E14

SP6E

10 E14

80bps100bps

E10

BA

CD

3 PH

Z SP

L

PCR

80bps80bps100bps100bps

E10

BA

CD

3 PH

Z SP

L

PCR

A)

B)

Page 206: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

196

DORSAL AORTA of E10.5 f /+ control mouse

DORSAL AORTA of E10.5 f /f mutant mouse

400X 400X

VIEW

Figure A-4. H&E stained sections of dorsal aorta from E10.5 f/f and f/+ littermates. (Top panel) Cross sections of embryos shown. Black arrow denotes position within the embryos. (Bottom panels) Blue arrows point to the ventral wall of the dorsal aorta. Hematopoietic clusters can be seen budding from the ventral wall in the f/+ mice, but these clusters are not as prominent in the f/f littermates at this stage.

Page 207: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

197

f / + control f / f mutant

Smad5

Probe

HPRT(loading control)

Figure A-5. Northern blot for Madh5 (Smad5) mRNA levels in spleen of f/+ and f/f adult mice following an acute anemia. RNA from spleen of f/f and f/+ adult mice D=4 post phenylhydrazine treatment (injection on D=0, and D=1) probed for HPRT as a loading control, and a probe specific to the coding region of Madh5.

Page 208: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

198

f / f f /+ +/+f / f f /+ +/+83 kDa

39 kDa

Figure A-6. Western blot for Madh5 (Smad5) protein levels in f/f, f/+ and +/+ mice. Whole cell extracts from 24 hr post phenylhydrazine treated adult spleens from f/f, f/+ and +/+ mice were immunoprecipitated with a Smad1/5/8 antibody (N-18 from Santa Cruz) and probed with a Smad5 specific antibody (D-20 from Santa Cruz). The molecular weight of Smad5 is 52kDa. The arrow denotes location of Smad5.

Page 209: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

A) 199

Figure A-7. Short-term rescue of f/f mutant spleen cells through retroviral transduction of full length Madh5 (Smad5) cDNA. f/f, f/+ and +/+ spleen cells were retrovirally transduced with a full length Smad5 transcript (Neo22) and then 1X105 cells plated in a BFU-E colony assays. Colonies were stained with acid benzidine for presence of hemoglobin. (A) Spleen cells incubated with viral supernatant containing Epo (3 U/mL), IL-3(10 ng/mL) and SCF (5 ng/mL) for 5 hrs prior to plating in methylcellulose media containing Epo (3 U/ml) or Epo + BMP4 (15 ng/mL). Shown is number of hemoglobin positive cells per 2X106 spleen cells plated. (B) Spleen cells incubated in viral supernatant containing Epo (3 U/mL), IL-3 (10 ng/mL) and BMP4 (15 ng/mL) for 5 hrs prior to plating in methylcellulose media containing Epo (3 U/mL) or Epo + BMP4 (15 ng/mL). Shown is the number of BFU-E per 2X106 cells plated.

0

20

40

60

80

100

120

140

f/f S d5f/f f/f+BMP4

C57 C57+BMP4

f/f+Smad5

f/f+SMad5+BMP4

= BFU-E = Mixed-Burst = Mixed-clump

0

50

100

150

200

250

300

+BMP4 +BMP4 +BMP4(-) (-) (-)

= f/f = f/f + Smad5= C57

0

20

40

60

80

100

120

140

f/f S d5f/f f/f+BMP4

C57 C57+BMP4

f/f+Smad5

f/f+SMad5+BMP4

f/f f/f+BMP4

C57 C57+BMP4

f/f+Smad5

f/f+SMad5+BMP4

= BFU-E = Mixed-Burst = Mixed-clump

0

50

100

150

200

250

300

+BMP4 +BMP4 +BMP4(-) (-) (-)

= f/f = f/f + Smad5= C57B)

Num

ber o

f BFU

-E p

er 2

X10

6

cells

pla

ted

Num

ber o

f hem

oglo

bini

zed

colo

nies

pe

r 2X

106

cells

pla

ted

Page 210: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

200

A)

E10

BA

C

(+) c

ontr

ol(-

) con

tro l

1434

1435

1436

1437

1438

*143

914

40

1441

1442

1443

1444

1445

1446

1447

1448

1449

1450

E10

BA

C

(+) c

ontr

ol(-

) con

tro l

1434

1435

1436

1437

1438

*143

914

40

1441

1442

1443

1444

1445

1446

1447

1448

1449

1450

E10

BA

C

(+) c

ontr

ol(-

) con

tro l

1434

1435

1436

1437

1438

*143

914

40

1441

1442

1443

1444

1445

1446

1447

1448

1449

1450

% H

emat

ocrit

Days post PHZ

% H

emat

ocrit

Days post PHZ

f/f BAC(+) n=8f/f BAC(-) n=8f/+ BAC(+) n=4f/+ BAC(-) n=4

f/f BAC(+) n=8f/f BAC(-) n=8f/+ BAC(+) n=4f/+ BAC(-) n=4

f/f BAC(+) n=8f/f BAC(-) n=8f/+ BAC(+) n=4f/+ BAC(-) n=4

Perc

ent h

emat

ocri

t

B)

Figure A-8. Long-term rescue of flexed-tail mutants: Madh5 transgenic f/fmice. (A) PCR for the chloramphenicol resistance gene carried on the E10 BAC used for transgenics. Founder #1439 shown in red. (B) The hematocrits of (C57BL/6J-f/f x B6D2F1 BAC Transgene [#1439])F1 x C57Bl/6J-f/f littermates were followed over the recovery from a phenylhydrazine induced acute anemia.

Page 211: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

201

Figu

re A

-9.

Gen

ome

Wal

king

res

ults

of M

adh5

geno

mic

reg

ion

from

f/f,

+/+

and

E10

BA

C.

Full

leng

th M

adh5

show

n at

top.

Sol

id c

olor

ed b

oxes

repr

esen

t exo

ns, w

hile

dot

ted

boxe

srep

rese

nt sk

ippe

d ex

ons i

n th

e f/f

trunc

ated

tran

scrip

t.

Bla

ck a

rrow

s den

ote

loca

tions

of g

enom

e w

alki

ng p

rimer

s, w

hile

whi

te a

rrow

s rep

rese

nt se

quen

cing

prim

ers.

Inte

rcis

tern

alA

par

ticle

foun

d in

intro

n 1

is re

pres

ente

d by

the

gold

hex

agon

.

150

1954

bps

AP2

-C52

7-R5

R3-

4R

9-1

0

RC

-AP2

-CIA

PA

P2-D

3500

bps

IAP

(300

0+bp

s)

AP2

-C52

7-R5

R3-

4

RC

-AP2

-C(1

800b

ps to

IAP)

R 9

-10

AP2

-DR

11-

12

260

Pres

ent i

n B

AC

and

f/f

R3-

4

260

Pres

ent i

n B

AC

and

f/f

R3-

4

260

Pres

ent i

n B

AC

and

f/f

R3-

4

333b

ps

F 3-

4

(OR

)?30

25 b

ps13

4bps

AP2

-AA

P2-B

IntR

3In

tR4

R1-

2

1036

bps

(wt)

F 5-

6F

5-6

1110

bps

( f )

F 7-

8F1

-2; F

5-6

+/+

f/f

+/+ f/f

Gen

ome

Wal

king

us

ing

E10

BA

C

R1-

2

F1-2

R 5

-6R

3-4

R1-

2

F 5-

6

R7-

8

F 7-

8F

3-4

Gen

ome

Wal

king

Res

ults

F1

-2

R 5

-6R

3-4

R1-

2

F 5-

6

R7-

8

F 7-

8F

3-4

Gen

ome

Wal

king

Res

ults

F 7-

8 500b

ps (250

0bps

)

150

1954

bps

AP2

-C52

7-R5

R3-

4R

9-1

0

RC

-AP2

-CIA

PA

P2-D

3500

bps

IAP

(300

0+bp

s)

AP2

-C52

7-R5

R3-

4

RC

-AP2

-C(1

800b

ps to

IAP)

R 9

-10

AP2

-DR

11-

12

260

Pres

ent i

n B

AC

and

f/f

R3-

4

260

Pres

ent i

n B

AC

and

f/f

R3-

4

260

Pres

ent i

n B

AC

and

f/f

R3-

4

333b

ps

F 3-

4

(OR

)?30

25 b

ps13

4bps

AP2

-AA

P2-B

IntR

3In

tR4

R1-

2

1036

bps

(wt)

F 5-

6F

5-6

1110

bps

( f )

F 7-

8F1

-2; F

5-6

+/+

f/f

+/+ f/f

Gen

ome

Wal

king

us

ing

E10

BA

C

R1-

2

F1-2

R 5

-6R

3-4

R1-

2

F 5-

6

R7-

8

F 7-

8F

3-4

Gen

ome

Wal

king

Res

ults

F1

-2

R 5

-6R

3-4

R1-

2

F 5-

6

R7-

8

F 7-

8F

3-4

Gen

ome

Wal

king

Res

ults

F 7-

8 500b

ps (250

0bps

)

Page 212: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

202

= exon unaltered in f/f truncated transcript

= exon altered in f/f truncated transcript = intron

= site of 14bp insertion

AlaskaBombay

ColoradoDublin

Egypt

Florence

Greece

Ireland

Hawaii

1680 bps

IAP

(3.57Kb)(3.97Kb)

(5.2Kb)(3.97Kb)

(3.36Kb)

(4.15Kb)

(4.63Kb)

(5.7Kb)

(3.16Kb)

(4.27Kb)

= exon unaltered in f/f truncated transcript

= exon altered in f/f truncated transcript = intron

= site of 14bp insertion

AlaskaBombay

ColoradoDublin

Egypt

Florence

Greece

Ireland

Hawaii

1680 bps

IAP

(3.57Kb)(3.97Kb)

(5.2Kb)(3.97Kb)

(3.36Kb)

(4.15Kb)

(4.63Kb)

(5.7Kb)

(3.16Kb)

(4.27Kb)

AlaskaBombay

ColoradoDublin

Egypt

Florence

Greece

Ireland

Hawaii

1680 bps

IAP

(3.57Kb)(3.97Kb)

(5.2Kb)(3.97Kb)

(3.36Kb)

(4.15Kb)

(4.63Kb)

(5.7Kb)

(3.16Kb)

(4.27Kb)

AlaskaBombay

ColoradoDublin

Egypt

Florence

Greece

Ireland

Hawaii

1680 bps

IAP

(3.57Kb)(3.97Kb)

(5.2Kb)(3.97Kb)

(3.36Kb)

(4.15Kb)

(4.63Kb)

(5.7Kb)

(3.16Kb)

(4.27Kb)

insertion

Figure A-10. Map of Madh5 subdivided for direct amplification, cloning and sequencing. Chr13: 56194401-56233706 Ensembl database (build 24) http://www.ensembl.org/Mus_musculus/. (10/2004 build 22: 55808625-55847352)

A) B)

Figure A-11. Secondary structure comparison between the difficult to amplify region of “Alaska” and the easily amplified comparable sized fragment of “Bombay” of the Madh5 gene. DNA Secondary Structure at 72°C (1187bps) using mFold (http://mfold.burnet.edu.au/). (A) Alaska fragment between the primers 6481F and 17937-A-R2-Amp which has a GC content of 68.8%. (B) 1-1186 bp of Bombay fragment which has a GC content of 43.1%.

Page 213: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

203

0%

5%

10%

15%

20%

25%

30%

35%

12T's 13T's 14T's 15T's 16T's 17T's 18T's

Number of T's

Perc

enta

ge o

f clo

nes

C57

f/ff/f

Figure A-12. Number of Ts in PolyT region (Greece Fragment) from flexed-tail and wild-type mice. The region around the polyT tract of the Greece fragment (see Figure A-10) was amplified and cloned for sequencing. Graphed is the percentage of clones with a particular numbers of Ts. There were a total of 11 clones from seven C57 wild-type mice, and 32 clones from 10 flexed-tail mice.

Page 214: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

204A)

Greece

3 4 5Tn2

Florence

B)

3 4 52 3 43 4 52 3 4 3 4 53 4 5

Greece constructFlorence/Greece construct

f/f wtC)

Figure A-13. In vitro splicing assays to determine the functional consequences of variations in the PolyT tract in flexed-tail and wild-type mice. A) Schematic of the region of Madh5 evaluated in the splicing assays (refer to Figure A-10). Numbers specify exons colored pink (aberrantly spliced exons in flexed-tail mice) or blue (unaffected exons). The PolyT tract is shown by the dotted vertical lines. B) Constructs of the pSPL3 exon trapping vector with regions of Madh5. In addition to labels from (A), yellow boxes represent the vector, while arrows show locations of primers for PCR to amplify splicing products. Clones used in assays were verified through sequences (14Ts in flexed-tail constructs; 16Ts in wild-type constructs). C) An example of the PCR amplification of splicing products from flexed-tail and wild-type mice using the pSPL3:Greece construct.

Page 215: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

205

Appendix B

Supplementary Information: Genotyping of the Sideroflexin1 (Sfxn1) locus

10% of our flexed-tail colony screened (4/40 mice) were heterozygous at the sideroflexin

locus (Fleming et al. Genes and Development, 2001.15: p.652-657) as determined by two different

genotyping methods that I developed (see Figure B-1, B-2, B-3). There was a 100%

agreement between both methods in the genotypes of a particular mouse when all steps of

the procedures were completed successfully. Currently we are breeding to produce a f/f

mouse that is wild-type at the sideroflexin locus to evaluate any modifying role

sideroflexin may have in the flexed-tail phenotype.

Page 216: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

206

Wildtype: GCGTCTTTACACACAGGT….CGAT CGGACCAGAGCACGTTCATT

Sfxn1 (mutant): GCGTCTTTACACACAGGT….CGAATCGGACCAGAGCACGTTCATT

GANTC Hinf1

CGRY CG BsiE1 R=A/G and Y=C/T

*

*

Mar

kers

Unc

ut

Bsi

E1

Hin

f1

Non specific amplification

Uncut

Cut

A.

Staining of a blood smear from a newborn f/f mouse with Prussian Blue stain to identify iron granule in the fetal erythrocytes. Blue dots in the erythrocytes indicate the presence of siderocytes.

B. Genotyping the Sideroflexin mutation in f/f mice from our colony.

(Top) Schematic of PCR primers and the diagnostic cut sites used to genotype the Sfxn1 mutation in f/f mice from our colony. The bold sequence is region around the A insertion mutant reported by Fleming et al.. Above and below this region are the sequences of the restriction sites used to differentiate between the alleles. The triangles mark the actual cut site. The * above the C base represents an inserted C nucleotide that generates the diagnostic restriction sites.

(Bottom) PCR analysis of DNA isolated from the mouse scored as a f/f in A. Both HinF1 and BsiE1 cut the PCR product indicating that this animal is heterozygous for the A insertion in Sfxn1. The lower band indicated by the arrow is a non-specific amplification.

Figure B-1. Genotyping for the Sideroflexin (Sfxn) mutation

Page 217: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

207

Figure B-2. Direct sequencing of the Sfxn1 exon 2 in f/f mice. Sfxn1 exon 2 from the mouse in B-1(A) was PCR amplified and cloned in pTOPO-10 (Invitrogen). Multiple independent plasmids were sequenced. Two sequences were obtained as marked in the boxes, mutant-GAATC and wildtype-GATC.

Page 218: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

Figure B-3. Analysis of Sfxn mutations in f/f mice by oligonucleotide ligationassay.

Allele specific reporter oligo

Common anchoring oligo

wild-type: ----ACGTGCTCTGGTCCCATCD

Sfxn mutant: ----ACGTGCTCTGGTCCCATTCF

(p)-GAGGCTCCTTGATGTTAATGTTG-- B

Allele specific reporter oligo

Common anchoring oligo

wild-type: ----ACGTGCTCTGGTCCCATCD

Sfxn mutant: ----ACGTGCTCTGGTCCCATTCF

(p)-GAGGCTCCTTGATGTTAATGTTG-- B

CTAC---------------- D

TAC-----------

CF

-----------GB

5’-------CGATG-------3’-

-----------GB

wild-type

T

5’-------CGATG-------3’-CTAC---------------- D

TAC-----------

CF

-----------GB

5’-------CGATG-------3’-

-----------GB

wild-type

T

5’-------CGATG-------3’-

Sfxn mutant

5’-------CGAATG--------3’

-----------GB

CTTAC----------- F

TAC---------------- D

5’-------CGAATG--------3’-----------GB

C

Sfxn mutant

5’-------CGAATG--------3’

-----------GB

CTTAC----------- F

TAC---------------- DTAC---------------- D

5’-------CGAATG--------3’-----------GB

C

y f f f y g g y

Streptavidin

F

B

AA

B

A

D

HRPAP

i.

ii.

iii. Oligonucleotide ligation assay (OLA) was used to confirm all PCR genotypes. (i)The technique utilizes a common anchoring oligo that is phosphorylated at its 5’ end and biotinylated at its 3’ end. Allele specific oligos that contain either a digoxigenin modification (wildtype) or a fluorescein (mutant) at its 5’ end. (ii) Ligation of the anchoring oligo to the allele specific oligo can only occur when the allele specific oligobinds the correct sequence. (iii) Anti-Dig conjugated to alkaline phosphatase and anti-Fitc conjugated to HRP are used to detect the ligatedproducts by ELISA.

We tested all f/f mice in our colony by the PCR assay and OLA. Both assays gave identical results. We conclude that in our colony, the mutation in Sfxn1 reported by Fleming et al. has been separated by recombination from the f/f mutant phenotype. We have identified numerous mice that exhibit the f/f phenotype, but are heterozygous for the Sfxn1 mutation. The f locus therefore cannot encode Sfxn1.

OLA was done as described in Single-well genotyping of diallelic sequence variations by a two-color ELISA-based oligonucleotide ligation assay. VO Tobe, SL Taylor and DA Nickerson. Nucleic Acids research 24:3728-2732, 1996.

208

Page 219: UNDERSTANDING OF EXPANSIVE ERYTHROPOIESIS

VITA

L. E. Lenox

EDUCATION Penn State University 1997-present Ph.D. in Biochemistry, Microbiology and Molecular Biology University of Scranton 1993-1997 Bachelor of Science in Biology—magna cum laude HONORS AND AWARDS Penn State University College of Agriculture Science 2003 Competitive Grant Program winner for the proposal: β-Globin Minor Expression in the flexed-tail (f) mutant mouse Purdue University Applied Management Principles (AMP) Program for Science and Engineering Doctoral Students--6 Continuing Education Units 2002 Representative from PSU Eberly College of Science 1998/1999 Paul M. Althouse Outstanding Graduate Teaching Assistant Award--Honorable Mention 1998 Pela Fay Braucher Scholarship Award COMMITTEES 2002 Representative for the Eberly College of Science (BMMB Department) to Daniel J.Larson Advisory Board (Dean of Eberly College of Science) Topic Focus: How to measure success in the educational process PUBLICATIONS Laurie E. Lenox, John Perry and Robert Paulson. BMP4 and Madh5 regulate the erythroid response to acute anemia. Blood, 2005. 105(7): p. 2741-8. Laurie E. Lenox and Robert Paulson (in preparation). Extramedullary hematopoiesis following an acute anemia in splenectomized mice. ABSTRACTS AND NATIONAL MEETINGS Laurie Lenox, J.M. Perry and R.F. Paulson. The recovery of splenectomized mice to acute anemia. (Talk and Poster). August 2004. 23rd Summer Symposium in Molecular Biology: Hematopoiesis and Immune Cell Function. University Park, PA. Laurie Lenox, John Perry, Robert Paulson. BMP4/Smad5 Signaling Regulates the Expansion of a Novel Erythroid Progenitor in the Adult Spleen in Response to Acute Anemia (Talk and Poster). July 2003 FASEB Summer Research Conference: TGF-β Superfamily: Signaling and Development. Tucson, AZ. Robert Paulson, Laurie Lenox, John Perry, Prashanth Porayette (2002). BMP4 and Smad5 Regulate the Erythroid Response to Acute Anemia in the Adult Spleen (Poster). 44th Annual Meeting and Exposition of the American Society of Hematology. Philadelphia, PA. Laurie E. Lenox and Robert F. Paulson (1999). Positional Cloning of the flexed-tail locus (Poster). Thirteenth Annual International Mammalian Genome Society. Philadelphia, PA.