(Review) (Biological Chemistry) Bernd Giebel Et Michael Punzel 2008

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Article in press - uncorrected proof

Biol. Chem., Vol. 389, pp. 813–824, July 2008 • Copyright  by Walter de Gruyter • Berlin • New York. DOI 10.1515/BC.2008.092

Review

Lineage development of hematopoietic stem and


progenitor cells

Bernd Giebel* and Michael Punzel* den and Le Blanc, 2005). Both exhaustion as well as
enforced expansion of HSC might result in clinical rele-
Institute for Transplantation Diagnostics and Cellular
vant cytopenia or leukemia, respectively. Therefore, the
Therapeutics, Heinrich Heine University, D-40225
size of the HSC-pool and thus the decision whether
Düsseldorf, Germany
HSCs self-renew or become committed to differentiate
* Corresponding authors needs to be biased and tightly regulated. Additionally, to
e-mail: giebel@itz.uni-duesseldorf.de; establish and maintain an appropriate composition of dif-
punzel@itz.uni-duesseldorf.de ferentiated blood cells, processes controlling lineage
specification and differentiation need to be highly con-
trolled as well. Thus, it is a central issue of modern stem
Abstract cell research to uncover the mechanisms governing the
development of HSCs and their committed progeny
Hematopoietic stem cells have the potential to develop
which are hierarchically organized.
into multipotent and different lineage-restricted progeni-
One fundamental requirement to uncover such mech-
tor cells that subsequently generate all mature blood cell
anisms is a reliable model which predicts the underlying
types. The classical model of hematopoietic lineage
hierarchical organization as well as the relationship of the
commitment proposes a first restriction point at which all
individual cell lineages to each other. For several years,
multipotent hematopoietic progenitor cells become com-
the classical hematopoietic commitment model was the
mitted either to the lymphoid or to the myeloid devel-
prevailing model according to which multipotent HSCs
opment, respectively. Recently, this model has been
give rise to multipotent hematopoietic progenitor cells
challenged by the identification of murine as well as
(HPCs), which then create progenitor cells either being
human hematopoietic progenitor cells with lymphoid dif-
restricted to the myeloid or lymphoid development (Reya
ferentiation capabilities that give rise to a restricted sub-
et al., 2001). As this classical hematopoietic model has
set of the myeloid lineages. As the classical model does
been challenged by several new data, we summarize
not include cells with such capacities, these findings
recent findings that imply the existence of additional
suggest the existence of alternative developmental path-
developmental branches within both the murine and the
ways that demand the existence of additional branches
human primitive hematopoietic system that are not
in the classical hematopoietic tree. Together with some
included in the classical model.
phenotypic criteria that characterize different subsets of
multipotent and lineage-restricted progenitor cells, we
summarize these recent findings here.
Murine hematopoietic differentiation
Keywords: hematopoietic progenitor cell (HPC);
hematopoietic stem cell (HSC); lineage specification. As already mentioned, all adult hematopoietic cells derive
from multipotent HSCs, which are capable to replenish
the peripheral blood system of lethally irradiated recipi-
Introduction ents after transplantation. In the murine system, the pool
of HSCs has been initially characterized by the Weissman
Hematopoietic stem cells (HSCs) are defined as clono- group as cells that do not express any lineage specific
genic cells that possess the dual abilities to self-renew antigens (Lin-), that are positive for the stem cell asso-
and to differentiate into mature blood cells of all lineages, ciated antigen (Sca-1) and that express low amounts of
and that contain the capacity to regenerate the whole the T-cell antigen Thy1 (Lin-Sca-1qThy-1low cells). Later,
blood system for a lifetime of an organism. These fea- mouse HSCs were grouped in the Lin-Sca-1qc-kitq (LSK)
tures became evident after transplantation of HSC-con- cell compartment (Spangrude et al., 1988; Ikuta and
taining cell fractions either obtained from adult bone Weissman, 1992; Morrison and Weissman, 1994; Osawa
marrow (BM), from peripheral blood of granulocyte-col- et al., 1996). As shown in Figure 1 and summarized in
ony stimulating factor (G-CSF) treated individuals or from Table 1, the LSK compartment contains different hier-
umbilical cord blood (CB) of newborns into immunocom- archical classes of stem cells, which can be distin-
promised hosts; apart from experiments in animals, clin- guished in clonogenic long-term self-renewing HSCs
ical applications demonstrate that the blood system of (LT-HSCs), transiently self-renewing HSCs (short-term
treated patients can fully be restored by donor HSC for HSCs), and non-self-renewing multipotent progenitors
the rest of their normal lasting life time (Bensinger et al., (MPPs) (Smith et al., 1991; Morrison and Weissman,
1997; Korbling and Anderlini, 2001; Ballen, 2005; Ring- 1994; Osawa et al., 1996; Randall et al., 1996; Akashi et

Brought to you by | BUPMC 2008/271


MIR (I894)
Authenticated
Download Date | 1/30/18 11:39 AM
Article in press - uncorrected proof

814 B. Giebel and M. Punzel

lineage, respectively (Kondo et al., 1997; Akashi et al.,


2000). Therefore, it was postulated that MPPs further dif-
ferentiate into either the common myeloid progenitors
(CMPs) or the common lymphoid progenitors (CLPs),
which sustain either granulocyte/macrophage progeni-
tors (GMPs), as well as megakaryocyte/erythrocyte pro-
genitors (MEPs) or T-cell, B-cell and Natural Killer (NK)
cell differentiation potentials, respectively (Kondo et al.,
1997; Akashi et al., 2000; Reya et al., 2001). Kondo et al.
(1997) described CLPs as a subpopulation of Lin-Sca-
1lowc-kitlowThy-1- cells that express the interleukin-7
receptor a chain (IL-7Ra) and are unable to differentiate
into myeloid cell types. In agreement with these findings,
the lymphopoiesis in IL-7Ra-deficient mice is severely
impaired (Peschon et al., 1994; He and Malek, 1996;
Maki et al., 1996; Kondo et al., 1997). The CMPs con-
taining fraction of Lin-Sca-1-c-kitq cells lacks lymphoid
gene expression and does not express IL-7Ra (Akashi et
al., 2000). Furthermore, and according to the expression
of the Fcg receptor (FcgR) and of CD34 the CMPs con-
taining fraction can be subdivided in three populations:
that of the (i) FcgRlowCD34q cells, (ii) the FcgRlowCD34-
cells, and (iii) the FcgRhighCD34q cells. In contrast to the
FcgRlowCD34q cell fraction which contains CMP activity,
Figure 1 Developmental route(s) in adult murine hematopoiesis. the developmental potential of the FcgRlowCD34- is
Within the murine hematopoietic system, HSCs are discriminat- restricted to the megakaryocyte/erythrocyte lineages and
ed in long-term self-renewing HSCs (LT-HSCs) and transiently defines MEPs, while that of the FcgRhighCD34q cells is
self-renewing HSCs (short-term HSCs) that give rise to non-self-
restricted to the granulocyte/macrophage development
renewing multipotent progenitors (MPPs). In contrast to the clas-
and defines GMPs (Figure 1) (Akashi et al., 2000).
sical model which predicts that MPPs become specified either
as common myeloid progenitors (CMPs) or as common lym- According to the discovery of CMPs and CLPs, it was
phoid progenitors (CLPs), recent findings suggest the existence widely accepted that at this developmental stage all
of additional developmental routes. The pool of MPPs can be hematopoietic cells become either restricted to the mye-
subdivided into cells containing the full CMP potential (FLT3lowV- loid or the lymphoid development, as it has been for-
CAM1q cells) and in cells containing some residual CMP activ- mulated in the ‘classical’ hematopoietic model (Reya et
ities that are primed to develop along the lymphoid routes al., 2001).
(FLT3qV-CAM1q cells). In addition to the classical model in
However, by using an additional set of markers, such
which granulocyte/macrophage progenitors (GMPs) and mega-
karyocyte/erythrocyte progenitors (MEPs) are derivatives of the as FMS-related tyrosine kinase 3 (Flt-3) and vascular cell
CMP, cells have been found that contain the CLP and GMP adhesion molecule 1 (VCAM-1) (Adolfsson et al., 2001;
potential but lack the MEP capacity. These cells are termed as Christensen and Weissman, 2001), Flt-3qV-CAM-1- cells
lymphoid primed multipotent progenitor cells (LMPPs) and are have been described which still contain lymphoid devel-
characterized as FLT3qV-CAM-1- cells. Within the group of the opmental capacities and give rise to granulocytes and
FLT3qV-CAM-1- cells, IL-7Rq cells seem to contain CLPs and macrophages but not to megakaryocytes or erythro-
no myeloid activities, and CCR9q cells still contain CLP capac-
cytes. Cells with these features seem to define a new
ities but are primed to develop along the T-cell lineage (ETP).
group of primitive hematopoietic cells termed as lym-
phoid primed multipotent progenitor cells (LMPPs) and
al., 2000). The LT-HSCs are highly enriched within the do not fit into the classical model of hematopoiesis in
LSK-Thy1.1low cell fraction, and a subset of them can which the first step of lineage commitment simply selects
nearly be purified as CD34- LSK cells that are located in between the lymphoid and the myeloid lineages (Adolfs-
the tip of the side population (‘Tip’-SP CD34- KSL) (Mor- son et al., 2005; Lai and Kondo, 2006). In this context,
rison and Weissman, 1994; Morrison et al., 1997; Mat- comparison of gene expression profiles of LMPPs with
suzaki et al., 2004). However, no sharp definition for that of HSCs revealed a significant down-regulation of
ST-HSCs or for MPPs has been described so far. More genes required for granulocyte/macrophage differentia-
recently, a set of new markers, cell surface receptors of tion, and an up-regulation of genes known to be impor-
the signaling lymphocyte activation molecule (SLAM) tantly involved in lymphoid differentiation, such as Ikaros
family, were found that also help to discriminate subpo- (Adolfsson et al., 2005; Lai and Kondo, 2006; Yoshida et
pulations. In this context, HSCs are described as al., 2006; Mansson et al., 2007). Therefore, these obser-
CD150qCD48-CD244- cells, MPP as CD244qCD48- vations led to the hypothesis that in addition to the clas-
CD150- cells and the most restricted progenitors as sically described development of the GMPs from CMPs,
CD48qCD244qCD150- cells (Kiel et al., 2005). GMPs can arise from LMPPs that are also predicted to
Downstream of the LSK compartment, cells lose their generate CLPs (Figure 1) (Adolfsson et al., 2005; Buza-
multipotent capacities, and at a single cell level progen- Vidas et al., 2007).
itors have been identified whose developmental capaci- Very recently, it was observed further that a subset of
ties are either restricted to the myeloid or to the lymphoid the MPPs expresses the chemokine receptor 9 (CCR9)

Brought to you by | BUPMC MIR (I894)


Authenticated
Download Date | 1/30/18 11:39 AM
Table 1 Murine hematopoietic stem and progenitor cell populations.

Defined mouse Phenotype Determination Frequency References


population function source
HSC Lin-Sca1qc-kitq (LSK) In vivo Tx 0.05% of mouse BM Spangrude et al., 1988; Morrison and Weissman, 1994;
Osawa et al., 1996
LT-HSC LSK Thy1.1low In vivo Tx BM Smith et al., 1991; Morrison and Weissman, 1994
Flt3-VCAM-1q Adolfsson et al., 2001; Christensen and Weissman, 2001
CD34-CD38q 1 LT-HSC in 5 LSK34- Osawa et al., 1996; Randall et al., 1996
‘Tip’-side-population 21% HSCs (0.008% of Matsuzaki et al., 2004
CD150qCD48-CD244- BM) Kiel et al., 2005
ST-HSC LSK Thy1.1int In vivo Tx BM Reya et al., 2001
Flt3lowVCAM-1q Adolfsson et al., 2001; Christensen and Weissman, 2001
CD34qCD38- Osawa et al., 1996; Randall et al., 1996
CD150qCD48-CD244- Kiel et al., 2005
MPP LSK Thy1.1- In vivo Tx BM
Flt3lowVCAM-1q Adolfsson et al., 2001; Christensen and Weissman, 2001
CD34qCD38- Osawa et al., 1996; Randall et al., 1996
CD150-CD48-CD244q Kiel et al., 2005
Flt3lowVCAM-1q MPP LSK Thy1.1-CD34q In vivo Tx BM Lai and Kondo, 2006
Flt3lowVCAM-1q Primitive CMP potential
Flt3qVCAM-1q MPP LSK Thy1.1-CD34q In vivo Tx 15% of CD34q Lai and Kondo, 2006
Flt3qVCAM-1q CLP 15% of Thy1.1-
(marginal CMP) LSK
BM
Subfraction of LMPP
LMPP LSK Flt3qVCAM-1- In vivo Tx 30% of CD34q Adolfsson et al., 2005; Lai and Kondo, 2006
IL-7Raq CLPqGMP (qMegE) 30% of Thy1.1- Forsberg et al., 2006; Mansson et al., 2007
LSK
Article in press - uncorrected proof

BM
ELP LSK Flt3q RAG1qCD34q In vivo Tx 5% of CD34q Igarashi et al., 2002
CLPqGMP 5% of Thy1.1-
LSK
BM
subfraction of LMPP
ETP precursor LSK Flt3qVCAM-1- CLPq(GMP) BM Lai and Kondo, 2007
CCR9q
CLP Lin-Sca1lowc-KitlowThy1.1- In vivo Tx BM Kondo et al., 1997
IL-7Raq
CMP Lin-Sca1-c-KitqIL-7Ra- In vivo Tx BM Akashi et al., 2000
FcgRlowCD34q Akashi et al., 2000
CMP/MEP FcgRlowCD34- Akashi et al., 2000
CMP/GMP FcgRhighCD34q Akashi et al., 2000

Download Date | 1/30/18 11:39 AM


Authenticated
Brought to you by | BUPMC MIR (I894)
Hematopoietic stem cell lineage development 815
Article in press - uncorrected proof

816 B. Giebel and M. Punzel

and preferentially home to the thymus. This together with Hematopoietic lineage specification
the concept that these cells are phenotypically related to
the most immature early T-lineage progenitors (ETPs) led The observations that hematopoietic cells can divide
to the suggestion that although these CCR9q MPPs pos- asymmetrically (Beckmann et al., 2007; Chang et al.,
sess limited GMP potential and can give rise to T-, B- 2007) and that their development depends on the sur-
and dendritic cells, they are already primed for T-cell rounding environment, the hematopoietic niches (Calvi et
development (Lai and Kondo, 2007). This implies that al., 2003; Zhang et al., 2003; Kiel et al., 2005) demon-
even cells that seem to be committed to a certain lineage strate that the development of hematopoietic cells is
are not necessarily determined and might retain a certain orchestrated by a combination of intrinsic and extrinsic
degree of developmental flexibility, sometimes called signals. These signals become integrated to finally alter
plasticity. In addition and compatible to this hypothesis, the genetic program of corresponding cells. Thus, it
it was found that pax5 mutant cells that already started becomes evident that transcription factors as key regu-
to rearrange their immunoglobulin genes cannot proceed lators of genetic programs play important roles in spec-
with the B-cell development and possess long-term and ifying the cell fate of hematopoietic cells during
multipotent reconstitution capacities (Schaniel et al., hematopoietic lineage development. Certain transcription
2002a,b). In this context, over-expression experiments factors, such as members of the GATA and SCL/Tal-
revealed that Pax5 inhibits T-cell development, but does 1gene families, act at different levels during hemato-
not block early myeloid-lineage development (Cotta et poietic development and seem to function as
al., 2003). It rather promotes maintenance of bipheno- ‘hematopoietic master regulators’. For example, during
typic myeloid progenitors which co-express myeloid and embryonal stages SCL/Tal-1 acts together with members
B-cell lineage-associated genes (Montecino-Rodriguez of the E-protein family (E2A) to induce mesodermal pro-
et al., 2001; Anderson et al., 2007). genitor cells to acquire the HSC fate. Mice lacking SCL/
Taken together, all these findings strongly suggest the Tal-1 activities are embryonic lethal and do not develop
existence of murine progenitor cells that follow alterna- a hematopoietic system (Robb et al., 1995, 1996; Shiv-
tive developmental pathways which do not correspond dasani et al., 1995; Porcher et al., 1996), whereas gained
to the classical model of hematopoiesis, and thus define SCL/Tal-1 activity increases the content of embryonic
new roadmaps during early hematopoiesis. hematopoiesis (Gering et al., 1998; Mead et al., 2001).
During adult hematopoiesis, SCL/Tal-1 is specifically
expressed in primitive hematopoietic cells as well as in
cells of the megakaryocyte/erythrocyte lineage and has
Unresolved controversies been found to control the differentiation of the latter cells
(Elefanty et al., 1998; Akashi et al., 2000; Ramalho-San-
tos et al., 2002; Mikkola et al., 2003; Zhang et al., 2005).
Although the latter data seem to be contradictory to the
In contrast to SCL/Tal-1, E2A expression now is required
classical hematopoietic model, the Weissman group
for governing the lymphoid differentiation (Bain et al.,
demonstrated that transplantation of approximately 500
1994, 1997, 1999).
green fluorescent protein (GFP)-expressing Flt-3qLSK A typical paradigm of how lineage decisions can be
cells that were designated to contain LMPP capacities controlled is the antagonism between PU.1 and Gata-1.
(Adolfsson et al., 2005; Lai and Kondo, 2006) result in a PU.1 directs the differentiation of hematopoietic progen-
robust reconstitution of GFPq platelets and GFPq ery- itors into macrophages, neutrophils and B-lymphocytes
throid progenitor cells (Forsberg et al., 2006). Conse- (DeKoter et al., 1998; DeKoter and Singh, 2000), while
quently, the hypothesis of alternative hematopoietic road Gata-1 promotes development along the MEP lineage
maps that might bypass the CMPs during GMP commit- (Pevny et al., 1991; Fujiwara et al., 1996; Shivdasani et
ment became questionable. According to Forsberg et al. al., 1997). It has been shown that PU.1 interacts with
(2006), this hypothesis might mistakenly be based on Gata proteins, i.e., PU.1 and Gata proteins antagonize
experimental artifacts probably related to the usage of activities of each other by direct protein-protein interac-
too little cell amounts and to the readouts that have been tions. Additionally, PU.1 can inhibit the expression of
performed. They suppose that the LMPPs are rather late Gata factor encoding genes (Nerlov and Graf, 1998;
MPPs with reduced self-renewal capacities, but still con- Rekhtman et al., 1999; Zhang et al., 1999, 2000; Nerlov
tain the full multi-lineage potentials (Forsberg et al., et al., 2000; Walsh et al., 2002). As it has been shown
2006). The Jacobsen group, however, argues against that that the amount of PU.1 decides upon whether hema-
and takes into account that the Flt-3qLSK-cell population topoietic progenitors follow the lymphoid development or
like any stem and progenitor cell population is hetero- whether they differentiate into macrophages, it becomes
geneous and that since the Weissman group used evident why enforced PU.1 or Gata expression can alter
approximately 500 cells for transplantation into single the development of hematopoietic progenitor cells
mice, they might have included some infrequent progen- (DeKoter and Singh, 2000; Iwasaki et al., 2003).
itor cells residing within this fraction that contained MEP Remarkably, over-expression experiments revealed
potentials (Mansson et al., 2007). To shed some more that enforced expression of PU.1 and Gata proteins as
light into this controversial discussion, they suggest sub- well as that of other myeloid transcription factors fre-
fractionation of the cell population thought to contain quently can override the lymphoid program and subse-
LMPP capacities and perform further analyses. quently reprogram these progenitors towards certain

Brought to you by | BUPMC MIR (I894)


Authenticated
Download Date | 1/30/18 11:39 AM
Article in press - uncorrected proof

Hematopoietic stem cell lineage development 817

myeloid differentiation pathways (Iwasaki et al., 2003, etic stem cell requires transplantation of donor stem cells
2006; Hsu et al., 2006). Transcription factors, whose into an isogeneic host. For ethical reasons, human cell
enforced expression is sufficient to enable committed fractions cannot be tested experimentally whether or not
CMPs to give rise to lymphoid cells, have to our knowl- they contain HSCs or more mature progenitor cells that
edge not been defined yet. This together with the obser- are unable to mediate long-term engraftment. Thus, it is
vation that HSCs express large numbers of myeloid but difficult to purify defined cell populations containing true
not lymphoid genes and that during lymphoid develop- human HSC activity.
ment myeloid gene expression progressively decreases Over the last two decades, surrogate xenogeneic in
and lymphoid gene expression subsequently increases vivo-transplantation models have been established to cir-
suggests that myelopoiesis and most likely the granulo- cumvent these problems. Nowadays, xenografts into
cyte macrophage development defines a default com- immunodeficient NOD-SCID mice are the gold standard
mitment pathway for multipotent hematopoietic pro- readout for human long-term repopulating hematopoietic
genitors that actively need to be maintained (Akashi et cells (Kamel-Reid and Dick, 1988; Lapidot et al., 1992;
al., 2003; Iwasaki and Akashi, 2007; Mansson et al., Larochelle et al., 1996), but also the fetal sheep model
2007). In this context, lymphopoiesis seems to become (Zanjani et al., 1992).
initiated when activities of pro-myelopoietic factors are Although, the usage of these animal surrogate models
inhibited or become modulated, e.g., by the activity of has rigorously increased the knowledge about human
the Notch signaling pathway (Franco et al., 2006; Laiosa primitive hematopoietic cells, it should be kept in mind
et al., 2006; Rothenberg, 2007). Depending on the inte- that the quality and frequency of human HSCs cannot
grated intrinsic and extrinsic signals, the default myeloid directly be compared to that obtained in syngeneic mice
commitment can theoretically differently be modulated transplants. There are several limitations, such as the for-
and thus might explain the existence of classical as well eign homing environment that skews lineage readout
as alternative lineage developmental pathways, i.e., that together with the usual short period (less than 7 weeks)
of CMPs, LMPPs and GMPs. between transplantation and engraftment analyses.
This concept further helps to explain the observation Indeed, it has been reported that most of the primitive
of Muller-Sieburg and colleagues, who found that indi- hematopoietic cells obtained from non-human primates,
vidual HSCs differ in their abilities to give rise to myeloid i.e., from baboons, that are able to engraft NOD-SCID
or lymphoid daughter cells. According to these obser- mice, only contain short-term but not long-term recon-
vations, HSCs can differently be biased, either towards stituting activities and therefore should not be designated
the lymphoid or to the myeloid lineage (Muller-Sieburg as true HSCs (Ramirez et al., 1998; Horn et al., 2003;
et al., 2002, 2004). Glimm et al., 2005; Horn and Blasczyk, 2007).
In addition to the xenotransplantation experiments, a
number of different in vitro readout systems had been
Human hematopoiesis established that have recently been extensively reviewed
(Coulombel, 2004). These assays also revealed, and
It is evident that progress in the discovery of lineage surely will reveal, comprehensive knowledge about
development and hematopoietic differentiation in the human primitive hematopoietic cells, which has been the
mouse model has mainly been achieved by two experi- basis for improvements of applications using human
mental approaches: (i) by analyses of genetic alterations HSC sources in clinical trials. Furthermore, this knowl-
that result in changes of the lineage composition and/or edge is important to increase our understanding of the
in alterations in the phenotype of distinct hematopoietic etiology of certain diseases, e.g., that of leukemia.
cells, and (ii) by the isolation of distinct subpopulations
containing cells with HSC/HPC potential and their sub- Phenotypic description of human primitive
sequent functional analyses, either performed in vivo hematopoietic cells
using transplantation models or in vitro at clonal levels.
Although, knowledge about biological processes As already pointed out, murine LT-HSCs are highly
obtained in mice can often be transferred to humans and enriched within the LSK cell fraction of cells which do
other mammals, some processes are not transferable not express the cell surface antigen CD34. However,
between species. In this context, essential differences most human cells with repopulating cell activities have
are found between the hematopoiesis in mouse and in been identified as CD34q cells. Hence, in contrast to
man. Before reviewing these differences, we would like mice, CD34 has become a surrogate marker of human
to add some comments about human HSCs. HSCs (Krause et al., 1996). Apart from CD34, there are
other substantial differences in the immunophenotype
between HSCs in man and mouse. Primitive hematopoi-
Caveats in human HSC research etic cells that have lost their repopulating capacities and
became committed to differentiate show higher expres-
The fact that stem cell grafts can fully restore the blood sion levels of the CD38 antigen. Therefore, the majority
system of immunocompromised patients is proof for of the most primitive human hematopoietic cells is found
human HSCs (Bensinger et al., 1997; Korbling and within the CD34qCD38- cell fraction (Terstappen et al.,
Anderlini, 2001; Ballen, 2005; Ringden and Le Blanc, 1991; Hao et al., 1996; Larochelle et al., 1996). In the
2005). However, the functional readout of a hematopoi- murine system, LT-HSCs have been characterized as

Brought to you by | BUPMC MIR (I894)


Authenticated
Download Date | 1/30/18 11:39 AM
Article in press - uncorrected proof

818 B. Giebel and M. Punzel

CD34-CD38q cells and more committed MPPs as criminate human multipotent HPCs from lineage-restrict-
CD34qCD38- cells (Osawa et al., 1996; Randall et al., ed CMPs or CLPs have been described as well. In this
1996). Therefore, CD38 is a primitive marker in mice and context, BM-derived Lin-CD34qCD38qCD45RAqCD10-
a differentiation marker in humans. wHLA-DRq c-kit-Thy1-x cells and CB-derived CD34qCD38-
In addition to CD34, the AC133 antigen, prominin-1/ CD7q wCD45RAqHLA-DRqc-kit-Thy1-x cells contain CLP
CD133, which is co-expressed on most of the CD34q activities (Galy et al., 1995; Hao et al., 2001; Hoebeke et
cells, was characterized as a stem cell surrogate marker al., 2007). Cells with CMP activities have been identified
in humans (Yin et al., 1997; de Wynter et al., 1998) and as BM-derived Lin-CD34qCD38qIL-3RalowCD45RA- cells
becomes localized on the uropod of cultivated CD34q that generate both Lin-CD34qCD38qIL-3RalowCD45RAq
cells (Giebel et al., 2004). As a very small fraction of cells containing the clonogenic potential for granulocyte
immature cells with repopulating activities has been iden- and macrophage (GMPs) development, and Lin-
tified to reside within the Lin-CD34-CD38- cell fraction CD34qCD38qIL-3Ra-CD45RA- cells containing that for
which expresses CD133 (Bhatia et al., 1998; Zanjani et megakaryocyte and erythrocyte (MEPs) development
al., 1998; Gallacher et al., 2000), CD133 seems to (Manz et al., 2002).
describe the most primitive human hematopoietic cells In controversy to the classical hematopoietic model,
more precisely than CD34. In agreement with this, we Haddad and colleagues described two human CB-
observed that upon cultivation CD34qCD133q cells gen- derived cell populations (Lin-CD34qCD45RAhighCD10q
erate a secondary CD34qCD133low/- population, which and Lin-CD34qCD45RAhighCD7q) that although they were
has lost the primitive characteristics of CD34qCD133q either primed to the T-cell and NK cell development or to
cells and increases over time (Giebel et al., 2006; von the B-cell lineage, respectively, still sustained robust gra-
Levetzow et al., 2006; Beckmann et al., 2007). Within the nulomonocytic differentiation potentials (Haddad et al.,
murine hematopoietic system, CD133 (Prominin) was 2004). Pointing toward the same direction, progeny of
found to be expressed on 36% of the BM-derived CD34q CD34qCD10-CD19- cells can follow the course of early
cells (Corbeil et al., 2000), but to our knowledge its B-cell development and perform B-cell specific DJH rear-
expression has not been systematically analyzed within rangements without losing their capacity to differentiate
the HSC-containing LSK fraction, yet. However, in con- into NK-cells, T-cells and macrophages (Reynaud et al.,
trast to primitive human hematopoietic cells, expanded 2003). Additionally, CD34qCD19q cells that belong to a
murine HSCs neither express CD133 nor CD62L (Zhang small CXCR4- B-cell progenitor subpopulation have been
and Lodish, 2005). According to our analyses, CD62L is found to contain capacities for granulocyte, macrophage
specifically expressed on cultivated human CD34q and red blood cell development (Hou et al., 2005).
CD133q cells (Beckmann et al., 2007). Similarly, in our own single cell analyses in which we
In addition to absence of lineage-specific antigens deposited individual CB-derived Lin-CD34qCD38- cells
(Lin-) and low-level expression of CD38 as one of the and studied the primitive hematopoietic re-initiation
most applied phenotypical criteria to describe human potential of their arising daughter cells in clonogenic
HSC candidates, other antigens have been used to dis- myeloid (long-term culture initiating cell; LTC-IC), as well
criminate more primitive from more mature CD34q cells as in clonogenic lymphoid assays (NK cell initiating cell;
(Table 2). Some of the combinations used are Lin- NK-IC), we retrospectively identified cells which gave rise
CD34qHLA-DR- (Verfaillie et al., 1990), CD34qc-kitqCD38- to NK cells and macrophages but not to granulocytes
CD33-Rho123- (Liu and Verfaillie, 2002), Lin-CD133q (Giebel et al., 2006). According to their colony formation
ALDHhi (Hess et al., 2006) or Lin-CD34qCD38-Rho123low potential, we termed such corresponding cells as
(McKenzie et al., 2007). Recently, together with CD71 macrophage NK-ICs (M-NK-ICs). Although we do not
and CD62L, which have been used to discriminate prim- know yet whether these cells retain other lymphoid
itive hematopoietic cell fractions before (Lansdorp and developmental capacities, cells that contain develop-
Dragowska, 1992; Mayani et al., 1993; Mohle et al., 1995; mental potentials for NK cells and macrophages while
Koenig et al., 1999), we qualified the tetraspanins CD53 simultaneously lacking granulocyte development should
and CD63 as new markers, that together with CD133
not exist according to the classical model of
help to discriminate more primitive cultured
hematopoiesis.
CD34qCD133qwCD53qCD62LqCD63-CD71lowx cells from
In addition, cells that gave rise to NK cells and also
more mature cultivated CD34qCD133low/-wCD53-CD62L-
generated granulocytes should – at least according to
CD63qCD71qx cells (Beckmann et al., 2007). It may be
the classical model – contain both the full myeloid as well
that this new combination of antigens will help to solve
as the full lymphoid developmental potentials. Accord-
the dilemma that upon cultivation certain cell surface
ingly, and even though it remains to be determined
antigens, such as c-kit or CD38, lose most of their pre-
whether such cells contain megakaryocytic and erythroid
dictive value for the identification of distinct functional
potentials, these primitive progenitors have been named
subpopulations of very primitive hematopoietic cells
myeloid lymphoid initiating cells or ML-ICs (Punzel et al.,
(Dorrell et al., 2000; Danet et al., 2001).
1999). Due to the fact that these cells are able to reinitiate
multiple myeloid as well as lymphoid long-term hema-
topoiesis in vitro, they are considered as one of the most
Pros and cons for the classical model of primitive in vitro measurable stem cell equivalent of
hematopoiesis in humans human hematopoietic cells (Punzel et al., 1999). Follow-
Comparable to the murine system and in agreement with ing the classical model, ML-ICs should give rise to
the classical model of hematopoiesis criteria that dis- daughter cells retaining the multipotent ML-IC capacity

Brought to you by | BUPMC MIR (I894)


Authenticated
Download Date | 1/30/18 11:39 AM
Table 2 Human hematopoietic stem and progenitor cell populations.

Defined human Phenotype Determination Frequency References


population function source

HSC CD34qCD38-c-kitqLin- Terstappen et al., 1991; Hao et al., 1996;


Larochelle et al., 1996
Additional marker CD133q(CD34-/q) Preimmune fetal sheep, Fetal liver, UCB, BM, PB Bhatia et al., 1998; de Wynter et al., 1998;
NOD/SCID mice; Zanjani et al., 1998; Yin et al., 1997
in vitro differentiation, Gallacher et al., 2000; Giebel et al., 2006;
flow cytometry Beckmann et al., 2007

ALDHhigh(CD133q) NOD/SCID/b2m-/- mice; UCB Hess et al., 2004, 2006


in vitro differentiation,
flow cytometry
HLA-DRlow(CD34q) In vitro differentiation, BM Verfaillie et al., 1990
flow cytometry
CD33-Rho123-(CD34q) NOD/SCID, UCB Liu and Verfaillie, 2002
In vitro differentiation, McKenzie et al., 2007
flow cytometry
CD71-CD62Lq(CD34q) In vitro differentiation, UCB, BM, PB, cultured Lansdorp and Dragowska, 1992; Mayani et al.,
flow cytometry UCB 1993; Mohle et al., 1995; Koenig et al., 1999;
Beckmann et al., 2007
CD53qCD63-(CD133q) In vitro differentiation, Cultured UCB Beckmann et al., 2007
flow cytometry

CLP CD34qCD38qLin-CD10q In vitro differentiation, BM Galy et al., 1995


HLA-DRqc-kit-Thy1.1- flow cytometry,
Gene expression
CD34qCD38-CD7q In vitro differentiation, UCB Hao et al., 2001; Hoebeke et al., 2007
HLA-DRqc-kit-Thy1.1- flow cytometry,
Gene expression
Article in press - uncorrected proof

CLP (B-cells)q CD34qLin-CD38qIL-3Ralow In vitro differentiation, UCB Haddad et al., 2004


GMP-potential CD45RAhighCD10q flow cytometry,
gene expression
CLP (T/NK-cells)q CD34qLin-CD38qIL-3Ralow UCB Haddad et al., 2004
GMP-potential CD45RAhighCD7q
CLPqmacrophages CD34qCD10-CD19- In vitro differentiation, UCB Reynaud et al., 2003
CD79aqIL-7Raq flow cytometry,
gene expression
BMP(B-cells)q CD34qCD19qCXCR4- In vitro differentiation, BM Hou et al., 2005
GMPqerythroid flow cytometry ;0.3% mononuclear cell fraction (MNC) BM
M-NK-IC CD34qCD38- In vitro differentiation UCB Giebel et al., 2004
CMP CD34qLin-CD38qIL-3Ralow NOD/SCID/b2m-/- mice, ;0.28% MNC BM; ;0.4% MNC UCB Manz et al., 2002
CD45RA-CD16-CD32- in vitro differentiation
GMP IL-3RalowCD45RAq ;0.35% MNC BM; ;0.3% MNC UCB Manz et al., 2002

Download Date | 1/30/18 11:39 AM


Authenticated
Brought to you by | BUPMC MIR (I894)
MeP IL-3Ra-CD45RA- ;0.13% MNC BM; ;0.05% MNC UCB Manz et al., 2002
Hematopoietic stem cell lineage development 819
Article in press - uncorrected proof

820 B. Giebel and M. Punzel

or to daughter cells, being committed to either the lym- fate. Remarkably, in all cases in which such daughter
phoid (CLPs) or the myeloid (CMPs) lineage develop- cells generated NK cells, we also observed the devel-
ment. Thus, in the ML-IC assay, CLPs should give rise opment of macrophages, suggesting that under the con-
to NK cells but not to myeloid colonies, CMPs should ditions used ML-ICs do not give rise to cells fulfilling the
generate myeloid colonies but not NK cells. However, in CLP criteria, but give rise to cells with M-NK-IC activities
our studies in which we analyzed the cell fate of individ- (Giebel et al., 2006). As cells with improved M-NK-IC
ual daughter cells, we never observed that CD34qCD38- activities frequently generated daughter cells, one of
cells originally containing the ML-IC capacity gave rise which retained the M-NK-IC potential and one which only
to one daughter cell fulfilling the CMP criteria and another generated NK cells and thus could in principal reflect
one fulfilling that of a CLP (Giebel et al., 2006). In all CLP activities (Giebel et al., 2006), we propose that under
cases studied, at least one daughter cell retained the full the conditions used in our experiments, CLPs arise from
ML-IC potential of its mother cell, and in more than 80% multipotent HPCs via progenitor cells containing M-NK-
of the cases, one daughter cell acquired a restricted cell IC activities. These findings within the early human
hematopoietic system strongly support the data obtained
from the groups of Jacobsen and Kondo which have
been described previously (Adolfsson et al., 2005; Lai
and Kondo, 2006). Taking the current literature into
account, a proposed model of human hematopoietic
lineage development is shown in Figure 2.

Conclusion

In summary, recent findings in mouse and in man predict


additional or alternative roadmaps within the model of
the hematopoietic development (Figures 1 and 2). There-
fore, the hypothesis that myeloid and lymphoid progen-
itors are strictly separated at a single branching point, as
proposed in the classical model of hematopoiesis, can-
not be kept in this form anymore. It seems more likely
that there are different developmental routes a multi-
potent hematopoietic cell can follow. As extrinsic factors
play important roles in governing the cell fate of hema-
topoietic progenitor cells and because primitive hema-
topoietic cells are motile and sometimes leave their
Figure 2 Developmental route(s) in neonatal and adult human niches enter the blood or lymph stream and return to
hematopoiesis. functional niches (Goodman and Hodgson, 1962; Wright
Clinical trials revealed that the heterogeneous pool of human et al., 2001; Abkowitz et al., 2003; Massberg et al., 2007),
CD34q cells contains HSC activities that can restore the hema-
progenitors will be influenced by a bundle of different
topoietic system of immunocompromised patients. However, for
ethical reasons subfractions should not be transplanted in places and different signals. In this context, the order
humans, their developmental potential can only be estimated in how progenitor cells receive certain signals should
in vitro assays or in xenogenic models, e.g., in NOD/SCID mice already influence their developmental potential. With an
or in fetal sheep. According to these surrogate HSC analyses, example of the developing external sensory organs of
the most primitive human hematopoietic cells seem to belong Drosophila melanogaster, it has been shown that cells
to the Lin-CD133qc-kitq cell fraction. Although a small propor-
which transduce the Notch-signaling in a first decision
tion of such cells lack CD34 expression – maybe the most prim-
itive human hematopoietic cells detected so far –, the vast
step but not in a second acquire a different cell fate than
majority of these cells express CD34. These cells can be dis- those cells that only transduce the Notch signal in the
criminated in more primitive CD34qCD38low/- cells and more second decision step (Campos-Ortega, 1996). Taking
mature CD34qCD38q cells. Similar to the mouse system, recent into account how primitive hematopoietic cells receive
findings have suggested that multipotent progenitors, enriched certain extrinsic signals, there should be many different
within the CD34qCD38low/- cell fraction, not only give rise to ways how the cell fate of developing hematopoietic cells
CMPs and CLPs, as predicted by the classical model of hema-
can be primed. Thus, the discovery of progenitor cells
topoiesis, but also to cells that display reduced developmental
capabilities of both the myeloid and the lymphoid developmental that are specified by different routes may be the tip of an
routes. In this context, T/NK-cell-primed and B-cell-primed cells iceberg, which is not compatible with the classical model
have been identified that still possess GMP abilities (T/NK GMP, of hematopoiesis.
B GMP), as well as cells with CLP and macrophage potentials In this context, the old discussion whether hemato-
(M CLP). Additionally, some putatively lineage restricted lym- poietic cell fates are specified in a deterministic or sto-
phocytes still contain capabilities to produce macrophages chastic manner can be repeated (McCulloch, 1983).
(M-NK, M-B, M-T). Abbreviations: ML-IC, myeloid-lymphoid ini-
According to our data in which most of the primitive
tiating cell; E-LTC-IC, extended long-term culture initiating cell;
CAFC, cobblestone forming cell; LTC-IC, long-term culture ini- hematopoietic cells which had ML-IC capacities gave
tiating cell; M-NK-IC, macrophage-NK cell initiating cell; NK-IC, rise to differently specified daughter cells (Giebel et al.,
NK cell initiating cell; CFU, colony-forming unit. 2006), and due to the identification of asymmetrically

Brought to you by | BUPMC MIR (I894)


Authenticated
Download Date | 1/30/18 11:39 AM
Article in press - uncorrected proof

Hematopoietic stem cell lineage development 821

segregating proteins (Beckmann et al., 2007), we suggest ages is hierarchically controlled during early hematopoiesis.
that very primitive hematopoietic cells are intrinsically Blood 101, 383–389.
Anderson, K., Rusterholz, C., Mansson, R., Jensen, C.T., Bacos,
determined to acquire different cell fates, which is sup-
K., Zandi, S., Sasaki, Y., Nerlov, C., Sigvardsson, M., and
ported by the finding that within the murine system clon- Jacobsen, S.E. (2007). Ectopic expression of PAX5 promotes
ally expanded HSCs maintain their tendency to give rise maintenance of biphenotypic myeloid progenitors coexpress-
to either myeloid-biased or lymphoid-biased HSC clones ing myeloid and B-cell lineage-associated genes. Blood 109,
(Muller-Sieburg et al., 2002, 2004). This concept would 3697–3705.
support the deterministic model of hematopoiesis (Mul- Bain, G., Maandag, E.C., Izon, D.J., Amsen, D., Kruisbeek, A.M.,
ler-Sieburg and Sieburg, 2006). We further hypothesize Weintraub, B.C., Krop, I., Schlissel, M.S., Feeney, A.J., van
Roon, M., et al. (1994). E2A proteins are required for proper
that as long as cells with identical developmental poten-
B cell development and initiation of immunoglobulin gene
tial receive cell fate controlling signals in the same order,
rearrangements. Cell 79, 885–892.
they should also obey the deterministic model and follow Bain, G., Engel, I., Robanus Maandag, E.C., te Riele, H.P.,
the same developmental routes. As soon as the order Voland, J.R., Sharp, L.L., Chun, J., Huey, B., Pinkel, D., and
and combination of signals is altered, cells will finally Murre, C. (1997). E2A deficiency leads to abnormalities in
adopt different developmental probabilities. A different alphabeta T-cell development and to rapid development of
combination of extrinsic signals might therefore account T-cell lymphomas. Mol. Cell. Biol. 17, 4782–4791.
for observed variabilities among different functional read- Bain, G., Quong, M.W., Soloff, R.S., Hedrick, S.M., and Murre,
C. (1999). Thymocyte maturation is regulated by the activity
out systems (Ramirez et al., 1998; Horn et al., 2003;
of the helix-loop-helix protein, E47. J. Exp. Med. 190,
Glimm et al., 2005; Horn and Blasczyk, 2007). Due to the 1605–1616.
multitude of different cell fate influencing factors, we Ballen, K.K. (2005). New trends in umbilical cord blood trans-
assume that there is a high degree of randomness within plantation. Blood 105, 3786–3792.
the order how cells receive such signals. Therefore, we Beckmann, J., Scheitza, S., Wernet, P., Fischer, J.C., and Giebel,
suggest that the process of hematopoietic cell fate spec- B. (2007). Asymmetric cell division within the human hema-
ification finally remains at least partially stochastic. topoietic stem and progenitor cell compartment: identifica-
tion of asymmetrically segregating proteins. Blood 109,
Future work and more detailed analyses with defined
5494–5501.
combinations of extrinsic signals will be required to finally Bensinger, W.I., Buckner, D., and Gahrton, G. (1997). Allogeneic
improve the hierarchical concept of the hematopoietic stem cell transplantation for multiple myeloma. Hematol.
tree and should define the basis for new functional anal- Oncol. Clin. North Am. 11, 147–157.
yses that help to dissect the mechanisms involved in Bhatia, M., Bonnet, D., Murdoch, B., Gan, O.I., and Dick, J.E.
governing self-renewing and differentiation processes (1998). A newly discovered class of human hematopoietic
within the primitive hematopoietic systems. cells with SCID-repopulating activity. Nat. Med. 4, 1038–
1045.
Buza-Vidas, N., Luc, S., and Jacobsen, S.E. (2007). Delineation
of the earliest lineage commitment steps of haematopoietic
Acknowledgments
stem cells: new developments, controversies and major
challenges. Curr. Opin. Hematol. 14, 315–321.
We thank Julia Beckmann, Verdon Taylor and Johannes Fischer
Calvi, L.M., Adams, G.B., Weibrecht, K.W., Weber, J.M., Olson,
for discussion and critics on the manuscript. Our studies were
D.P., Knight, M.C., Martin, R.P., Schipani, E., Divieti, P., Bring-
supported by grants from the Deutsche Forschungsgemein-
hurst, F.R., et al. (2003). Osteoblastic cells regulate the
schaft (SPP1109 GI 336/1-4), the Forschungskommission of the
haematopoietic stem cell niche. Nature 425, 841–846.
HHU-Düsseldorf und the Leukämieliga e.V.
Campos-Ortega, J.A. (1996). Numb diverts notch pathway off
the tramtrack. Neuron 17, 1–4.
Chang, J.T., Palanivel, V.R., Kinjyo, I., Schambach, F., Intlekofer,
References A.M., Banerjee, A., Longworth, S.A., Vinup, K.E., Mrass, P.,
Oliaro, J., et al. (2007). Asymmetric T lymphocyte division in
the initiation of adaptive immune responses. Science 315,
Abkowitz, J.L., Robinson, A.E., Kale, S., Long, M.W., and Chen,
J. (2003). Mobilization of hematopoietic stem cells during 1687–1691.
homeostasis and after cytokine exposure. Blood 102, Christensen, J.L. and Weissman, I.L. (2001). Flk-2 is a marker in
1249–1253. hematopoietic stem cell differentiation: a simple method to
Adolfsson, J., Borge, O.J., Bryder, D., Theilgaard-Monch, K., isolate long-term stem cells. Proc. Natl. Acad. Sci. USA 98,
Astrand-Grundstrom, I., Sitnicka, E., Sasaki, Y., and Jacob- 14541–14546.
sen, S.E. (2001). Upregulation of Flt3 expression within the Corbeil, D., Roper, K., Hellwig, A., Tavian, M., Miraglia, S., Watt,
bone marrow Lin-Sca1qc-kitq stem cell compartment is S.M., Simmons, P.J., Peault, B., Buck, D.W., and Huttner,
accompanied by loss of self-renewal capacity. Immunity 15, W.B. (2000). The human AC133 hematopoietic stem cell anti-
659–669. gen is also expressed in epithelial cells and targeted to plas-
Adolfsson, J., Mansson, R., Buza-Vidas, N., Hultquist, A., Liuba, ma membrane protrusions. J. Biol. Chem. 275, 5512–5520.
K., Jensen, C.T., Bryder, D., Yang, L., Borge, O.J., Thoren, Cotta, C.V., Zhang, Z., Kim, H.G., and Klug, C.A. (2003). Pax5
L.A., et al. (2005). Identification of Flt3q lympho-myeloid determines B- versus T-cell fate and does not block early
stem cells lacking erythro-megakaryocytic potential a revised myeloid-lineage development. Blood 101, 4342–4346.
road map for adult blood lineage commitment. Cell 121, Coulombel, L. (2004). Identification of hematopoietic stem/pro-
295–306. genitor cells: strength and drawbacks of functional assays.
Akashi, K., Traver, D., Miyamoto, T., and Weissman, I.L. (2000). Oncogene 23, 7210–7222.
A clonogenic common myeloid progenitor that gives rise to Danet, G.H., Lee, H.W., Luongo, J.L., Simon, M.C., and Bonnet,
all myeloid lineages. Nature 404, 193–197. D.A. (2001). Dissociation between stem cell phenotype and
Akashi, K., He, X., Chen, J., Iwasaki, H., Niu, C., Steenhard, B., NOD/SCID repopulating activity in human peripheral blood
Zhang, J., Haug, J., and Li, L. (2003). Transcriptional acces- CD34q cells after ex vivo expansion. Exp. Hematol. 29,
sibility for genes of multiple tissues and hematopoietic line- 1465–1473.

Brought to you by | BUPMC MIR (I894)


Authenticated
Download Date | 1/30/18 11:39 AM
Article in press - uncorrected proof

822 B. Giebel and M. Punzel

de Wynter, E.A., Buck, D., Hart, C., Heywood, R., Coutinho, L.H., Hao, Q.L., Thiemann, F.T., Petersen, D., Smogorzewska, E.M.,
Clayton, A., Rafferty, J.A., Burt, D., Guenechea, G., Bueren, and Crooks, G.M. (1996). Extended long-term culture reveals
J.A., et al. (1998). CD34qAC133q cells isolated from cord a highly quiescent and primitive human hematopoietic pro-
blood are highly enriched in long-term culture-initiating cells, genitor population. Blood 88, 3306–3313.
NOD/SCID-repopulating cells and dendritic cell progenitors. Hao, Q.L., Zhu, J., Price, M.A., Payne, K.J., Barsky, L.W., and
Stem Cells 16, 387–396. Crooks, G.M. (2001). Identification of a novel, human multi-
DeKoter, R.P. and Singh, H. (2000). Regulation of B lymphocyte lymphoid progenitor in cord blood. Blood 97, 3683–3690.
and macrophage development by graded expression of He, Y.W. and Malek, T.R. (1996). Interleukin-7 receptor a is
PU.1. Science 288, 1439–1441. essential for the development of gdq T cells, but not natural
DeKoter, R.P., Walsh, J.C., and Singh, H. (1998). PU.1 regulates killer cells. J. Exp. Med. 184, 289–293.
both cytokine-dependent proliferation and differentiation of Hess, D.A., Meyerrose, T.E., Wirthlin, L., Craft, T.P., Herrbrich,
granulocyte/macrophage progenitors. EMBO J. 17, 4456– P.E., Creer, M.H., and Nolta, J.A. (2004). Functional charac-
4468. terization of highly purified human hematopoietic repopula-
Dorrell, C., Gan, O.I., Pereira, D.S., Hawley, R.G., and Dick, J.E. ting cells isolated according to aldehyde dehydrogenase
(2000). Expansion of human cord blood CD34qCD38- cells activity. Blood 104, 1648–1655.
in ex vivo culture during retroviral transduction without a cor- Hess, D.A., Wirthlin, L., Craft, T.P., Herrbrich, P.E., Hohm, S.A.,
responding increase in SCID repopulating cell (SRC) fre- Lahey, R., Eades, W.C., Creer, M.H., and Nolta, J.A. (2006).
quency: dissociation of SRC phenotype and function. Blood Selection based on CD133 and high aldehyde dehydrogen-
95, 102–110. ase activity isolates long-term reconstituting human hema-
Elefanty, A.G., Begley, C.G., Metcalf, D., Barnett, L., Kontgen, F., topoietic stem cells. Blood 107, 2162–2169.
and Robb, L. (1998). Characterization of hematopoietic pro- Hoebeke, I., De Smedt, M., Stolz, F., Pike-Overzet, K., Staal, F.J.,
genitor cells that express the transcription factor SCL, using Plum, J., and Leclercq, G. (2007). T-, B- and NK-lymphoid,
a lacZ ‘knock-in’ strategy. Proc. Natl. Acad. Sci. USA 95, but not myeloid cells arise from human CD34qCD38-CD7q
11897–11902. common lymphoid progenitors expressing lymphoid-specific
genes. Leukemia 21, 311–319.
Forsberg, E.C., Serwold, T., Kogan, S., Weissman, I.L., and Pas-
Horn, P.A. and Blasczyk, R. (2007). Severe combined immuno-
segue, E. (2006). New evidence supporting megakaryocyte-
deficiency-repopulating cell assay may overestimate long-
erythrocyte potential of flk2/flt3q multipotent hematopoietic
term repopulation ability. Stem Cells 25, 3271–3272.
progenitors. Cell 126, 415–426.
Horn, P.A., Thomasson, B.M., Wood, B.L., Andrews, R.G., Mor-
Franco, C.B., Scripture-Adams, D.D., Proekt, I., Taghon, T.,
ris, J.C., and Kiem, H.P. (2003). Distinct hematopoietic stem/
Weiss, A.H., Yui, M.A., Adams, S.L., Diamond, R.A., and
progenitor cell populations are responsible for repopulating
Rothenberg, E.V. (2006). Notch/Delta signaling constrains
NOD/SCID mice compared with nonhuman primates. Blood
reengineering of pro-T cells by PU.1. Proc. Natl. Acad. Sci.
102, 4329–4335.
USA 103, 11993–11998.
Hou, Y.H., Srour, E.F., Ramsey, H., Dahl, R., Broxmeyer, H.E.,
Fujiwara, Y., Browne, C.P., Cunniff, K., Goff, S.C., and Orkin, S.H.
and Hromas, R. (2005). Identification of a human B-cell/mye-
(1996). Arrested development of embryonic red cell precur- loid common progenitor by the absence of CXCR4. Blood
sors in mouse embryos lacking transcription factor GATA-1. 105, 3488–3492.
Proc. Natl. Acad. Sci. USA 93, 12355–12358. Hsu, C.L., King-Fleischman, A.G., Lai, A.Y., Matsumoto, Y.,
Gallacher, L., Murdoch, B., Wu, D.M., Karanu, F.N., Keeney, M., Weissman, I.L., and Kondo, M. (2006). Antagonistic effect of
and Bhatia, M. (2000). Isolation and characterization of CCAAT enhancer-binding protein-a and Pax5 in myeloid or
human CD34-Lin- and CD34qLin- hematopoietic stem cells lymphoid lineage choice in common lymphoid progenitors.
using cell surface markers AC133 and CD7. Blood 95, Proc. Natl. Acad. Sci. USA 103, 672–677.
2813–2820. Igarashi, H., Gregory, S.C., Yokota, T., Sakaguchi, N., and Kin-
Galy, A., Travis, M., Cen, D., and Chen, B. (1995). Human T, B, cade, P.W. (2002). Transcription from the RAG1 locus marks
natural killer, and dendritic cells arise from a common bone the earliest lymphocyte progenitors in bone marrow. Immu-
marrow progenitor cell subset. Immunity 3, 459–473. nity 17, 117–130.
Gering, M., Rodaway, A.R., Gottgens, B., Patient, R.K., and Ikuta, K. and Weissman, I.L. (1992). Evidence that hematopoietic
Green, A.R. (1998). The SCL gene specifies haemangioblast stem cells express mouse c-kit but do not depend on steel
development from early mesoderm. EMBO J. 17, factor for their generation. Proc. Natl. Acad. Sci. USA 89,
4029–4045. 1502–1506.
Giebel, B., Corbeil, D., Beckmann, J., Hohn, J., Freund, D., Gie- Iwasaki, H. and Akashi, K. (2007). Myeloid lineage commitment
sen, K., Fischer, J., Kogler, G., and Wernet, P. (2004). Seg- from the hematopoietic stem cell. Immunity 26, 726–740.
regation of lipid raft markers including CD133 in polarized Iwasaki, H., Mizuno, S., Wells, R.A., Cantor, A.B., Watanabe, S.,
human hematopoietic stem and progenitor cells. Blood 104, and Akashi, K. (2003). GATA-1 converts lymphoid and mye-
2332–2338. lomonocytic progenitors into the megakaryocyte/erythrocyte
Giebel, B., Zhang, T., Beckmann, J., Spanholtz, J., Wernet, P., lineages. Immunity 19, 451–462.
Ho, A.D., and Punzel, M. (2006). Primitive human hemato- Iwasaki, H., Mizuno, S., Arinobu, Y., Ozawa, H., Mori, Y., Shi-
poietic cells give rise to differentially specified daughter cells gematsu, H., Takatsu, K., Tenen, D.G., and Akashi, K. (2006).
upon their initial cell division. Blood 107, 2146–2152. The order of expression of transcription factors directs hier-
Glimm, H., Schmidt, M., Fischer, M., Schwarzwaelder, K., Wiss- archical specification of hematopoietic lineages. Genes Dev.
ler, M., Klingenberg, S., Prinz, C., Waller, C.F., Lange, W., 20, 3010–3021.
Eaves, C.J., and von Kalle, C. (2005). Efficient marking of Kamel-Reid, S. and Dick, J.E. (1988). Engraftment of immune-
human cells with rapid but transient repopulating activity in deficient mice with human hematopoietic stem cells. Science
autografted recipients. Blood 106, 893–898. 242, 1706–1709.
Goodman, J.W. and Hodgson, G.S. (1962). Evidence for stem Kiel, M.J., Yilmaz, O.H., Iwashita, T., Terhorst, C., and Morrison,
cells in the peripheral blood of mice. Blood 19, 702–714. S.J. (2005). SLAM family receptors distinguish hematopoietic
Haddad, R., Guardiola, P., Izac, B., Thibault, C., Radich, J., Dele- stem and progenitor cells and reveal endothelial niches for
zoide, A.L., Baillou, C., Lemoine, F.M., Gluckman, J.C., Pflu- stem cells. Cell 121, 1109–1121.
mio, F., and Canque, B. (2004). Molecular characterization of Koenig, J.M., Baron, S., Luo, D., Benson, N.A., and Deisseroth,
early human T/NK and B-lymphoid progenitor cells in umbil- A.B. (1999). L-selectin expression enhances clonogenesis of
ical cord blood. Blood 104, 3918–3926. CD34q cord blood progenitors. Pediatr. Res. 45, 867–870.

Brought to you by | BUPMC MIR (I894)


Authenticated
Download Date | 1/30/18 11:39 AM
Article in press - uncorrected proof

Hematopoietic stem cell lineage development 823

Kondo, M., Weissman, I.L., and Akashi, K. (1997). Identification in the absence of stem-cell leukaemia SCL/tal-1 gene.
of clonogenic common lymphoid progenitors in mouse bone Nature 421, 547–551.
marrow. Cell 91, 661–672. Mohle, R., Murea, S., Kirsch, M., and Haas, R. (1995). Differential
Korbling, M. and Anderlini, P. (2001). Peripheral blood stem cell expression of L-selectin, VLA-4, and LFA-1 on CD34q pro-
versus bone marrow allotransplantation: does the source of genitor cells from bone marrow and peripheral blood during
hematopoietic stem cells matter? Blood 98, 2900–2908. G-CSF-enhanced recovery. Exp. Hematol. 23, 1535–1542.
Krause, D.S., Fackler, M.J., Civin, C.I., and May, W.S. (1996). Montecino-Rodriguez, E., Leathers, H., and Dorshkind, K.
CD34: structure, biology, and clinical utility. Blood 87, 1–13. (2001). Bipotential B-macrophage progenitors are present in
Lai, A.Y. and Kondo, M. (2006). Asymmetrical lymphoid and adult bone marrow. Nat. Immunol. 2, 83–88.
myeloid lineage commitment in multipotent hematopoietic Morrison, S.J. and Weissman, I.L. (1994). The long-term repo-
progenitors. J. Exp. Med. 203, 1867–1873. pulating subset of hematopoietic stem cells is deterministic
Lai, A.Y. and Kondo, M. (2007). Identification of a bone marrow and isolatable by phenotype. Immunity 1, 661–673.
precursor of the earliest thymocytes in adult mouse. Proc. Morrison, S.J., Wandycz, A.M., Hemmati, H.D., Wright, D.E., and
Natl. Acad. Sci. USA 104, 6311–6316. Weissman, I.L. (1997). Identification of a lineage of multi-
Laiosa, C.V., Stadtfeld, M., Xie, H., de Andres-Aguayo, L., and potent hematopoietic progenitors. Development 124, 1929–
Graf, T. (2006). Reprogramming of committed T cell progen- 1939.
itors to macrophages and dendritic cells by C/EBP a and Muller-Sieburg, C.E. and Sieburg, H.B. (2006). The GOD of
PU.1 transcription factors. Immunity 25, 731–744. hematopoietic stem cells: a clonal diversity model of the
Lansdorp, P.M. and Dragowska, W. (1992). Long-term erythro- stem cell compartment. Cell Cycle 5, 394–398.
poiesis from constant numbers of CD34q cells in serum-free Muller-Sieburg, C.E., Cho, R.H., Thoman, M., Adkins, B., and
cultures initiated with highly purified progenitor cells from Sieburg, H.B. (2002). Deterministic regulation of hemato-
human bone marrow. J. Exp. Med. 175, 1501–1509. poietic stem cell self-renewal and differentiation. Blood 100,
Lapidot, T., Pflumio, F., Doedens, M., Murdoch, B., Williams, 1302–1309.
D.E., and Dick, J.E. (1992). Cytokine stimulation of multi- Muller-Sieburg, C.E., Cho, R.H., Karlsson, L., Huang, J.F., and
lineage hematopoiesis from immature human cells engrafted Sieburg, H.B. (2004). Myeloid-biased hematopoietic stem
in SCID mice. Science 255, 1137–1141. cells have extensive self-renewal capacity but generate
Larochelle, A., Vormoor, J., Hanenberg, H., Wang, J.C., Bhatia, diminished lymphoid progeny with impaired IL-7 responsive-
M., Lapidot, T., Moritz, T., Murdoch, B., Xiao, X.L., Kato, I., ness. Blood 103, 4111–4118.
et al. (1996). Identification of primitive human hematopoietic Nerlov, C. and Graf, T. (1998). PU.1 induces myeloid lineage
cells capable of repopulating NOD/SCID mouse bone mar- commitment in multipotent hematopoietic progenitors.
row: implications for gene therapy. Nat. Med. 2, 1329–1337. Genes Dev. 12, 2403–2412.
Liu, H. and Verfaillie, C.M. (2002). Myeloid-lymphoid initiating Nerlov, C., Querfurth, E., Kulessa, H., and Graf, T. (2000). GATA-
cells (ML-IC) are highly enriched in the rhodamine-c- 1 interacts with the myeloid PU.1 transcription factor and
kitqCD33-CD38- fraction of umbilical cord CD34q cells. Exp. represses PU.1-dependent transcription. Blood 95,
Hematol. 30, 582–589. 2543–2551.
Maki, K., Sunaga, S., Komagata, Y., Kodaira, Y., Mabuchi, A., Osawa, M., Hanada, K., Hamada, H., and Nakauchi, H. (1996).
Karasuyama, H., Yokomuro, K., Miyazaki, J.I., and Ikuta, K. Long-term lymphohematopoietic reconstitution by a single
(1996). Interleukin 7 receptor-deficient mice lack gammadelta CD34-low/negative hematopoietic stem cell. Science 273,
T cells. Proc. Natl. Acad. Sci. USA 93, 7172–7177. 242–245.
Mansson, R., Hultquist, A., Luc, S., Yang, L., Anderson, K., Kha- Peschon, J.J., Morrissey, P.J., Grabstein, K.H., Ramsdell, F.J.,
razi, S., Al-Hashmi, S., Liuba, K., Thoren, L., Adolfsson, J., Maraskovsky, E., Gliniak, B.C., Park, L.S., Ziegler, S.F., Wil-
et al. (2007). Molecular evidence for hierarchical transcrip-
liams, D.E., Ware, C.B., et al. (1994). Early lymphocyte
tional lineage priming in fetal and adult stem cells and
expansion is severely impaired in interleukin 7 receptor-defi-
multipotent progenitors. Immunity 26, 407–419.
cient mice. J. Exp. Med. 180, 1955–1960.
Manz, M.G., Miyamoto, T., Akashi, K., and Weissman, I.L. (2002).
Pevny, L., Simon, M.C., Robertson, E., Klein, W.H., Tsai, S.F.,
Prospective isolation of human clonogenic common myeloid
D’Agati, V., Orkin, S.H., and Costantini, F. (1991). Erythroid
progenitors. Proc. Natl. Acad. Sci. USA 99, 11872–11877.
differentiation in chimaeric mice blocked by a targeted muta-
Massberg, S., Schaerli, P., Knezevic-Maramica, I., Köllnberger,
tion in the gene for transcription factor GATA-1. Nature 349,
M., Tubo, N., Moseman, E.A., Huff, I.V., Junt, T., Wagers, A.J.,
257–260.
Mazo, I.B., and von Andrian, U.H. (2007). Immunosurveil-
Porcher, C., Swat, W., Rockwell, K., Fujiwara, Y., Alt, F.W., and
lance by hematopoietic progenitor cells trafficking through
Orkin, S.H. (1996). The T cell leukemia oncoprotein SCL/
blood, lymph, and peripheral tissues. Cell 131, 994–1008.
Matsuzaki, Y., Kinjo, K., Mulligan, R.C., and Okano, H. (2004). tal-1 is essential for development of all hematopoietic line-
Unexpectedly efficient homing capacity of purified murine ages. Cell 86, 47–57.
hematopoietic stem cells. Immunity 20, 87–93. Punzel, M., Wissink, S.D., Miller, J.S., Moore, K.A., Lemischka,
Mayani, H., Dragowska, W., and Lansdorp, P.M. (1993). Cyto- I.R., and Verfaillie, C.M. (1999). The myeloid-lymphoid initi-
kine-induced selective expansion and maturation of erythroid ating cell (ML-IC) assay assesses the fate of multipotent
versus myeloid progenitors from purified cord blood precur- human progenitors in vitro. Blood 93, 3750–3756.
sor cells. Blood 81, 3252–3258. Ramalho-Santos, M., Yoon, S., Matsuzaki, Y., Mulligan, R.C.,
McCulloch, E.A. (1983). Stem cells in normal and leukemic and Melton, D.A. (2002). ‘‘Stemness’’: transcriptional profiling
hemopoiesis (Henry Stratton Lecture, 1982). Blood 62, 1–13. of embryonic and adult stem cells. Science 298, 597–600.
McKenzie, J.L., Takenaka, K., Gan, O.I., Doedens, M., and Dick, Ramirez, M., Rottman, G.A., Shultz, L.D., and Civin, C.I. (1998).
J.E. (2007). Low rhodamine 123 retention identifies long-term Mature human hematopoietic cells in donor bone marrow
human hematopoietic stem cells within the Lin-CD34qCD38- complicate interpretation of stem/progenitor cell assays in
population. Blood 109, 543–545. xenogeneic hematopoietic chimeras. Exp. Hematol. 26,
Mead, P.E., Deconinck, A.E., Huber, T.L., Orkin, S.H., and Zon, 332–344.
L.I. (2001). Primitive erythropoiesis in the Xenopus embryo: Randall, T.D., Lund, F.E., Howard, M.C., and Weissman, I.L.
the synergistic role of LMO-2, SCL and GATA-binding pro- (1996). Expression of murine CD38 defines a population of
teins. Development 128, 2301–2308. long-term reconstituting hematopoietic stem cells. Blood 87,
Mikkola, H.K., Klintman, J., Yang, H., Hock, H., Schlaeger, T.M., 4057–4067.
Fujiwara, Y., and Orkin, S.H. (2003). Haematopoietic stem Rekhtman, N., Radparvar, F., Evans, T., and Skoultchi, A.I.
cells retain long-term repopulating activity and multipotency (1999). Direct interaction of hematopoietic transcription fac-

Brought to you by | BUPMC MIR (I894)


Authenticated
Download Date | 1/30/18 11:39 AM
Article in press - uncorrected proof

824 B. Giebel and M. Punzel

tors PU.1 and GATA-1: functional antagonism in erythroid von Levetzow, G., Spanholtz, J., Beckmann, J., Fischer, J.,
cells. Genes Dev. 13, 1398–1411. Kogler, G., Wernet, P., Punzel, M., and Giebel, B. (2006).
Reya, T., Morrison, S.J., Clarke, M.F., and Weissman, I.L. (2001). Nucleofection, an efficient nonviral method to transfer genes
Stem cells, cancer, and cancer stem cells. Nature 414, into human hematopoietic stem and progenitor cells. Stem
105–111. Cells Dev. 15, 278–285.
Reynaud, D., Lefort, N., Manie, E., Coulombel, L., and Levy, Y. Walsh, J.C., DeKoter, R.P., Lee, H.J., Smith, E.D., Lancki, D.W.,
(2003). In vitro identification of human pro-B cells that give Gurish, M.F., Friend, D.S., Stevens, R.L., Anastasi, J., and
rise to macrophages, natural killer cells, and T cells. Blood Singh, H. (2002). Cooperative and antagonistic interplay
101, 4313–4321. between PU.1 and GATA-2 in the specification of myeloid cell
Ringden, O. and Le Blanc, K. (2005). Allogeneic hematopoietic fates. Immunity 17, 665–676.
stem cell transplantation: state of the art and new perspec- Wright, D.E., Wagers, A.J., Gulati, A.P., Johnson, F.L., and
tives. APMIS 113, 813–830. Weissman, I.L. (2001). Physiological migration of hemato-
Robb, L., Rasko, J.E., Bath, M.L., Strasser, A., and Begley, C.G. poietic stem and progenitor cells. Science 294, 1933–1936.
(1995). scl, a gene frequently activated in human T cell leu- Yin, A.H., Miraglia, S., Zanjani, E.D., Almeida-Porada, G., Oga-
kaemia, does not induce lymphomas in transgenic mice. wa, M., Leary, A.G., Olweus, J., Kearney, J., and Buck, D.W.
Oncogene 10, 205–209. (1997). AC133, a novel marker for human hematopoietic stem
Robb, L., Elwood, N.J., Elefanty, A.G., Kontgen, F., Li, R., Bar- and progenitor cells. Blood 90, 5002–5012.
nett, L.D., and Begley, C.G. (1996). The scl gene product is Yoshida, T., Ng, S.Y., Zuniga-Pflucker, J.C., and Georgopoulos,
required for the generation of all hematopoietic lineages in K. (2006). Early hematopoietic lineage restrictions directed by
the adult mouse. EMBO J. 15, 4123–4129. Ikaros. Nat. Immunol. 7, 382–391.
Rothenberg, E.V. (2007). Negotiation of the T lineage fate deci- Zanjani, E.D., Pallavicini, M.G., Ascensao, J.L., Flake, A.W., Lan-
sion by transcription-factor interplay and microenvironmental glois, R.G., Reitsma, M., MacKintosh, F.R., Stutes, D., Har-
signals. Immunity 26, 690–702.
rison, M.R., and Tavassoli, M. (1992). Engraftment and
Schaniel, C., Bruno, L., Melchers, F., and Rolink, A.G. (2002a).
long-term expression of human fetal hemopoietic stem cells
Multiple hematopoietic cell lineages develop in vivo from
in sheep following transplantation in utero. J. Clin. Invest. 89,
transplanted Pax5-deficient pre-B I-cell clones. Blood 99,
1178–1188.
472–478.
Zanjani, E.D., Almeida-Porada, G., Livingston, A.G., Flake, A.W.,
Schaniel, C., Gottar, M., Roosnek, E., Melchers, F., and Rolink,
and Ogawa, M. (1998). Human bone marrow CD34- cells
A.G. (2002b). Extensive in vivo self-renewal, long-term
engraft in vivo and undergo multilineage expression that
reconstitution capacity, and hematopoietic multipotency of
includes giving rise to CD34q cells. Exp. Hematol. 26,
Pax5-deficient precursor B-cell clones. Blood 99,
353–360.
2760–2766.
Zhang, C.C. and Lodish, H.F. (2005). Murine hematopoietic stem
Shivdasani, R.A., Mayer, E.L., and Orkin, S.H. (1995). Absence
of blood formation in mice lacking the T-cell leukaemia onco- cells change their surface phenotype during ex vivo expan-
protein tal-1/SCL. Nature 373, 432–434. sion. Blood 105, 4314–4320.
Shivdasani, R.A., Fujiwara, Y., McDevitt, M.A., and Orkin, S.H. Zhang, P., Behre, G., Pan, J., Iwama, A., Wara-Aswapati, N.,
(1997). A lineage-selective knockout establishes the critical Radomska, H.S., Auron, P.E., Tenen, D.G., and Sun, Z.
role of transcription factor GATA-1 in megakaryocyte growth (1999). Negative cross-talk between hematopoietic regula-
and platelet development. EMBO J. 16, 3965–3973. tors: GATA proteins repress PU.1. Proc. Natl. Acad. Sci. USA
Smith, L.G., Weissman, I.L., and Heimfeld, S. (1991). Clonal 96, 8705–8710.
analysis of hematopoietic stem-cell differentiation in vivo. Zhang, P., Zhang, X., Iwama, A., Yu, C., Smith, K.A., Mueller,
Proc. Natl. Acad. Sci. USA 88, 2788–2792. B.U., Narravula, S., Torbett, B.E., Orkin, S.H., and Tenen,
Spangrude, G.J., Heimfeld, S., and Weissman, I.L. (1988). Puri- D.G. (2000). PU.1 inhibits GATA-1 function and erythroid dif-
fication and characterization of mouse hematopoietic stem ferentiation by blocking GATA-1 DNA binding. Blood 96,
cells. Science 241, 58–62. 2641–2648.
Terstappen, L.W., Huang, S., Safford, M., Lansdorp, P.M., and Zhang, J., Niu, C., Ye, L., Huang, H., He, X., Tong, W.G., Ross,
Loken, M.R. (1991). Sequential generations of hematopoietic J., Haug, J., Johnson, T., Feng, J.Q., et al. (2003). Identifi-
colonies derived from single nonlineage-committed cation of the haematopoietic stem cell niche and control of
CD34qCD38- progenitor cells. Blood 77, 1218–1227. the niche size. Nature 425, 836–841.
Verfaillie, C., Blakolmer, K., and McGlave, P. (1990). Purified Zhang, W.J., Park, C., Arentson, E., and Choi, K. (2005). Mod-
primitive human hematopoietic progenitor cells with long- ulation of hematopoietic and endothelial cell differentiation
term in vitro repopulating capacity adhere selectively to irra- from mouse embryonic stem cells by different culture con-
diated bone marrow stroma. J. Exp. Med. 172, 502–509. ditions. Blood 105, 111–114.

Brought to you by | BUPMC MIR (I894)


Authenticated
Download Date | 1/30/18 11:39 AM

You might also like