You are on page 1of 23

BioDrugs

https://doi.org/10.1007/s40259-020-00439-6

REVIEW ARTICLE

Innovative Therapies for Hemoglobin Disorders


Karine Sii‑Felice1 · Olivier Negre1,2,3 · Christian Brendel2 · Alisa Tubsuwan4 · Eglantine Morel‑à‑l’Huissier1 ·
Camille Filardo1 · Emmanuel Payen1 

© Springer Nature Switzerland AG 2020

Abstract
β-Globin gene transfer has been used as a paradigm for hematopoietic stem cell (HSC) gene therapy, but is subject to major
difficulties, such as the lack of selection of genetically corrected HSCs, the need for high-level expression of the therapeutic
gene, and cell-specific transgene expression. It took more than 40 years for scientists and physicians to advance from the
cloning of globin gene and discovering globin gene mutations to improving our understanding of the pathophysiological
mechanisms involved, the detection of genetic modifiers, the development of animal models and gene transfer vectors, com-
prehensive animal testing, and demonstrations of phenotypic improvement in clinical trials, culminating in the authorization
of the first gene therapy product for β-thalassemia in 2019. Research has focused mostly on the development of lentiviral
gene therapy vectors expressing variants of the β-globin gene or, more recently, targeting a γ-globin repressor, some of which
have entered clinical testing and should soon diversify the available treatments and promote price competition. These results
are encouraging, but we have yet to reach the end of the story. New molecular and cellular tools, such as gene editing or the
development of induced pluripotent stem cells, are being developed, heralding the emergence of alternative products, the
efficacy and safety of which are being studied. Hemoglobin disorders constitute an important model for testing the pros and
cons of these advanced technologies, some of which are already in the clinical phase. In this review, we focus on the develop-
ment of the advanced products and recent technological innovations that could lead to clinical trials in the near future, and
provide hope for a definitive cure of these severe conditions.

Key Points 

Advances in the development of lentiviral gene therapies


in hematopoietic stem cells have led to clinical benefits
in several patients with β-thalassemia and sickle cell
disease.
Gene editing approaches for hemoglobinopathies are
exciting and have already started to transition from lab to
Karine Sii-Felice and Olivier Negre contributed equally.
clinic.
* Emmanuel Payen
emmanuel.payen@cea.fr
1
Division of Innovative Therapies and UMR 1184 IMVA‑HB,
Institute of Biology François Jacob, CEA, INSERM, Paris-
Saclay University, CEA Fontenay aux Roses, 18 route du
Panorama, 92260 Fontenay aux Roses, France 1 Introduction
2
Dana‑Farber/Boston Children’s Cancer and Blood Disorders
Center, Harvard Medical School, Boston, MA, USA 1.1 Hemoglobin Disorders
3
Biotherapy Partners SAS, Paris, France
4
β-Hemoglobin disorders are the most common single-gene
Institute of Molecular Biosciences, Mahidol University,
disorders worldwide. They affect hundreds of thousands
Salaya, Phutthamonthon District 73170, Nakhon Pathom,
Thailand of newborns each year [1], not only in regions of endemic

Vol.:(0123456789)
K. Sii‑Felice et al.

malaria [2], but throughout the world [3, 4] due to popula- of human leukocyte antigen (HLA)-matched sibling donor
tion movements [5]. cells in pediatric recipients, resulting in an event-free sur-
In patients with β-thalassemia, an imbalance between the vival of more than 80% and an overall survival of 90% [21].
α- and β-globin chains causes the precipitation of unstable For both disorders, the long-term quality of life is better
α-globin/heme complexes in erythroid precursors, ineffec- in patients successfully treated by curative HSCT than in
tive erythropoiesis, and hemolytic anemia [6], resulting in patients treated by noncurative symptomatic therapy [22,
erythroid hyperplasia and the dysregulation of iron homeo- 23]. High-resolution HLA-typing has made it possible to
stasis, which worsens the impairment of erythropoiesis [7]. achieve satisfactory outcomes in a limited number of TDT
A distinction is made between subjects producing small patients receiving cells from matched unrelated donors [24],
amounts of the β-globin chain (non-β0/β0 or β+-thalassemia) but the risk of graft-versus-host disease is much higher than
and patients entirely lacking this chain (β0/β0-thalassemia), that in the matched sibling donor setting and is considered
but the various forms of β-thalassemia are best graded on the unacceptable in SCD [25]. Thus, despite the advantages of
basis of hemoglobin levels, clinical tolerance, and transfu- its curative nature, HSCT entails greater short-term risks
sion dependence. The most severe form of β-thalassemia, than conservative therapy. For patients without a sibling
transfusion-dependent thalassemia (TDT), requires life- donor and those at high risk of transplantation-related
long red blood cell transfusions, resulting in iron overload morbidity, gene therapy is a promising alternative medical
and complications, such as heart failure, liver fibrosis. and approach.
endocrine disease [8]. The leading cause of early death in
these patients is cardiac disease [9, 10], but increases in 1.2 Goal of Gene Therapies
lifespan will probably reveal the effects of risk factors on
other causes of premature morbidity and mortality, such as Improvement of the β-thalassemia phenotype requires the
infection [11], liver disease [12], hepatocarcinoma [13], and expression of sufficient amounts of β like-globin chain to
various other complications [14]. restore the α/β-globin chain balance. This can be achieved
Sickle cell disease (SCD) results from the substitu- with the β-globin gene itself, which can be transferred or
tion of a valine for a glutamic acid in the β-globin chain corrected, the γ-globin gene, which can be transferred or
of sickle hemoglobin tetramers (HbS, α2βS2). This amino induced, or any engineered variant producing a protein capa-
acid interacts with a hydrophobic pocket on the surface of ble of associating with α-globin chains through a mechanism
another tetramer, initiating the aggregation of deoxy-HbS similar to that operating in adult (HbA, α2β2) or fetal (HbF,
into long fibers that modify the structure and rheological α2γ2) hemoglobin. In SCD, the objective is to express a β
properties of red blood cells [15]. HbS polymerization and like-globin gene sufficiently strongly to dilute the βS-globin
erythrocyte injury cause vaso-occlusive disease, hemolytic chain effectively and to disrupt the HbS polymerization
anemia, endothelial dysfunction, and chronic inflammation process. HbF impairs HbS polymerization, and its concen-
[16, 17]. All these processes are mutually self-sustaining tration in erythrocytes is correlated with an improvement
and act in synergy to cause acute pain and end-organ failure. of clinical symptoms [26]. Red blood cells containing HbS
Sickle cell syndromes include sickle cell anemia (SCA), the together with HbF sickle less readily than cells with a simi-
genotype of which is homozygosity for βS (p.Glu6Val in lar proportion of HbA [27]. Thus, the transfer of the fetal
HBB), and other forms resulting from the coinheritance of γ-globin gene into hematopoietic stem cells (HSCs) or the
one βS allele and another HBB variant. These include sickle reactivation of its expression in adult erythroid cells is a
β0-thalassemia (HbSβ0-thal), sickle β+-thalassemia (HbSβ+- valuable therapeutic approach for this disease. HbF and the
thal), and sickle-hemoglobin C disease (HbSC). The clinical minor component of adult erythrocytes HbA2 (α2δ2) have
features of HbSβ+-thal and HbSC are generally less severe similar sparing effects on intracellular HbS polymerization
than those in SCA [18]. HbSβ0-thal is generally considered [28]. The hydrophobic interaction between βS and γ/δ-globin
clinically indistinguishable from SCA, although some stud- chains is destabilized by the glutamine residue at position
ies have reported a slightly lower incidence of SCD comor- 87 (Gln-γ/δ87) [29–31], which is a threonine residue in
bidities [19]. β-globin chains. A mutated βT87Q-globin variant has been
Until recently, allogeneic hematopoietic stem cell trans- shown to inhibit HbS polymerization as efficiently as the
plantation (HSCT) was the only consolidated curative γ-globin chain in vitro [32]. The βT87Q-globin gene can thus
approach for severe β-hemoglobinopathies. In a recent be expressed via gene therapy vectors, both to compensate
international study including 1000 SCD patients receiving for the β-globin chain deficiency in β-thalassemia and to
matched sibling donor cells, 5-year overall survival was destabilize the HbS polymer in SCD. It has two key advan-
92.9% and event-free survival was 91.4% [20]. Over the last tages over the γ- and δ-globin genes: (1) γ-globin mRNAs
30 years, HSCT has been used to treat β-thalassemia patients are not very stable in cells containing significant numbers
with severe forms of the disease, mostly through the infusion of β-globin transcripts [33, 34] and (2) the affinity of the β
Innovative Therapies for Hemoglobin Disorders

subunit for the α-globin chain is greater than that of the γ- or 47], but long-term expression levels in erythroid cells after
δ-chains [35, 36]. These properties make the formation of transplantation remained low [44, 48]. It was ultimately the
the HbA tetramer, and most likely of ­HbAT87Q (α2βA-T87Q2), advent of LVVs that made it possible to consider treating
more likely than that of HbF or HbA2 at similar levels of hemoglobinopathies by ex vivo gene transfer into HSCs.
gene transcription and/or vector copy numbers (VCNs). Side-by-side comparisons of LVVs containing either the
Another potentially interesting intermolecular contact within minimal LCR sequence, such as those that could be intro-
the polymer is the axial interaction between the Glu-β22 and duced into γ-RVVs (≈ 1 kb), or longer genomic fractions
His-α20 residues of adjacent tetramers. The replacement of of the β-globin LCR confirmed the superiority of LVVs for
Glu-β22 with an alanine residue has been shown to interfere exporting complex sequences and expressing the β-globin
with HbS polymerization in vitro [30], and the incorporation gene in mice [49]. Another advantage of LVVs is their abil-
of this substitution into anti-sickling β-globin variants to ity to transduce nondividing cells [50], which is of para-
produce βAS2 (βA-T87Q/E22A) has been proposed [37]. How- mount importance because HSCs are mostly quiescent [51].
ever, the E22A mutation decreases the negative charge on The first studies demonstrating the value of LVVs for
the β-subunit, which may decrease its ability to dimerize the treatment of hemoglobin disorders by the autologous
with α-globin chains [38]. A third mutation (β16 Gly−Asp), transplantation of genetically modified HSCs were published
known as HbJ-Baltimore, was added to compensate for the in the early 2000s by the groups of Leboulch and Sadelain
loss of the negatively charged amino acid [39]. Single- and [32, 49, 52, 53]. The HPV403, HPV436, and TNS9 vec-
triple-mutated β-globin variants (βT87Q and βAS3) have anti- tors were essentially similar (Fig. 1). The β-globin gene was
sickling effects similar to that of HbF in vitro [32, 39]. Their introduced in its entirety, with the exception of a deletion
relative effectiveness is yet to be determined. of an AT-rich region in intron 2, potentially responsible for
The most advanced gene therapy technology is based on vector instability [46, 47]. The DNase I-hypersensitive sites
the ex vivo modification of HSCs and infusion of the geneti- 2, 3, and 4, corresponding to about 3 kb of the β-globin LCR
cally modified cells after myeloablative conditioning. These (Fig. 1), were included to ensure high expression levels. One
therapeutic methods, at various stages of development into of the differences between these vectors lay in the thera-
clinical applications, are based on the addition of genes via peutic genes, βA-globin in TNS9 and HPV403, and βA-T87Q-
lentiviral vectors (LVVs), the use of guided nucleases to globin in HPV436, which had a glutamine residue in place
modify genome elements, or gene repair through homolo- of the threonine in position β87 [32]. The central polypurine
gous recombination. Other technologies are being studied tract/central termination sequence (cPPT/CTS) of HIV-1
in animal models, including the use of genetically modified was incorporated into the HPV403 and HPV436 constructs
induced pluripotent stem cells (IPSCs), the nucleotide modi- to increase viral titers and HSC transduction [32].
fication of specific target sites, and the genetic modification
of cells in vivo. 2.1.1 β‑Thalassemia

Mice reproducing the cellular defects observed in humans


2 Preclinical Studies with β-thalassemia intermedia (classified as non-transfu-
sion–dependent thalassemia [NTDT]) [54, 55], were used
2.1 Globin Gene Addition to study the efficacy of human β-globin–expressing vectors
to correct the disease. Long-term high-level human β-chain
Gene therapy requires the development of tools (1) mediat- expression was achieved with the LVVs TNS9 and HPV403
ing a long-lasting effect through gene integration into the [49, 52, 53]. The level of expression was high enough to
HSC genome, (2) transferring a sequence long enough, reduce the free α-chain imbalance, to restore hemoglobin
including a large distal enhancer element, for expression of levels, to improve red blood cell morphology, to correct
the globin gene to sustained and therapeutic levels in eryth- dyserythropoiesis, to prevent extramedullary hematopoiesis,
roid cells, and (3) minimizing the risk of insertional onco- and to correct iron overload. Incorporation of the cPPT/CTS
genesis. The combination of core elements of the β-globin element [56] into the HPV403 vector may have made it pos-
locus control region (LCR) made it possible to generate sible to obtain a higher titer and to transduce the cells more
sequences capable of promoting high levels of gene expres- efficiently. Indeed, the mean VCN per leukocyte was ≈ 3
sion in erythroid cells [40, 41], but their introduction into rather than ≈ 1 (copies per diploid genome) in peripheral
gamma-retroviral vectors (γ-RVVs) resulted in low vector leukocytes, and the human β-globin chain accounted for a
titers [42, 43], limited expression [43], position effect varie- mean of 32.4% of all β-globin chains with HPV403, rather
gation [44], and poor transduction of the mouse HSCs used than the 21.0% observed with TNS9. A number of similar
for transplantation [45]. Removal of the DNA structures LVVs have since been developed (Fig. 1), all capable of pro-
responsible for the low titers increased vector stability [46, ducing high levels of β-like globin chains in mouse models
K. Sii‑Felice et al.
Innovative Therapies for Hemoglobin Disorders

◂Fig. 1  Lentiviral vector expressing β-globin variants. a The β-globin γ-globin LVV-transduced cells, 30–50% HbF was sufficient
locus (NG_000007.3), including DNAse-1 hypersensitive sites 1, to protect mice against organ damage [62]. The use of a vec-
2, 3, 4, and 5 (HS1, HS2, HS3, HS4, and HS5), embryonic (HBE),
fetal (HBG1 and HBG2), and adult (HBD and HBB) globin genes,
tor encoding the triple anti-sickling variant βAS3 in a mouse
and a pseudogene (HBBP1). b The HBG1 and HBB genes including model resembling BERK mice [63] led to the correction
exons (E1, E2, and E3), introns, untranslated regions, and regula- of red blood cell morphology, a decrease in the frequency
tory sequences used to construct the lentiviral vectors shown in (c). c of vaso-occlusive events, and an improvement in splenic,
Structure of the lentiviral vectors, in their plasmid form, described in
this review. Each group, separated by a dotted line, consists of a pro-
liver, and kidney diseases in mice with ≈ 25% ­HbAS3. This
totype (top) and its derivatives. The sizes of the locus control region percentage is similar to that of HbF in SCD patients with
(LCR) HS and of the promoter, indicated for the prototypes, are the hereditary persistence of fetal hemoglobin (HPFH) and
same in the derived vectors. Amino-acid modifications (relative to mild disease [64]. As the anti-polymerization activities of
the open reading phase in the HBB gene) are indicated in red, and
the insulating elements are indicated by purple ellipses. U3, R, U5:
the βAS3 and βA-T87Q-globin subunits are similar to those of
HIV long terminal repeat (LTR), CMV, R, U3: CMV/HIV-hybrid 5′ the γ-globin chain, these results suggest that 25% ­HbAS3 or
LTR, ΔU3: 3′ self-inactivating (SIN) LTR. cPPT/CTS central poly- ­HbAT87Q would be sufficient for the successful treatment of
purine track/central termination sequence, Ψ packaging sequence, SCD patients with gene therapy, provided that expression is
RRE Rev-responsive element, WPRE Woodchuck hepatitis virus post-
transcriptional regulatory element, Δ and Δ* 374- and 593-bp dele-
evenly distributed among a majority of erythrocytes [65].
tions, respectively, within HBB intron 2, Δ** 720-bp deletion within
HBG1 intron 2, FB minimal CTCF-binding site (FII) of the 250-bp 2.1.3 Chromosomal Position Effect
core of the 1.2-kb chicken β-globin HS4 and the analogous region of
the human T-cell receptor δ/α BEAD-1 insulator, G and A regulatory
elements from the erythroid cell-specific murine GATA1 and human
These pioneering studies also showed that the elements used
Ank1 genes, respectively. HPV403 [52], HPV436 [32], HPV524 to control β-globin gene expression in LVVs were sensitive
[238], HPV569 [98], BB305 [74], TNS9 [49], AnkT9W [100], to silencing [32, 49, 52]. An analysis of individual clones
mLARβΔγV5 [57], V5m3 [62], V5m3-400 [116], BG [58], BG-I obtained upon the transduction of erythroid cells revealed
[68], ­SGbG [104], DL-βaS3 [64], CCL-βaS3-FB [93], CoreGA-AS3-
FB [99], UV-AS3-FB [99], GLOBE [59], GLOBE-AS3 [108],
considerable variation in the levels of β-globin gene expres-
GLOBE-AS3-FB [94] sion. High gene transfer efficacy and the integration of mul-
tiple copies resulted in the pancellular erythroid expression
of the human gene, whereas, in mice with a mean VCN of
of β-thalassemia and/or in primary human cells [57–59]. one per cell, heterocellular expression persisted [32, 52, 66].
Overall, the preclinical studies demonstrated that LVVs can Position-dependent expression was also observed in eryth-
be used to transfer complex DNA sequences, including the roid mouse cells transduced with mLARβΔγV5 and V5m3,
β-globin LCR, to achieve expression levels much higher than two γ-globin LVVs containing 3.2 kb of LCR elements [57,
those achieved with γ-RVVs, and they demonstrated the fea- 62]. Thus, despite the promising results obtained with LVVs
sibility of complete correction for β-thalassemia. containing extensive LCR sequences, the position effect
remained an issue, and the mean lentiviral VCN required
2.1.2 Sickle Cell Disease for consistent correction in human red blood cells remained
unclear at this stage of preclinical studies. As a means of
The βT87Q-globin chain is as potent as the γ-globin chain for protecting the β/γ-globin LVVs against the encroachment of
inhibiting HbS polymerization in vitro, and the oxygen affin- adjacent inactive chromatin, insulators with heterochromatin
­ BAT87Q is within the range observed for HbA [32].
ity of H barrier activity were introduced into their 3′ long terminal
The first LVV tested in mouse models of SCD thus encoded repeat (LTR) [67]. Insertion of the 1.2 kb chromatin insula-
the βT87Q-globin protein (HPV436, Fig. 1). High titers and a tor of the chicken β-globin 5′ DNAse I-hypersensitive site 4
high VCN (mean value of ≈ 3) per leukocyte were obtained (cHS4) into the BG LVV 3′LTR (BG-I) doubled the prob-
upon HSC transduction and transplantation, resulting in pan- ability of β-globin gene expression in erythroid cell lines
cellular erythroid expression. Two models were used: the and mouse red blood cells [68]. However, it also resulted
SAD [60] and Berkeley (BERK) [61] mouse models. The in a tenfold decrease in vector titer [69]. Two copies of the
BERK mice closely model HbSβ0-thal patients, a popula- cHS4 250-bp core element were also inserted into the BG
tion of severely affected individuals also likely to benefit and HPV569 LVVs, but characterization of these vectors
from gene therapy. Inhibition of sickling at low oxygen pres- showed that the tandem repeats recombined at a high fre-
sure and the disappearance of irreversibly sickled cells were quency [69, 70], resulting in the loss of one of the two cores.
observed in mice receiving marrow cells transduced with Unfortunately, one copy was reported to have only mild
the HPV436 vector [32]. In red blood cells harvested from barrier activity in several studies [71–73]. Flanking with
the recipient mice, HbS polymerization was delayed in vitro a 650-bp fragment combining the 5′ 250-bp core and the
and was similar to that measured in cells from asymptomatic 3′ 400-bp sequences of the 1.2-kb cHS4 insulator resulted
AS heterozygous human subjects. In BERK mice receiving in the maintenance of barrier activity in red blood cells of
K. Sii‑Felice et al.

thalassemic mice [72], with only a two- to threefold loss of The enhancer-blocking activity of insulators has been
viral titer [69, 72]. This may, nevertheless, pose a significant shown to be context dependent [95], variable according to
problem with regard to the therapeutic potential of the vec- the promoters activated [96], and only partially effective
tor, as decreases in titer are known to be accompanied by an in vitro and in vivo [73]. LVV-mediated genotoxicity may
additional decrease in the infectivity of hematopoietic cells also occur through mechanisms other than simple enhancer/
[36, 73, 74]. Thus, the small effect of barrier insulators on promoter interaction, and may interfere with the splicing of
the probability of cells expressing the transgene at a constant primary transcripts [97, 98]. For these reasons, and those
level is counterbalanced by the costs, in terms of titer and discussed above, few of the viral vectors developed to date
transduction efficiency [75, 76]. contain these elements. It was decided to remove the insu-
lator sequence from the HPV569 LVV [74], after this vec-
tor had been used clinically [98], but to retain it in Lenti/
2.1.4 Insertional Mutagenesis βAS3-FB LVVs [93, 99].

LVVs carrying active LTR enhancer sequences efficiently 2.1.5 Comparison of Globin LVVs
activate oncogenes in vivo [77, 78] and disturb cellular
gene regulation within a large window around the insertion A strict inter-vector comparison between studies, based on
site [79], suggesting that they are not inherently safer than preclinical data and the proportion of transgenic hemoglobin
γ-RVVs. Nevertheless, in side-by-side comparisons, the normalized according to VCN, would not be appropriate,
genotoxic potential of LVVs is lower than that of γ-RVVs for several reasons: (1) the correlation between transgenic
[77, 80], due to the preferential targeting of promoter regions hemoglobin levels and VCN is not linear; (2) red blood
by γ-RVVs and of intragenic regions by LVVs [81]. Self- cells with the highest levels of therapeutic globin expres-
inactivating (SIN) vectors are safer [77], especially when sion appear in the bloodstream due to in vivo selection; (3)
gene promoters with low enhancer promiscuity are used to in vivo selection depends on the levels of dyserythropoie-
control transgene expression [82]. Due to their restricted sis and/or red blood cell survival; and (4) the competition
expression, tissue-specific promoters can decrease the risk between endogenous and exogenous β-like globin genes and
of oncogene activation in non-target cells [83]. Thus, in the proteins, in terms of expression, translation, dimerization
LVVs currently being developed for gene therapy, strong with α-globin, and/or protein stability, remains unknown
LTR enhancers are deleted, and internal cellular gene pro- and may depend on the mouse strain. In vitro, the human
moters are preferred over strong ubiquitous transcriptional cell genotype and the culture system may also make a dif-
elements for the control of therapeutic gene expression ference. Therefore, the only way to compare vector efficien-
[84]. The β-globin LCR is a long-range enhancer that can cies is to test the vectors in parallel. Studies of this type
generate high levels of erythroid-specific expression for have been performed with vectors containing identical LCR
cis-linked genes [85], affects chromatin structure over a elements, making it possible to improve vector prototypes
distance of 200 kb [86], and has an active chromatin con- [62, 74, 93, 100], but rarely with LVVs developed by dif-
figuration before commitment to the erythroid lineage [87, ferent research groups. One exception is the CCL-βAS3-FB
88]. The ≈ 3-kb LCR enhancer elements present in the glo- vector from Kohn’s team, which has been compared with
bin LVVs mLARβΔγV5 and BG have been shown to dys- both V5m3-400 [101] and GLOBE-AS3 [94], these three
regulate cellular genes within at least 600 kb around the LVVs being derived from three different prototype vectors
site of insertion in primary erythroid cells [89], and have (Fig. 1). The V5m3-400 and CCL-βAS3-FB LVVs yielded
a residual capacity to immortalize primary murine hemat- equivalent infectious titers and transgene expression levels
opoietic cells in vitro [80]. As a means of shielding the [101], whereas GLOBE-AS3 had a titer about four times
genome from the internal enhancer, and further minimizing higher than that of CCL-βAS3-FB [102]. Furthermore, at a
genotoxic impact, the addition of insulator elements, such similar multiplicity of infection (MOI), the CCL-βAS3-FB
as the cHS4 insulator element, has been proposed [90, 91]. had a 2.4-fold lower transduction efficiency in hematopoi-
Following its introduction into the 3′LTR of BG LVV, the etic ­CD34+ cells than GLOBE-AS3. However, these weak-
1.2-kb element, or its 650-bp derivative, has been shown to nesses were counterbalanced by a higher level of transgene
decrease genotoxicity slightly in vitro [80]. The FII/BEAD-1 expression with CCL-βAS3-FB than GLOBE-AS3 at equal
(FB) composite 77-bp fragment has been shown to decrease VCN [103]. The two vectors may therefore be considered
the genotoxic potential of strong viral enhancer elements to work equally well on a per infectious particle per ­CD34+
[92], with a minimal effect on β-globin LVV titers [93, 94]. cell basis. The CCL-βAS3-FB makes it possible to obtain
FB is, thus, a potentially interesting insulator, although its therapeutic gene expression with a lower VCN, making the
blocking activity in the globin LVVs into which it has been vector safer. Conversely, the GLOBE-AS3 vector is pro-
inserted has not been formally evaluated. duced at a higher titer and transduces primary C ­ D34+ cells
Innovative Therapies for Hemoglobin Disorders

more efficiently. The production cost is therefore likely to be genome sequencing and population genetics in the twenty-
lower. Recent refinement of the LCR core HS elements from first century, together with the availability of high-quality
the CCL-βAS3-FB vector have made it possible to reduce phenotypic information, has led to the identification of
the total size of the vector by 2.1 kb and to increase its titer several crucial factors involved in the fetal-to-adult hemo-
and transduction rate, but at the expense of decreasing thera- globin switch [109–113]. The transcription factor B cell
peutic gene expression in erythroid cells [99], bringing the CLL/lymphoma 11A (BCL11A), a potent developmental
properties of this new vector (CoreGA-AS3-FB) closer to stage-specific repressor of γ-globin, emerged from genome-
those of GLOBE-AS3. Overall, a number of LVVs, encod- wide association studies (GWAS) of individuals with high
ing βA, βA-T87Q, βAS3, or γ-globin chains and developed baseline levels of HbF. BCL11A is unusual among the fac-
and tested in vitro and in animal models [70, 74, 93, 102, tors initially discovered in that it seems to have no crucial
104–108] —HPV569, TNS9.3.55, BB305, GLOBE, ­sGbG, function in erythrocytes other than regulation of the hemo-
CCL-βAS3-FB, and GLOBE-AS3 (Fig. 1)—have been or globin switch, whereas it has major indispensable functions
are being evaluated in clinical trials in humans (Fig. 2). in other hematopoietic lineages, such as B-cell progenitors
[114, 115]. The first therapeutic approach to harness this
2.2 Artificial MicroRNA Targeting a Suppressor knowledge by targeting and downregulating BCL11A to
of Fetal Hemoglobin induce HbF production was developed shortly after these
discoveries were made. A LV gene therapy vector express-
An alter native strategy for t he treatment of ing a small hairpin RNA (shRNA) for the downregulation
β-hemoglobinopathies is based on reactivation of the endog- of BCL11A was initially generated. It yielded high levels
enous fetal γ-globin gene, effectively reversing the fetal-to- of HbF induction in vitro, but transplantation experiments
adult hemoglobin switch. Given that the hemoglobin switch in mice were unsuccessful due to the near-complete loss
functions like a rheostat, activation of the fetal globin gene of HSCs carrying the modified gene [116–118]. Multiple
leads to concomitant downregulation of the disease-causing modifications were introduced to address several shortcom-
βS chain. This basic strategy is decades old, but the lack ings of this initial construct. BCL11A knockdown-related
of knowledge concerning the cellular factors governing the toxicity in non-erythroid lineages, specifically in HSCs and
switch precluded the development of therapeutic approaches the B-cell lineage, was prevented by replacing the simple
based on this mechanism. The rise of high-throughput ubiquitously active polymerase III–driven shRNA with a

Fig. 2  Clinical trials of gene therapy for β-hemoglobinopathies. homology-directed repair, MSKCC Memorial Sloan Kettering Cancer
APHP Assistance Publique – Hôpitaux de Paris, CRISPR/Cas9 clus- Center, NHEJ non-homologous end joining, SCD sickle cell disease,
tered regularly interspaced short palindromic repeats/CRISPR-associ- THAL β-thalassemia, UCLA University of California, Los Angeles,
ated protein 9, GmbH Gesellschaft mit beschränkter Haftung, HDR ZFN zinc finger nuclease
K. Sii‑Felice et al.

microRNA mimic ­(shRNAmiR) under the control of eryth- One study reported a very high frequency of Glu6Val cor-
roid cell-specific polymerase II–driven transcriptional regu- rection, ranging from 30% to 70% in C ­ D34+ cells, following
latory elements identical to those used for LVV-mediated electroporation with the Cas9/guide RNA (gRNA) complex
β-like globin expression (HS2/HS3/β-globin promoter). The followed by the delivery of a homologous DNA template
resulting construct downregulated BCL11A selectively in by adeno-associated virus serotype 6 (AAV6) [131, 133].
erythroid cells, efficiently induced HbF, and largely attenu- However, the level of correction in long-term HSCs was
ated the hematologic effects of SCD [117, 119]. This vector lower than that in C ­ D34+ cell pools, by a factor of 5–20
entered a phase I clinical study in 2018 at Boston Children’s [131], confirming that HSCs are more resistant to HDR than
Hospital, and highly promising initial results were presented committed progenitor cells [128, 131, 134].
at the annual American Society of Hematology (ASH) meet- A single editing system is required for SCD, but a num-
ing: therapeutic efficacy exceeded that reported for other ber of specific nuclease/gRNA complexes must be designed
approaches at similar VCNs [120] (see clinical trial results and optimized to deal with the diversity of point mutations
below). and deletions in β-thalassemia. The effort required may be
Combination of ­shRNAmiR-mediated BCL11A knock- too great for clinical applications aimed to reverse point
down with the recombinant expression of anti-sickling glo- mutations. The targeted integration of a functional copy
bin variants, such as βAS3 or βT87Q, encoded by the same of the β-globin gene into the β-globin locus has, there-
vector is an appealing idea. Hairpin shRNAs and globin fore, been studied as a universal therapeutic approach for
transgenes can, potentially, be expressed under the con- β-hemoglobinopathies. Dever et al. reported a gene addi-
trol of the same promoter, and the presence of intervening tion system, in which the therapeutic HBB complementary
sequences 1 and 2 in globin vectors facilitates the position- DNA (cDNA) was inserted within exon 1 of a mutated HBB
ing of the hairpin sequence within the introns of the globin allele, such that expression was controlled by the endog-
transgene. The resulting vector structure and the molecular enous β-globin LCR and promoter [131]. The mean target-
processing steps of the mRNA are complex, but theoreti- ing efficiency was 11% in adult C ­ D34+ cells following the
cally, the intronic ­shRNAmiR can be excised and processed AAV6-mediated delivery of a donor cassette carrying the
by the microprocessor complex in the nucleus with no nega- functional HBB cDNA plus a selection gene and a gRNA/
tive effect on globin transgene expression. This approach Cas9 RNP complex targeting the HBB gene. The initial cor-
will probably prove to be the most effective yet, in terms rection efficiency was not high, but an enrichment in cells
of therapeutic efficacy per integrated vector copy, but the with edited genes of up to ≈ 90% could be obtained in long-
complex structure and size of the vector may limit the gen- term HSCs, with the selection marker [131]. This system has
eration of high viral titers, as reported for LVVs harboring the potential to enrich HSC preparations in cells with edited
large transcriptional control segments [99]. genes, but it remains to be determined whether engraftment
levels will be sufficiently high to sustain hematopoietic
2.3 Gene Editing reconstitution permanently in humans.
The coinheritance of HPHF with β-globin gene mutations
Recent advances in genome editing technologies, based on is associated with mild clinical symptoms of β-hemoglobin
clustered regularly interspaced short palindromic repeats disorders. Therefore, the introduction of a natural HPFH
(CRISPR) RNA-guided nucleases (CRISPR/Cas) or cus- mutation by HDR, mimicking the HPFH phenotype, is
tomized engineered zinc-finger nucleases (ZFNs) or tran- being considered as an alternative genome editing option
scription activator-like effector nucleases (TALENs), have applicable to both SCD and β-thalassemia. Several point
made it possible to modify the genome in a targeted man- mutations and small deletions in the Aγ- or Gγ-globin gene
ner in many cell types and organisms. The nucleases intro- promoters resulting in an increase in HbF levels to more than
duce double-strand breaks (DSBs) at a specific site, which 30% in adults have been reported [135–147]. Introduction of
are then repaired via two competitive cellular repair path- the naturally occurring − 175T > C HPFH mutation into the
A
ways: non-homologous end joining (NHEJ) and homology- γ-globin promoter has been shown to increase HbF produc-
directed repair (HDR) [121]. tion significantly in the K562 erythroid cell line [148]. The
Theoretically, the correction of disease-causing mutations introduction of point mutations 115 and 200 bp upstream
through HDR with a designer nuclease and a DNA tem- from the transcription start site, within binding sites for the
plate is the ideal approach, as it can restore normal β-globin fetal globin gene repressors BCL11A or zinc-finger—and
function while ensuring physiological and high levels of BTB-domain–containing 7A (ZBTB7A, also known as
β-globin gene expression [122–132]. This strategy avoids LRF), respectively, led to high levels of HbF production in
the risk of insertional mutagenesis associated with LVV- the HUDEP-2 immortalized erythroid progenitor cell line
mediated gene transfer, and is particularly suitable for the [149]. A recent study revealed that introducing the natu-
Glu6Val mutation affecting millions of patients with SCD. ral -113A > G HPFH mutation into the Aγ-globin promoter
Innovative Therapies for Hemoglobin Disorders

increased HbF levels in HUDEP-2 cells to potentially thera- erythroid cell-specific enhancer in the BCL11A gene. As
peutically relevant levels. The increase in HbF production in the ­shRNAmiR-mediated knockdown approach targeting
did not disrupt BCL11A binding, but probably resulted from BCL11A, the resulting downregulation of the BCL11A pro-
the de novo creation of a binding site for the GATA-1 master tein is limited to the erythroid lineage and spares all other
erythroid transcription activator [150]. Further investigations cell types, thereby preventing undesirable side effects.
of the efficacy of these approaches are required, in clinically Multiple critical enhancer sites have been identified within
relevant cell types. Furthermore, as NHEJ predominates over intron 2 of BCL11A, initially by GWAS, with subsequent
HDR in quiescent HSCs, a careful evaluation of correction characterization at single-nucleotide resolution with high-
will be required in these cells. The various editing systems density functional CRISPR screens [160, 161]. A specific
using nucleases as well as the methods for co-delivery of the GATA-1 binding site at the +58-kb position has been shown
DNA repair template should be compared in terms of their to be responsible for the erythroid cell-specific upregula-
ability to trigger HDR, the frequency of which will need to tion of BCL11A, and even single-base pair deletions within
exceed that of the more error-prone NHEJ [133]. The iden- this binding site strongly decrease BCL11A expression and
tification of compounds that enhance HDR and/or inhibit induce γ-globin production as effectively as larger dele-
NHEJ will be a great challenge in attempts to improve gene tions, demonstrating that most of the function is contained
editing via HDR in HSCs and to bring HDR-based thera- within this small functional DNA element [162]. CRISPR/
peutic applications closer to clinical practice. An additional Cas9-mediated disruption of this site in human HSCs, with
challenge for clinical development will be the expansion of ribonucleoprotein (RNP) delivery, resulted in very high
HSCs with edited genes if the selection of edited cells is editing rates (about 90%) and an induction of fetal globin
required due to limited HDR efficacy. This limited efficacy to levels of ~ 50 to 60% total hemoglobin, well above the
of HDR in HSCs can be accounted for by the limitation of range required for therapeutic efficacy, and similar to those
the repair mechanism to the S/G2 phases [151]. By using reported for shRNA-mediated knockdown in in vitro assays
a modified version of Cas9 with lower levels of nuclease [117].
activity in the G1 phase of the cell cycle, in which HDR can- Several gene editing approaches for treating hemoglobi-
not occur, and transiently increasing the proportion of cells nopathies are already being explored by biotech and phar-
in the phases in which HDR preferentially occurs (S/G2), maceutical companies: Sangamo Therapeutics is assessing
Lomova and coworkers managed to increase the HDR/NHEJ ST-400, Bioverativ BIVV003, Vertex/CRISPR Therapeutics
ratio fourfold over the control condition [152]. It remains CTX001, Editas Medicine EDIT-301, Intellia Therapeutics
to be determined whether this treatment does not affect the OTQ923, and Shanghai Bioray Laboratory is assessing
ability of cells to reconstitute human hematopoiesis. autologous HSCs in which γ-globin has been reactivated
The introduction of insertions and/or deletions (indels) and autologous HSCs in which β-globin expression has been
through mechanisms involving DSB repair by the NHEJ restored. The base editing strategy (see Sect. 4) is being
pathway can be used as an indirect approach for creating tested by Beam Therapeutics for SCD and TDT. Induced
naturally occurring HPFH deletions in HSPCs and induc- hematopoietic stem cell (iHSC) technology is being devel-
ing HbF, to improve SCD and β-thalassemia phenotypes oped by Allife Medicine.
[153–156]. This approach yielded high level γ-globin
expression in human primary cells grown in vitro [157]. The
transplantation of a CRISPR-Cas9–edited HSC-enriched 3 Clinical Studies
population with a targeted deletion of the BCL11A bind-
ing site in the Aγ-globin gene promoter in nonhuman pri- Several gene therapy programs for β-hemoglobin disorders
mates was recently shown to lead to sustained HbF reac- are already implemented in clinical practice. A brief descrip-
tivation in 6–18% of the red blood cells of these animals tion of each clinical trial is available from ClinicalTrials.
[158]. However, it remains to be seen whether the survival gov, and data from some of the studies have been published
advantage conferred on corrected erythroid cells and the in peer-reviewed journals. A brief overview of the clini-
amount of HbF produced per cell are sufficient to improve cal applications is provided in Fig. 2. The approach used
the phenotype of SCD or TDT patients. Indeed, due to the involves the autologous transplantation of HSCs geneti-
strong homology and the proximity of the two γ-globin cally modified ex  vivo with LVVs carrying a transgene
genes, HBG1 and HBG2, the simultaneous DSBs of both under the control of an erythroid cell-specific promoter and
promoters by nucleases can result in the deletion of HBG2 enhancer. The transgene can encode the wild-type β-globin
[159] and the HBG intergenic region, and thus reduce the or mutated versions to provide particular properties, such as
overall expression of therapeutic HbF in edited cells. The anti-sickling (βA-T87Q, βAS3). Alternatively, the transgene may
most advanced pipeline currently available for NHEJ-medi- encode a γ-globin chain or be replaced by a shRNA targeting
ated disruption to treat β-hemoglobinopathies targets the a repressor of γ-globin (e.g., s­ hRNAmiR BCL11A). A phase
K. Sii‑Felice et al.

I clinical trial for the treatment of TDT with autologous Another LVV encoding the normal human β-globin
­CD34+ cells transduced with the TNS9.3.55 LVV encod- (GLOBE) was developed by the group of Giualina Ferrari
ing the normal human β-globin was initiated in New York in Italy. A phase I/II clinical trial sponsored by IRCCS San
City, at Memorial Sloan Kettering Cancer Center, in July Raffaele was initiated in May 2015 (NCT02453477). The
2012, by the groups of Michel Sadelain and Isabelle Riviere design of this study involved the successive enrollment of
(NCT01639690). The first patient was a 23-year-old woman three groups of TDT patients: three adults (≥ 18 years old),
who underwent peripheral blood stem cell mobilization with three older children (8–17  years old), and four younger
filgrastim. Her ­CD34+ cells were transduced with the LVV, children (under the age of 8 years). Intraosseous autologous
and the mean VCN in bulk ­CD34+ cells was 0.39 copies per stem cell transplantation was performed. Interim data were
cell. The non-myeloablative conditioning was well tolerated, reported in February 2019, after the treatment of three adults
and rapid engraftment was observed [163]. Three additional and six children [169]. Rapid hematopoietic recovery with
subjects were treated and displayed stable engraftment. The multilineage polyclonal engraftment was achieved. The
treatment was well tolerated, but the patients did not achieve mean VCN ranged from 0.1 to 1.97 in erythroid precursors
transfusion independence [164]. 1 year post-treatment. No clonal dominance was observed.
T h e f i r st s u c c e s s f u l ge n e t h e ra py fo r Transfusion requirement decreased in adults, and three of the
β-hemoglobinopathies in France was performed with the four younger children treated were transfusion independent
vector developed by the group of Philippe Leboulch in the at the time of the publication. The NCT03275051 long-term
LG001 clinical trial [98]. An adult patient with severe βE/ follow-up study is sponsored by Orchard Therapeutics.
β0 thalassemia became transfusion independent after full A trial of an LVV encoding a γ-globin chain, sponsored
myeloablative conditioning with busulfan and the autolo- by Aruvant Sciences GmbH, was initiated in July 2014. For
gous transplantation of HSCs genetically modified with eligibility, subjects had to have an HbSβ0-thal, HbSβ+-thal,
the SIN LVV HPV569 encoding the βA-T87Q-globin gene. or HBSS SCD genotype. Bone marrow or mobilized ­CD34+
The 2 × 250-bp insulator was unstable, and one of the two cells were transduced ex vivo and infused into the patient
cores was lost in a large proportion of integrated vectors. after reduced-intensity chemotherapy conditioning with sin-
Partial clonal dominance, which subsided over time and gle-dose melphalan. The treatment of two patients with the
subsequently declined, was observed upon vector integra- HbSβ0-thal genotype was reported in 2018. The first patient
tion into the HMGA2 gene [165]. An analysis of the inte- received 1 × 106 ­CD34+ cells/kg with a VCN of 0.22, and
grated vector revealed transcriptional activation of the the second patient was treated with 6.9 × 106 ­CD34+ cells/kg
HMGA2 gene in erythroid cells and abnormal splicing with a VCN of 0.46 [170]. The time to neutrophil engraft-
of the endogenous HMGA2 RNA. Retrospective analy- ment was 9 days for the first patient and 7 days for the sec-
sis showed that blood hemoglobin levels remained stable, ond patient, and reconstitution was highly polyclonal. Only
at about 8.5 g/dL, for more than 8 years [166]. A next- a few results are available: a proportion of 20% was reported
generation LentiGlobin vector (BB305) was designed to for the modified HbF (produced from the transgene) in the
remove the unstable insulator sequence, and the titer was blood of the first patient (2.1 g/dL modified HbF in 10.6 g/
increased with a CMV hybrid LTR [74]. These changes dL total hemoglobin) 1 year after treatment.
increased vector titers (three- to fourfold) and trans- Two clinical trials are being conducted with βAS3-globin
duction efficacy (two- to threefold). Patients in France LVVs (NCT02247843 [171] and NCT03964792), but no
were treated with this new LVV (NCT02151526), and data have been published yet.
clinical trials were extended to other European countries A LVV carrying an anti-BCL11A s­ hRNAmiR (micro-
(NCT02906202), the USA (NCT02140554, NCT03207009), RNA-adapted shRNA) expressed under the control of the
Australia (NCT01745120), and Thailand (NCT01745120, β-globin promoter and LCR elements (HS2 and HS3) has
NCT02906202). The first improvement in SCD phenotype been developed by the group of David Williams in Bos-
following gene therapy was reported in 2017 for the first ton. This vector is designed to induce the expression of
patient treated with BB305 [167]. The level of therapeutic endogenous γ-globin genes, by downregulating the tran-
anti-sickling β-globin remained high (approximately 50% scriptional repressor BCL11A. In preclinical studies,
of β-like globin chains), with correction of the biological ­C D34 + cells from healthy donors or donors with SCD
hallmarks of the disease. In 2018, the potential of BB305 to were transduced with GMP-grade LVV. A high VCN (> 5)
treat TDT was demonstrated in 22 patients [168]. Autolo- and high levels of gene marking (> 80%) resulted in a
gous transplantation with ­CD34+ cells transduced with the three- to fivefold induction of HbF with no effect on cell
BB305 LVV eliminated or reduced the need for red blood growth or differentiation [119]. A clinical trial was initi-
cell transfusions. Patients from the various clinical trials will ated in 2018, on the infusion of autologous C ­ D34+ cells
be monitored for 15 years, in a long-term follow-up protocol transduced with the LVV after near-myeloablative condi-
(NCT02633943). tioning with busulfan. Eight patients, aged 7–36 years, all
Innovative Therapies for Hemoglobin Disorders

with an HbSS genotype, were enrolled in the trial [172]. 4 Alternative Directions
Initial results for the first five patients, with a maximum
follow-up of 18 months, were reported at the ASH meet- 4.1 Induced Pluripotent Stem Cells
ing in 2019 [172]. Rapid engraftment was observed in all
patients; there were no severe adverse events related to Gene therapy protocols would be greatly improved if an
the vector, and HbF levels in this patient cohort ranged expansion phase were possible for the corrected cells
from 23% to 43% of total hemoglobin. The fraction before their transplantation, but the challenge of expand-
of F cells was consistently between 65% and 75% and ing HSC populations ex vivo or, at least, maintaining their
remained stable over time at VCNs of 0.4–1.6 per cell. stemness properties during their correction, has slowed
Clinical trials of gene editing for TDT and SCD are progress. IPSCs have unlimited self-renewal and prolif-
also underway and will provide data relating to the eration capacities, they can be genetically modified and
safety and efficacy of this non-viral strategy. ST-400 is selected, and they can differentiate into various mature cell
an ex vivo gene-edited cell therapy for TDT developed types [176]. They therefore represent a potential alterna-
by Sangamo Therapeutics. This autologous cell therapy tive source of cells for regenerative medicine.
involves the editing of the BCL11A gene in HSCs from The first proof-of-concept for the treatment of SCD
the patient with ZFN technology (NCT03432364). Eryth- with IPSCs was obtained in 2007 [177]. IPSC genera-
roid colony genotyping showed biallelic modification of tion and gene electroporation were combined, and the
the BCL11A erythroid enhancer in > 50% of progenitors, SCD mouse model was corrected by the transplantation
resulting in an increase in γ-globin mRNA and protein of hematopoietic progenitors derived from the corrected
levels to values more than four times those in the controls IPSCs. Progress has since been made toward the targeting
[173]. No patient follow-up data have been published yet. of point mutations in IPSCs derived from patients suffer-
Four clinical trials involving the use of the CRISPR/Cas9 ing from SCD, with ZFNs [178], TALENs [179], CRISPR/
system to reactivate γ-globin through a similar strategy Cas [127, 180, 181]. However, none of these IPSCs were
are still underway. CRISPR Therapeutics initiated clini- evaluated for clinical potential after transplantation. IPSCs
cal trials in 2018 in Europe (NCT03655678) for TDT and from β-thalassemia patients have been successfully gener-
in the USA (NCT03745287) for SCD. Shanghai Bioray ated and corrected by homologous recombination [182]
Laboratory is sponsoring two clinical trials in China. The or gene addition [183, 184], and the functionality of these
NCT04205435 trial is a study including up to 12 par- cells has been demonstrated by in vivo transplantation in
ticipants with TDT. Its objective is to evaluate the safety immunodeficient mice. However, these studies reported
and efficacy of treatment with autologous HSCs in which very low rates of hematopoietic reconstitution. More
γ-globin has been reactivated with the CRISPR/Cas9 gene recently, integration-free IPSCs have been obtained from
editing system. The NCT04211480 trial is designed to patients with β-thalassemia using non-integrative vectors,
determine the safety and efficacy of autologous HSCs and HBB mutations have been corrected by a TALEN-
in which β-globin expression has been restored in TDT based approach [124], ZFN-mediated gene therapy [185],
patients with the IVS-654 mutation. β-Globin expression or CRISPR/Cas9 technology [122, 186–188]. In these
will be restored in autologous HSCs with the CRISPR/ studies, evaluations of the potential therapeutic effect
Cas9 gene editing system. No results have been reported of the correction were mostly based on the capacity of
yet. corrected IPSCs to differentiate in vitro but not in vivo.
One clinical trial is using the CRISPR/Cas9 system to The transplantation into immunodeficient mice of human-
correct the genome of iHSCs (see Sect.  4.1, third par- derived and CRISPR/Cas9-corrected IPSCs led to the pro-
agraph). It is sponsored by Allife Medical Science and duction of therapeutic hemoglobin, but only low levels of
Technology (NCT03728322) and was initiated in January hematopoietic reconstitution [187].
2019 [174]. No results have been published yet. IPSC reprogramming is based on the modification
The first gene therapy product reached the mar- of somatic cells by transcription factors. Great progress
ket in June 2019 [175]. In a press release, bluebird bio has been made towards limiting the risk of tumors asso-
announced EU conditional marketing authorization for ciated with this process through the use of non-integra-
Zynteglo™ (autologous ­C D34 + cells encoding the β A- tive strategies [124, 189] or screening for safe harbor
T87Q
-globin gene) gene therapy for patients aged 12 years integration sites [183], but IPSC generation can lead to
or over with transfusion-dependent β-thalassemia and a genomic instability [185, 186]. The mutation status and
genotype other than β0/β0. The large number of clinical tumorigenic potential of these cells will therefore need to
trials underway or soon to be initiated is a source of hope be thoroughly tested before their use in clinical practice
that new gene therapy products will become available very [190]. The near-impossibility of generating large numbers
soon.
K. Sii‑Felice et al.

of transplantable HSCs from IPSCs greatly hampers the DNA. The UGI domain prevents the action of the base exci-
use of these cells in clinical practice. Daley’s group has sion repair (BER) pathway, thereby preventing the reversion
worked on re-specification, a strategy for reverting IPSC- of the obtained U–G mismatch to C–G. C ­ as9D10A cleaves
derived hematopoiesis-committed progenitors to a more the complementary strand, thereby guiding the mismatch
immature state for the generation of iHSCs, for expan- repair (MMR) system to the unedited nicked strand, using
sion in vitro and with multipotent engraftment potential information from the edited template. Replication or repair
in vivo. Depending on the combination of transcription mechanisms then mediate the conversion of uridine into thy-
factors used, the iHSCs support short- [191] or long-term midine, and permanently convert the C–G pair into a T–A
[192] hematopoietic reconstitution. Respecifying IPSCs, pair [200–202] (Fig. 3b). Gaudelli and coworkers created an
to generate engraftable iHSCs, is a promising approach adenosine base editor (ABE), in which the cytosine deami-
to the testing of gene therapy strategies through assess- nase is replaced by an engineered enzyme that can transform
ments of the correction of human cells in in vivo models of adenine into inosine, facilitating the highly efficient conver-
human diseases, but the generation of functional and safe sion of A–T to G–C pairs in human cells [199]. These tar-
HSC populations that can be expanded to the clinical scale geting strategies are promising for the treatment of diseases
for engraftment in patients with no loss of differentiation caused by single-nucleotide polymorphisms. However, the
capacity or safety may still be a long way off. use of these base editors induces indels, albeit at a lower
frequency than CRISPR/Cas9 associated with DNA donor
4.2 Base and Prime Editing template [199, 201]. In mouse embryos, unexpected conver-
sions or indel mutations are frequently observed at the target
Engineered genome-editing nucleases create DSBs that are site of CBE [203]. Recent findings have shown that off-target
then repaired by one of the two main repair mechanisms single-nucleotide variants (SNVs) are limited but present in
of cellular DNA: NHEJ or HDR. At the target site, in the cell lines [204], mouse embryos [205], and rice cells [206]
presence of the correcting homologous DNA template, in experiments using CBE [205, 206], and ABEs [204]. Fur-
HDR competes with NHEJ during the resolution of DSBs thermore, thousands of SNVs have been identified in RNA
(Fig. 3a). This is particularly problematic in non-dividing transcripts. They have been detected in oncogenes and tumor
cells, such as HSCs, because HDR is mostly restricted to suppressor genes, raising concerns about the oncogenic risk
the S/G2 phases of the cell cycle [193]. As a result, gene associated with the use of base editors [207–209]. CBE var-
correction rates are low in primary HSCs. DSBs can, thus, iants involving alternative and engineered cytidine deami-
generate small insertions, deletions or point mutations [184], nases, developed to increase the precision and efficiency of
large deletions [194], and chromosomal truncations [195], at on-target editing and to reduce off-target mutations, have
the target site and at more distant locations. These undesired been tested for the correction of a mutation frequently found
on- and off-target effects constitute a major challenge to the in the β-globin promoter of Southeast Asian patients with
efficient and safe use of CRISPR/Cas systems for correcting β-thalassemia. They efficiently corrected the targeted muta-
genetic diseases in clinical practice. In this context, attempts tion, but unwanted on- and off-target SNVs were nevertheless
to circumvent DSBs during gene editing have resulted in the detectable in human cells [210, 211]. Highly efficient CBE-
recent development of alternative approaches that do not rely mediated editing (> 90% in some cases) has been achieved
on HDR (discussed in [196–198]). in the erythroid cell-specific enhancer of the BCL11A gene
and a β-thalassemia-associated promoter mutation in the
4.2.1 Base Editing HBB gene in human HSCs. This editing was associated with
the induction of therapeutic levels of HbF (30–40% of total
Engineered base editors can convert one base pair into hemoglobin) in unselected cells in vitro and in humanized
another at a target site without generating DNA DSBs, result- mice in vivo, with the detection of minimal off-target edit-
ing in lower rates of indel mutations than Cas9-based homol- ing. These results highlight the rapid progress made and the
ogous-recombination (HR) methods at both on- and off-target therapeutic potential of this technology [162, 212].
sites [199–202]. The cytosine base editor (CBE) converts
the C–G pair into a T–A pair. It has four components: a 4.2.2 Prime Editing
gRNA, a catalytically impaired C ­ as9D10A mutant, a cytidine
deaminase domain from apolipoprotein B editing complex A number of base conversions cannot be performed with
1 (APOBEC1) or activation-induced deaminase (AID), and base editors, including the T–A to A–T conversion required
a uracil glycosylase inhibitor (UGI). The gRNA-Cas9 RNP to correct SCD. The non–DSB-mediated prime editing strat-
complex binds to its complementary DNA, and the cytidine egy can generate insertions, deletions, and the 12 possible
deaminase domain then converts cytidine residues into uri- base-to-base conversions [213]. The prime editing complex
dines within a window of five nucleotides around the target has two components: the prime editor (PE), which consists
Innovative Therapies for Hemoglobin Disorders

Fig. 3  CRISPR-Cas9–mediated genome editing. a DNA double- tary strand upstream from the PAM sequence and mismatch repair
strand break (DSB) repair. Guide RNA (gRNA) directs the Cas9- systems use the modified U base to correct the nicked strand. After
RNA complex to its genomic target site and Cas9 breaks DNA three replication and repair, the G–C is thus converted into A–T. c Prime
nucleotides upstream from the protospacer adjacent motif (PAM). editing allows all 12 base-to-base conversions. The prime editor
The DSB is repaired by homologous recombination when a donor (PE) complex consists of a reverse transcriptase (RT), a C ­ as9H840A
sequence with homology arms is provided, or by non-homologous enzyme, and a guide RNA (prime editing guide RNA [pegRNA]) that
end joining. Random insertions or deletions are then introduced. b binds to the non-PAM strand, and specifically cleaves the PAM-con-
Base-editing system converting one base pair into another. The cyto- taining strand. The primer binding site (PBS) of pegRNA binds to the
sine base editor (CBE) converts the C–G pair into T–A. The gRNA cleaved strand, and the RT elongates DNA from the nicked 3′ end,
binds its complementary DNA, and cytidine deaminase (C-deami- using its template. Flap equilibration allows the excision of one of the
nase) then converts cytidine into uridine, leading to a U–G mismatch. flaps by nucleases. When the 3′ DNA flap is not excised, ligation and
The uracil glycosylase inhibitor (UGI) domain prevents reversion via DNA repair result in gene editing. CRISPR clustered regularly inter-
the base excision repair pathway. C ­ as9D10A cleaves the complemen- spaced short palindromic repeats, Cas9 CRISPR-associated protein 9

of a reverse transcriptase (RT) fused to the C terminus of edited flap is excised by endonuclease, the sequence remains
the mutant ­Cas9H840A enzyme, and the prime editing guide unchanged, but if the 5′ non-edited flap is cleaved and the
RNA (pegRNA), which contains a binding site for the non- 3′ edited flap is ligated, edited DNA is obtained after DNA
complementary protospacer adjacent motif (PAM) strand repair (Fig. 3c). Any mutation desired can be generated, with
(PBS), an RT template providing the desired edit, and a an efficiency of ≈ 33%, with the most advanced prime editing
spacer sequence able to bind the non-PAM DNA target. The technology. This system provides a solution complementary
PAM strand is nicked by ­Cas9H840, and the PBS hybridizes to base editors, with strengths and weaknesses specific to
to the resulting 3′ end of the PAM strand. The RT elon- the targeted mutation. For example, the SCD mutation may
gates the nicked DNA strand from its 3′ end, using the RT be edited very efficiently in a cell line, but indel mutations
template with the desired edit. There is then competition are still observed at a high frequency, although lower than
between the 3′ edited and 5′ non-edited DNA flaps. If the 3′
K. Sii‑Felice et al.

that observed with Cas9-mediated HDR. Further studies are 4.3.2 Hybrid Adenovirus/Transposon System
required to evaluate this system and to improve its accuracy.
The access of HSCs to the bone marrow may be limited by
4.3 In Vivo Gene Therapy the rigidity of the stroma and the limited space available
for vector injection. Lieber’s group aimed to increase trans-
In vivo gene therapy approaches would constitute a very duction by developing an approach based on HSC mobi-
interesting alternative to current gene therapy procedures, lization and the intravenous injection of helper-dependent
as they would avoid the need for both the myeloablative adenovirus (HD-Ad). Mobilization was achieved with a
conditioning of patients and ex vivo correction of HSCs. combination of granulocyte colony-stimulating factor
Here, we discuss several in vivo gene therapy approaches (G-CSF) and CXCR4 antagonists. The cells were then
that have given promising results in preclinical models, but transduced with an HD-Ad5/35 ++ vector system. Cell
for which long-term safety studies are required before trans- transduction with the Ad5/35 vector does not require cell
fer into clinical practice. division [223]. Chimeric Ad5 vectors carrying fibers from
Ad serotype 35 further mutated into HD-Ad5/35++ have
4.3.1 Intraosseous Injection of LVVs no viral genes, preventing HSC cytotoxicity. Furthermore,
they target cells expressing CD46 with high affinity [224].
Intraosseous injection targets vector delivery to the bone This vector therefore combines efficient DNA delivery
marrow, thereby increasing the likelihood of HSC trans- through binding to CD46, a receptor expressed on HSCs
duction. In addition, the direct modification of HSCs has and highly abundant in the blood after mobilization, with
the advantage of keeping the cells in their physiological the transduction of non-dividing cells. The integrating
microenvironment. However, the lack of a specific marker HD-Ad5/35++ hybrid system developed by Lieber’s group
for HSCs to allow their targeting, the phase of the cell cycle consists of an adenoviral helper vector providing the inte-
in which HSCs reside, the limited volume per injection, and grating machinery of the Sleeping Beauty (SB) transposon
the number of HSCs required for targeting pose obstacles to (HD-Ad-SB) and a transposon-donor adenoviral vector
the development of this technique. In mice, the intraosseous carrying the transgene (HD-Ad-Tg). An advantage of this
injection of vesicular stomatitis virus glycoprotein (VSV- transposase-dependent integration system is that it does not
G) pseudotyped LV particles has been reported to lead to require the cellular DNA repair machinery and results in
only 2–3% of the peripheral blood cells being genetically random integration. This system has been shown to modify
modified [214]. This may be a sufficiently high level for 5–10% of mouse HSCs in the long term, after peripheral
hemophilia A [215] or Fanconi anemia [216], but it is too blood cell mobilization and intravenous injection [225].
low for the correction of hemoglobin disorders [217, 218]. Vector genomes have been detected in the lung, liver, heart,
Viral particles pseudotyped with VSV-G can enter almost and kidney, but this detection may reflect the infiltration
any cell type, due to the omnipresence of VSV-G receptors of these organs with blood cells or the presence of resi-
[219], and these particles are highly sensitive to comple- dential macrophages. Further analyses are required to test
ment-mediated degradation mechanisms [220]. LVVs should this hypothesis.
therefore be engineered to target HSCs specifically and to be
resistant to complement. In addition, activation of the G0/G1 4.3.3 In Vivo Selection
transition would undoubtedly facilitate HSC transduction.
To this end, LV particles carrying the complement-resistant Given the high level of gene marking required for hemo-
engineered fusion protein RDTR instead of VSV-G, together globin disorders, this approach has been associated with
with stem cell factor (SCF), have been produced [221]. They the in vivo selection of corrected cells based on insertion
promoted highly efficient ­CD34+ cell-targeted gene trans- of the mutant ­O6-methylguanine-DNA methyltransferase
fer in whole cord blood (CB) samples. In immunodeficient ­MGMTP140K gene together with the transgene [226]. The
mice into which human CB ­CD34+ cells had previously been ­MGMTP140K gene allows for the in vivo selection of gene-
injected, following inoculation of the femur with LV parti- bearing cells with low doses of ­O6-benzylguanine ­(O6BG)
cles, the percentage of transduced human cells was much plus bis-chloroethylnitrosourea (BCNU). A 12-kb DNA
higher than with classical vector preparations, but neverthe- fragment consisting of the γ-globin HBG1 gene, 4.3 kb
less remained insufficient. LVV preparations contain pro- of the β-globin LCR, a 0.7-kb promoter element, and an
teins captured from the packaging cell lines that can induce ­MGMTP140k expression cassette is inserted into the trans-
alloimmune reactivity and limit vector survival. Genome poson-donor vector. Following intravenous injection of
editing in packaging cell lines for the production of alloanti- the HD-Ad-γ-globin/MGMT and the HD-Ad-SB vectors
gen-free LVVs has been shown to enhance LVV stability and into mobilized animals [227], γ-globin expressing red
may help to increase efficacy after in vivo injection [222]. blood cells were hardly detectable. However, after in vivo
Innovative Therapies for Hemoglobin Disorders

selection, up to 60% of red blood cells expressed the human 4.3.5 CRISPR/Cas9‑Mediated DSBs
γ-globin chain in the long term. In the ­Hbbth3/th3 mouse
model of thalassemia, even in the absence of MGMT-medi- HSC mobilization followed by non-integrative
ated selection, the percentage of γ-globin-expressing red HD-Ad5/35++ vector injection has also been used for the
blood cells was higher than that in normal mice, thanks temporary expression of CRISPR/Cas9 to edit BCL11A
to the survival advantage conferred on the corrected red binding sites and reactivate γ-globin gene expression [231].
blood cells. After several cycles of ­O6BG and BCNU treat- In mice transgenic for the human β-globin locus, the per-
ment, the level of corrected red blood cells reached 60%, centage of edited cells was limited to 5–10%, but the advan-
resulting in the near-complete correction of splenomegaly, tage provided by the expression of γ-globin in the erythroid
hemosiderosis, and red blood cell parameters, despite low cells of patients with β-thalassemia or SCD should make
levels of γ-globin transgene expression (10% that in adult this approach more efficient in this context. The prolonged
mouse globin genes). This system is of great interest, as it expression and/or activity of nucleases may be detrimental
makes it possible to increase the size of the LCR further, to HSCs in the long term [131, 134, 232, 233]. Solving the
thanks to the high cargo capacity of the vector (≈ 30 kb). problem of nuclease toxicity to HSC will be important not
Studies performed to date in large-animal models with the only in the context of in vivo gene therapy, but also for the
MGMT selection system have provided no evidence of a ongoing HSC gene therapy approaches involving CRISPR/
risk of carcinogenesis [228]. Nevertheless, the risks of trig- Cas9-mediated DSBs.
gering mutations due to the alkylating agent BCNU and the
DNA repair inhibitor O ­ 6BG in surviving selected HSCs, 4.3.6 PNA‑Mediated Gene Editing
and those associated with the random integration of the
vector, will require more complete evaluation in long-term Another in vivo gene editing approach is based on triplex-
studies. forming peptide nucleic acids (PNAs). PNAs bind DNA in
a site-specific manner, via strand invasion and the formation
4.3.4 CRISPR/Cas9‑Mediated HDR of a PNA/DNA/PNA triplex, through both Watson–Crick
and Hoogsteen binding. The PNAs then recruit DNA repair
With a view to improving the safety profile of in vivo HSC systems to correct the genome with the desired sequence,
transduction, intravenous HD-Ad5/35++ vectors have been co-delivered as a template. As PNAs involve high-fidelity
used to provide a CRISPR/Cas9 complex generating DSBs DNA repair pathways and have no intrinsic nuclease activ-
at the AAVS1 locus, together with a donor template incor- ity, their use is less prone to indel mutations and/or chro-
porating the γ-globin-MGMT DNA cassette flanked by mosome rearrangements than current approaches based on
AAVS1 homology arms and target sites for the AAVS1 nucleases. However, their in vivo delivery remains challeng-
CRISPR/Cas9 gRNA [229]. The use of this helper- ing. The extent of in vivo correction was recently evalu-
dependent HD-Ad5/35 ++ vector with a modified capsid ated in a mouse model of β-thalassemia transduced with
to transduce mobilized HSCs, followed by in vivo selec- a human β-globin gene with a splicing mutation in intron
tion, resulted in a high frequency of targeted integration in 2 (IVS2-654) [234]. The intravenous administration of
MP
long-term HSCs. Interestingly, γ-globin expression levels γPNA (PNA with a mini polyethylene-glycol group) in
were twice those obtained with the transposase-dependent poly(lactic-co-glycolic acid) nanoparticles (PLGA NPs),
system for similar numbers of integrated transgene cop- associated with in vivo cKit ligand (SCF) treatment, led to
ies per cell. However, on-target deletion and off-target gene editing in as many as ≈ 7% of early bone marrow pro-
integration events were detected in this study. Given the genitor cells, a level sufficiently high for partial reversal of
potential rearrangements induced by CRISPR/Cas9 at the the β-thalassemia phenotype. It remained unclear why SCF
target site or distal to the cutting site [230], and the edit- stimulated gene editing, and there is a need to determine
ing of large numbers of cells, such lesions may constitute whether this mechanism will operate in humans. γPNAs
a first carcinogenic hit. Long-term studies are required to loaded into PLGA NPs have also been successfully used for
determine whether cells with such deleterious chromo- genome editing in utero. There are several advantages to in
somal changes are removed by counterselection or become utero gene therapy: HSCs are mostly proliferative and the
neoplastic over time. Furthermore, the need to eliminate fetal immune system is immature, allowing a better toler-
a large part of the stem cell reservoir due to the low infec- ance of delivery strategies. In utero gene editing would also
tion efficiency before selection raises questions about the fix the molecular defect before the appearance of the first
capacity of the remaining cells to ensure hematopoiesis symptoms and before the occurrence of irreversible dam-
for the rest of the patient’s life. This is also an issue for age in SCD or thalassemia. Furthermore, PLGA NPs have
patients treated with ex vivo gene therapy after myeloabla- been approved for drug delivery applications [235] and their
tive conditioning. intravenous injection leads to their accumulation in the fetal
K. Sii‑Felice et al.

liver [236], the site of hematopoiesis in the fetus. In this because it increases the risk of insertional mutagenesis
study, after a single treatment with PNAs and a donor tem- and, thus, of uncontrolled cell expansion and cell transfor-
plate, in the absence of SCF injection, an editing frequency mation. Insertional oncogenesis has not yet been observed
of 6% was observed in adult bone marrow cells, resulting in with SIN LVVs, but remains a risk for which long-term
an improvement of the disease phenotype and the postnatal assessment is required. Patients treated by gene therapy
survival of β-thalassemic mice [236]. Almost no off-target will be followed carefully for many years, with close scru-
mutations were detected, rendering this strategy potentially tiny of the risks of cell transformation.
more attractive than nuclease-based technologies. More One of the major challenges in the last few years has been
studies are required to assess the lack of specificity of non- the delivery of LVVs to genomic sites devoid of oncogenes
viral delivery and the risk of germ cell mutation. In addition, or sequences involved in the control of cell fate. The ques-
further improvements in editing efficiency, to attain 20–30% tion of targeting LVVs to safe chromatin harbors was not
modified cells, are thought to be required for the correction addressed in this review, due to the lack of solid evidence
of hemoglobin disorders. that such an approach can work efficiently. Another possible
In vivo gene therapies limit the risk of losing the stemness way of reducing the risk of cell transformation would be to
and homing capability of HSCs due to the handling of these decrease the number of vector copies required per cell by
cells, and eliminate the need for myeloablative condition- increasing therapeutic efficacy. It may, therefore, be useful
ing in patients with chronic diseases and preexisting organ to develop vectors capable of expressing transgenes more
damage. However, the translation of these technologies into strongly, or of delivering several genes together, each pro-
clinical practice poses significant challenges. In vivo gene viding a specific and complementary therapeutic solution.
therapy protocols can activate innate and adaptive immune Alternatives to gene addition include targeted modifica-
responses through the intravenous injection of the vector, tions through the use of nucleases or derivatives. The in vitro
and can lead to the transduction of cells other than HSCs. results obtained to date are very promising, and several clini-
Many ethical issues have been raised regarding the use of cal trials have already begun. The lack of hindsight makes it
these technologies, partly due to the risk of “off-target” difficult to draw any firm conclusions about the efficacy and
effects on target cells, but also due to the risk of germ cell safety of these approaches as yet. They are also associated
modification. with an undeniable risk of mutagenesis. The challenge in
the coming years will be to benchmark these methods and
weigh up their respective risks and benefits [237]. There is
5 Conclusion no doubt that much research remains to be done in the field
of gene therapy for hemoglobin disorders.
The results of recent clinical trials using LVVs to trans-
duce HSCs for the treatment of β-thalassemia or SCD Author Contributions  This review article was written at the request of
the editor. All authors participated in the literature review and writing
show very promising therapeutic efficacy. Several patients of the manuscript. Emmanuel Payen critically revised and edited the
with no other hope of definitive treatment have been suc- manuscript.
cessfully cured of their chronic and debilitating hemo-
globin disorder. These extremely convincing results have Availability of Data and Material  Not applicable.
led the European Medicines Agency (EMA) to grant a
conditional marketing authorization for Zynteglo, a gene Declarations 
therapy developed by the pharmaceutical company blue-
bird bio Inc., for the treatment of adult and adolescent Funding  This work was supported with European Research Projects
on Rare Diseases (GETHERTHAL) funding from the European Com-
TDT patients with the non-β0/β0 genotype, who are eli- mission.
gible for HSC transplantation, and for whom an HLA-
matched donor is unavailable. One difficulty remains to be Conflict of interest  Olivier Negre is a consultant for bluebird bio. He
resolved: obtaining strictly reproducible and controllable was a former employee and share holder of bluebird bio. The other
results in all patients, including those with the most severe authors have no conflict of interest to declare.
conditions (β0-thal for example), while maintaining a satis-
Ethics approval  Not applicable.
factory safety profile. Efforts to achieve therapeutic benefit
have, to date, focused on optimizing the rate of transduc- Consent to participate  Not applicable.
tion and the intensity of myeloablative conditioning before
the infusion of genetically modified HSCs, to maximize Consent for publication  All authors have read and approved the entire
contents of the manuscript.
the number of vector copies per cell, and reach a mean
VCN sufficient to achieve disease correction. This strong Code availability  Not applicable.
pressure on the transduction rate is a double-edged sword,
Innovative Therapies for Hemoglobin Disorders

References 21. Angelucci E. Hematopoietic stem cell transplantation


in thalassemia. Hematol Am Soc Hematol Educ Progr.
2010;2010:456–62.
1. Weatherall D. The inherited disorders of haemoglobin: an
22. Bhatia M, Kolva E, Cimini L, Jin Z, Satwani P, Savone M, et al.
increasingly neglected global health burden. Indian J Med Res.
Health-related quality of life after allogeneic hematopoietic
2011;134:493–7.
stem cell transplantation for sickle cell disease. Biol Blood
2. Allison AC. Protection afforded by sickle-cell trait against sub-
Marrow Transplant. 2015;21:666–72.
tertian malareal infection. BMJ. 1954;1:290–4.
23. La Nasa G, Caocci G, Efficace F, Dessi C, Vacca A, Piras E,
3. Piel FB, Tatem AJ, Huang Z, Gupta S, Williams TN, Weather-
et al. Long-term health-related quality of life evaluated more
all DJ. Global migration and the changing distribution of sickle
than 20 years after hematopoietic stem cell transplantation for
haemoglobin: a quantitative study of temporal trends between
thalassemia. Blood. 2013;122:2262–70.
1960 and 2000. Lancet Glob Health. 2014;2:e80–9.
24. Locatelli F, Merli P, Strocchio L. Transplantation for thalas-
4. Shook LM, Ware RE. Sickle cell screening in Europe: the time
semia major: alternative donors. Curr Opin Hematol.
has come. Br J Haematol. 2018;183:534–5.
2016;23:515–23.
5. Flint J, Harding RM, Boyce AJ, Clegg JB. The population
25. Shenoy S, Eapen M, Panepinto JA, Logan BR, Wu J, Abra-
genetics of the haemoglobinopathies. Bailliere’s Clin Haema-
ham A, et al. A trial of unrelated donor marrow transplan-
tol. 1998;11:1–51.
tation for children with severe sickle cell disease. Blood.
6. Finch CA, Sturgeon P. Erythrokinetics in Cooley’s anemia.
2016;128:2561–7.
Blood. 1957;12:64–73.
26. Platt OS, Brambilla DJ, Rosse WF, Milner PF, Castro O,
7. Camaschella C, Nai A. Ineffective erythropoiesis and regula-
Steinberg MH, et  al. Mortality in sickle cell disease. Life
tion of iron status in iron loading anaemias. Br J Haematol.
expectancy and risk factors for early death. N Engl J Med.
2016;172:512–23.
1994;330:1639–44.
8. Taher AT, Saliba AN. Iron overload in thalassemia: differ-
27. Sunshine HR, Hofrichter J, Eaton WA. Gelation of sickle cell
ent organs at different rates. Hematol Am Soc Hematol Educ
hemoglobin in mixtures with normal adult and fetal hemoglob-
Progr. 2017;2017:265–71.
ins. J Mol Biol. 1979;133:435–67.
9. Aydinok Y, Porter JB, Piga A, Elalfy M, El-Beshlawy A, Kilinc
28. Poillon WN, Kim BC, Rodgers GP, Noguchi CT, Schechter
Y, et al. Prevalence and distribution of iron overload in patients
AN. Sparing effect of hemoglobin F and hemoglobin A2 on the
with transfusion-dependent anemias differs across geographic
polymerization of hemoglobin S at physiologic ligand satura-
regions: results from the CORDELIA study. Eur J Haematol.
tions. Proc Natl Acad Sci USA. 1993;90:5039–43.
2015;95:244–53.
29. Adachi K, Konitzer P, Surrey S. Role of gamma 87 Gln in the
10. Kremastinos DT, Farmakis D, Aessopos A, Hahalis G, Ham-
inhibition of hemoglobin S polymerization by hemoglobin F. J
odraka E, Tsiapras D, et al. Beta-thalassemia cardiomyopathy:
Biol Chem. 1994;269:9562–7.
history, present considerations, and future perspectives. Circ
30. Nagel RL, Bookchin RM, Johnson J, Labie D, Wajcman H, Isaac-
Heart Fail. 2010;3:451–8.
Sodeye WA, et al. Structural bases of the inhibitory effects of
11. Modell B, Khan M, Darlison M, Westwood MA, Ingram D,
hemoglobin F and hemoglobin A2 on the polymerization of
Pennell DJ. Improved survival of thalassaemia major in the
hemoglobin S. Proc Natl Acad Sci USA. 1979;76:670–2.
UK and relation to T2* cardiovascular magnetic resonance. J
31. Reddy LR, Reddy KS, Surrey S, Adachi K. Role of beta87 Thr
Cardiovasc Magn Reson. 2008;10:42.
in the beta6 Val acceptor site during deoxy Hb S polymerization.
12. Voskaridou E, Ladis V, Kattamis A, Hassapopoulou E, Econo-
Biochemistry. 1997;36:15992–8.
mou M, Kourakli A, et al. A national registry of haemoglobi-
32. Pawliuk R, Westerman KA, Fabry ME, Payen E, Tighe R, Bou-
nopathies in Greece: deducted demographics, trends in mortal-
hassira EE, et al. Correction of sickle cell disease in transgenic
ity and affected births. Ann Hematol. 2012;91:1451–8.
mouse models by gene therapy. Science. 2001;294:2368–71.
13. Marsella M, Ricchi P. Thalassemia and hepatocellular carci-
33. Chakalova L, Osborne CS, Dai YF, Goyenechea B, Metaxotou-
noma: links and risks. J Blood Med. 2019;10:323–34.
Mavromati A, Kattamis A, et al. The Corfu deltabeta thalas-
14. Pinto VM, Poggi M, Russo R, Giusti A, Forni GL. Manage-
semia deletion disrupts gamma-globin gene silencing and
ment of the aging beta-thalassemia transfusion-dependent pop-
reveals post-transcriptional regulation of HbF expression. Blood.
ulation - the Italian experience. Blood Rev. 2019;38:100594.
2005;105:2154–60.
15. Eaton WA, Hofrichter J. Hemoglobin S gelation and sickle cell
34. Russell JE. A post-transcriptional process contributes to efficient
disease. Blood. 1987;70:1245–66.
gamma-globin gene silencing in definitive erythroid cells. Eur J
16. Kato GJ, Steinberg MH, Gladwin MT. Intravascular hemolysis
Haematol. 2007;79:516–25.
and the pathophysiology of sickle cell disease. J Clin Investig.
35. Kutlar F, Gonzalez-Redondo JM, Kutlar A, Gurgey A, Altay C,
2017;127:750–60.
Efremov GD, et al. The levels of zeta, gamma, and delta chains
17. Sundd P, Gladwin MT, Novelli EM. Pathophysiology of Sickle
in patients with Hb H disease. Hum Genet. 1989;82:179–86.
Cell Disease. Annu Rev Pathol. 2019;14:263–92.
36. Martinez G, Menendez R. Differences in affinity of beta and delta
18. Thein SL. Genetic modifiers of sickle cell disease. Hemo-
hemoglobin chains for alpha chains. A possible explanation for
globin. 2011;35:589–606.
the variation in the percentages of hemoglobin A2 in thalassemia
19. Day ME, Rodeghier M, DeBaun MR. Children with HbS-
and other disorders. Biochem Biophys Acta. 1983;743:256–9.
beta(0) thalassemia have higher hemoglobin levels and lower
37. McCune SL, Reilly MP, Chomo MJ, Asakura T, Townes TM.
incidence rate of acute chest syndrome compared to children
Recombinant human hemoglobins designed for gene therapy of
with HbSS. Pediatr Blood Cancer. 2018;65:e27352.
sickle cell disease. Proc Natl Acad Sci USA. 1994;91:9852–6.
20. Gluckman E, Cappelli B, Bernaudin F, Labopin M, Volt F,
38. Bunn HF, McDonald MJ. Electrostatic interactions in the assem-
Carreras J, et al. Sickle cell disease: an international survey
bly of haemoglobin. Nature. 1983;306:498–500.
of results of HLA-identical sibling hematopoietic stem cell
39. Levasseur DN, Ryan TM, Reilly MP, McCune SL, Asakura
transplantation. Blood. 2017;129:1548–56.
T, Townes TM. A recombinant human hemoglobin with anti-
sickling properties greater than fetal hemoglobin. J Biol Chem.
2004;279:27518–24.
K. Sii‑Felice et al.

40. Forrester WC, Novak U, Gelinas R, Groudine M. Molecular a gamma-globin lentiviral vector in murine beta-thalassemia.
analysis of the human beta-globin locus activation region. Proc Blood. 2004;104:2281–90.
Natl Acad Sci USA. 1989;86:5439–43. 58. Puthenveetil G, Scholes J, Carbonell D, Qureshi N, Xia P, Zeng L,
41. Talbot D, Collis P, Antoniou M, Vidal M, Grosveld F, Greaves et al. Successful correction of the human beta-thalassemia major
DR. A dominant control region from the human beta-globin phenotype using a lentiviral vector. Blood. 2004;104:3445–53.
locus conferring integration site-independent gene expression. 59. Miccio A, Cesari R, Lotti F, Rossi C, Sanvito F, Ponzoni M, et al.
Nature. 1989;338:352–5. In vivo selection of genetically modified erythroblastic progeni-
42. Gelinas R, Frazier A, Harris E. A normal level of beta-globin tors leads to long-term correction of beta-thalassemia. Proc Natl
expression in erythroid cells after retroviral cells transfer. Bone Acad Sci USA. 2008;105:10547–52.
Marrow Transplant. 1992;9(Suppl 1):154–7. 60. Trudel M, De Paepe ME, Chretien N, Saadane N, Jacmain J,
43. Novak U, Harris EA, Forrester W, Groudine M, Gelinas R. Sorette M, et al. Sickle cell disease of transgenic SAD mice.
High-level beta-globin expression after retroviral transfer of Blood. 1994;84:3189–97.
locus activation region-containing human beta-globin gene 61. Paszty C, Brion CM, Manci E, Witkowska HE, Stevens ME,
derivatives into murine erythroleukemia cells. Proc Natl Acad Mohandas N, et al. Transgenic knockout mice with exclusively
Sci USA. 1990;87:3386–90. human sickle hemoglobin and sickle cell disease. Science.
44. Rivella S, Sadelain M. Genetic treatment of severe hemoglo- 1997;278:876–8.
binopathies: the combat against transgene variegation and 62. Pestina TI, Hargrove PW, Jay D, Gray JT, Boyd KM, Persons
transgene silencing. Semin Hematol. 1998;35:112–25. DA. Correction of murine sickle cell disease using gamma-globin
45. Plavec I, Papayannopoulou T, Maury C, Meyer F. A human lentiviral vectors to mediate high-level expression of fetal hemo-
beta-globin gene fused to the human beta-globin locus control globin. Mol Ther. 2009;17:245–52.
region is expressed at high levels in erythroid cells of mice 63. Ryan TM, Ciavatta DJ, Townes TM. Knockout-transgenic mouse
engrafted with retrovirus-transduced hematopoietic stem cells. model of sickle cell disease. Science. 1997;278:873–6.
Blood. 1993;81:1384–92. 64. Levasseur DN, Ryan TM, Pawlik KM, Townes TM. Correction of
46. Leboulch P, Huang GM, Humphries RK, Oh YH, Eaves CJ, a mouse model of sickle cell disease: lentiviral/antisickling beta-
Tuan DY, et al. Mutagenesis of retroviral vectors transducing globin gene transduction of unmobilized, purified hematopoietic
human beta-globin gene and beta-globin locus control region stem cells. Blood. 2003;102:4312–9.
derivatives results in stable transmission of an active transcrip- 65. Akinsheye I, Alsultan A, Solovieff N, Ngo D, Baldwin CT,
tional structure. EMBO J. 1994;13:3065–76. Sebastiani P, et al. Fetal hemoglobin in sickle cell anemia. Blood.
47. Sadelain M, Wang CH, Antoniou M, Grosveld F, Mulligan RC. 2011;118:19–27.
Generation of a high-titer retroviral vector capable of express- 66. Rivella S, May C, Chadburn A, Riviere I, Sadelain M. A novel
ing high levels of the human beta-globin gene. Proc Natl Acad murine model of Cooley anemia and its rescue by lentiviral-medi-
Sci USA. 1995;92:6728–32. ated human beta-globin gene transfer. Blood. 2003;101:2932–9.
48. Raftopoulos H, Ward M, Leboulch P, Bank A. Long-term trans- 67. Li CL, Emery DW. The cHS4 chromatin insulator reduces gam-
fer and expression of the human beta-globin gene in a mouse maretroviral vector silencing by epigenetic modifications of inte-
transplant model. Blood. 1997;90:3414–22. grated provirus. Gene Ther. 2008;15:49–53.
49. May C, Rivella S, Callegari J, Heller G, Gaensler KM, Luz- 68. Arumugam PI, Scholes J, Perelman N, Xia P, Yee JK, Malik P.
zatto L, et al. Therapeutic haemoglobin synthesis in beta-tha- Improved human beta-globin expression from self-inactivating
lassaemic mice expressing lentivirus-encoded human beta- lentiviral vectors carrying the chicken hypersensitive site-4
globin. Nature. 2000;406:82–6. (cHS4) insulator element. Mol Ther. 2007;15:1863–71.
50. Naldini L, Blomer U, Gallay P, Ory D, Mulligan R, Gage FH, 69. Urbinati F, Arumugam P, Higashimoto T, Perumbeti A, Mitts
et al. In vivo gene delivery and stable transduction of nondivid- K, Xia P, et al. Mechanism of reduction in titers from lenti-
ing cells by a lentiviral vector. Science. 1996;272:263–7. virus vectors carrying large inserts in the 3′LTR. Molecular
51. Nakamura-Ishizu A, Takizawa H, Suda T. The analysis, roles therapy: the journal of the American Society of Gene Therapy.
and regulation of quiescence in hematopoietic stem cells. 2009;17:1527–36.
Development. 2014;141:4656–66. 70. Ronen K, Negre O, Roth S, Colomb C, Malani N, Denaro M,
52. Imren S, Payen E, Westerman KA, Pawliuk R, Fabry ME, et al. Distribution of lentiviral vector integration sites in mice
Eaves CJ, et  al. Permanent and panerythroid correction following therapeutic gene transfer to treat beta-thalassemia. Mol
of murine beta thalassemia by multiple lentiviral integra- Ther. 2011;19:1273–86.
tion in hematopoietic stem cells. Proc Natl Acad Sci USA. 71. Aker M, Tubb J, Groth AC, Bukovsky AA, Bell AC, Felsenfeld
2002;99:14380–5. G, et al. Extended core sequences from the cHS4 insulator are
53. May C, Rivella S, Chadburn A, Sadelain M. Successful treatment necessary for protecting retroviral vectors from silencing position
of murine beta-thalassemia intermedia by transfer of the human effects. Hum Gene Ther. 2007;18:333–43.
beta-globin gene. Blood. 2002;99:1902–8. 72. Arumugam PI, Urbinati F, Velu CS, Higashimoto T, Grimes
54. Rouyer-Fessard P, Leroy-Viard K, Domenget C, Mrad A, Beu- HL, Malik P. The 3′ region of the chicken hypersensitive site-4
zard Y. Mouse beta thalassemia, a model for the membrane insulator has properties similar to its core and is required for full
defects of erythrocytes in the human disease. J Biol Chem. insulator activity. PLoS One. 2009;4:e6995.
1990;265:20247–51. 73. Uchida N, Washington KN, Lap CJ, Hsieh MM, Tisdale JF.
55. Yang B, Kirby S, Lewis J, Detloff PJ, Maeda N, Smithies O. A Chicken HS4 insulators have minimal barrier function among
mouse model for beta 0-thalassemia. Proc Natl Acad Sci USA. progeny of human hematopoietic cells transduced with an HIV1-
1995;92:11608–12. based lentiviral vector. Mol Ther. 2011;19:133–9.
56. Zennou V, Petit C, Guetard D, Nerhbass U, Montagnier L, 74. Negre O, Bartholomae C, Beuzard Y, Cavazzana M, Christiansen
Charneau P. HIV-1 genome nuclear import is mediated by a L, Courne C, et al. Preclinical evaluation of efficacy and safety
central DNA flap. Cell. 2000;101:173–85. of an improved lentiviral vector for the treatment of beta-thalas-
57. Hanawa H, Hargrove PW, Kepes S, Srivastava DK, Nien- semia and sickle cell disease. Curr Gene Ther. 2015;15:64–81.
huis AW, Persons DA. Extended beta-globin locus control 75. Ramezani A, Hawley TS, Hawley RG. Performance- and safety-
region elements promote consistent therapeutic expression of enhanced lentiviral vectors containing the human interferon-beta
Innovative Therapies for Hemoglobin Disorders

scaffold attachment region and the chicken beta-globin insulator. 5′HS4 and human T-cell receptor alpha/delta BEAD-1 insula-
Blood. 2003;101:4717–24. tors in self-inactivating retroviral vectors reduces their genotoxic
76. Hanawa H, Yamamoto M, Zhao H, Shimada T, Persons DA. potential. Stem cells. 2008;26:3257–66.
Optimized lentiviral vector design improves titer and transgene 93. Romero Z, Urbinati F, Geiger S, Cooper AR, Wherley J, Kauf-
expression of vectors containing the chicken beta-globin locus man ML, et al. Beta-globin gene transfer to human bone marrow
HS4 insulator element. Mol Ther. 2009;17:667–74. for sickle cell disease. J Clin Investig. 2013;123:3317–30.
77. Montini E, Cesana D, Schmidt M, Sanvito F, Bartholomae CC, 94. Urbinati F, Campo Fernandez B, Masiuk KE, Poletti V, Hollis
Ranzani M, et al. The genotoxic potential of retroviral vectors is RP, Koziol C, et al. Gene therapy for sickle cell disease: a lentivi-
strongly modulated by vector design and integration site selec- ral vector comparison study. Hum Gene Ther. 2018;29:1153–66.
tion in a mouse model of HSC gene therapy. J Clin Investig. 95. Gaszner M, Felsenfeld G. Insulators: exploiting transcriptional
2009;119:964–75. and epigenetic mechanisms. Nat Rev Genet. 2006;7:703–13.
78. Ranzani M, Cesana D, Bartholomae CC, Sanvito F, Pala M, Ben- 96. Desprat R, Bouhassira EE. Gene specificity of suppression of
edicenti F, et al. Lentiviral vector-based insertional mutagen- transgene-mediated insertional transcriptional activation by the
esis identifies genes associated with liver cancer. Nat Methods. chicken HS4 insulator. PLoS One. 2009;4:e5956.
2013;10:155–61. 97. Moiani A, Paleari Y, Sartori D, Mezzadra R, Miccio A, Cat-
79. Maruggi G, Porcellini S, Facchini G, Perna SK, Cattoglio C, Sar- toglio C, et al. Lentiviral vector integration in the human genome
tori D, et al. Transcriptional enhancers induce insertional gene induces alternative splicing and generates aberrant transcripts. J
deregulation independently from the vector type and design. Mol Clin Investig. 2012;122:1653–66.
Ther. 2009;17:851–6. 98. Cavazzana-Calvo M, Payen E, Negre O, Wang G, Hehir K,
80. Arumugam PI, Higashimoto T, Urbinati F, Modlich U, Nestheide Fusil F, et al. Transfusion independence and HMGA2 activa-
S, Xia P, et al. Genotoxic potential of lineage-specific lentivirus tion after gene therapy of human beta-thalassaemia. Nature.
vectors carrying the beta-globin locus control region. Mol Ther. 2010;467:318–22.
2009;17:1929–37. 99. Morgan RA, Unti MJ, Aleshe B, Brown D, Osborne KS, Koziol
81. Bushman F, Lewinski M, Ciuffi A, Barr S, Leipzig J, Hannenhalli C, et al. Improved titer and gene transfer by lentiviral vectors
S, et al. Genome-wide analysis of retroviral DNA integration. Nat using novel, small beta-globin locus control region elements.
Rev Microbiol. 2005;3:848–58. Mol Ther. 2020;28:328–40.
82. Zychlinski D, Schambach A, Modlich U, Maetzig T, Meyer J, 100. Breda L, Casu C, Gardenghi S, Bianchi N, Cartegni L, Narla M,
Grassman E, et al. Physiological promoters reduce the genotoxic et al. Therapeutic hemoglobin levels after gene transfer in beta-
risk of integrating gene vectors. Mol Ther. 2008;16:718–25. thalassemia mice and in hematopoietic cells of beta-thalassemia
83. Chang AH, Sadelain M. The genetic engineering of hematopoi- and sickle cells disease patients. PLoS One. 2012;7:e32345.
etic stem cells: the rise of lentiviral vectors, the conundrum of 101. Urbinati F, Hargrove PW, Geiger S, Romero Z, Wherley J, Kauf-
the ltr, and the promise of lineage-restricted vectors. Mol Ther. man ML, et al. Potentially therapeutic levels of anti-sickling glo-
2007;15:445–56. bin gene expression following lentivirus-mediated gene transfer
84. Modlich U, Navarro S, Zychlinski D, Maetzig T, Knoess S, Brug- in sickle cell disease bone marrow CD34 + cells. Exp Hematol.
man MH, et al. Insertional transformation of hematopoietic cells 2015;43:346–51.
by self-inactivating lentiviral and gammaretroviral vectors. Mol 102. Urbinati F, Wherley J, Geiger S, Fernandez BC, Kaufman ML,
Ther. 2009;17:1919–28. Cooper A, et al. Preclinical studies for a phase 1 clinical trial of
85. Grosveld F, van Assendelft GB, Greaves DR, Kollias G. Posi- autologous hematopoietic stem cell gene therapy for sickle cell
tion-independent, high-level expression of the human beta-globin disease. Cytotherapy. 2017;19:1096–112.
gene in transgenic mice. Cell. 1987;51:975–85. 103. Navas PA, Peterson KR, Li Q, McArthur M, Stamatoyannopou-
86. Forrester WC, Epner E, Driscoll MC, Enver T, Brice M, Papay- los G. The 5′HS4 core element of the human beta-globin locus
annopoulou T, et al. A deletion of the human beta-globin locus control region is required for high-level globin gene expression
activation region causes a major alteration in chromatin structure in definitive but not in primitive erythropoiesis. J Mol Biol.
and replication across the entire beta-globin locus. Genes Dev. 2001;312:17–26.
1990;4:1637–49. 104. Perumbeti A, Higashimoto T, Urbinati F, Franco R, Meiselman
87. Jimenez G, Griffiths SD, Ford AM, Greaves MF, Enver T. Acti- HJ, Witte D, et al. A novel human gamma-globin gene vector for
vation of the beta-globin locus control region precedes com- genetic correction of sickle cell anemia in a humanized sickle
mitment to the erythroid lineage. Proc Natl Acad Sci USA. mouse model: critical determinants for successful correction.
1992;89:10618–22. Blood. 2009;114:1174–85.
88. Papayannopoulou T, Priestley GV, Rohde A, Peterson KR, Naka- 105. Roselli EA, Mezzadra R, Frittoli MC, Maruggi G, Biral E,
moto B. Hemopoietic lineage commitment decisions: in vivo Mavilio F, et al. Correction of beta-thalassemia major by gene
evidence from a transgenic mouse model harboring micro LCR- transfer in haematopoietic progenitors of pediatric patients.
betapro-LacZ as a transgene. Blood. 2000;95:1274–82. EMBO Mol Med. 2010;2:315–28.
89. Hargrove PW, Kepes S, Hanawa H, Obenauer JC, Pei D, Cheng 106. Boulad F, Wang X, Qu J, Taylor C, Ferro L, Karponi G, et al.
C, et  al. Globin lentiviral vector insertions can perturb the Safe mobilization of CD34 + cells in adults with beta-thalassemia
expression of endogenous genes in beta-thalassemic hematopoi- and validation of effective globin gene transfer for clinical inves-
etic cells. Mol Ther. 2008;16:525–33. tigation. Blood. 2014;123:1483–6.
90. Bell AC, West AG, Felsenfeld G. The protein CTCF is required 107. Kiem HP, Arumugam PI, Burtner CR, Fox CF, Beard BC, Dex-
for the enhancer blocking activity of vertebrate insulators. Cell. heimer P, et al. Pigtailed macaques as a model to study long-term
1999;98:387–96. safety of lentivirus vector-mediated gene therapy for hemoglobi-
91. Recillas-Targa F, Pikaart MJ, Burgess-Beusse B, Bell AC, Litt nopathies. Mol Ther Methods Clin Dev. 2014;1:14055.
MD, West AG, et al. Position-effect protection and enhancer 108. Poletti V, Urbinati F, Charrier S, Corre G, Hollis RP, Campo
blocking by the chicken beta-globin insulator are separable Fernandez B, et al. Pre-clinical development of a lentiviral vec-
activities. Proc Natl Acad Sci USA. 2002;99:6883–8. tor expressing the anti-sickling betaAS3 globin for gene therapy
92. Ramezani A, Hawley TS, Hawley RG. Combinatorial incorpora- for sickle cell disease. Molecular therapy Methods & clinical
tion of enhancer-blocking components of the chicken beta-globin development. 2018;11:167–79.
K. Sii‑Felice et al.

109. Lettre G, Sankaran VG, Bezerra MA, Araujo AS, Uda M, Sanna 125. Sun N, Liang J, Abil Z, Zhao H. Optimized TAL effector nucle-
S, et al. DNA polymorphisms at the BCL11A, HBS1L-MYB, ases (TALENs) for use in treatment of sickle cell disease. Mol
and beta-globin loci associate with fetal hemoglobin levels BioSyst. 2012;8:1255–63.
and pain crises in sickle cell disease. Proc Natl Acad Sci USA. 126. Ramalingam S, Annaluru N, Kandavelou K, Chandrasegaran S.
2008;105:11869–74. TALEN-mediated generation and genetic correction of disease-
110. Uda M, Galanello R, Sanna S, Lettre G, Sankaran VG, Chen specific human induced pluripotent stem cells. Curr Gene Ther.
W, et al. Genome-wide association study shows BCL11A asso- 2014;14:461–72.
ciated with persistent fetal hemoglobin and amelioration of 127. Huang X, Wang Y, Yan W, Smith C, Ye Z, Wang J, et  al.
the phenotype of beta-thalassemia. Proc Natl Acad Sci USA. Production of gene-corrected adult beta globin protein in
2008;105:1620–5. human erythrocytes differentiated from patient iPSCs after
111. Sankaran VG, Menne TF, Xu J, Akie TE, Lettre G, Van Han- genome editing of the sickle point mutation. Stem cells.
del B, et al. Human fetal hemoglobin expression is regulated by 2015;33:1470–9.
the developmental stage-specific repressor BCL11A. Science. 128. Hoban MD, Cost GJ, Mendel MC, Romero Z, Kaufman ML,
2008;322:1839–42. Joglekar AV, et al. Correction of the sickle cell disease muta-
112. Menzel S, Garner C, Gut I, Matsuda F, Yamaguchi M, Heath tion in human hematopoietic stem/progenitor cells. Blood.
S, et al. A QTL influencing F cell production maps to a gene 2015;125:2597–604.
encoding a zinc-finger protein on chromosome 2p15. Nat Genet. 129. Zou J, Mali P, Huang X, Dowey SN, Cheng L. Site-specific gene
2007;39:1197–9. correction of a point mutation in human iPS cells derived from an
113. Thein SL, Menzel S, Peng X, Best S, Jiang J, Close J, et al. adult patient with sickle cell disease. Blood. 2011;118:4599–608.
Intergenic variants of HBS1L-MYB are responsible for a 130. Antony JS, Latifi N, Haque A, Lamsfus-Calle A, Daniel-
major quantitative trait locus on chromosome 6q23 influenc- Moreno A, Graeter S, et al. Gene correction of HBB mutations
ing fetal hemoglobin levels in adults. Proc Natl Acad Sci USA. in CD34(+) hematopoietic stem cells using Cas9 mRNA and
2007;104:11346–51. ssODN donors. Mol Cell Pediatr. 2018;5:9.
114. Liu P, Keller JR, Ortiz M, Tessarollo L, Rachel RA, Nakamura 131. Dever DP, Bak RO, Reinisch A, Camarena J, Washington G,
T, et al. Bcl11a is essential for normal lymphoid development. Nicolas CE, et  al. CRISPR/Cas9 β-globin gene targeting in
Nat Immunol. 2003;4:525–32. human haematopoietic stem cells. Nature. 2016;539:384–9.
115. Tsang JC, Yu Y, Burke S, Buettner F, Wang C, Kolodziejczyk 132. Romero Z, Lomova A, Said S, Miggelbrink A, Kuo CY, Campo-
AA, et al. Single-cell transcriptomic reconstruction reveals cell Fernandez B, et al. Editing the sickle cell disease mutation in
cycle and multi-lineage differentiation defects in Bcl11a-deficient human hematopoietic stem cells: comparison of endonucleases
hematopoietic stem cells. Genome Biol. 2015;16:178. and homologous donor templates. Molecular therapy: the journal
116. Wilber A, Hargrove PW, Kim YS, Riberdy JM, Sankaran VG, of the American Society of Gene Therapy. 2019;27:1389–406.
Papanikolaou E, et al. Therapeutic levels of fetal hemoglobin in 133. Pattabhi S, Lotti SN, Berger MP, Singh S, Lux CT, Jacoby K,
erythroid progeny of beta-thalassemic CD34 + cells after lenti- et al. In vivo outcome of homology-directed repair at the HBB
viral vector-mediated gene transfer. Blood. 2011;117:2817–26. gene in HSC using alternative donor template delivery methods.
117. Brendel C, Guda S, Renella R, Bauer DE, Canver MC, Kim Molr Ther Nucl Acids. 2019;17:277–88.
YJ, et  al. Lineage-specific BCL11A knockdown circum- 134. Genovese P, Schiroli G, Escobar G, Di Tomaso T, Firrito C,
vents toxicities and reverses sickle phenotype. J Clin Investig. Calabria A, et al. Targeted genome editing in human repopulating
2016;126:3868–78. haematopoietic stem cells. Nature. 2014;510:235–40.
118. Guda S, Brendel C, Renella R, Du P, Bauer DE, Canver 135. Fessas P, Stamatoyannopoulos G. Hereditary persistence of
MC, et  al. miRNA-embedded shRNAs for lineage-specific fetal hemoglobin in Greece. A study and a comparison. Blood.
BCL11A Knockdown and hemoglobin F induction. Mol Ther. 1964;24:223–40.
2015;23:1465–74. 136. Collins FS, Metherall JE, Yamakawa M, Pan J, Weissman SM,
119. Brendel C, Negre O, Rothe M, Guda S, Parsons G, Harris C, Forget BG. A point mutation in the A gamma-globin gene pro-
et al. Preclinical evaluation of a novel lentiviral vector driving moter in Greek hereditary persistence of fetal haemoglobin.
lineage-specific BCL11A knockdown for sickle cell gene therapy. Nature. 1985;313:325–6.
Mol Ther Methods Clin Dev. 2020;17:589–600. 137. Oner R, Kutlar F, Gu LH, Huisman TH. The Georgia type of
120. Esrick EB, Brendel C, Manis JP, Armant MA, Negre H, nondeletional hereditary persistence of fetal hemoglobin has a
Dansereau C, et al. Flipping the switch: initial results of genetic C—T mutation at nucleotide-114 of the A gamma-globin gene.
targeting of the Fetal to adult globin switch in sickle cell patients. Blood. 1991;77:1124–5.
Blood. 2018;132:1023. 138. Fucharoen S, Shimizu K, Fukumaki Y. A novel C-T transition
121. Shrivastav M, De Haro LP, Nickoloff JA. Regulation of within the distal CCAAT motif of the G gamma-globin gene in
DNA double-strand break repair pathway choice. Cell Res. the Japanese HPFH: implication of factor binding in elevated
2008;18:134–47. fetal globin expression. Nucleic Acids Res. 1990;18:5245–53.
122. Wattanapanitch M, Damkham N, Potirat P, Trakarnsanga K, 139. Gilman JG, Mishima N, Wen XJ, Stoming TA, Lobel J, Huisman
Janan M, U-Pratya Y, Kheolamai P, Klincumhom N, Issaragrisil TH. Distal CCAAT box deletion in the A gamma globin gene
S. One-step genetic correction of hemoglobin E/beta-thalassemia of two black adolescents with elevated fetal A gamma globin.
patient-derived iPSCs by the CRISPR/Cas9 system. Stem Cell Nucleic Acids Res. 1988;16:10635–42.
Res Ther. 2018;9:46. 140. Zertal-Zidani S, Merghoub T, Ducrocq R, Gerard N, Satta D,
123. Xie F, Ye L, Chang JC, Beyer AI, Wang J, Muench MO, et al. Krishnamoorthy R. A novel C– > A transversion within the dis-
Seamless gene correction of β-thalassemia mutations in patient- tal CCAAT motif of the Ggamma-globin gene in the Algerian
specific iPSCs using CRISPR/Cas9 and piggyBac. Genome Res. Ggammabeta + -hereditary persistence of fetal hemoglobin.
2014;24:1526–33. Hemoglobin. 1999;23:159–69.
124. Ma N, Liao B, Zhang H, Wang L, Shan Y, Xue Y, et al. Tran- 141. Giglioni B, Casini C, Mantovani R, Merli S, Comi P, Ottolenghi
scription activator-like effector nuclease (TALEN)-mediated S, et al. A molecular study of a family with Greek hereditary
gene correction in integration-free β-thalassemia induced pluri- persistence of fetal hemoglobin and beta-thalassemia. EMBO J.
potent stem cells. J Biol Chem. 2013;288:34671–9. 1984;3:2641–5.
Innovative Therapies for Hemoglobin Disorders

142. Tasiopoulou M, Boussiou M, Sinopoulou K, Moraitis G, 158. Humbert O, Radtke S, Samuelson C, Carrillo RR, Perez
Loutradi-Anagnostou A, Karababa P. G gamma-196 C– > T, AM, Reddy SS, et  al. Therapeutically relevant engraftment
A gamma-201 C– > T: two novel mutations in the promoter of a CRISPR-Cas9-edited HSC-enriched population with
region of the gamma-globin genes associated with nondele- HbF reactivation in nonhuman primates. Sci Transl Med.
tional hereditary persistence of fetal hemoglobin in Greece. 2019;11:eaaw3768.
Blood Cells Mol Dis. 2008;40:320–2. 159. Metais JY, Doerfler PA, Mayuranathan T, Bauer DE, Fowler SC,
143. Amato A, Cappabianca MP, Perri M, Zaghis I, Grisanti P, Hsieh MM, et al. Genome editing of HBG1 and HBG2 to induce
Ponzini D, et  al. Interpreting elevated fetal hemoglobin in fetal hemoglobin. Blood Adv. 2019;3:3379–92.
pathology and health at the basic laboratory level: new and 160. Bauer DE, Kamran SC, Lessard S, Xu J, Fujiwara Y, Lin C, et al.
known γ- gene mutations associated with hereditary persis- An erythroid enhancer of BCL11A subject to genetic variation
tence of fetal hemoglobin. Int J Lab Hematol. 2014;36:13–9. determines fetal hemoglobin level. Science. 2013;342:253–7.
144. Collins FS, Stoeckert CJ Jr, Serjeant GR, Forget BG, Weiss- 161. Canver MC, Smith EC, Sher F, Pinello L, Sanjana NE, Shalem
man SM. G gamma beta + hereditary persistence of fetal O, et al. BCL11A enhancer dissection by Cas9-mediated in situ
hemoglobin: cosmid cloning and identification of a specific saturating mutagenesis. Nature. 2015;527:192–7.
mutation 5′ to the G gamma gene. Proc Natl Acad Sci USA. 162. Wu Y, Zeng J, Roscoe BP, Liu P, Yao Q, Lazzarotto CR, et al.
1984;81:4894–8. Highly efficient therapeutic gene editing of human hematopoietic
145. Hattori Y, Kutlar F, Kutlar A, McKie VC, Huisman TH. Hap- stem cells. Nat Med. 2019;25:776–83.
lotypes of beta S chromosomes among patients with sickle cell 163. Boulad F, Riviere I, Wang X, Bartido S, Prockop SE, Barone R,
anemia from Georgia. Hemoglobin. 1986;10:623–42. et al. Fisrt US phase I clinical trial of globin gene transfer for the
146. Motum PI, Deng ZM, Huong L, Trent RJ. The Australian treatment of beta-thalassemia major. Blood. 2013;122:716.
type of nondeletional G gamma-HPFH has a C– > G substitu- 164. Mansilla-Soto J, Riviere I, Boulad F, Sadelain M. Cell and gene
tion at nucleotide -114 of the G gamma gene. Br J Haematol. therapy for the beta-thalassemias: advances and prospects. Hum
1994;86:219–21. Gene Ther. 2016;27:295–304.
147. Gilman JG, Mishima N, Wen XJ, Kutlar F, Huisman TH. 165. Negre O, Eggimann AV, Beuzard Y, Ribeil JA, Bourget P, Bor-
Upstream promoter mutation associated with a modest eleva- wornpinyo S, et al. Gene therapy of the beta-hemoglobinopathies
tion of fetal hemoglobin expression in human adults. Blood. by lentiviral transfer of the beta(A(T87Q))-globin gene. Hum
1988;72:78–81. Gene Ther. 2016;27:148–65.
148. Wienert B, Funnell AP, Norton LJ, Pearson RC, Wilkinson-White 166. Cavazzana M, Bushman FD, Miccio A, Andre-Schmutz I, Six
LE, Lester K, et al. Editing the genome to introduce a beneficial E. Gene therapy targeting haematopoietic stem cells for inher-
naturally occurring mutation associated with increased fetal glo- ited diseases: progress and challenges. Nat Rev Drug Discov.
bin. Nat Commun. 2015;6:7085. 2019;18:447–62.
149. Martyn GE, Wienert B, Yang L, Shah M, Norton LJ, Burdach J, 167. Ribeil JA, Hacein-Bey-Abina S, Payen E, Magnani A, Semeraro
et al. Natural regulatory mutations elevate the fetal globin gene M, Magrin E, et al. Gene therapy in a patient with sickle cell
via disruption of BCL11A or ZBTB7A binding. Nat Genet. disease. N Engl J Med. 2017;376:848–55.
2018;50:498–503. 168. Thompson AA, Walters MC, Kwiatkowski J, Rasko JEJ, Ribeil
150. Martyn GE, Wienert B, Kurita R, Nakamura Y, Quinlan KGR, JA, Hongeng S, et al. Gene therapy in patients with transfusion-
Crossley M. A natural regulatory mutation in the proximal pro- dependent beta-thalassemia. N Engl J Med. 2018;378:1479–93.
moter elevates fetal globin expression by creating a de novo 169. Marktel S, Scaramuzza S, Cicalese MP, Giglio F, Galimberti
GATA1 site. Blood. 2019;133:852–6. S, Lidonnici MR, et al. Intrabone hematopoietic stem cell gene
151. Branzei D, Foiani M. Regulation of DNA repair throughout the therapy for adult and pediatric patients affected by transfusion-
cell cycle. Nat Rev Mol Cell Biol. 2008;9:297–308. dependent ss-thalassemia. Nat Med. 2019;25:234–41.
152. Lomova A, Clark DN, Campo-Fernandez B, Flores-Bjurstrom 170. Malik P, Grimley M, Quinn CT, Shova A, Courtney L, Lutzko
C, Kaufman ML, Fitz-Gibbon S, et al. Improving gene editing C, et al. Gene therapy for sickle cell anemia using a modified
outcomes in human hematopoietic stem and progenitor cells by gamma globin lentivirus vector and reduced intensity condition-
temporal control of DNA repair. Stem cells. 2019;37:284–94. ing transplant shows promising correction of the disease pheno-
153. Lux CT, Pattabhi S, Berger M, Nourigat C, Flowers DA, Negre type. Blood. 2018;132(Suppl. 1):Abstract 1021.
O, et al. TALEN-mediated gene editing of HBG in human hemat- 171. Bueren JA, Quintana-Bustamante O, Almarza E, Navarro S, Rio
opoietic stem cells leads to therapeutic fetal hemoglobin induc- P, Segovia JC, et al. Advances in the gene therapy of monogenic
tion. Mol Ther. 2019;12:175–83. blood cell diseases. Clin Genet. 2020;97:89–102.
154. Traxler EA, Yao Y, Wang YD, Woodard KJ, Kurita R, Naka- 172. Esrick E, Achebe M, Armant M, Bartolucci P, Ciuculescu MF,
mura Y, et al. A genome-editing strategy to treat beta-hemoglo- Daley H, et al. Validation of BCL11A as therapeutic target in
binopathies that recapitulates a mutation associated with a benign sickle cell disease: results from the adult cohort of a pilot/fea-
genetic condition. Nat Med. 2016;22:987–90. sibility gene therapy trail inducing sustained expression of fetal
155. Ye L, Wang J, Tan Y, Beyer AI, Xie F, Muench MO, et  al. hemoglobin using post-transcriptional gene silencing. Blood.
Genome editing using CRISPR-Cas9 to create the HPFH geno- 2019;134(Suppl. 2):Abstract LBA-5.
type in HSPCs: an approach for treating sickle cell disease and 173. Holmes MC, Reik A, Rebar EJ, Miller JC, Zhou Y, Zhang L, et al.
beta-thalassemia. Proc Natl Acad Sci USA. 2016;113:10661–5. A potential therapy for beta-thalassemia (ST-400) and sickle cell
156. Lattanzi A, Meneghini V, Pavani G, Amor F, Ramadier S, Felix disease (BIVV003). Blood. 2017;130(Suppl. 1):Abstract 2066.
T, et al. Optimization of CRISPR/Cas9 delivery to human hemat- 174. You L, Tong R, Li M, Liu Y, Xue J, Lu Y. Advancements and
opoietic stem and progenitor cells for therapeutic genomic rear- obstacles of CRISPR-Cas9 technology in translational research.
rangements. Mol Ther. 2019;27:137–50. Mol Ther MethodsClin Dev. 2019;13:359–70.
157. Lamsfus-Calle A, Daniel-Moreno A, Antony JS, Epting T, 175. Harrison C. First gene therapy for beta-thalassemia approved.
Heumos L, Baskaran P, et  al. Comparative targeting analy- Nat Biotechnol. 2019;37:1102–3.
sis of KLF1, BCL11A, and HBG1/2 in CD34(+) HSPCs by 176. Takahashi K, Yamanaka S. Induction of pluripotent stem cells
CRISPR/Cas9 for the induction of fetal hemoglobin. Sci Rep. from mouse embryonic and adult fibroblast cultures by defined
2020;10:10133. factors. Cell. 2006;126:663–76.
K. Sii‑Felice et al.

177. Hanna J, Wernig M, Markoulaki S, Sun CW, Meissner A, Cas- 194. Adikusuma F, Piltz S, Corbett MA, Turvey M, McColl SR,
sady JP, et al. Treatment of sickle cell anemia mouse model Helbig KJ, et al. Large deletions induced by Cas9 cleavage.
with iPS cells generated from autologous skin. Science. Nature. 2018;560:E8–9.
2007;318:1920–3. 195. Cullot G, Boutin J, Toutain J, Prat F, Pennamen P, Rooryck C,
178. Sebastiano V, Maeder ML, Angstman JF, Haddad B, Khayter et al. CRISPR-Cas9 genome editing induces megabase-scale
C, Yeo DT, et al. In situ genetic correction of the sickle cell chromosomal truncations. Nat Commun. 2019;10:1136.
anemia mutation in human induced pluripotent stem cells using 196. Broeders M, Herrero-Hernandez P, Ernst MPT, van der Ploeg
engineered zinc finger nucleases. Stem Cells. 2011;29:1717–26. AT, Pijnappel W. Sharpening the molecular scissors: advances
179. Sun N, Zhao H. Seamless correction of the sickle cell disease in gene-editing technology. iScience. 2020;23:100789.
mutation of the HBB gene in human induced pluripotent stem 197. Hess GT, Tycko J, Yao D, Bassik MC. Methods and applica-
cells using TALENs. Biotechnol Bioeng. 2014;111:1048–53. tions of CRISPR-mediated base editing in eukaryotic genomes.
180. Cai L, Bai H, Mahairaki V, Gao Y, He C, Wen Y, et al. A univer- Mol Cell. 2017;68:26–43.
sal approach to correct various HBB gene mutations in human 198. Molla KA, Yang Y. CRISPR/Cas-mediated base editing: tech-
stem cells for gene therapy of beta-thalassemia and sickle cell nical considerations and practical applications. Trends Bio-
disease. Stem Cells Transl Med. 2018;7:87–97. technol. 2019;37:1121–42.
181. Martin RM, Ikeda K, Cromer MK, Uchida N, Nishimura T, 199. Gaudelli NM, Komor AC, Rees HA, Packer MS, Badran
Romano R, et al. Highly efficient and marker-free genome edit- AH, Bryson DI, et  al. Programmable base editing of A*T
ing of human pluripotent stem cells by CRISPR-Cas9 RNP and to G*C in genomic DNA without DNA cleavage. Nature.
AAV6 donor-mediated homologous recombination. Cell Stem 2017;551:464–71.
Cell. 2019;24(821–828):e825. 200. Kim YB, Komor AC, Levy JM, Packer MS, Zhao KT, Liu DR.
182. Wang Y, Zheng CG, Jiang Y, Zhang J, Chen J, Yao C, et al. Increasing the genome-targeting scope and precision of base
Genetic correction of beta-thalassemia patient-specific iPS cells editing with engineered Cas9-cytidine deaminase fusions. Nat
and its use in improving hemoglobin production in irradiated Biotechnol. 2017;35:371–6.
SCID mice. Cell Res. 2012;22:637–48. 201. Komor AC, Kim YB, Packer MS, Zuris JA, Liu DR. Program-
183. Papapetrou EP, Lee G, Malani N, Setty M, Riviere I, Tiruna- mable editing of a target base in genomic DNA without double-
gari LM, et al. Genomic safe harbors permit high beta-globin stranded DNA cleavage. Nature. 2016;533:420–4.
transgene expression in thalassemia induced pluripotent stem 202. Nishida K, Arazoe T, Yachie N, Banno S, Kakimoto M, Tabata
cells. Nat Biotechnol. 2011;29:73–8. M, et al. Targeted nucleotide editing using hybrid prokaryotic and
184. Tubsuwan A, Abed S, Deichmann A, Kardel MD, Bartholoma vertebrate adaptive immune systems. Science. 2016;353:aaf8729.
C, Cheung A, et al. Parallel assessment of globin lentiviral trans- 203. Kim K, Ryu SM, Kim ST, Baek G, Kim D, Lim K, et al. Highly
fer in induced pluripotent stem cells and adult hematopoietic efficient RNA-guided base editing in mouse embryos. Nat Bio-
stem cells derived from the same transplanted beta-thalassemia technol. 2017;35:435–7.
patient. Stem Cells. 2013;31:1785–94. 204. Kim D, Kim DE, Lee G, Cho SI, Kim JS. Genome-wide target
185. Ma N, Shan Y, Liao B, Kong G, Wang C, Huang K, et  al. specificity of CRISPR RNA-guided adenine base editors. Nat
Factor-induced reprogramming and zinc finger nuclease-aided Biotechnol. 2019;37:430–5.
gene targeting cause different genome instability in beta-thalas- 205. Zuo E, Sun Y, Wei W, Yuan T, Ying W, Sun H, et al. Cytosine
semia induced pluripotent stem cells (iPSCs). J Biol Chem. base editor generates substantial off-target single-nucleotide vari-
2015;290:12079–89. ants in mouse embryos. Science. 2019;364:289–92.
186. Niu X, He W, Song B, Ou Z, Fan D, Chen Y, et al. Combining 206. Jin S, Zong Y, Gao Q, Zhu Z, Wang Y, Qin P, et al. Cytosine, but
single strand oligodeoxynucleotides and CRISPR/Cas9 to cor- not adenine, base editors induce genome-wide off-target muta-
rect gene mutations in beta-thalassemia-induced pluripotent stem tions in rice. Science. 2019;364:292–5.
cells. J Biol Chem. 2016;291:16576–85. 207. Grunewald J, Zhou R, Garcia SP, Iyer S, Lareau CA, Aryee MJ,
187. Ou Z, Niu X, He W, Chen Y, Song B, Xian Y, et al. The combina- et al. Transcriptome-wide off-target RNA editing induced by
tion of CRISPR/Cas9 and iPSC technologies in the gene therapy CRISPR-guided DNA base editors. Nature. 2019;569:433–7.
of human beta-thalassemia in mice. Sci Rep. 2016;6:32463. 208. Zhou C, Sun Y, Yan R, Liu Y, Zuo E, Gu C, et al. Off-target RNA
188. Song B, Fan Y, He W, Zhu D, Niu X, Wang D, et al. Improved mutation induced by DNA base editing and its elimination by
hematopoietic differentiation efficiency of gene-corrected beta- mutagenesis. Nature. 2019;571:275–8.
thalassemia induced pluripotent stem cells by CRISPR/Cas9 209. Rees HA, Wilson C, Doman JL, Liu DR. Analysis and minimiza-
system. Stem Cells Dev. 2015;24:1053–65. tion of cellular RNA editing by DNA adenine base editors. Sci
189. Martins GLS, Paredes BD, Azevedo CM, Sampaio GLA, Nonaka Adv. 2019;5:eaax5717.
CKV, Cavalcante BRR, et al. Generation of integration-free iPS 210. Gehrke JM, Cervantes O, Clement MK, Wu Y, Zeng J, Bauer
cell lines from three sickle cell disease patients from the state of DE, et  al. An APOBEC3A-Cas9 base editor with mini-
Bahia, Brazil. Stem cell Res. 2018;33:10–4. mized bystander and off-target activities. Nat Biotechnol.
190. Haapaniemi E, Botla S, Persson J, Schmierer B, Taipale J. 2018;36:977–82.
CRISPR-Cas9 genome editing induces a p53-mediated DNA 211. Liang P, Ding C, Sun H, Xie X, Xu Y, Zhang X, et al. Correction
damage response. Nat Med. 2018;24:927–30. of beta-thalassemia mutant by base editor in human embryos.
191. Doulatov S, Vo LT, Chou SS, Kim PG, Arora N, Li H, et al. Protein Cell. 2017;8:811–22.
Induction of multipotential hematopoietic progenitors from 212. Zeng J, Wu Y, Ren C, Bonanno J, Shen AH, Shea D, et al. Thera-
human pluripotent stem cells via respecification of lineage- peutic base editing of human hematopoietic stem cells. Nat Med.
restricted precursors. Cell Stem Cell. 2013;13:459–70. 2020;26:535–41.
192. Sugimura R, Jha DK, Han A, Soria-Valles C, da Rocha EL, Lu 213. Anzalone AV, Randolph PB, Davis JR, Sousa AA, Koblan LW,
YF, et al. Haematopoietic stem and progenitor cells from human Levy JM, et al. Search-and-replace genome editing without dou-
pluripotent stem cells. Nature. 2017;545:432–8. ble-strand breaks or donor DNA. Nature. 2019;576:149–57.
193. Ranjha L, Howard SM, Cejka P. Main steps in DNA double- 214. Worsham DN, Schuesler T, von Kalle C, Pan D. In vivo gene
strand break repair: an introduction to homologous recombina- transfer into adult stem cells in unconditioned mice by in situ
tion and related processes. Chromosoma. 2018;127:187–214. delivery of a lentiviral vector. Mol Ther. 2006;14:514–24.
Innovative Therapies for Hemoglobin Disorders

215. Wang X, Shin SC, Chiang AF, Khan I, Pan D, Rawlings DJ, et al. ameliorates murine thalassemia intermedia. J Clin Investig.
Intraosseous delivery of lentiviral vectors targeting factor VIII 2019;129:598–615.
expression in platelets corrects murine hemophilia A. Mol Ther. 228. Beard BC, Trobridge GD, Ironside C, McCune JS, Adair JE,
2015;23:617–26. Kiem HP. Efficient and stable MGMT-mediated selection of
216. Habi O, Girar J, Bourdages V, Delisle MC, Carreau M. Correc- long-term repopulating stem cells in nonhuman primates. J Clin
tion of Fanconi anemia group C hematopoietic stem cells follow- Investig. 2010;120:2345–54.
ing intrafemoral gene transfer. Anemia. 2010;2010:947816. 229. Li C, Mishra AS, Gil S, Wang M, Georgakopoulou A, Papa-
217. Andreani M, Manna M, Lucarelli G, Tonucci P, Agostinelli F, yannopoulou T, et  al. Targeted integration and high-level
Ripalti M, et al. Persistence of mixed chimerism in patients trans- transgene expression in AAVS1 transgenic mice after in vivo
planted for the treatment of thalassemia. Blood. 1996;87:3494–9. HSC transduction with HDAd5/35 ++  vectors. Mol Ther.
218. Persons DA, Allay ER, Sabatino DE, Kelly P, Bodine DM, Nien- 2019;27:2195–212.
huis AW. Functional requirements for phenotypic correction of 230. Kosicki M, Tomberg K, Bradley A. Repair of double-strand
murine beta-thalassemia: implications for human gene therapy. breaks induced by CRISPR-Cas9 leads to large deletions and
Blood. 2001;97:3275–82. complex rearrangements. Nat Biotechnol. 2018;36:765–71.
219. Finkelshtein D, Werman A, Novick D, Barak S, Rubinstein 231. Li C, Psatha N, Sova P, Gil S, Wang H, Kim J, et al. Reacti-
M. LDL receptor and its family members serve as the cellular vation of gamma-globin in adult beta-YAC mice after ex vivo
receptors for vesicular stomatitis virus. Proc Natl Acad Sci USA. and in vivo hematopoietic stem cell genome editing. Blood.
2013;110:7306–11. 2018;131:2915–28.
220. DePolo NJ, Reed JD, Sheridan PL, Townsend K, Sauter SL, 232. Saydaminova K, Ye X, Wang H, Richter M, Ho M, Chen H,
Jolly DJ, et al. VSV-G pseudotyped lentiviral vector particles et al. Efficient genome editing in hematopoietic stem cells with
produced in human cells are inactivated by human serum. Mol helper-dependent Ad5/35 vectors expressing site-specific endo-
Ther. 2000;2:218–22. nucleases under microRNA regulation. Mol Ther Methods Clin
221. Frecha C, Costa C, Negre D, Amirache F, Trono D, Rio P, Dev. 2015;1:14057.
et al. A novel lentiviral vector targets gene transfer into human 233. Li C, Psatha N, Gil S, Wang H, Papayannopoulou T, Lieber A.
hematopoietic stem cells in marrow from patients with bone mar- HDAd5/35(++) adenovirus vector expressing anti-CRISPR pep-
row failure syndrome and in vivo in humanized mice. Blood. tides decreases CRISPR/Cas9 toxicity in human hematopoietic
2012;119:1139–50. stem cells. Mol Ther Methods Clin Dev. 2018;9:390–401.
222. Milani M, Annoni A, Bartolaccini S, Biffi M, Russo F, Di 234. Bahal R, Ali McNeer N, Quijano E, Liu Y, Sulkowski P, Turchick
Tomaso T, et al. Genome editing for scalable production of A, et al. In vivo correction of anaemia in beta-thalassemic mice
alloantigen-free lentiviral vectors for in  vivo gene therapy. by gammaPNA-mediated gene editing with nanoparticle deliv-
EMBO Mol Med. 2017;9:1558–73. ery. Nat Commun. 2016;7:13304.
223. Nilsson M, Karlsson S, Fan X. Functionally distinct subpopula- 235. Makadia HK, Siegel SJ. Poly lactic-co-glycolic acid (PLGA)
tions of cord blood CD34 + cells are transduced by adenoviral as biodegradable controlled drug delivery carrier. Polymers.
vectors with serotype 5 or 35 tropism. Mol Ther. 2004;9:377–88. 2011;3:1377–97.
224. Wang H, Liu Y, Li Z, Tuve S, Stone D, Kalyushniy O, et al. 236. Ricciardi AS, Bahal R, Farrelly JS, Quijano E, Bianchi AH, Luks
In  vitro and in  vivo properties of adenovirus vectors with VL, et al. In utero nanoparticle delivery for site-specific genome
increased affinity to CD46. J Virol. 2008;82:10567–79. editing. Nat Commun. 2018;9:2481.
225. Richter M, Saydaminova K, Yumul R, Krishnan R, Liu J, Nagy 237. Lamsfus-Calle A, Daniel-Moreno A, Urena-Bailen G, Raju
EE, et al. In vivo transduction of primitive mobilized hematopoi- J, Antony JS, Handgretinger R, et al. Hematopoietic stem cell
etic stem cells after intravenous injection of integrating adenovi- gene therapy: the optimal use of lentivirus and gene editing
rus vectors. Blood. 2016;128:2206–17. approaches. Blood Rev. 2020;40:100641.
226. Wang H, Richter M, Psatha N, Li C, Kim J, Liu J, et al. A com- 238. Imren S, Fabry ME, Westerman KA, Pawliuk R, Tang P, Ros-
bined in vivo HSC transduction/selection approach results in ten PM, et al. High-level beta-globin expression and preferred
efficient and stable gene expression in peripheral blood cells in intragenic integration after lentiviral transduction of human cord
mice. Mol Ther Methods Clin Dev. 2018;8:52–64. blood stem cells. J Clin Investig. 2004;114:953–62.
227. Wang H, Georgakopoulou A, Psatha N, Li C, Capsali C,
Samal HB, et al. In vivo hematopoietic stem cell gene therapy

You might also like