Effective medical management for sickle cell disease (SCD) remains elusive. As a prevalent and severe monogenic disorder, SCD has been long considered a logical candidate for gene therapy. Significant progress has been made in moving toward this goal. These efforts have provided substantial insight into the natural regulation of the globin genes and illuminated challenges for genetic manipulation of the hematopoietic system. The initial γ-retroviral vectors, next-generation lentiviral vectors, and novel genome engineering and gene regulation approaches each share the goal of preventing erythrocyte sickling. After years of preclinical studies, several clinical trials for SCD gene therapies are now open. This review focuses on progress made toward achieving gene therapy, the current state of the field, consideration of factors that may determine clinical success, and prospects for future development.

The single base substitution A-T (β6Glu→Val) in the first exon of the β-globin gene is the defining mutation of the sickle allele. Individuals homozygous for the mutation have the classical sickle cell disease (SCD) genotype. All pathophysiological consequences ensue from this monogenic defect. The amino acid substitution in βS-globin allows formation of defective hemoglobin tetramers that polymerize upon deoxygenation.1  Hemoglobin polymerization causes affected red blood cells (RBCs) to lose normal deformability and adopt the archetypal sickle shape. Rigid sickle RBCs prematurely breakdown, damage endothelia, and occlude vasculature, leading to a cascade of hemolysis, ischemia, inflammation, and endothelial injury.2  Patients are afflicted with intensely painful episodes, susceptibility to infection, end-organ injury, and early mortality among other sequelae.3,4  Despite its status as a “monogenic” disease, SCD is surprisingly variable in its clinical severity.

At present, the only curative treatment of SCD is allogeneic hematopoietic stem cell (HSC) transplantation. 6-year disease-free survival of >90% has been reported for transplants from HLA-matched sibling donors.5  However, in the United States, <14% of patients have a matched sibling donor.6  Transplants with matched unrelated donors are limited by donor availability and immunologic barriers, such as graft rejection and graft-versus-host disease.3  Attempts to extend allogeneic transplant for SCD to alternative donor sources is an area of ongoing effort.7  The SCD community has been cautious to embrace allogeneic HSC transplant in part given its short-term morbidity and mortality risks, though nonmyeloablative preparative regimens may help mitigate these risks.8-11  Given that the current clinical approach to SCD is largely reliant upon supportive care and hydroxyurea,12  the development of definitive therapies based on genetic manipulation of autologous HSCs would constitute a major advance.

Gene therapy has long been proposed as a potential cure for SCD.13  The permanent delivery of a corrective or antisickling gene cassette into long-term, repopulating HSCs could allow for the production of corrected RBCs for the life of the patient. More recent advances in manipulating the human genome suggest nonviral forms of genome editing could also be therapeutically relevant. This review will focus on the initial development of viral vectors for SCD gene therapy via gene addition, discuss current vectors in detail including those in clinical trials, highlight novel approaches to SCD genetic manipulation, and discuss key parameters for achieving clinical success. Although outside the scope of this article, it is notable that many of the approaches discussed below are also relevant to β-thalassemia, a disorder of inadequate β-globin production.

Over the last 3 decades, several groups have worked toward achieving efficient and safe gene transfer to HSCs for SCD as well as other genetic disorders.14  In order for gene therapy for SCD to become a reality, 2 main objectives must be achieved: (1) safe and efficient gene transfer or correction of long-term repopulating HSCs and (2) high-level, appropriately regulated, stable gene expression. With current progress at the bench and in the clinic, these goals now appear within reach. The long path to the clinic for SCD gene therapies has been paved by landmark discoveries that have provided important insights into the developmental regulation of the β-globin gene cluster.

β-globin–expressing vectors

Initial attempts to develop integrating γ-retroviral vectors for β-globin gene transfer revealed challenges to attaining efficient, high-level erythroid-specific expression following HSC transduction. Long terminal repeat (LTR)-driven transgene expression of β-globin resulted in little or no β-globin following hematopoietic reconstitution in mice.15-18  The principal finding that allowed for high-level erythroid-specific expression in vivo was the discovery and characterization of the β-globin locus control region (LCR). This essential regulatory element 40 to 60 kb upstream of the β-globin gene, which is demarcated by 5 DNase I hypersensitive (HS) sites, contains potent erythroid-specific enhancers that function in concert with elements flanking the downstream globin genes.19-21  Careful mapping of the positions and activities of these HS sites established the need for powerful erythroid-specific enhancer elements for high-level erythroid expression.19,22-27 

A second challenge encountered with retroviral vectors was the recognition that they are subject to expression variegation and silencing.28,29  Furthermore, their powerful regulatory elements may alter expression of endogenous genes, particularly those adjacent to their insertion site.30  Insulators that shield internal sequences from repressive effects of neighboring chromatin (barrier activity) or decrease trans-activation by enhancers (enhancer-blocking activity), such as the chicken hypersensitive site-4 (cHS4), ankyrin, and FB (second DNaseI hypersensitive site from 5′ of cHS4 [FII] and blocking element alpha/delta 1 [BEAD]) insulators, may promote the safety and efficacy of integrating vectors.31-38  Balancing the benefits of expression control with the negative impact of large insulators on vector titer has been an ongoing issue. Recently, small enhancer-blocking insulators have been described that do not diminish viral titer in reporter constructs.39 

As evidence accumulated that lentiviral vectors offer important advantages over γ-retroviral vectors (including the ability to transduce nondividing HSCs, a larger capacity for DNA, stable transmission of complicated cargo, and a safer integration profile), the gene therapy field has largely transitioned to the use of these vectors.14,40-45  The tendency for γ-retroviral vectors to integrate near the transcriptional start sites of genes exacerbates their potential to alter gene expression. In contrast, lentiviruses tend to insert more randomly with a bias toward integration in gene bodies of expressed genes.43,46  In addition, vectors with a self-inactivating (SIN) design have become standard. SIN vectors have a deletion in the U3 region of the 3′ LTR, which is copied into the 5′ LTR upon reverse transcription. This modification minimizes transactivation of nearby genes, thereby limiting potential for insertional oncogenesis47  (Figure 1). Various lentiviral vectors carrying human β-globin have been used to correct disease models in both β-thalassemia and SCD.48-51 

Figure 1

Vector schematic: general organization of proviral form of lentiviral vectors for gene therapy for SCD. (A) Antisickling β-globin vector containing the T87Q mutation. The SIN lentiviral vector is noninsulated and contains a modified LCR with HS2, HS3, and HS4. (B) Antisickling β-globin vector containing the G16D, E22A, and T87Q mutations (AS3). The SIN lentiviral vector has an FB element for enhancer-blocking activity and contains a modified LCR with HS2, HS3, and HS4. (C) A γ/β-globin hybrid vector containing the coding regions of γ-globin and the noncoding regions of β-globin. γ-globin (dark blue boxes), β-globin, or antisickling β-globin gene cassettes (blue boxes) with β-globin 5′ and 3′ untranslated regions (short blue boxes) under the control of a β-globin promoter (blue arrow) and modified β-globin LCR (red). Gene cassette is in reverse with respect to viral transcription to avoid aberrant splicing during packaging due to presence of globin intronic sequences (light blue boxes). All vectors are SIN (ΔU3) lentiviral vectors (green). Vectors are intended as basic schematics of gene therapy clinical vectors for SCD and are neither drawn to scale nor are all details included. ψ, packaging signal; cPPT, central polypurine tract; HS, DNase I hypersensitive site; LTR, long terminal repeats (U3, R, and U5); RRE, rev-responsive element; WPRE, woodchuck hepatitis virus posttranscriptional regulator element.

Figure 1

Vector schematic: general organization of proviral form of lentiviral vectors for gene therapy for SCD. (A) Antisickling β-globin vector containing the T87Q mutation. The SIN lentiviral vector is noninsulated and contains a modified LCR with HS2, HS3, and HS4. (B) Antisickling β-globin vector containing the G16D, E22A, and T87Q mutations (AS3). The SIN lentiviral vector has an FB element for enhancer-blocking activity and contains a modified LCR with HS2, HS3, and HS4. (C) A γ/β-globin hybrid vector containing the coding regions of γ-globin and the noncoding regions of β-globin. γ-globin (dark blue boxes), β-globin, or antisickling β-globin gene cassettes (blue boxes) with β-globin 5′ and 3′ untranslated regions (short blue boxes) under the control of a β-globin promoter (blue arrow) and modified β-globin LCR (red). Gene cassette is in reverse with respect to viral transcription to avoid aberrant splicing during packaging due to presence of globin intronic sequences (light blue boxes). All vectors are SIN (ΔU3) lentiviral vectors (green). Vectors are intended as basic schematics of gene therapy clinical vectors for SCD and are neither drawn to scale nor are all details included. ψ, packaging signal; cPPT, central polypurine tract; HS, DNase I hypersensitive site; LTR, long terminal repeats (U3, R, and U5); RRE, rev-responsive element; WPRE, woodchuck hepatitis virus posttranscriptional regulator element.

Close modal

γ-Globin and other antisickling vectors

In addition to vectors containing β-globin expression cassettes, vectors containing alternative globin genes, such as γ-globin or hybrid β/γ-globin, have been developed. The rationale for γ-globin gene addition is based on the observation that fetal hemoglobin (HbF, α2γ2) is a more potent antisickling hemoglobin as compared with adult hemoglobin (α2β2).52  It has long been appreciated that SCD patients with increased levels of HbF have an attenuated clinical course.53,54  There appears to be a dose-response relationship, where even the upper quartile of SCD patients with HbF levels >8.6% have extended survival, Arab-Indian haplotype patients with ∼20% HbF have mild clinical phenotypes, and sickle hemoglobin/hereditary persistence of HbF patients with ∼30% HbF are essentially free of disease manifestations. Therefore, gene addition with γ-globin may lead to a therapeutic effect with a relatively lower requirement for gene expression as compared with β-globin.

Several early configurations of γ-globin vectors, like contemporaneous β-globin vectors, were prone to low erythroid expression, genetic recombination, expression variegation, and silencing despite several different designs.29,55  Newer-generation lentiviral vectors carrying γ-globin cassettes have been described,22,56,57  including 1 with γ-globin coding sequences and β-globin regulatory elements which could effectively correct the Berkeley SCD mouse model following reduced-intensity conditioning.58,59 

Several groups have developed antisickling globin vectors that express modified β-globin capable of conferring resistance to sickle hemoglobin polymerization. An advantage of these vectors is the ease with which the contribution of various globin chains can be assessed. Because the erythropoietic stress associated with HSC transplant may lead to elevated endogenous γ-globin expression, the capacity to distinguish exogenous from endogenous globin chains facilitates evaluation of gene addition efficacy. Synthetic β-globin variants with mutations affecting axial and lateral contacts in the sickle fiber show that substitutions such as E22A and T87Q confer potent antisickling activity.60  Expression of the T87Q mutation alone from a lentiviral vector resulted in pancellular, erythroid-specific expression in a murine model of SCD at levels capable of correcting the disease pathophysiology.51  Lentiviral expression of a triple-mutant β-globin (G16D, E22A, and T87Q) with increased affinity for α-globin subunits (so-called βAS3) also corrected SCD in a murine model.61,62  An insulated βAS3 SIN lentiviral vector (βAS3-FB) efficiently transduces human SCD patient bone marrow CD34+ cells and directs exogenous globin levels within an anticipated therapeutic range in in vitro erythroid differentiation and xenograft transplant models.63 

A comparison of the βAS3-FB vector and the V5m3-400 vector (a γ-globin–expressing lentiviral vector) found equivalent antisickling hemoglobin levels and phenotypic RBC correction, suggesting that either vector might be suitable for clinical development.64  This type of head-to-head comparison is an important example of collaborative effort. A limitation within the gene therapy field has been that individual groups often take different approaches in terms of vector, gene cargo, regulatory elements, insulator elements, viral production, cellular transduction, experimental models, and other variables. Although in the end this may help accelerate innovation, it complicates isolating the effects of individual parameters.

Three groups have open clinical trials for SCD gene therapy registered on clinicaltrials.gov65  (Table 1; Figure 1). Bluebird Bio, a biotechnology company, is the first to treat a SCD patient with gene therapy. Their vector, LentiGlobin BB305,66  expresses an antisickling β-globin (T87Q). The vector utilizes modified LCR and β-globin promoter regulatory elements and does not include insulator sequences. Treatment of the first SCD gene therapy patient, a 13-year-old subject, resulted in a vector copy number of 2.4 copies per peripheral blood leukocyte and 24% antisickling (exogenous) hemoglobin 4.5 months following autologous hematopoietic transplantation with transduced CD34+ cells with no adverse events reported.67  Two other trials are open: 1 at the University of California, Los Angeles63,64  and the other at Cincinnati Children’s Hospital Medical Center,14,65  though neither group has yet reported patient results.

Table 1

Open sickle cell gene therapy trials

GroupVector*ConditioningEnrollmentClinicalTrials.gov identifierStatus
Gene cassetteInsulator
Bluebird Bio βA-T87Q-globin None Busulfan: 12.8 mg/kg IV, pK adjusted Ages 5-37 y, up to 7 subjects; adult, up to 8 subjects NCT02140554; NCT02151526 Ongoing: at least 1 patient treated67  
University of California, Los Angeles βAS3-globin FB Busulfan: 12.8 mg/kg IV, pK adjusted Adult, up to 6 subjects NCT02247843 Open 
Cincinnati Children’s Hospital Medical Center γ-globin None Melphalan: 140 mg/m2 Adult to age 35 y, up to 10 subjects NCT02186418 Open 
GroupVector*ConditioningEnrollmentClinicalTrials.gov identifierStatus
Gene cassetteInsulator
Bluebird Bio βA-T87Q-globin None Busulfan: 12.8 mg/kg IV, pK adjusted Ages 5-37 y, up to 7 subjects; adult, up to 8 subjects NCT02140554; NCT02151526 Ongoing: at least 1 patient treated67  
University of California, Los Angeles βAS3-globin FB Busulfan: 12.8 mg/kg IV, pK adjusted Adult, up to 6 subjects NCT02247843 Open 
Cincinnati Children’s Hospital Medical Center γ-globin None Melphalan: 140 mg/m2 Adult to age 35 y, up to 10 subjects NCT02186418 Open 
*

All vectors described are SIN lentiviruses.

Enhancer blocker only.

With these trials now in the clinic, it is imperative to establish robust means for assessing efficacy. Although vector marking, transgene expression, and hematopoietic reconstitution have been established as important measures in other lentiviral-based gene therapy trials,68  determinants of success for SCD gene therapy should also include objective clinical criteria, such as survival, hospitalizations, painful crises, and other SCD-associated events (such as acute chest syndrome, stroke, etc) as well as patient-reported outcomes. Although these clinical outcomes will require evaluation of numerous patients over an extended period of time, they will constitute the ultimate evidence for safety and efficacy of gene therapy.

The advent of genome engineering has expanded possible genetic therapy options for many hematopoietic diseases. Given their prevalence and severity, the hemoglobinopathies have attracted considerable attention. Targeted nucleases, such as zinc-finger nucleases (ZFNs), transcription-activator like effector nucleases (TALENs), meganucleases, and the CRISPR (clustered, regularly interspaced palindromic repeats)-associated nuclease Cas9, each have programmable DNA-binding modules that allow introduction of site-specific double-strand breaks into the human genome69-71  (Table 2). Unlike the other nucleases, Cas9 utilizes RNA to guide target sequence recognition. Following cleavage, subsequent endogenous eukaryotic repair mechanisms may be simplified into 2 major pathways: nonhomologous end joining (NHEJ) and homology-directed repair (HDR). The ability to direct either of these pathways to produce predictable modifications could be of therapeutic relevance for SCD69,72,73  (Figure 2). Unlike gene addition strategies, genome engineering approaches only rely on a transient ex vivo intervention and do not result in permanent insertion of foreign DNA into the genome.

Table 2

Targeted nucleases

NucleaseDNA-binding domainActive configurationCleaving enzyme; break characteristicTarget range
Homing endonucleases (meganucleases) 5 families: LAGLIDADG, GIY-YIG, HNH, His-Cys box, and PD- (D/E)XK Homodimeric or monomeric 3′ overhang 14-40 bases 
ZFN Zinc-finger motif Protein dimer FokI; 5′ overhang 9-18 bases per monomer 
TALEN Repeat variable diresidue Protein dimer FokI; 5′ overhang 12-31 bases per monomer 
CRISPR/Cas9* CRISPR-RNA and trans-activating RNA RNA/protein complex Cas9; Blunt 17-20 bases 
NucleaseDNA-binding domainActive configurationCleaving enzyme; break characteristicTarget range
Homing endonucleases (meganucleases) 5 families: LAGLIDADG, GIY-YIG, HNH, His-Cys box, and PD- (D/E)XK Homodimeric or monomeric 3′ overhang 14-40 bases 
ZFN Zinc-finger motif Protein dimer FokI; 5′ overhang 9-18 bases per monomer 
TALEN Repeat variable diresidue Protein dimer FokI; 5′ overhang 12-31 bases per monomer 
CRISPR/Cas9* CRISPR-RNA and trans-activating RNA RNA/protein complex Cas9; Blunt 17-20 bases 
*

Characteristics listed here describe Streptococcus pyogenes Cas9. Other Cas9 variants and orthologs have related but not identical properties.

Figure 2

Strategies for gene therapy for SCD: schematic overview of various approaches for correcting the sickle phenotype via gene therapy. Gene correction: targeted genome engineering leads to correction of the sickle mutation such that βS is repaired as βA. HbF induction: multiple strategies for induction of γ-globin expression include shRNA-mediated knockdown of BCL11A, targeted disruption of the +58 DNase I HS site in the BCL11A erythroid-specific enhancer, and forced chromatin looping to promote association of the β-globin LCR with the γ-globin genes. Gene addition: integrating lentiviral vector carrying a β-globin, γ-globin, or antisickling β-globin cassette. Ldb1, transcription factor; ZF/SA, zinc-finger self-association domain.

Figure 2

Strategies for gene therapy for SCD: schematic overview of various approaches for correcting the sickle phenotype via gene therapy. Gene correction: targeted genome engineering leads to correction of the sickle mutation such that βS is repaired as βA. HbF induction: multiple strategies for induction of γ-globin expression include shRNA-mediated knockdown of BCL11A, targeted disruption of the +58 DNase I HS site in the BCL11A erythroid-specific enhancer, and forced chromatin looping to promote association of the β-globin LCR with the γ-globin genes. Gene addition: integrating lentiviral vector carrying a β-globin, γ-globin, or antisickling β-globin cassette. Ldb1, transcription factor; ZF/SA, zinc-finger self-association domain.

Close modal

Gene correction

Site-specific gene correction of the causal sickle β-globin mutation in HSCs would seem the most straightforward strategy for gene therapy. In principle, this approach only requires transient delivery of nuclease and repair template to achieve correction, thereby avoiding introduction of foreign DNA and associated risk of insertional oncogenesis as occurs with viral vectors. In addition, direct correction of the sickle mutation has the advantage of preserving endogenous globin regulatory mechanisms to achieve the appropriate balance of globin chains during erythropoiesis.

Correction of the sickle mutation by targeted nucleases followed by HDR in various cell types has been demonstrated,74,75  including reports of correction of induced pluripotent stem cells from both mice and humans.75-78  In addition, oligonucleotide-based gene therapy strategies, such as triplex-forming peptide nucleic acids which rely on HDR but not on the initial formation of a double-stranded break, have achieved low-frequency correction of the SCD mutation.79-82  Although these approaches offer the possibility to determine genome modification specificity on a clonal level, derivation of functional HSCs from pluripotent cells remains a great challenge.14,83,84  Recently, correction in human HSCs was reported. However, the rates of correction in long-term HSCs were well below levels necessary for therapeutic benefit.85  A similar finding of preferential utilization of NHEJ in HSCs (despite relatively robust HDR repair in unfractionated CD34+ hematopoietic stem and progenitor cells) has been observed in experiments attempting to correct the SCID-X1 mutation in human HSCs.86  A simple explanation of this observation may be that the HDR pathway is restricted to the S and G2 phases of the cell cycle when sister chromatids are available as donor repair template sequences. In contrast, HSCs, which are largely quiescent cells, rely mainly on NHEJ.87 

Reactivation of fetal hemoglobin

Although technical advances may be required to realize therapeutic HDR in HSCs, NHEJ appears to be a robust repair pathway in these cells. A clinical trial of genetic disruption by NHEJ in T cells has already been conducted using ZFN targeting of CCR5 in HIV-infected subjects demonstrating safety and genetic efficacy.88  A growing understanding of mechanisms of developmental globin gene regulation suggests alternate strategies that could be appropriate for genetic disruption rather than repair.89  These approaches are neither gene addition nor gene correction per se but rather target a modifier of disease severity. Although this strategy of modifying a modifier might seem somewhat indirect, it actually capitalizes on knowledge of naturally occurring protective mechanisms in SCD (ie, elevated HbF level).

Genome-wide association studies have implicated 3 key loci (BCL11A, the intergenic region between HBS1L and MYB, and the β-globin cluster itself) in regulation of HbF level.90,91  BCL11A is critically required for HbF repression in primary human erythroid cells and in transgenic mice.92-94  The fact that RBC production remains ostensibly normal in Bcl11a knockouts is surprising for a critical transcription factor and distinguishes BCL11A from many other potential targets that have broad requirements in erythropoiesis. Erythroid-specific loss of BCL11A in mouse models of SCD is sufficient to reverse the hematologic and pathologic manifestations of disease, thereby validating BCL11A as a therapeutic target.94  As BCL11A also has important roles in B-cell and dendritic cell development, its complete knockout in HSCs could impair production of immune cells.95-97  Furthermore, BCL11A is expressed in HSCs and early progenitor cells, suggesting it might play additional roles in homeostasis of the hematopoietic system.96  The identification of the intronic erythroid-specific BCL11A enhancer as the site of common genetic variants associated with HbF level has suggested that disruption of the erythroid enhancer could circumvent negative effects of impairment of BCL11A in nonerythroid contexts by only affecting its expression in erythroid precursors.98  Recently, a systematic CRISPR screen saturating the DNase I HS sites with guide RNAs identified critical minimal sequences within the +58 DHS required for erythroid expression of BCL11A and adult-stage repression of HbF. Single cleavages targeting these crucial enhancer sequences were found to be sufficient for robust HbF induction in primary human erythroid precursors.99  A complementary approach targeting DNase I footprints within the BCL11A erythroid enhancer revealed similar critical minimal sequences that could be disrupted by ZFNs.100  Production of single cleavages by targeted nucleases is an appealing genome engineering strategy in that it is relatively simple and takes advantage of the endogenous error-prone NHEJ repair mechanism that is robust in HSCs.

An alternative approach to BCL11A knockout is reduction of BCL11A expression below a critical threshold through RNA interference.92,101  A novel short hairpin RNA (shRNA) vector was recently reported, capable of potent BCL11A knockdown in human erythroid precursors. This vector used a Pol II microRNA architecture, suggesting that knockdown controlled by lineage-specific regulatory elements could be a strategy for restricting gene inhibition to erythroid cells.102  Similar to gene addition approaches, this strategy would rely on both HSC delivery by integrating vectors and sustained erythroid expression, in this case of the inhibitory RNA.

The ability to program DNA-binding specificity of synthetic factors also allows for targeted gene regulation.103  Expression of the self-association domain of Ldb1 (a transcription factor involved in chromatin looping of the LCR) linked to artificial zinc fingers recognizing the γ-globin promoter in adult human erythroblasts resulted in robust increases in γ-globin.104  These experiments suggest that expression of such a construct could lead to therapeutic increases in γ-globin expression at the expense of βS-globin in SCD. By necessity, this strategy would rely on long-term expression from integrating vectors in HSCs. Perhaps the requirements for lower-level expression could be a safety advantage over classic globin gene addition. The potency of this effect, the continuous expression of foreign protein, as well as specificity for globin gene expression as compared with other transcriptional impacts, are important issues to be considered before further clinical development.

Combined approaches (globin gene addition plus HbF induction) for SCD gene therapy have also been explored.105,106  These multipronged approaches might prove superior to single modality vectors, but their effects and potential off-target issues would need to be investigated thoroughly. As knowledge expands of mechanisms repressing HbF in the adult stage, additional targets may emerge. Although not explicitly discussed here, similar strategies as above could be contemplated to target other HbF repressors or to mimic naturally occurring variation at the β-globin cluster associated with hereditary persistence of HbF.107-109 

Irrespective of which gene therapy strategy is considered for SCD, there are several common critical factors that will determine success or failure in the clinical setting.

Adequate cell dose

Autologous engraftment is highly dependent on the dose of CD34+ cells.110  Mobilization of CD34+ cells by granulocyte colony-stimulating factor treatment followed by peripheral blood apheresis is most commonly used in conventional bone marrow transplant and in hematopoietic cell gene therapy. However, granulocyte colony-stimulating factor mobilization has been associated with severe adverse events and even mortality in SCD patients.111,112  Thus, gene therapy approaches for SCD have focused on relatively more invasive bone marrow harvest as cell source of CD34+ cells. Even this procedure is not completely without risk, given anesthesia dangers in SCD. Use of plerixafor for CD34+ cell mobilization for gene therapy has been widely investigated, including in β-thalassemia patients. Clinical trials for plerixafor mobilization in SCD are planned but results have not yet been reported.113  Safe and effective collection of HSCs will be essential for the subsequent repopulation of the bone marrow.

Assessment of true HSCs and their modification

CD34+ cells, either mobilized or harvested from marrow, are a mixture of many progenitors and few true HSCs. A current challenge in the field is that it is not yet possible to determine the fraction of modified HSCs based on the level of marking of unfractionated CD34+ cells. The current laboratory standard is to perform engraftment of immunodeficient mice, which is both time-consuming and costly. Of greater concern is that these xenograft systems do not fully support human hematopoiesis (and in particular do not support human erythropoiesis). Secondary transplant can be very difficult to achieve, limiting the ability to truly test long-term self-renewal activity. A promising approach is to test modification within subpopulations of CD34+ cells enriched for HSCs as defined by immunophenotype, such as by expression of CD38, CD133, CD90, and CD45RA,114-117  though this may require short-term culture.118  Targeted transfer into subpopulations of CD34+ cells enriched for HSCs have also been investigated using a vector with tropism for CD133.119  Efforts at in vitro expansion of human HSCs, while preserving their stem cell qualities, might advance efforts at HSC modification.120-123  Careful correlation of laboratory parameters of the modified graft with clinical outcomes in gene therapy subjects may eventually inform optimal surrogate markers to predict HSC activity.

High levels of modified-HSC engraftment

Another important consideration for SCD gene therapy is the fraction of modified cells as compared with the entire hematopoietic compartment. Based on allogeneic HSC transplant for SCD patients, stable donor chimerism of 10% to 30% may provide significant clinical improvement.6,124-126  These results reflect that the survival advantage of nonsickle cells both at the level of erythrocytes (reduced hemolysis) and erythroblasts (reduced ineffective erythropoiesis) leads to selective survival of corrected RBCs out of proportion to overall hematopoietic engraftment.127  Stable engraftment of 10% to 60% gene-modified HSCs has been observed in other clinical trials with fully cytoablative conditioning.44,45,128,129  This degree of engraftment would appear to approximate or exceed the threshold required for substantial clinical benefit. Despite the above inferences, there remains some uncertainty as to the minimal level of modification required to derive curative outcome for SCD. It will be critically important for careful evaluation of the fraction of modified cells within erythroid and nonerythroid progenitor cells in subjects of SCD gene therapies to empirically derive the quantitative relationship between genetic modification of HSCs and hematologic and clinical outcomes.

Although the degree of myeloablation has been shown to be proportional to the degree of autologous engraftment and thus to correction efficiency,68,110,130  fully myeloablative regimens with alkylating chemotherapy carry nontrivial risks to patients, especially to SCD patients who may already have experienced some disease-related end-organ compromise. A goal for the field would be to develop novel approaches to achieve engraftment while limiting preparative conditioning-related toxicity. One approach that appears promising in mouse models is transient inhibition of stem cell factor/c-Kit signaling.131-133 

Safety

Independent of the type of gene therapy, safety must be paramount. For integrating viral vectors, safety evaluation focuses on insertional mutagenesis with risk for clonal dominance or cellular transformation. Clinical gene therapy trials using γ-retroviral vectors for other diseases have been accompanied unfortunately by severe adverse events due to negative effects of the integration site of the vector.129,134-136 

It is estimated that successful gene therapy with current methods results in thousands of viral insertions per recipient, each insertion measured as the hematopoietic output of a uniquely marked clone.137-139  These many insertions carry an inherent risk for insertional oncogenesis or loss of appropriate gene regulation by mechanisms such as aberrant splicing.128,136  Although safety may be modulated by vector design, including the use of insulators, optimal cellular promoters, and lentiviral vectors,140,141  it may not be possible to fully eliminate risk from thousands of semirandom genomic viral integrations.

A gene therapy clinical trial for β-thalassemia using a γ-retroviral vector resulted in the outgrowth of a myeloid-biased hematopoietic clone associated with integration-mediated transcriptional activation of HMGA2 due to aberrant splicing. The affected patient achieved transfusion independence without leukemia development, so for this patient the clonal dominance may have provided some advantage in that the corrected cells were present at a high frequency.128,142  Of note, the causal cryptic 3′ splice signal was introduced by rearranged sequences from a tandem repeat of the cHS4 insulator region of the vector, turning an intended protective feature into one apparently compromising safety. The myeloid bias of the long-lived corrected hematopoietic progenitor provocatively suggests that modification of long-lived lineage-restricted progenitors (in addition to true HSCs) could prove therapeutically beneficial in certain contexts.143,144 

For genome engineering approaches, evaluation of off-target mutagenesis will be important preclinical criteria prior to embarking on clinical studies. Typically these approaches have used computational predictions and/or empirically determined sites of chromatin binding to prioritize a limited set of susceptible genomic sites for analysis.145  More recently, unbiased methods have been developed that allow for genome-wide assessment of off-target mutagenesis.146-151  Cellular products for autologous HSC transplant would typically include 108 or more cells, so it is virtually impossible to measure very rare off-target mutations. The significance of modifications at most genomic sites is uncertain, and likely the vast majority of modifications would be functionally neutral. The occasional somatic mutations that accompany normal aging and environmental exposures also raise the question of baseline level of mutagenesis.152  These challenges to assaying and interpreting genetic perturbations emphasize the importance of functional readouts of safety. In the end, investigators need to maintain humility that any laboratory safety testing will likely be incompletely predictive of in vivo behavior of modified cells in humans.153  As with all novel therapeutic approaches, meticulous clinical trials will be essential to ensure the success and safety of genetic approaches to SCD.

Although several of the initial hurdles to SCD gene therapy appear to have been overcome, it is prudent to recognize barriers that remain. Efficient transduction of HSCs with lentiviral vectors has become increasingly reliable, but the complicated components of many globin vectors present unique challenges for production of high-titer virus capable of robust transduction. Scaling up procedures to multiple patients is a nontrivial challenge. Safety and efficacy can only be established by careful clinical trials with extended patient follow-up. Gene engineering methods are rapidly evolving and should facilitate development of “second-generation” gene therapy approaches in the coming years. After many years of preclinical laboratory investigation, gene therapy options are now on the horizon for patients with SCD.

The authors thank Marina Cavazzana, Donald Kohn, and Punam Malik for personal communications regarding open gene therapy trials, and Brian Sorrentino and Christian Brendel for helpful comments.

This work was supported by National Institutes of Health, National Heart, Lung, and Blood Institute grants T32HL007574 (M.D.H.), R01HL032259, and P01HL032262, and National Institute of Diabetes and Digestive and Kidney Diseases grants P30DK049216 (Center of Excellence in Molecular Hematology) (S.H.O.) and K08DK093705 (Career Development Award) (D.E.B.). This work was also supported by the Doris Duke Charitable Foundation, the Charles H. Hood Foundation, and the Cooley’s Anemia Foundation (D.E.B.).

Contribution: M.D.H., S.H.O., and D.E.B. wrote and edited the manuscript.

Conflict-of-interest disclosure: D.E.B. and S.H.O. are inventors on a patent application related to therapeutic genome editing of BCL11A. D.E.B. has served as a consultant to Editas Medicine. S.H.O. was previously on the Scientific Advisory Board of Editas Medicine. D.E.B. and S.H.O. have received research support from Biogen. M.D.H. declares no competing financial interests.

Correspondence: Stuart H. Orkin, Department of Pediatrics, Harvard Medical, Harvard Stem Cell Institute, Dana-Farber Cancer Institute, 44 Binney St, Boston, MA 02115; email: orkin@bloodgroup.tch.harvard.edu.

1
Stamatoyannopoulos
 
G
The molecular basis of hemoglobin disease.
Annu Rev Genet
1972
, vol. 
6
 (pg. 
47
-
70
)
2
Platt
 
OS
Rosenstock
 
W
Espeland
 
MA
Influence of sickle hemoglobinopathies on growth and development.
N Engl J Med
1984
, vol. 
311
 
1
(pg. 
7
-
12
)
3
Madigan
 
C
Malik
 
P
Pathophysiology and therapy for haemoglobinopathies. Part I: sickle cell disease.
Expert Rev Mol Med
2006
, vol. 
8
 
9
(pg. 
1
-
23
)
4
Creary
 
M
Williamson
 
D
Kulkarni
 
R
Sickle cell disease: current activities, public health implications, and future directions.
J Womens Health (Larchmt)
2007
, vol. 
16
 
5
(pg. 
575
-
582
)
5
Locatelli
 
F
Kabbara
 
N
Ruggeri
 
A
, et al. 
Eurocord and European Blood and Marrow Transplantation (EBMT) group
Outcome of patients with hemoglobinopathies given either cord blood or bone marrow transplantation from an HLA-identical sibling.
Blood
2013
, vol. 
122
 
6
(pg. 
1072
-
1078
)
6
Walters
 
MC
Patience
 
M
Leisenring
 
W
, et al. 
Multicenter Investigation of Bone Marrow Transplantation for Sickle Cell Disease
Stable mixed hematopoietic chimerism after bone marrow transplantation for sickle cell anemia.
Biol Blood Marrow Transplant
2001
, vol. 
7
 
12
(pg. 
665
-
673
)
7
Fitzhugh
 
CD
Abraham
 
AA
Tisdale
 
JF
Hsieh
 
MM
Hematopoietic stem cell transplantation for patients with sickle cell disease: progress and future directions.
Hematol Oncol Clin North Am
2014
, vol. 
28
 
6
(pg. 
1171
-
1185
)
8
Walters
 
MC
Stem cell therapy for sickle cell disease: transplantation and gene therapy.
Hematology Am Soc Hematol Educ Program
2005
, vol. 
2005
 (pg. 
66
-
73
)
9
Walters
 
MC
Patience
 
M
Leisenring
 
W
, et al. 
Barriers to bone marrow transplantation for sickle cell anemia.
Biol Blood Marrow Transplant
1996
, vol. 
2
 
2
(pg. 
100
-
104
)
10
Thompson
 
AL
Bridley
 
A
Twohy
 
E
, et al. 
An educational symposium for patients with sickle cell disease and their families: results from surveys of knowledge and factors influencing decisions about hematopoietic stem cell transplant.
Pediatr Blood Cancer
2013
, vol. 
60
 
12
(pg. 
1946
-
1951
)
11
Hsieh
 
MM
Fitzhugh
 
CD
Weitzel
 
RP
, et al. 
Nonmyeloablative HLA-matched sibling allogeneic hematopoietic stem cell transplantation for severe sickle cell phenotype.
JAMA
2014
, vol. 
312
 
1
(pg. 
48
-
56
)
12
Yawn
 
BP
Buchanan
 
GR
Afenyi-Annan
 
AN
, et al. 
Management of sickle cell disease: summary of the 2014 evidence-based report by expert panel members.
JAMA
2014
, vol. 
312
 
10
(pg. 
1033
-
1048
)
13
Bank
 
A
Markowitz
 
D
Lerner
 
N
Gene transfer. A potential approach to gene therapy for sickle cell disease.
Ann N Y Acad Sci
1989
, vol. 
565
 
1
(pg. 
37
-
43
)
14
Chandrakasan
 
S
Malik
 
P
Gene therapy for hemoglobinopathies: the state of the field and the future.
Hematol Oncol Clin North Am
2014
, vol. 
28
 
2
(pg. 
199
-
216
)
15
Novak
 
U
Harris
 
EA
Forrester
 
W
Groudine
 
M
Gelinas
 
R
High-level beta-globin expression after retroviral transfer of locus activation region-containing human beta-globin gene derivatives into murine erythroleukemia cells.
Proc Natl Acad Sci USA
1990
, vol. 
87
 
9
(pg. 
3386
-
3390
)
16
Karlsson
 
S
Bodine
 
DM
Perry
 
L
Papayannopoulou
 
T
Nienhuis
 
AW
Expression of the human beta-globin gene following retroviral-mediated transfer into multipotential hematopoietic progenitors of mice.
Proc Natl Acad Sci USA
1988
, vol. 
85
 
16
(pg. 
6062
-
6066
)
17
Dzierzak
 
EA
Papayannopoulou
 
T
Mulligan
 
RC
Lineage-specific expression of a human beta-globin gene in murine bone marrow transplant recipients reconstituted with retrovirus-transduced stem cells.
Nature
1988
, vol. 
331
 
6151
(pg. 
35
-
41
)
18
Cone
 
RD
Weber-Benarous
 
A
Baorto
 
D
Mulligan
 
RC
Regulated expression of a complete human beta-globin gene encoded by a transmissible retrovirus vector.
Mol Cell Biol
1987
, vol. 
7
 
2
(pg. 
887
-
897
)
19
Grosveld
 
F
van Assendelft
 
GB
Greaves
 
DR
Kollias
 
G
Position-independent, high-level expression of the human β-globin gene in transgenic mice.
Cell
1987
, vol. 
51
 
6
(pg. 
975
-
985
)
20
Tuan
 
D
Solomon
 
W
Li
 
Q
London
 
IM
The “β-like-globin” gene domain in human erythroid cells.
Proc Natl Acad Sci USA
1985
, vol. 
82
 
19
(pg. 
6384
-
6388
)
21
Forrester
 
WC
Takegawa
 
S
Papayannopoulou
 
T
Stamatoyannopoulos
 
G
Groudine
 
M
Evidence for a locus activation region: the formation of developmentally stable hypersensitive sites in globin-expressing hybrids.
Nucleic Acids Res
1987
, vol. 
15
 
24
(pg. 
10159
-
10177
)
22
Persons
 
DA
Hargrove
 
PW
Allay
 
ER
Hanawa
 
H
Nienhuis
 
AW
The degree of phenotypic correction of murine beta -thalassemia intermedia following lentiviral-mediated transfer of a human gamma-globin gene is influenced by chromosomal position effects and vector copy number.
Blood
2003
, vol. 
101
 
6
(pg. 
2175
-
2183
)
23
Plavec
 
I
Papayannopoulou
 
T
Maury
 
C
Meyer
 
F
A human beta-globin gene fused to the human beta-globin locus control region is expressed at high levels in erythroid cells of mice engrafted with retrovirus-transduced hematopoietic stem cells.
Blood
1993
, vol. 
81
 
5
(pg. 
1384
-
1392
)
24
Sadelain
 
M
Wang
 
CH
Antoniou
 
M
Grosveld
 
F
Mulligan
 
RC
Generation of a high-titer retroviral vector capable of expressing high levels of the human beta-globin gene.
Proc Natl Acad Sci USA
1995
, vol. 
92
 
15
(pg. 
6728
-
6732
)
25
Leboulch
 
P
Huang
 
GM
Humphries
 
RK
, et al. 
Mutagenesis of retroviral vectors transducing human beta-globin gene and beta-globin locus control region derivatives results in stable transmission of an active transcriptional structure.
EMBO J
1994
, vol. 
13
 
13
(pg. 
3065
-
3076
)
26
Collis
 
P
Antoniou
 
M
Grosveld
 
F
Definition of the minimal requirements within the human beta-globin gene and the dominant control region for high level expression.
EMBO J
1990
, vol. 
9
 
1
(pg. 
233
-
240
)
27
Ellis
 
J
Tan-Un
 
KC
Harper
 
A
, et al. 
A dominant chromatin-opening activity in 5′ hypersensitive site 3 of the human beta-globin locus control region.
EMBO J
1996
, vol. 
15
 
3
(pg. 
562
-
568
)
28
Rivella
 
S
Sadelain
 
M
Genetic treatment of severe hemoglobinopathies: the combat against transgene variegation and transgene silencing.
Semin Hematol
1998
, vol. 
35
 
2
(pg. 
112
-
125
)
29
Emery
 
DW
Morrish
 
F
Li
 
Q
Stamatoyannopoulos
 
G
Analysis of gamma-globin expression cassettes in retrovirus vectors.
Hum Gene Ther
1999
, vol. 
10
 
6
(pg. 
877
-
888
)
30
Evans-Galea
 
MV
Wielgosz
 
MM
Hanawa
 
H
Srivastava
 
DK
Nienhuis
 
AW
Suppression of clonal dominance in cultured human lymphoid cells by addition of the cHS4 insulator to a lentiviral vector.
Mol Ther
2007
, vol. 
15
 
4
(pg. 
801
-
809
)
31
Emery
 
DW
Yannaki
 
E
Tubb
 
J
Stamatoyannopoulos
 
G
A chromatin insulator protects retrovirus vectors from chromosomal position effects.
Proc Natl Acad Sci USA
2000
, vol. 
97
 
16
(pg. 
9150
-
9155
)
32
Ramezani
 
A
Hawley
 
TS
Hawley
 
RG
Combinatorial incorporation of enhancer-blocking components of the chicken β-globin 5'HS4 and human T-cell receptor α/δ BEAD-1 insulators in self-inactivating retroviral vectors reduces their genotoxic potential.
Stem Cells
2008
, vol. 
26
 
12
(pg. 
3257
-
3266
)
33
Emery
 
DW
Yannaki
 
E
Tubb
 
J
Nishino
 
T
Li
 
Q
Stamatoyannopoulos
 
G
Development of virus vectors for gene therapy of β chain hemoglobinopathies: flanking with a chromatin insulator reduces γ-globin gene silencing in vivo.
Blood
2002
, vol. 
100
 
6
(pg. 
2012
-
2019
)
34
Urbinati
 
F
Arumugam
 
P
Higashimoto
 
T
, et al. 
Mechanism of reduction in titers from lentivirus vectors carrying large inserts in the 3'LTR.
Mol Ther
2009
, vol. 
17
 
9
(pg. 
1527
-
1536
)
35
Arumugam
 
PI
Scholes
 
J
Perelman
 
N
Xia
 
P
Yee
 
JK
Malik
 
P
Improved human β-globin expression from self-inactivating lentiviral vectors carrying the chicken hypersensitive site-4 (cHS4) insulator element.
Mol Ther
2007
, vol. 
15
 
10
(pg. 
1863
-
1871
)
36
Romero
 
Z
Campo-Fernandez
 
B
Wherley
 
J
, et al. 
The human ankyrin 1 promoter insulator sustains gene expression in a β-globin lentiviral vector in hematopoietic stem cells.
Mol Ther Methods Clin Dev
2015
, vol. 
2
 pg. 
15012
 
37
Ryu
 
BY
Evans-Galea
 
MV
Gray
 
JT
Bodine
 
DM
Persons
 
DA
Nienhuis
 
AW
An experimental system for the evaluation of retroviral vector design to diminish the risk for proto-oncogene activation.
Blood
2008
, vol. 
111
 
4
(pg. 
1866
-
1875
)
38
Koldej
 
RM
Carney
 
G
Wielgosz
 
MM
, et al. 
Comparison of insulators and promoters for expression of the Wiskott-Aldrich syndrome protein using lentiviral vectors.
Hum Gene Ther Clin Dev
2013
, vol. 
24
 
2
(pg. 
77
-
85
)
39
Liu
 
M
Maurano
 
MT
Wang
 
H
, et al. 
Genomic discovery of potent chromatin insulators for human gene therapy.
Nat Biotechnol
2015
, vol. 
33
 
2
(pg. 
198
-
203
)
40
Naldini
 
L
Blömer
 
U
Gage
 
FH
Trono
 
D
Verma
 
IM
Efficient transfer, integration, and sustained long-term expression of the transgene in adult rat brains injected with a lentiviral vector.
Proc Natl Acad Sci USA
1996
, vol. 
93
 
21
(pg. 
11382
-
11388
)
41
Hargrove
 
PW
Kepes
 
S
Hanawa
 
H
, et al. 
Globin lentiviral vector insertions can perturb the expression of endogenous genes in beta-thalassemic hematopoietic cells.
Mol Ther
2008
, vol. 
16
 
3
(pg. 
525
-
533
)
42
Arumugam
 
PI
Higashimoto
 
T
Urbinati
 
F
, et al. 
Genotoxic potential of lineage-specific lentivirus vectors carrying the beta-globin locus control region.
Mol Ther
2009
, vol. 
17
 
11
(pg. 
1929
-
1937
)
43
Cattoglio
 
C
Facchini
 
G
Sartori
 
D
, et al. 
Hot spots of retroviral integration in human CD34+ hematopoietic cells.
Blood
2007
, vol. 
110
 
6
(pg. 
1770
-
1778
)
44
Aiuti
 
A
Biasco
 
L
Scaramuzza
 
S
, et al. 
Lentiviral hematopoietic stem cell gene therapy in patients with Wiskott-Aldrich syndrome.
Science
2013
, vol. 
341
 
6148
pg. 
1233151
 
45
Biffi
 
A
Montini
 
E
Lorioli
 
L
, et al. 
Lentiviral hematopoietic stem cell gene therapy benefits metachromatic leukodystrophy.
Science
2013
, vol. 
341
 
6148
pg. 
1233158
 
46
Schröder
 
ARW
Shinn
 
P
Chen
 
H
Berry
 
C
Ecker
 
JR
Bushman
 
F
HIV-1 integration in the human genome favors active genes and local hotspots.
Cell
2002
, vol. 
110
 
4
(pg. 
521
-
529
)
47
Miyoshi
 
H
Blömer
 
U
Takahashi
 
M
Gage
 
FH
Verma
 
IM
Development of a self-inactivating lentivirus vector.
J Virol
1998
, vol. 
72
 
10
(pg. 
8150
-
8157
)
48
May
 
C
Rivella
 
S
Callegari
 
J
, et al. 
Therapeutic haemoglobin synthesis in beta-thalassaemic mice expressing lentivirus-encoded human beta-globin.
Nature
2000
, vol. 
406
 
6791
(pg. 
82
-
86
)
49
May
 
C
Rivella
 
S
Chadburn
 
A
Sadelain
 
M
Successful treatment of murine beta-thalassemia intermedia by transfer of the human beta-globin gene.
Blood
2002
, vol. 
99
 
6
(pg. 
1902
-
1908
)
50
Rivella
 
S
May
 
C
Chadburn
 
A
Rivière
 
I
Sadelain
 
M
A novel murine model of Cooley anemia and its rescue by lentiviral-mediated human beta-globin gene transfer.
Blood
2003
, vol. 
101
 
8
(pg. 
2932
-
2939
)
51
Pawliuk
 
R
Westerman
 
KA
Fabry
 
ME
, et al. 
Correction of sickle cell disease in transgenic mouse models by gene therapy.
Science
2001
, vol. 
294
 
5550
(pg. 
2368
-
2371
)
52
Sunshine
 
HR
Hofrichter
 
J
Eaton
 
WA
Requirement for therapeutic inhibition of sickle haemoglobin gelation.
Nature
1978
, vol. 
275
 
5677
(pg. 
238
-
240
)
53
Platt
 
OS
Thorington
 
BD
Brambilla
 
DJ
, et al. 
Pain in sickle cell disease. Rates and risk factors.
N Engl J Med
1991
, vol. 
325
 
1
(pg. 
11
-
16
)
54
Platt
 
OS
Brambilla
 
DJ
Rosse
 
WF
, et al. 
Mortality in sickle cell disease. Life expectancy and risk factors for early death.
N Engl J Med
1994
, vol. 
330
 
23
(pg. 
1639
-
1644
)
55
Li
 
Q
Emery
 
DW
Fernandez
 
M
Han
 
H
Stamatoyannopoulos
 
G
Development of viral vectors for gene therapy of beta-chain hemoglobinopathies: optimization of a gamma-globin gene expression cassette.
Blood
1999
, vol. 
93
 
7
(pg. 
2208
-
2216
)
56
Pestina
 
TI
Hargrove
 
PW
Jay
 
D
Gray
 
JT
Boyd
 
KM
Persons
 
DA
Correction of murine sickle cell disease using gamma-globin lentiviral vectors to mediate high-level expression of fetal hemoglobin.
Mol Ther
2009
, vol. 
17
 
2
(pg. 
245
-
252
)
57
Perumbeti
 
A
Malik
 
P
Therapy for beta-globinopathies: a brief review and determinants for successful and safe correction.
Ann N Y Acad Sci
2010
, vol. 
1202
 (pg. 
36
-
44
)
58
Perumbeti
 
A
Higashimoto
 
T
Urbinati
 
F
, et al. 
A novel human gamma-globin gene vector for genetic correction of sickle cell anemia in a humanized sickle mouse model: critical determinants for successful correction.
Blood
2009
, vol. 
114
 
6
(pg. 
1174
-
1185
)
59
Moreau-Gaudry
 
F
Xia
 
P
Jiang
 
G
, et al. 
High-level erythroid-specific gene expression in primary human and murine hematopoietic cells with self-inactivating lentiviral vectors.
Blood
2001
, vol. 
98
 
9
(pg. 
2664
-
2672
)
60
McCune
 
SL
Reilly
 
MP
Chomo
 
MJ
Asakura
 
T
Townes
 
TM
Recombinant human hemoglobins designed for gene therapy of sickle cell disease.
Proc Natl Acad Sci USA
1994
, vol. 
91
 
21
(pg. 
9852
-
9856
)
61
Levasseur
 
DN
Ryan
 
TM
Pawlik
 
KM
Townes
 
TM
Correction of a mouse model of sickle cell disease: lentiviral/antisickling beta-globin gene transduction of unmobilized, purified hematopoietic stem cells.
Blood
2003
, vol. 
102
 
13
(pg. 
4312
-
4319
)
62
Levasseur
 
DN
Ryan
 
TM
Reilly
 
MP
McCune
 
SL
Asakura
 
T
Townes
 
TM
A recombinant human hemoglobin with anti-sickling properties greater than fetal hemoglobin.
J Biol Chem
2004
, vol. 
279
 
26
(pg. 
27518
-
27524
)
63
Romero
 
Z
Urbinati
 
F
Geiger
 
S
, et al. 
β-globin gene transfer to human bone marrow for sickle cell disease.
J Clin Invest
2013
, vol. 
123
 
8
(pg. 
3317
-
3330
)
64
Urbinati
 
F
Hargrove
 
PW
Geiger
 
S
, et al. 
Potentially therapeutic levels of anti-sickling globin gene expression following lentivirus-mediated gene transfer in sickle cell disease bone marrow CD34+ cells.
Exp Hematol
2015
, vol. 
43
 
5
(pg. 
346
-
351
)
65
Archer
 
N
Galacteros
 
F
Brugnara
 
C
2015 Clinical trials update in sickle cell anemia.
Am J Hematol
2015
, vol. 
90
 
10
(pg. 
934
-
950
)
66
Negre
 
O
Bartholomae
 
C
Beuzard
 
Y
, et al. 
Preclinical evaluation of efficacy and safety of an improved lentiviral vector for the treatment of β-thalassemia and sickle cell disease.
Curr Gene Ther
2015
, vol. 
15
 
1
(pg. 
64
-
81
)
67
Cavazzana
 
M
Ribeil
 
JA
Payen
 
E
, et al. 
Outcomes of gene therapy for B-thalassemia major and severe sickle cell disease via transplantation of autologous hematopoietic stem cells transduced ex vivo with a lentiviral beta globin vector.
Haematologica
2015
, vol. 
100
 
s1
pg. 
171
 
68
Aiuti
 
A
Slavin
 
S
Aker
 
M
, et al. 
Correction of ADA-SCID by stem cell gene therapy combined with nonmyeloablative conditioning.
Science
2002
, vol. 
296
 
5577
(pg. 
2410
-
2413
)
69
Carroll
 
D
Genome engineering with targetable nucleases.
Annu Rev Biochem
2014
, vol. 
83
 (pg. 
409
-
439
)
70
Gaj
 
T
Gersbach
 
CA
Barbas
 
CF
ZFN, TALEN, and CRISPR/Cas-based methods for genome engineering.
Trends Biotechnol
2013
, vol. 
31
 
7
(pg. 
397
-
405
)
71
Arnould
 
S
Delenda
 
C
Grizot
 
S
, et al. 
The I-CreI meganuclease and its engineered derivatives: applications from cell modification to gene therapy.
Protein Eng Des Sel
2011
, vol. 
24
 
1-2
(pg. 
27
-
31
)
72
Mao
 
Z
Bozzella
 
M
Seluanov
 
A
Gorbunova
 
V
DNA repair by nonhomologous end joining and homologous recombination during cell cycle in human cells.
Cell Cycle
2008
, vol. 
7
 
18
(pg. 
2902
-
2906
)
73
Perez
 
EE
Wang
 
J
Miller
 
JC
, et al. 
Establishment of HIV-1 resistance in CD4+ T cells by genome editing using zinc-finger nucleases.
Nat Biotechnol
2008
, vol. 
26
 
7
(pg. 
808
-
816
)
74
Voit
 
RA
Hendel
 
A
Pruett-Miller
 
SM
Porteus
 
MH
Nuclease-mediated gene editing by homologous recombination of the human globin locus.
Nucleic Acids Res
2014
, vol. 
42
 
2
(pg. 
1365
-
1378
)
75
Sun
 
N
Zhao
 
H
Seamless correction of the sickle cell disease mutation of the HBB gene in human induced pluripotent stem cells using TALENs.
Biotechnol Bioeng
2014
, vol. 
111
 
5
(pg. 
1048
-
1053
)
76
Sebastiano
 
V
Maeder
 
ML
Angstman
 
JF
, et al. 
In situ genetic correction of the sickle cell anemia mutation in human induced pluripotent stem cells using engineered zinc finger nucleases.
Stem Cells
2011
, vol. 
29
 
11
(pg. 
1717
-
1726
)
77
Zou
 
J
Mali
 
P
Huang
 
X
Dowey
 
SN
Cheng
 
L
Site-specific gene correction of a point mutation in human iPS cells derived from an adult patient with sickle cell disease.
Blood
2011
, vol. 
118
 
17
(pg. 
4599
-
4608
)
78
Hanna
 
J
Wernig
 
M
Markoulaki
 
S
, et al. 
Treatment of sickle cell anemia mouse model with iPS cells generated from autologous skin.
Science
2007
 
318(5858):1920-1923
79
Xu
 
XS
Glazer
 
PM
Wang
 
G
Activation of human gamma-globin gene expression via triplex-forming oligonucleotide (TFO)-directed mutations in the gamma-globin gene 5′ flanking region.
Gene
2000
, vol. 
242
 
1-2
(pg. 
219
-
228
)
80
Chin
 
JY
Reza
 
F
Glazer
 
PM
Triplex-forming peptide nucleic acids induce heritable elevations in gamma-globin expression in hematopoietic progenitor cells.
Mol Ther
2013
, vol. 
21
 
3
(pg. 
580
-
587
)
81
Rogers
 
FA
Lin
 
SS
Hegan
 
DC
Krause
 
DS
Glazer
 
PM
Targeted gene modification of hematopoietic progenitor cells in mice following systemic administration of a PNA-peptide conjugate.
Mol Ther
2012
, vol. 
20
 
1
(pg. 
109
-
118
)
82
Goncz
 
KK
Prokopishyn
 
NL
Abdolmohammadi
 
A
, et al. 
Small fragment homologous replacement-mediated modification of genomic beta-globin sequences in human hematopoietic stem/progenitor cells.
Oligonucleotides
2006
, vol. 
16
 
3
(pg. 
213
-
224
)
83
Vanhee
 
S
Vandekerckhove
 
B
Pluripotent stem cell based gene therapy for hematological diseases.
Crit Rev Oncol Hematol
2016
, vol. 
97
  
238-246
84
Vo
 
LT
Daley
 
GQ
De novo generation of HSCs from somatic and pluripotent stem cell sources.
Blood
2015
, vol. 
125
 
17
(pg. 
2641
-
2648
)
85
Hoban
 
MD
Cost
 
GJ
Mendel
 
MC
, et al. 
Correction of the sickle cell disease mutation in human hematopoietic stem/progenitor cells.
Blood
2015
, vol. 
125
 
17
(pg. 
2597
-
2604
)
86
Genovese
 
P
Schiroli
 
G
Escobar
 
G
, et al. 
Targeted genome editing in human repopulating haematopoietic stem cells.
Nature
2014
, vol. 
510
 
7504
(pg. 
235
-
240
)
87
Mohrin
 
M
Bourke
 
E
Alexander
 
D
, et al. 
Hematopoietic stem cell quiescence promotes error-prone DNA repair and mutagenesis.
Cell Stem Cell
2010
, vol. 
7
 
2
(pg. 
174
-
185
)
88
Tebas
 
P
Stein
 
D
Tang
 
WW
, et al. 
Gene editing of CCR5 in autologous CD4 T cells of persons infected with HIV.
N Engl J Med
2014
, vol. 
370
 
10
(pg. 
901
-
910
)
89
Bauer
 
DE
Kamran
 
SC
Orkin
 
SH
Reawakening fetal hemoglobin: prospects for new therapies for the β-globin disorders.
Blood
2012
, vol. 
120
 
15
(pg. 
2945
-
2953
)
90
Menzel
 
S
Garner
 
C
Gut
 
I
, et al. 
A QTL influencing F cell production maps to a gene encoding a zinc-finger protein on chromosome 2p15.
Nat Genet
2007
, vol. 
39
 
10
(pg. 
1197
-
1199
)
91
Uda
 
M
Galanello
 
R
Sanna
 
S
, et al. 
Genome-wide association study shows BCL11A associated with persistent fetal hemoglobin and amelioration of the phenotype of beta-thalassemia.
Proc Natl Acad Sci USA
2008
, vol. 
105
 
5
(pg. 
1620
-
1625
)
92
Sankaran
 
VG
Menne
 
TF
Xu
 
J
, et al. 
Human fetal hemoglobin expression is regulated by the developmental stage-specific repressor BCL11A.
Science
2008
, vol. 
322
 
5909
(pg. 
1839
-
1842
)
93
Sankaran
 
VG
Xu
 
J
Ragoczy
 
T
, et al. 
Developmental and species-divergent globin switching are driven by BCL11A.
Nature
2009
, vol. 
460
 
7259
(pg. 
1093
-
1097
)
94
Xu
 
J
Peng
 
C
Sankaran
 
VG
, et al. 
Correction of sickle cell disease in adult mice by interference with fetal hemoglobin silencing.
Science
2011
, vol. 
334
 
6058
(pg. 
993
-
996
)
95
Liu
 
P
Keller
 
JR
Ortiz
 
M
, et al. 
Bcl11a is essential for normal lymphoid development.
Nat Immunol
2003
, vol. 
4
 
6
(pg. 
525
-
532
)
96
Yu
 
Y
Wang
 
J
Khaled
 
W
, et al. 
Bcl11a is essential for lymphoid development and negatively regulates p53.
J Exp Med
2012
, vol. 
209
 
13
(pg. 
2467
-
2483
)
97
Ippolito
 
GC
Dekker
 
JD
Wang
 
Y-H
, et al. 
Dendritic cell fate is determined by BCL11A.
Proc Natl Acad Sci USA
2014
, vol. 
111
 
11
(pg. 
E998
-
E1006
)
98
Bauer
 
DE
Kamran
 
SC
Lessard
 
S
, et al. 
An erythroid enhancer of BCL11A subject to genetic variation determines fetal hemoglobin level.
Science
2013
, vol. 
342
 
6155
(pg. 
253
-
257
)
99
Canver
 
MC
Smith
 
EC
Sher
 
F
, et al. 
BCL11A enhancer dissection by Cas9-mediated in situ saturating mutagenesis.
Nature
2015
, vol. 
527
 
7577
 
192-197
100
Vierstra
 
J
Reik
 
A
Chang
 
K-H
, et al. 
Functional footprinting of regulatory DNA.
Nat Methods
2015
, vol. 
12
 
10
(pg. 
927
-
930
)
101
Wilber
 
A
Hargrove
 
PW
Kim
 
Y-S
, et al. 
Therapeutic levels of fetal hemoglobin in erythroid progeny of β-thalassemic CD34+ cells after lentiviral vector-mediated gene transfer.
Blood
2011
, vol. 
117
 
10
(pg. 
2817
-
2826
)
102
Guda
 
S
Brendel
 
C
Renella
 
R
, et al. 
miRNA-embedded shRNAs for lineage-specific BCL11A knockdown and hemoglobin F induction.
Mol Ther
2015
, vol. 
23
 
9
(pg. 
1465
-
1474
)
103
Deng
 
W
Lee
 
J
Wang
 
H
, et al. 
Controlling long-range genomic interactions at a native locus by targeted tethering of a looping factor.
Cell
2012
, vol. 
149
 
6
(pg. 
1233
-
1244
)
104
Deng
 
W
Rupon
 
JW
Krivega
 
I
, et al. 
Reactivation of developmentally silenced globin genes by forced chromatin looping.
Cell
2014
, vol. 
158
 
4
(pg. 
849
-
860
)
105
Zuccato
 
C
Breda
 
L
Salvatori
 
F
, et al. 
A combined approach for β-thalassemia based on gene therapy-mediated adult hemoglobin (HbA) production and fetal hemoglobin (HbF) induction.
Ann Hematol
2012
, vol. 
91
 
8
(pg. 
1201
-
1213
)
106
Samakoglu
 
S
Lisowski
 
L
Budak-Alpdogan
 
T
, et al. 
A genetic strategy to treat sickle cell anemia by coregulating globin transgene expression and RNA interference.
Nat Biotechnol
2006
, vol. 
24
 
1
(pg. 
89
-
94
)
107
Wienert
 
B
Funnell
 
AP
Norton
 
LJ
, et al. 
Editing the genome to introduce a beneficial naturally occurring mutation associated with increased fetal globin.
Nat Commun
2015
, vol. 
6
 pg. 
7085
 
108
Zhou
 
D
Liu
 
K
Sun
 
C-W
Pawlik
 
KM
Townes
 
TM
KLF1 regulates BCL11A expression and gamma- to beta-globin gene switching.
Nat Genet
2010
, vol. 
42
 
9
(pg. 
742
-
744
)
109
Borg
 
J
Papadopoulos
 
P
Georgitsi
 
M
, et al. 
Haploinsufficiency for the erythroid transcription factor KLF1 causes hereditary persistence of fetal hemoglobin.
Nat Genet
2010
, vol. 
42
 
9
(pg. 
801
-
805
)
110
Candotti
 
F
Shaw
 
KL
Muul
 
L
, et al. 
Gene therapy for adenosine deaminase-deficient severe combined immune deficiency: clinical comparison of retroviral vectors and treatment plans.
Blood
2012
, vol. 
120
 
18
(pg. 
3635
-
3646
)
111
Fitzhugh
 
CD
Hsieh
 
MM
Bolan
 
CD
Saenz
 
C
Tisdale
 
JF
Granulocyte colony-stimulating factor (G-CSF) administration in individuals with sickle cell disease: time for a moratorium?
Cytotherapy
2009
, vol. 
11
 
4
(pg. 
464
-
471
)
112
Adler
 
BK
Salzman
 
DE
Carabasi
 
MH
Vaughan
 
WP
Reddy
 
VV
Prchal
 
JT
Fatal sickle cell crisis after granulocyte colony-stimulating factor administration.
Blood
2001
, vol. 
97
 
10
(pg. 
3313
-
3314
)
113
Civriz Bozdag
 
S
Tekgunduz
 
E
Altuntas
 
F
The current status in hematopoietic stem cell mobilization.
J Clin Apher
2015
, vol. 
30
 
5
(pg. 
273
-
280
)
114
Baldwin
 
K
Urbinati
 
F
Romero
 
Z
, et al. 
Enrichment of human hematopoietic stem/progenitor cells facilitates transduction for stem cell gene therapy.
Stem Cells
2015
, vol. 
33
 
5
(pg. 
1532
-
1542
)
115
Majeti
 
R
Park
 
CY
Weissman
 
IL
Identification of a hierarchy of multipotent hematopoietic progenitors in human cord blood.
Cell Stem Cell
2007
, vol. 
1
 
6
(pg. 
635
-
645
)
116
Lansdorp
 
PM
Sutherland
 
HJ
Eaves
 
CJ
Selective expression of CD45 isoforms on functional subpopulations of CD34+ hemopoietic cells from human bone marrow.
J Exp Med
1990
, vol. 
172
 
1
(pg. 
363
-
366
)
117
Wisniewski
 
D
Affer
 
M
Willshire
 
J
Clarkson
 
B
Further phenotypic characterization of the primitive lineage- CD34+CD38-CD90+CD45RA- hematopoietic stem cell/progenitor cell sub-population isolated from cord blood, mobilized peripheral blood and patients with chronic myelogenous leukemia.
Blood Cancer J
2011
, vol. 
1
 
9
pg. 
e36
 
118
Dorrell
 
C
Gan
 
OI
Pereira
 
DS
Hawley
 
RG
Dick
 
JE
Expansion of human cord blood CD34+CD38− cells in ex vivo culture during retroviral transduction without a corresponding increase in SCID repopulating cell (SRC) frequency: dissociation of SRC phenotype and function.
Blood
2000
, vol. 
95
 
1
(pg. 
102
-
110
)
119
Brendel
 
C
Goebel
 
B
Daniela
 
A
, et al. 
CD133-targeted gene transfer into long-term repopulating hematopoietic stem cells.
Mol Ther
2015
, vol. 
23
 
1
(pg. 
63
-
70
)
120
Fares
 
I
Chagraoui
 
J
Gareau
 
Y
, et al. 
Cord blood expansion. Pyrimidoindole derivatives are agonists of human hematopoietic stem cell self-renewal.
Science
2014
, vol. 
345
 
6203
(pg. 
1509
-
1512
)
121
Boitano
 
AE
Wang
 
J
Romeo
 
R
, et al. 
Aryl hydrocarbon receptor antagonists promote the expansion of human hematopoietic stem cells.
Science
2010
, vol. 
329
 
5997
(pg. 
1345
-
1348
)
122
North
 
TE
Goessling
 
W
Walkley
 
CR
, et al. 
Prostaglandin E2 regulates vertebrate haematopoietic stem cell homeostasis.
Nature
2007
, vol. 
447
 
7147
(pg. 
1007
-
1011
)
123
Zhang
 
CC
Kaba
 
M
Iizuka
 
S
Huynh
 
H
Lodish
 
HF
Angiopoietin-like 5 and IGFBP2 stimulate ex vivo expansion of human cord blood hematopoietic stem cells as assayed by NOD/SCID transplantation.
Blood
2008
, vol. 
111
 
7
(pg. 
3415
-
3423
)
124
Andreani
 
M
Testi
 
M
Gaziev
 
J
, et al. 
Quantitatively different red cell/nucleated cell chimerism in patients with long-term, persistent hematopoietic mixed chimerism after bone marrow transplantation for thalassemia major or sickle cell disease.
Haematologica
2011
, vol. 
96
 
1
(pg. 
128
-
133
)
125
Wu
 
CJ
Gladwin
 
M
Tisdale
 
J
, et al. 
Mixed haematopoietic chimerism for sickle cell disease prevents intravascular haemolysis.
Br J Haematol
2007
, vol. 
139
 
3
(pg. 
504
-
507
)
126
Krishnamurti
 
L
Kharbanda
 
S
Biernacki
 
MA
, et al. 
Stable long-term donor engraftment following reduced-intensity hematopoietic cell transplantation for sickle cell disease.
Biol Blood Marrow Transplant
2008
, vol. 
14
 
11
(pg. 
1270
-
1278
)
127
Wu
 
CJ
Krishnamurti
 
L
Kutok
 
JL
, et al. 
Evidence for ineffective erythropoiesis in severe sickle cell disease.
Blood
2005
, vol. 
106
 
10
(pg. 
3639
-
3645
)
128
Cavazzana-Calvo
 
M
Payen
 
E
Negre
 
O
, et al. 
Transfusion independence and HMGA2 activation after gene therapy of human β-thalassaemia.
Nature
2010
, vol. 
467
 
7313
(pg. 
318
-
322
)
129
Cartier
 
N
Hacein-Bey-Abina
 
S
Bartholomae
 
CC
, et al. 
Hematopoietic stem cell gene therapy with a lentiviral vector in X-linked adrenoleukodystrophy.
Science
2009
, vol. 
326
 
5954
(pg. 
818
-
823
)
130
Kohn
 
DB
Eliminating SCID row: new approaches to SCID.
Hematology Am Soc Hematol Educ Program
2014
, vol. 
2014
 (pg. 
475
-
480
)
131
Czechowicz
 
A
Kraft
 
D
Weissman
 
IL
Bhattacharya
 
D
Efficient transplantation via antibody-based clearance of hematopoietic stem cell niches.
Science
2007
, vol. 
318
 
5854
(pg. 
1296
-
1299
)
132
Derderian
 
SC
Togarrati
 
PP
King
 
C
, et al. 
In utero depletion of fetal hematopoietic stem cells improves engraftment after neonatal transplantation in mice.
Blood
2014
, vol. 
124
 
6
(pg. 
973
-
980
)
133
Cosgun
 
KN
Rahmig
 
S
Mende
 
N
, et al. 
Kit regulates HSC engraftment across the human-mouse species barrier.
Cell Stem Cell
2014
, vol. 
15
 
2
(pg. 
227
-
238
)
134
Hacein-Bey-Abina
 
S
Von Kalle
 
C
Schmidt
 
M
, et al. 
LMO2-associated clonal T cell proliferation in two patients after gene therapy for SCID-X1.
Science
2003
, vol. 
302
 
5644
(pg. 
415
-
419
)
135
Stein
 
S
Ott
 
MG
Schultze-Strasser
 
S
, et al. 
Genomic instability and myelodysplasia with monosomy 7 consequent to EVI1 activation after gene therapy for chronic granulomatous disease.
Nat Med
2010
, vol. 
16
 
2
(pg. 
198
-
204
)
136
Hacein-Bey-Abina
 
S
von Kalle
 
C
Schmidt
 
M
, et al. 
A serious adverse event after successful gene therapy for X-linked severe combined immunodeficiency.
N Engl J Med
2003
, vol. 
348
 
3
(pg. 
255
-
256
)
137
Montini
 
E
Cesana
 
D
Schmidt
 
M
, et al. 
Hematopoietic stem cell gene transfer in a tumor-prone mouse model uncovers low genotoxicity of lentiviral vector integration.
Nat Biotechnol
2006
, vol. 
24
 
6
(pg. 
687
-
696
)
138
Cesana
 
D
Ranzani
 
M
Volpin
 
M
, et al. 
Uncovering and dissecting the genotoxicity of self-inactivating lentiviral vectors in vivo.
Mol Ther
2014
, vol. 
22
 
4
(pg. 
774
-
785
)
139
Cornils
 
K
Bartholomae
 
CC
Thielecke
 
L
, et al. 
Comparative clonal analysis of reconstitution kinetics after transplantation of hematopoietic stem cells gene marked with a lentiviral SIN or a γ-retroviral LTR vector.
Exp Hematol
2013
, vol. 
41
 
1
(pg. 
28.e3
-
38.e3
)
140
Zhou
 
S
Mody
 
D
DeRavin
 
SS
, et al. 
A self-inactivating lentiviral vector for SCID-X1 gene therapy that does not activate LMO2 expression in human T cells.
Blood
2010
, vol. 
116
 
6
(pg. 
900
-
908
)
141
Montini
 
E
Cesana
 
D
Schmidt
 
M
, et al. 
The genotoxic potential of retroviral vectors is strongly modulated by vector design and integration site selection in a mouse model of HSC gene therapy.
J Clin Invest
2009
, vol. 
119
 
4
(pg. 
964
-
975
)
142
Breda
 
L
Casu
 
C
Gardenghi
 
S
, et al. 
Therapeutic hemoglobin levels after gene transfer in β-thalassemia mice and in hematopoietic cells of β-thalassemia and sickle cells disease patients.
PLoS One
2012
, vol. 
7
 
3
pg. 
e32345
 
143
Sun
 
J
Ramos
 
A
Chapman
 
B
, et al. 
Clonal dynamics of native haematopoiesis.
Nature
2014
, vol. 
514
 
7522
(pg. 
322
-
327
)
144
Busch
 
K
Klapproth
 
K
Barile
 
M
, et al. 
Fundamental properties of unperturbed haematopoiesis from stem cells in vivo.
Nature
2015
, vol. 
518
 
7540
(pg. 
542
-
546
)
145
Fu
 
Y
Foden
 
JA
Khayter
 
C
, et al. 
High-frequency off-target mutagenesis induced by CRISPR-Cas nucleases in human cells.
Nat Biotechnol
2013
, vol. 
31
 
9
(pg. 
822
-
826
)
146
Frock
 
RL
Hu
 
J
Meyers
 
RM
Ho
 
YJ
Kii
 
E
Alt
 
FW
Genome-wide detection of DNA double-stranded breaks induced by engineered nucleases.
Nat Biotechnol
2015
, vol. 
33
 
2
(pg. 
179
-
186
)
147
Gabriel
 
R
Lombardo
 
A
Arens
 
A
, et al. 
An unbiased genome-wide analysis of zinc-finger nuclease specificity.
Nat Biotechnol
2011
, vol. 
29
 
9
(pg. 
816
-
823
)
148
Tsai
 
SQ
Zheng
 
Z
Nguyen
 
NT
, et al. 
GUIDE-seq enables genome-wide profiling of off-target cleavage by CRISPR-Cas nucleases.
Nat Biotechnol
2015
, vol. 
33
 
2
(pg. 
187
-
197
)
149
Pattanayak
 
V
Guilinger
 
JP
Liu
 
DR
Determining the specificities of TALENs, Cas9, and other genome-editing enzymes.
Methods Enzymol
2014
, vol. 
546
 (pg. 
47
-
78
)
150
Kim
 
D
Bae
 
S
Park
 
J
, et al. 
Digenome-seq: genome-wide profiling of CRISPR-Cas9 off-target effects in human cells.
Nat Methods
2015
, vol. 
12
 
3
(pg. 
237
-
243
)
151
Gabriel
 
R
von Kalle
 
C
Schmidt
 
M
Mapping the precision of genome editing.
Nat Biotechnol
2015
, vol. 
33
 
2
(pg. 
150
-
152
)
152
Steensma
 
DP
Bejar
 
R
Jaiswal
 
S
, et al. 
Clonal hematopoiesis of indeterminate potential and its distinction from myelodysplastic syndromes.
Blood
2015
, vol. 
126
 
1
(pg. 
9
-
16
)
153
Corrigan-Curay
 
J
O’Reilly
 
M
Kohn
 
DB
, et al. 
Genome editing technologies: defining a path to clinic.
Mol Ther
2015
, vol. 
23
 
5
(pg. 
796
-
806
)
Sign in via your Institution