Skip to content
BY 4.0 license Open Access Published by De Gruyter Open Access May 3, 2021

Gene therapy in PIDs, hemoglobin, ocular, neurodegenerative, and hemophilia B disorders

  • Arome Solomon Odiba EMAIL logo , Nkwachukwu Oziamara Okoro , Olanrewaju Ayodeji Durojaye and Yanjun Wu
From the journal Open Life Sciences

Abstract

A new approach is adopted to treat primary immunodeficiency disorders, such as the severe combined immunodeficiency (SCID; e.g., adenosine deaminase SCID [ADA-SCID] and IL-2 receptor X-linked severe combined immunodeficiency [SCID-X1]). The success, along with the feasibility of gene therapy, is undeniable when considering the benefits recorded for patients with different classes of diseases or disorders needing treatment, including SCID-X1 and ADA-SCID, within the last two decades. β-Thalassemia and sickle cell anemia are two prominent monogenic blood hemoglobin disorders for which a solution has been sought using gene therapy. For instance, transduced autologous CD34+ HSCs via a self-inactivating (SIN)-Lentivirus (LV) coding for a functional copy of the β-globin gene has become a feasible procedure. adeno-associated virus (AAV) vectors have found application in ocular gene transfer in retinal disease gene therapy (e.g., Leber’s congenital amaurosis type 2), where no prior treatment existed. In neurodegenerative disorders, successes are now reported for cases involving metachromatic leukodystrophy causing severe cognitive and motor damage. Gene therapy for hemophilia also remains a viable option because of the amount of cell types that are capable of synthesizing biologically active FVIII and FIX following gene transfer using AAV vectors in vivo to correct hemophilia B (FIX deficiency), and it is considered an ideal target, as proven in preclinical studies. Recently, the clustered regularly interspaced palindromic repeats (CRISPR)/CRISPR-associated protein 9 gene-editing tool has taken a center stage in gene therapy research and is reported to be efficient and highly precise. The application of gene therapy to these areas has pushed forward the therapeutic clinical application.

1 Background

Molecular biology and biotechnology tools remained important elements in gene therapy. Gene editing/modification (replacement, insertion, and deletion) largely characterizes this field of biological sciences. The idea of gene therapy was implemented clinically about three decades ago as an alternative to the limitations of pharmacotherapy. Approximately 3,000 known clinical trials are on record (Table 1). Some limitations are associated with gene therapy; and hence, the need to improve on the current strategies. This has resulted in sophisticated tools using viral and nonviral vectors (Figure 1). Although most of the gene therapy studies are directed toward cancer worldwide (http://www.wiley.com/legacy/wileychi/genmed/clinical/), other areas of notable disease require the gene therapy approach; and these include primary immunodeficiency disorders (PIDs), hemoglobin, hemophilia B, ocular, and neurodegenerative disorders. Gene therapy encompasses the introduction of new therapeutic genes, as DNA segments, modification of existing genes, or the introduction of RNA into cells, with the aim of preventing, treating, or curing disorders and diseases in order to restore or add gene expression. Diseases such as diabetes, Parkinson’s disease, heart failure, cancers, and neurodegenerative and metabolic disorders have been well managed through gene therapy [1].

Table 1

Indications (disease conditions) addressed by gene therapy clinical trials in order of descending percentage hierarchy

Indications Gene therapy clinical trials
Number Percentage
Cancer diseases 1,590 64.6
Monogenic diseases 259 10.5
Infectious diseases 182 7.4
Cardiovascular diseases 178 7.2
Healthy volunteers 54 2.2
Neurodegenerative diseases 45 1.8
Others 56 2.3
Gene marking 50 2
Ocular diseases 34 1.4
Inflammatory diseases 15 0.6
Total 2,463

Data sourced from: www.wiley.co.uk/genmed/clinical on 2nd June, 2017.

Figure 1 
               Viral and nonviral vectors used in gene therapy. Data sourced from: www.wiley.co.uk/genmed/clinical, 2018.
Figure 1

Viral and nonviral vectors used in gene therapy. Data sourced from: www.wiley.co.uk/genmed/clinical, 2018.

These strides are, however, not devoid of challenges that include recognizing the protein on the viral capsid by host immune system; delineating it as an antigen, though this has been managed through the development of elaborate technologies; and shielding the protein sites from eliciting inflammatory activities. Hence, this produces excellent safety profiles even for in vivo gene therapy [2]. A key advantage of this protocol is that adverse reactions and side effects are reduced to a feasible minimum, and it also provides room for the administration of the desired concentration. Therapeutic gene treatment also holds healthier potentials for use in managing diseases such as hemoglobinopathies, cancer immunotherapies, hemophilia B, ocular diseases, neurodegenerative diseases [3], immunological and metabolic disorders, and hematological diseases that have eluded cure, in resistance to the conventional treatment options [4,5]. Rapid improvements are recorded every year. Many phases of clinical and experimental trials are in the pipeline to improve on all the factors surrounding the anticipated accomplishments in gene therapy. It is, however, very pertinent to keep the medical and research community abreast of the most recent developments and redirections discussed under some of the systematic headings in the selected areas of PIDs, hemoglobin, ocular, neurodegenerative, and hemophilia B disorders as addressed in this study. We further briefly highlight the application of the more recent gene-editing tool, clustered regularly interspaced palindromic repeats (CRISPR)/CRISPR-associated protein 9 (CRISPR-cas9), but will be discussed in greater detail in the future reviews.

2 The path to treating PIDs

Records have been documented for many monogenic disorders (inheritable disorders that are a result of a single-defective gene on the autosomes) that require treatment using gene therapy. These are caused by a mutation in a single gene. Examples include cystic fibrosis, sickle cell disease, Wiskott–Aldrich syndrome (WAS), chronic granulomatous disease (CGD), Tay–Sachs disease, polycystic kidney disease, and PIDs, which comprise a class of rare, inheritable disorders of the immune system. A typical instance includes the severe combined immunodeficiency ([SCID] e.g., adenosine deaminase SCID [ADA-SCID] and IL-2 receptor X-linked severe combined immunodeficiency [SCID-X1]) [6]. The single most common strategy in dealing with these forms of diseases has been the genetic modification of hematopoietic stem cell (HSC), which has proven efficacious over the years. For example, HSC can be transplanted from a donor that has a similar and compatible HLA to a beneficiary (patient) to treat PID. Over 90% success has been recorded in such trials [7]. Substantial feasibility has been documented for the ex vivo delivery of a single gene into the isolated HSC, followed by transplant, with very well-established protocols. However, because of the complications associated with mismatch and incompatibility with some donors and their recipients, it is considered more expedient to modify the patient’s own HSC and readminister to the patient. Therefore, the donor is the recipient, and this factor rules out compatibility complications.

The phenotypic manifestation of the CGD is that the obviously mature phagocytes cannot carry out their function of killing the ingested microorganisms, potentially leading to serious inflammation. p22phox, gp91phox, p67phox, p47phox, and p40phox are all genes coding for the NADPH oxidase complex of the phagocytes, and it is in this set of genes that mutations occur, thereby impairing their function [8]. The catalytic subunit gp91phox coded by the cytochrome b (558) gene (CYBB) is located on the X-chromosome (X-CGD) and is usually affected (mutated) in most disease cases. Hence, treatment by delivering gp91phox as a transgene cures patients without major side effects. Prominent locations where ADA deficiency has been treated include the US, UK, and Italy, with commendable successes [9], though not 100% of the time. For example, in the case of some SCID-X1 patients treated by gene therapy, the enthusiasm was dampened as 2–6 years later an onset of acute T-cell lymphoblastic leukemia (T-ALL) was reported in 5 of 60 patients who received HSC as gene therapy in the past [10]. However, the situation was resolved or managed by the activation of LIN-11, Isl-1 and MEC-3 (LIM) domain only 2 (LMO2), a proto-oncogene, and a transcriptional cofactor capable of promoting the self-renewal of committed T cells [11]. The previous intended therapeutic use of gamma retroviral vectors (usually containing intact 50 long terminal repeats [LTRs]) produced severe adverse effects that were noticed in early implementation of gene therapy trials, calling for a search into the process of retroviral incorporation into human cell lines as well as primary human HSCs (CD34þ), which ultimately elucidated the need to search for better alternatives [12]. Even when using gene therapy methods, different outcomes have been recorded with different and distinct PIDs. This necessitates the establishment of protocols that consider the uniqueness of the different diseases, especially as they relate to vector design. This is actually dependent on the fundamental understanding of the molecular basis for interaction.

It is important to note that viral vector integration is actually an active process catalyzed by the tethering of the viral preintegration complex to the open chromatin regions in the genome of the host cell characterized by DNaseI hypersensitive sites as well as epigenetic marks [12]. Considering all the previous points so far, vector–chromatin interaction and its consequences have turned out to be appreciably more predictable; however, vector-induced leukemogenesis remains an unpredictable factor as it pertains to gene therapy, owing to its multifactorial nature. Nevertheless, there is a need to balance the glaring potential adverse effects against the clinical benefits for a particular patient, while putting the clinical complications associated with alternative treatments into perspective (e.g., allogeneic hematopoietic stem cell transplantation (HSCT) from a mismatched donor). The success along with the feasibility of gene therapy is undeniable, taking into account the benefits recorded with patients of different classes of diseases or disorders needing cure, including SCID-X1 and ADA-SCID, within the last two decades. PIDs (manifesting in the so-called bubble boys and girls) are rare without any doubt; nevertheless, they are life-threatening genetic diseases capable of severely compromising the integrity and functions of the immune system. This forces such individuals to live in a pathogen-free environment due to an inefficient immune system. PIDs targeted by gene therapy, which include ADA-SCID, SCID-X1, WAS, and CGD, occur in children with symptoms that include recurrent infections, system failure, and often death, most likely soon after birth. Bone marrow transplantation (using HSC) from human leukocyte antigen (HLA)-matched donors (<20% availability) has been a good alternative; however, the risks and complications associated with such strategies typically outweigh the benefits [13,14,15]. The alternative strategies that address these setbacks/limitations of scarce HLA-matched donors are found in the therapeutic potentials of autologous HSCs. As a result, this helps to avoid complications associated with immune suppression (IS) as well as the hurdle of graft against host infections [16,17,18,19].

Major setbacks and safety concerns are involved in the use of the existing murine γ-retroviral vectors designed for ex vivo gene transfer, leading to the onset of leukemia in a good number of patients who received treatment [20,21,22]. However, SCID-X1, which actually involves immunodeficiency disorders with T cells absent, natural killer (NK) cell deficiency, impaired function of B cells, and γ-chain (γc)-dependent cytokines, can be corrected by the reconstitution of immune competence in patients (for instance, via the application of SIN viral vectors). One of the core advantages of SIN is that they are devoid of LTR promoter/enhancer function, thereby reducing the risk of clonal dominance as well as insertion-related mutagenesis [23,24,25]. A similar application of the γ-retroviral vectors (murine) is present in ADA-SCID, a fatal PID with impaired T-, B-, and NK-cell growth, characterized by mutations in the gene coding for ADA (responsible for detoxifying the body of toxic purine metabolites), thus exposing patients to severe infections [9,16,17,18,26].

A number of ongoing clinical trials (#NCT01175239 (London), #NCT01410019 (Paris), #NCT01129544 (United States), #NCT01852071, #NCT01380990) are aimed at infusing interleukin-2 receptor γ-chain gene (IL2RG) to autologous HSCs in order to be transduced into SCID-X1 recipients to restore T-cell efficiency [26,27]. The same approach applies to WAS (#NCT01515462). There are also other areas where SIN-LVs are beneficially used. Autologous CD34+ HSCs derived from bone marrow, transduced ex vivo by means of LV-w1.6W vectors that convey a corrected copy of WASp gene, is one of the specific applications of LVs to WAS therapy. Compared with the γ-retroviral strategy for gene therapy in WAS, no clonal dominance or insertional mutagenesis was observed, and there is no need for a germ-free environment [28,29,30]. In addition to the gene correction in multiple cell lineages as well as the noticeably faster immunological reconstitution, the absence of detectable vector integration-associated adverse reactions is observed up to 3 years after close monitoring [31,32,33]. Approaches, such as employing a myeloid-specific promoter to avoid insertion-related mutation complications, are aimed at achieving better gene-corrected cell reconstitution and are also currently a critical area of focus. In these trials, however, some conditions could not be monitored effectively due to some other influencing conditions such as the metabolic effects of enzyme replacement therapy [9,16,34,35]. Low engraftment of gene-corrected cells in attempts to manage CGD (an uncommon genetic disorder because of a mutation in gp91phox subunit of nicotinamide adenine dinucleotide phosphate oxidase [NADPH]) has proven relatively unsuccessful due to the transient clinical benefits that it provides, as evidenced in the recurrence of the disease. The disease results in the failure of phagocytic cells (macrophages and neutrophils) to generate reactive oxygen species, hampering the efficient clearance of fungal and bacterial infections [36]. The use of γ-retroviral vectors for this disease also results in clonal dominance of gene-corrected cells capable of producing monosomy and myelodysplastic syndrome [27,36,37]. Most importantly, now is the time to put a spotlight on the most recent developments in this field as well as putting forth a road map for the future breakthroughs.

Vaccines are most commonly made up from the attenuated pathogens in forms administrable into a host’s system. However, some vaccines fail, and two main reasons could account for why vaccines fail after their preliminary development. The first is failure of the immune response, and the second is pathogenicity. The potency of a vaccine heavily relies on its ability to elicit an immune response just by recognition, while remaining nonpathogenic. However, the main cause of why vaccines fail, is that the vaccine candidate may not induce an immune response, as we lack full and comprehensive knowledge of the complexities of the host immune system. Second cause is the ability of the vaccine candidate to illicit an immune response by actually becoming pathogenic. In most PID cases, administration of subcutaneous or intravenous immunoglobulin is the primary choice of treatment. The main challenges with the use of vaccines in PIDs are the impairment of immune response, and impediment the immune system’s antigen stimulation, which results in very little or no protection caused by the immunodeficiency in some PID cases. In addition, adverse effects often result from vaccines as well as possible emergence of pathogenesis from vaccine strains. For example, in patients with phagocytic cell defects or defective T cells or NK cells (leukocyte adhesion deficiency), the immune response could be altered as a result of the live viral as well as live attenuated bacterial vaccines. This can create severe disorders linked to the vaccine strains, with the Chediak–Higashi syndrome and CGD being a common example.

2.1 Current developments en route for managing hemoglobin disorders such as sickle cell disease and β-thalassemia

β-Thalassemia and sickle cell anemia are two prominent monogenic blood hemoglobin disorders responsible for early mortality and morbidity. Mutation in the β-globin of hemoglobin is what gives rise to hemoglobin S, otherwise known as sickle hemoglobin, resulting in a “sickle-shaped” erythrocytes that lack flexibility, thereby obstructing blood flow by sticking to blood vessels and depriving cells of oxygen. Progress reports show that, where hydroxyurea with the purpose of increasing the expression of the fetal globin gene failed, HSC gene transfer of a β-globin mediated by LV has proven successful in the treatment of a child (13-year-old), producing a 47% β-globin expression when monitored for more than a year. The strategy is that the β-globin carrying a missense mutation that confers “anti-sickling” properties (βA-T87Q) on it [38,39]. β-Thalassemia stems from the loss of β-globin, a functional component of hemoglobin in red blood cells. The previously available treatment options (with the exception of bone marrow transplantation), such as frequent blood transfusions and chelation therapy to avert the amassing of iron, are not curative and have associated side effects. Complications of a matched donor arise with the bone marrow transplant option, in addition to the high risk.

Although challenges are involved in acquiring an adequate amount of corrected globin protein expressing genes in erythrocytes, curative gene transfer to HSC seems to be a good alternative. For instance, transduced autologous CD34+ HSCs via an SIN-LV coding for a functional copy of the β-globin gene has become a feasible procedure. However, myeloablative conditioning was necessary before reinfusing the gene-corrected HSCs. In addition to the phase 1/2 studies (#NCT01639690, #NCT02151526, #NCT02453477, and #NCT01745120), where LV transfer of an engineered β-globin gene to HSC has been used to treat patients (e.g., those with βE form) without the need for multiple transfusions and iron chelation therapy, recipients of such treatments are progressing healthwise for up to 2 years and beyond [38,40]. This treatment option is not as straightforward as in the case of the β 0/β 0 genotype, due to the lack of endogenous expression, resulting in only a partial correction of the disorder. Relentless efforts toward making a cure in the near future for many genetic diseases, including globin disorder, are also ongoing in different locations, including Thailand, the US, Italy, Australia, France, and other countries (Table 2).

Table 2

Various countries involved in the geographical distribution of gene therapy clinical trials in order of descending percentage hierarchy

Country Gene therapy clinical trials
Number Percentage
USA 1,550 62.9
UK 219 8.9
Multicountry 120 4.9
Germany 92 3.7
China 68 2.8
France 57 2.3
Switzerland 50 2
Japan 42 1.7
The Netherlands 36 1.5
Australia 32 1.3
Spain 29 1.2
Canada 27 1.1
Italy 26 1.1
Belgium 22 0.9
South Korea 20 0.8
Sweden 12 0.5
Russia 10 0.4
Israel 8 0.3
Poland 6 0.2
Finland 6 0.2
Norway 4 0.2
Austria 3 0.1
Singapore 3 0.1
Czech Republic 2 0.1
Denmark 2 0.1
Ireland 2 0.1
Mexico 2 0.1
New Zealand 2 0.1
Taiwan 2 0.1
Kenya 1 0
Kuwait 1 0
Gambia 1 0
Romania 1 0
Uganda 1 0
Burkina Faso 1 0
Egypt 1 0
Saudi Arabia 1 0
Senegal 1 0
Total 2,463

Data sourced from: www.wiley.co.uk/genmed/clinical on June, 2, 2017.

2.2 Improvements for ocular disease

The AAV vectors find application in ocular gene transfer and have been extensively used in retinal diseases gene therapy (NCT02161380 and #NCT01267422), including inborn forms of blindness (e.g., Leber’s congenital amaurosis type 2 [LCA2]) for which no prior treatment existed [41,42]. Using a murine model in gene therapy efforts on Leber hereditary optic neuropathy has shown that a single intravitreal inoculation of scAAV2, having triple Y-F mutations in AAV2 capsid and conveying a wild-type human ND4 gene, is able to further halt the degeneration of the optic nerve, accompanied by a significantly elevated proportion of complex-I-dependent ATP synthesis. This suggests the correction of the compromised electron transport chain (ETC). Mutations in the RPE65 gene (expressing a 65-kilodalton protein) lead to retinal pigment epithelium inexpression, thereby impairing visual phototransduction. It is a mitochondrial gene affecting the complex I of ETC. Clinical studies (#NCT00643747, #NCT01208389, and #NCT00481546) have demonstrated that a single-dose subretinal injection of AAV2 vector to deliver the therapeutic gene (RPE65) can improve vision [41,43,44,45,46,47,48]. Through well-modulated final formulation, vector design, and immunomodulatory regimens, long-term follow-up on these clinical studies suggests an improvement in the visual acuity, sensitivity of the retina, and gain of function over time [49]. Canine and murine models have proven that gene amplification could limit the progression of degeneration if the intervention starts early [50,51,52,53]. Accomplishment with the LCA2 gene therapy is a milestone for clinical trials in other retinal diseases (#NCT01461213; for choroideremia) [54]. This genetic disease results from a nonfunctional copy of the CHM gene. Obvious diseased manifestations are slow and advance deterioration of the patient’s choroid, photoreceptors, and retinal-pigmented epithelium. This could result in the complete loss of sight at an intermediate age. However, a subretinal administration of the AAV2 vector conveying the CHM gene significantly improves the vision.

2.3 Revisiting neurodegenerative disorders

The central nervous system is the most complex of all the systems in the body, and as such, treating neurodegenerative disorders with conventional pharmacological medications is very difficult. The complexity is made obvious in the blood–brain barrier (BBB), which could restrict even the most therapeutically potent agent from getting access to the target site. Although molecular gene therapy seems to solve this difficulty, the challenges of vector delivery to the CNS-specific target site persist. This notwithstanding, successes have now been reported for cases involving metachromatic leukodystrophy (in which patients are deficient in arylsulfatase A [ARSA], leading to the accumulation of sulfatide in myelin-producing cells that produce severe cognitive and motor damage), aromatic l-amino acid decarboxylase (AADC) deficiency, and adrenoleukodystrophy (ALD, with mutated ABCD1 gene, affecting the adrenal cortex). These procedures employ LV (integrating) and nonintegrating AAV without any reported major safety concern [55]. Unexpectedly, neither HSC nor bone marrow transplant can effectively solve this challenge. It is, however, currently hypothesized that overexpression of ARSA in genetically modified hematopoietic cells could possibly remove these barriers associated with bone marrow transplant through the delivery of ARSA to stop the progress of demyelination (#NCT01560182), employing SIN-LV-mediated gene transfer in autologous CD34+ HSCs. A continuous production of the functional ARSA protein with normal cognitive and motor development without established insertion-associated mutagenesis has been reported. AADC deficiency results in impairment of the synthesis as well as secretion of neurotransmitters, including dopamine and serotonin. This leads to setbacks characterized by dystonia, oculogyric crises, truncal hypotonia, severe movement disorders, sweating, tongue protrusion, neurodegenerative impairment in children, and jaw spasms.

A gene therapy clinical trial (#NCT01395641) on AADC through direct injection of AAV2 vector to deliver AADC gene within the bilateral putamen, with obvious restoration in the patients, is currently in progress. Commendable scores on the Peabody Developmental Motor Scale, Alberta Infant Motor Scale, toddlers scores, and comprehensive developmental inventory support this for infants after 15–24 months of gene therapy [56,57]. Transient increases in dyskinesia and frequent episodes of apnea were noticeable adverse effects. Other clinical trials (e.g., #NCT00229736 and #NCT02122952), along this line in relation to Parkinson’s disease, spinal muscular atrophy type 1, and Canavan disease, are ongoing [58]. Other limitations associated with gene therapy applications to neurodegenerative disorders, such as unintended binding to extracellular matrix components and profuse spread from the injection site, have been observed. Capsid engineering, cisterna magna, and serotype alternatives are viable means of alleviating these common challenges [59,60].

2.4 Better approach to hemophilia B coagulation disorder

A very common well-known blood disease is hemophilia, a hematological disorder caused by mutations in the gene that codes for coagulation factor VIII or IX. The disorder is X-linked and occurs in about 1 in 5,000 (hemophilia A) or 1 in 30,000 (hemophilia B) male births throughout the world. In very simple and straightforward terms, the blood refuses to clot. Many possible consequences of these mutation errors could be deduced clinically. Contemporary treatment available for hemophilia involves very tedious and regular (about 2–3 times per week) intravenous infusion of recombinant or FVIII/FIX proteins (both of which are secreted in inactive form at plasma levels of 200 and 5,000 ng/mL, respectively) derived from the plasma. Such lifelong disease management is burdensome as well as expensive, and the third-world countries often cannot afford them. Gene therapy, on the other hand, is curative and sustained (greater than 10 years in canine models) following a single round of gene transfer. The gene therapy option for hemophilia remains a viable alternative because of the amount of cell types that are capable of synthesizing biologically active FVIII (synthesized basically in endothelial cells, including liver sinusoidal) and FIX (fundamentally synthesized in the hepatocytes) following gene transfer.

By using AAV vectors in vivo, molecular gene transfer through the hepatic artery of the liver can correct hemophilia B (FIX deficiency) and is considered an ideal strategy as demonstrated in preclinical studies [61,62]. However, additional studies showed that memory CD8+ T cells and neutralizing antibodies guard against the AAV capsid (particularly serotype 2) in humans who have had previous natural exposure to AAV, thereby eliminating the therapeutic transduced hepatocytes and obstructing gene transfer to the liver beyond a particular titer [63]. These points also highlight the fact that the immune system remains a challenge for in vivo gene transfer, noting that, in contrast to adenovirus, unaided AAV vectors do not incite strong immune responses. Clinical trials (e.g., #NCT00979238) have been initiated and executed on AAV8 (a different AAV serotype) with an abated production frequency of AAV neutralizing antibodies [64,65]. This strategy also provides for a less invasive peripheral vein administration by taking advantage of the self-complementary genome, a codon-optimized F9 sequence. This option provides the advantage of a transient IS regimen using prednisolone if the patients are showing mild transaminitis or loss of circulating FIX [66,67]. Hyperactive FIX variant (R338L, also FIX-Padua) occurring naturally in a self-complementary AAV8 vector is also in the pipeline of phase 1/2 molecular gene therapy testing (#NCT01687608) for treating hemophilia using animal models [68,69,70,71,72,73]. A long-term expression with enhanced catalytic action of FIX variant (FIXR338L) at reduced curative intended vector doses (scAAV8-FIXR338L) was observed [74,75]. Compared to the IS using prednisolone, which could not salvage expression, the subjects who were administered the highest vector dose lost expression, expressing IFN-ϒ producing T cells and transaminitis in response to the viral capsid antigen, though no liver toxicity was observed. Over time, no signs of patients’ emerging immune response were observed as a countermechanism against the FIX protein in more than 20 hemophilia B patients who have received AAV-F9 vector treatments, with additional research backing up these claims. However, some inconsistencies are observed with different clinical trials (Table 1), potentially related to issues such as vector design, vector doses, use of suppressive steroid drugs, higher transduction efficiency, the occurrence of immune stimulatory CpG motifs, or other factors for which answers and explanations are necessary [76,77,78,79,80,81,82,83].

3 CRISPR-Cas9 and the future of gene therapy

The CRISPR-Cas9 genome editing tool has recently become a common preference for gene therapy studies in basic research as well as in clinical trials. Though discussions are still ongoing concerning ethics in human applications, the CRISPR-Cas9 technology is undoubtedly a powerful tool for gene therapy when fully refined. CRISPR-Cas9 is an efficient site-directed genome editing tool mainly composed of Cas9 RNA-guided nuclease and CRISPR RNA (crRNA), which directs the Cas9 enzyme to the sequence of interest in the genome. This specificity is determined by a 20-nucleotide sequence complementary to the crRNA, following a protospacer adjacent motif sequence [84,85]. The CRISPR system uses the host cell’s innate error-prone nonhomologous end joining to achieve the aim. Many concerns regarding off-target and associated complications are still under review by experts to ascertain the safety of this genetic tool before it can be implemented in human trials. Genetic engineering and biotechnology companies are keeping a close look at the CRISPR-Cas9 tool as the technology has been used very recently in basic science research, and clinical trials are commencing in pharmaceutical research to use the CRISPR-Cas9 to treat patients with β-thalassemia [86].

Although CRISPR-Cas9 has proven to be an enormous scientific breakthrough, it has many limitations that have made it a frontline topic of discussion in research and medical applications, with legal implications. In 2015, an international summit on human gene editing supported the establishment of an international committee to assess the implications of CRISPR-Cas9. The Committee on Ethics, Law and Society of the Human Genome Organization on CRISPR-Cas9 produced some vital ethical points to consider in the application of CRISPR-Cas9, most particularly in its application to human biology and medicine [87]. On the bright side, CRISPR targets almost any nucleotide sequence with a short length of specific genes; however, it does not faithfully insert new target DNA sequences. Off-targets often occur, creating deleterious effects that are sometimes irreversible. This error leads to serious mutations such as in Friedreich ataxia, fragile X syndrome, and Huntington disease. As effective as it is, the CRISPR system can fail in identifying the right target, including introns, and as a result cause further damage. Challenges and ethical considerations about the use of CRISPR technology are not just peculiar to CRISPR but common to all gene editing tools previously employed. Experiences and challenges with gene editing raise discussion on delaying the application of the CRISPR-Cas9 in medical applications until well proven and tested using highly supported basic science [88,89,90,91]. For instance, an example is the emergence of childhood leukemia arising when the viral vector activated a latent human oncogene, as Double Strand Breaks (DSBs) generated by CRISPR-Cas9 often induces apoptosis. Some of the components of the CRISPR-Cas9 system, such as the cas9 protein, are of microbial origin and may trigger immune responses in the human system, which may be undetected for a longtime. The CRISPR-Cas9 system has been around only for a short time and requires more time to investigate all the implications associated with its use in genetic engineering.

4 Conclusions

In recent times, the field of gene therapy has recorded clinical achievements in various disease types by employing different methodologies and improved vectors. Recently, numerous gene therapy trials have validated the clinical benefits of the application of molecular biology and biotechnology tools in treating diseases or disorders that have proven elusive to cure using conventional pharmacotherapy. Such areas of application include hereditary immune system disorders, congenital eye diseases, hemophilia B, lipoprotein lipase deficiency, and adoptive transfer of genetically engineered T cells for cancer. Although gene therapy possesses obvious scientific, ethical, and technological challenges, many efforts have been made to facilitate the effectual translation of gene therapy into clinical practice. Both recent and noteworthy is the application of gene therapy to PIDs, hemoglobin, hemophilia B, ocular, and neurodegenerative disorders. Most recently, but still under strict refining, is the CRISPR-cas9 gene-editing tool; although it currently has some setbacks, it is promising at providing solutions to many genetic diseases. The application of gene therapy to these areas has pushed medical research and other applications toward becoming the remedy we have sought concerning such disease areas.


tel: +234-706-767-7773, +86-156-7714-2215

  1. Funding information: The authors state no funding involved.

  2. Conflict of interest: The authors state no conflict of interest.

  3. Data availability statement: Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.

References

[1] Elsner M, Terbish T, Jörns A, Naujok O, Wedekind D, Hedrich H-J, et al. Reversal of diabetes through gene therapy of diabetic rats by hepatic insulin expression via lentiviral transduction. Mol Ther. 2012;20:918–26.10.1038/mt.2012.8Search in Google Scholar PubMed PubMed Central

[2] Russell SJ, Peng K-W, Bell JC. Oncolytic virotherapy. Nat Biotechnol. 2012;30:658–70.10.1038/nbt.2287Search in Google Scholar PubMed PubMed Central

[3] Kaufman HL, Kohlhapp FJ, Zloza A. Oncolytic viruses: a new class of immunotherapy drugs. Nat Rev Drug Discov. 2015;14:642–62.10.1038/nrd4663Search in Google Scholar PubMed PubMed Central

[4] Naldini L. Medicine. A comeback for gene therapy. Science. 2009;326:805–6.10.1126/science.1181937Search in Google Scholar PubMed

[5] Herzog RW, Cao O, Srivastava A. Two decades of clinical gene therapy–success is finally mounting. Discov Med. 2010;9:105–11.Search in Google Scholar

[6] Fischer A, Hacein-Bey-Abina S, Cavazzana-Calvo M. Gene therapy of primary T cell immunodeficiencies. Gene. 2013;525:170–3.10.1016/j.gene.2013.03.092Search in Google Scholar PubMed

[7] Railey MD, Lokhnygina Y, Buckley RH. Long-term clinical outcome of patients with severe combined immunodeficiency who received related donor bone marrow transplants without pretransplant chemotherapy or post-transplant GVHD prophylaxis. J Pediatr. 2009;155:834–40.10.1016/j.jpeds.2009.07.049Search in Google Scholar PubMed PubMed Central

[8] van den Berg JM, van Koppen E, Ahlin A, Belohradsky BH, Bernatowska E, Corbeel L, et al. Chronic granulomatous disease: the European experience. PLoS One. 2009;4:e5234.10.1371/journal.pone.0005234Search in Google Scholar PubMed PubMed Central

[9] Candotti F, Shaw KL, Muul L, Carbonaro D, Sokolic R, Choi C, et al. Gene therapy for adenosine deaminase-deficient severe combined immune deficiency: clinical comparison of retroviral vectors and treatment plans. Blood. 2012;120:3635–46.10.1182/blood-2012-02-400937Search in Google Scholar PubMed PubMed Central

[10] Cavazzana-Calvo M, Fischer A, Hacein-Bey-Abina S, Aiuti A. Gene therapy for primary immunodeficiencies: part 1. Curr Opin Immunol. 2012;24:580–4.10.1016/j.coi.2012.08.008Search in Google Scholar PubMed

[11] McCormack MP, Young LF, Vasudevan S, de Graaf CA, Codrington R, Rabbitts TH, et al. The Lmo2 oncogene initiates leukemia in mice by inducing thymocyte self-renewal. Science. 2010;327:879–83.10.1126/science.1182378Search in Google Scholar PubMed

[12] Deichmann A, Brugman MH, Bartholomae CC, Schwarzwaelder K, Verstegen MM, Howe SJ, et al. Insertion sites in engrafted cells cluster within a limited repertoire of genomic areas after gammaretroviral vector gene therapy. Mol Ther. 2011;19:2031–9.10.1038/mt.2011.178Search in Google Scholar PubMed PubMed Central

[13] Boztug K, Schmidt M, Schwarzer A, Banerjee PP, Diez IA, Dewey RA, et al. Stemcell gene therapy for the Wiskott–Aldrich syndrome. N Engl J Med. 2010;363:1918–27.10.1056/NEJMoa1003548Search in Google Scholar PubMed PubMed Central

[14] Griffith LM, Cowan MJ, Notarangelo LD, Kohn DB, Puck JM, Pai SY, et al. Primary immune deficiency treatment consortium (PIDTC) report. J Allergy Clin Immunol. 2014;133:335–47.10.1016/j.jaci.2013.07.052Search in Google Scholar PubMed PubMed Central

[15] Braun CJ, Boztug K, Paruzynski A, Witzel M, Schwarzer A, Rothe M, et al. Gene therapy for Wiskott–Aldrich syndrome – long-term efficacy and genotoxicity. Sci Transl Med. 2014;6:227–33.10.1126/scitranslmed.3007280Search in Google Scholar PubMed

[16] Aiuti A, Slavin S, Aker M, Ficara F, Deola S, Mortellaro A, et al. Correction of ADASCID by stem cell gene therapy combined with nonmyeloablative conditioning. Science. 2002;296:2410–3.10.1126/science.1070104Search in Google Scholar PubMed

[17] Aiuti A, Cattaneo F, Galimberti S, Benninghoff U, Cassani B, Callegaro L, et al. Gene therapy for immunodeficiency due to adenosine deaminase deficiency. N Engl J Med. 2009;360:447–58.10.1056/NEJMoa0805817Search in Google Scholar PubMed

[18] Gaspar HB, Cooray S, Gilmour KC, Parsley KL, Zhang F, Adams S, et al. Hematopoietic stem cell gene therapy for adenosine deaminase-deficient severe combined immunodeficiency leads to long-term immunological recovery and metabolic correction. Sci Transl Med. 2011;3:97ra80.10.1126/scitranslmed.3002716Search in Google Scholar PubMed

[19] Aiuti A, Bacchetta R, Seger R, Villa A, Cavazzana-Calvo M. Gene therapy for primary immunodeficiencies: part 2. Curr Opin Immunol. 2012;24:585–91.10.1016/j.coi.2012.07.012Search in Google Scholar PubMed

[20] Hacein-Bey-Abina S, von Kalle C, Schmidt M, Le Deist F, Wulffraat N, McIntyre E, et al. A serious adverse event after successful gene therapy for X-linked severe combined immunodeficiency. N Engl J Med. 2003;348:255–6.10.1056/NEJM200301163480314Search in Google Scholar PubMed

[21] Hacein-Bey-Abina S, Garrigue A, Wang GP, Soulier J, Lim A, Morillon E, et al. Insertional oncogenesis in 4 patients after retrovirus-mediated gene therapy of SCID-X1. J Clin Invest. 2008;118:3132–42.10.1172/JCI35700Search in Google Scholar PubMed PubMed Central

[22] Howe SJ, Mansour MR, Schwarzwaelder K, Bartholomae C, Hubank M, Kempski H, et al. Insertional mutagenesis combined with acquired somatic mutations causes leukemogenesis following gene therapy of SCID-X1 patients. J Clin Invest. 2008;118:3143–50.10.1172/JCI35798Search in Google Scholar PubMed PubMed Central

[23] Montini E, Cesana D, Schmidt M, Sanvito F, Ponzoni M, Bartholomae C, et al. Hematopoietic stem cell gene transfer in a tumor-prone mouse model uncovers low genotoxicity of lentiviral vector integration. Nat Biotechnol. 2006;24:687–96.10.1038/nbt1216Search in Google Scholar PubMed

[24] Modlich U, Navarro S, Zychlinski D, Maetzig T, Knoess S, Brugman MH, et al. Insertional transformation of hematopoietic cells by self-inactivating lentiviral and gammaretroviral vectors. Mol Ther. 2009;17:1919–28.10.1038/mt.2009.179Search in Google Scholar PubMed PubMed Central

[25] Xu W, Russ JL, Eiden MV. Evaluation of residual promoter activity in γ-retroviral self-inactivating (SIN) vectors. Mol Ther. 2012;20:84–90.10.1038/mt.2011.204Search in Google Scholar PubMed PubMed Central

[26] Gaspar BB, Rivat C, Himoudi N, Gilmour K, Booth C, Xu-Bayford J. Immunological and metabolic correction after lentiviral vector mediated haematopoietic stem cell gene therapy for ADA deficiency. J Clin Immunol. 2014;34(Suppl2):S167–8.Search in Google Scholar

[27] Cicalese MP, Aiuti A. Clinical applications of gene therapy for primary immunodeficiencies. Hum Gene Ther. 2015;26:210–9.10.1089/hum.2015.047Search in Google Scholar PubMed PubMed Central

[28] Qasim W, Gennery AR. Gene therapy for primary immunodeficiencies: current status and future prospects. Drugs. 2014;74:963–9.10.1007/s40265-014-0223-7Search in Google Scholar PubMed

[29] Ferrua FM, Galimberti S, Scaramuzza S, Giannelli S, Pajno R, Dionisio F, et al. Safety and clinical benefit of lentiviral hematopoietic stem cell gene therapy for Wiskott–Aldrich syndrome. Blood. 2015;126:259.10.1182/blood.V126.23.259.259Search in Google Scholar

[30] Chu JI, Myriam Armant L, Male F, Dansereau CH, MacKinnon B, Burke CJ, et al. Gene therapy using a self-inactivating lentiviral vector improves clinical and laboratory manifestations of Wiskott–Aldrich syndrome. Blood. 2015;126:260.10.1182/blood.V126.23.260.260Search in Google Scholar

[31] Hacein-Bey-Abina S, Pai SY, Gaspar HB, Armant M, Berry CC, Blanche S, et al. A modified γ-retrovirus vector for X-linked severe combined immunodeficiency. N Engl J Med. 2014;371:1407–17.10.1056/NEJMoa1404588Search in Google Scholar PubMed PubMed Central

[32] Touzot F, Moshous D, Creidy R, Neven B, Frange P, Cros G, et al. Faster T-cell development following gene therapy compared with haploidentical HSCT in the treatment of SCID-X1. Blood. 2015;125:3563–9.10.1182/blood-2014-12-616003Search in Google Scholar PubMed

[33] Otsu M, Yamada M, Nakajima S, Kida M, Maeyama Y, Hatano N, et al. Outcomes in two Japanese adenosine deaminase-deficiency patients treated by stem cell gene therapy with no cytoreductive conditioning. J Clin Immunol. 2015;35:384–98.10.1007/s10875-015-0157-1Search in Google Scholar PubMed

[34] Aiuti A, Cassani B, Andolfi G, Mirolo M, Biasco L, Recchia A, et al. Multilineage hematopoietic reconstitution without clonal selection in ADA-SCID patients treated with stem cell gene therapy. J Clin Invest. 2007;117:2233–40.10.1172/JCI31666Search in Google Scholar PubMed PubMed Central

[35] Gaspar HB. Gene therapy for ADA-SCID: defining the factors for successful outcome. Blood. 2012;120:3628–9.10.1182/blood-2012-08-446559Search in Google Scholar PubMed

[36] Grez M, Reichenbach J, Schwable J, Seger R, Dinauer MC, Thrasher AJ. Gene therapy of chronic granulomatous disease: the engraftment dilemma. Mol Ther. 2011;19:28–35.10.1038/mt.2010.232Search in Google Scholar PubMed PubMed Central

[37] Stein S, Ott MG, Schultze-Strasser S, Jauch A, Burwinkel B, Kinner A, et al. Genomic instability and myelodysplasia with monosomy 7 consequent to EVI1 activation after gene therapy for chronic granulomatous disease. Nat Med. 2010;16:198–204.10.1038/nm.2088Search in Google Scholar PubMed

[38] Cavazzana M, Emmanuel Payen J-A, Suarez F, Beuzard Y, Touzot F, Cavallesco R, et al. Outcomes of gene therapy for severe sickle disease and beta-thalassemia major via transplantation of autologous hematopoietic stem cells transduced ex vivo with alentiviral beta AT87Q-globin vector. Blood. 2015;126:202.10.1182/blood.V126.23.202.202Search in Google Scholar

[39] Malik P. Gene therapy for hemoglobinopathies: tremendous successes and remaining caveats. Mol Ther. 2016;24:68–670.10.1038/mt.2016.57Search in Google Scholar PubMed PubMed Central

[40] Walters MC, Suradej Hongeng J, Kwiatkowski J, Schiller GJ, Kletzel M, Ho PJ, et al. Update of results from the Northstar Study (HGB-204): a phase 1/2 study of gene therapy for beta-thalassemia major via transplantation of autologous hematopoietic stem cells transduced ex-vivo with a lentiviral beta AT87Q-globin vector (LentiGlobinBB305 Drug Product). Blood. 2015;126:201.10.1182/blood.V126.23.201.201Search in Google Scholar

[41] Cideciyan AV, Hauswirth WW, Aleman TS, Kaushal S, Schwartz SB, Boye SL, et al. Human RPE65 gene therapy for Leber congenital amaurosis: persistence of early visual improvements and safety at 1 year. Hum Gene Ther. 2009;20:999–1004.10.1089/hum.2009.086Search in Google Scholar PubMed PubMed Central

[42] Boye SE, Boye SL, Lewin AS, Hauswirth WW. A comprehensive review of retinal gene therapy. Mol Ther. 2013;21:509–19.10.1038/mt.2012.280Search in Google Scholar PubMed PubMed Central

[43] Cideciyan AV, Aleman TS, Boye SL, Schwartz SB, Kaushal S, Roman AJ, et al. Human gene therapy for RPE65 isomerase deficiency activates the retinoid cycle of vision but with slow rod kinetics. Proc Natl Acad Sci. 2008;105:15112–7.10.1073/pnas.0807027105Search in Google Scholar PubMed PubMed Central

[44] Simonelli F, Maguire AM, Testa F, Pierce EA, Mingozzi F, Bennicelli JL, et al. Gene therapy for Leber’s congenital amaurosis is safe and effective through 1.5 years after vector administration. Mol Ther. 2010;18:643–50.10.1038/mt.2009.277Search in Google Scholar PubMed PubMed Central

[45] Jacobson SG, Cideciyan AV, Ratnakaram R, Heon E, Schwartz SB, Roman AJ, et al. Gene therapy for Leber congenital amaurosis caused by RPE65 mutations: safety and efficacy in 15 children and adults followed up to 3 years. Arch Ophthalmol. 2012;130:9–24.10.1001/archophthalmol.2011.298Search in Google Scholar PubMed PubMed Central

[46] Bennett J, Ashtari M, Wellman J, Marshall KA, Cyckowski LL, Chung DC, et al. AAV2 gene therapy readministration in three adults with congenital blindness. Sci Transl Med. 2012;4:120–35.10.1126/scitranslmed.3002865Search in Google Scholar PubMed PubMed Central

[47] Testa F, Maguire AM, Rossi S, Pierce EA, Melillo P, Marshall K, et al. Three- year follow-up after unilateral subretinal delivery of adeno-associated virus in patients with Leber congenital amaurosis type 2. Ophthalmology. 2013;120:1283–91.10.1016/j.ophtha.2012.11.048Search in Google Scholar PubMed PubMed Central

[48] Schimmer J, Breazzano S. Investor outlook: significance of the positive LCA2gene therapy phase III results. Hum Gene Ther Clin Dev. 2015;26:208–10.10.1089/humc.2015.29004.schSearch in Google Scholar PubMed

[49] Bainbridge JW, Mehat MS, Sundaram V, Robbie SJ, Barker SE, Ripamonti C, et al. Long-term effect of gene therapy on Leber’s congenital amaurosis. N Engl J Med. 2015;372:1887–97.10.1056/NEJMoa1414221Search in Google Scholar PubMed PubMed Central

[50] Kostic C, Crippa SV, Pignat V, Bemelmans AP, Samardzija M, Grimm C, et al. Gene therapy regenerates protein expression in cone photoreceptors in Rpe65 (R91W/R91W) mice. PLoS One. 2011;6:e16588.10.1371/journal.pone.0016588Search in Google Scholar PubMed PubMed Central

[51] Li X, Li W, Dai X, Kong F, Zheng Q, Zhou X, et al. Gene therapy rescues cone structure and function in the 3-month-old rd12 mouse: a model for midcourse RPE65 Leber congenital amaurosis. Invest Ophthalmol Vis Sci. 2011;52:7–15.10.1167/iovs.10-6138Search in Google Scholar PubMed PubMed Central

[52] Mowat FM, Breuwer AR, Bartoe JT, Annear MJ, Zhang Z, Smith AJ, et al. RPE65 gene therapy slows cone loss in Rpe65-deficient dogs. Gene Ther. 2013;20:545–55.10.1038/gt.2012.63Search in Google Scholar PubMed

[53] Shanab AY, Mysona BA, Matragoon S, El-Remessy AB. Silencing p75 (NTR) prevents proNGF-induced endothelial cell death and development of acellular capillaries in rat retina. Mol Ther Methods Clin Dev. 2015;2:15013.10.1038/mtm.2015.13Search in Google Scholar

[54] MacLaren RE, Groppe M, Barnard AR, Cottriall CL, Tolmachova T, Seymour L, et al. Retinal gene therapy in patients with choroideremia: initial findings from a phase ½ clinical trial. Lancet. 2014;383:1129–37.10.1016/S0140-6736(13)62117-0Search in Google Scholar

[55] Cartier N, Hacein-Bey-Abina S, Bartholomae CC, Bougneres P, Schmidt M, Kalle CV, et al. Lentiviral hematopoietic cell gene therapy for X-linked adrenoleukodystrophy. Methods Enzymol. 2012;507:187–98.10.1016/B978-0-12-386509-0.00010-7Search in Google Scholar

[56] Hwu WL, Muramatsu S, Tseng SH, Tzen KY, Lee NC, Chien YH, et al. Gene therapy for aromatic L-amino acid decarboxylase deficiency. Sci Transl Med. 2012;4:134–61.10.1126/scitranslmed.3003640Search in Google Scholar

[57] Lee DW, Kochenderfer JN, Stetler-Stevenson M, Cui YK, Delbrook C, Feldman SA, et al. T cells expressing CD19 chimeric antigen receptors for acute lymphoblastic leukaemia in children and young adults: a phase 1 dose-escalation trial. Lancet. 2015;385:517–28.10.1016/S0140-6736(14)61403-3Search in Google Scholar

[58] Mittermeyer G, Christine CW, Rosenbluth KH, Baker SL, Starr P, Larson P, et al. Long-term evaluation of a phase 1 study of AADC gene therapy for Parkinson’s disease. Hum Gene Ther. 2012;23:377–81.10.1089/hum.2011.220Search in Google Scholar PubMed PubMed Central

[59] Samaranch L, Salegio EA, San Sebastian W, Kells AP, Bringas JR, Forsayeth J, et al. Strong cortical and spinal cord transduction after AAV7 and AAV9 delivery into the cerebrospinal fluid of nonhuman primates. Hum Gene Ther. 2013;24:526–32.10.1089/hum.2013.005Search in Google Scholar PubMed PubMed Central

[60] Choudhury SR, Hudry E, Maguire CA, Sena-Esteves M, Breakefield XO, Grandi P. Viral vectors for therapy of neurologic diseases. Neuropharmacology. 2017;120:63–80.10.1016/j.neuropharm.2016.02.013Search in Google Scholar PubMed PubMed Central

[61] Manno CS, Chew AJ, Hutchison S, Larson PJ, Herzog RW, Arruda VR, et al. AAV mediated factor IX gene transfer to skeletal muscle in patients with severe hemophilia B. Blood. 2003;101:2963–72.10.1182/blood-2002-10-3296Search in Google Scholar PubMed

[62] Mingozzi F, Maus MV, Hui DJ, Sabatino DE, Murphy SL, Rasko JE, et al. CD8+ T-cell responses to adeno-associated virus capsid in humans. Nat Med. 2007;13:419–22.10.1038/nm1549Search in Google Scholar PubMed

[63] Hui DJ, Podsakoff GM, Pein GC, Ivanciu L, Camire RM, Ertl H, et al. AAV capsid CD8+ T cell epitopes are highly conserved across AAV serotypes. Mol Ther Methods Clin Dev. 2007;2:15029.10.1038/mtm.2015.29Search in Google Scholar PubMed PubMed Central

[64] Nathwani AC, Reiss UM, Tuddenham EG, Rosales C, Chowdary P, McIntosh J, et al. Long-term safety and efficacy of factor IX gene therapy in hemophilia B. N Engl J Med. 2014;371:1994–2004.10.1056/NEJMoa1407309Search in Google Scholar PubMed PubMed Central

[65] Rogers GL, Herzog RW. Gene therapy for hemophilia. Front Biosci (LandmarkEd). 2015;20:556–603.10.2741/4324Search in Google Scholar PubMed PubMed Central

[66] Nathwani AC, Gray JT, Ng CY, Zhou J, Spence Y, Waddington SN, et al. Self-complementary adeno-associated virus vectors containing a novel liver- specific human factor IX expression cassette enable highly efficient transduction of murine and nonhuman primate liver. Blood. 2006;107:2653–261.10.1182/blood-2005-10-4035Search in Google Scholar PubMed PubMed Central

[67] Nathwani AC, Gray JT, McIntosh J, Ng CY, Zhou J, Spence Y, et al. Safe and efficient transduction of the liver after peripheral vein infusion of self-complementary AAV vector results in stable therapeutic expression of human FIX in nonhuman primates. Blood. 2007;109:1414–21.10.1182/blood-2006-03-010181Search in Google Scholar PubMed PubMed Central

[68] Schuettrumpf J, Herzog RW, Schlachterman A, Kaufhold A, Stafford DW, Arruda VR. Factor IX variants improve gene therapy efficacy for hemophilia B. Blood. 2005;105:2316–23.10.1182/blood-2004-08-2990Search in Google Scholar PubMed

[69] Brunetti-Pierri N, Grove NC, Zuo Y, Edwards R, Palmer D, Cerullo V, et al. Bioengineered factor IX molecules with increased catalytic activity improve the therapeutic index of gene therapy vectors for hemophilia B. Hum Gene Ther. 2009;20:479–85.10.1089/hum.2008.084Search in Google Scholar PubMed PubMed Central

[70] Cantore A, Nair N, Della VP, Di Matteo M, Matrai J, Sanvito F, et al. Hyperfunctional coagulation factor IX improves the efficacy of gene therapy in hemophilic mice. Blood. 2012;120:4517–20.10.1182/blood-2012-05-432591Search in Google Scholar PubMed

[71] Finn JD, Nichols TC, Svoronos N, Merricks EP, Bellenger DA, Zhou S, et al. The efficacy and the risk of immunogenicity of FIX Padua (R338L) in hemophilia B dogs treated by AAV muscle gene therapy. Blood. 2012;120:4521–3.10.1182/blood-2012-06-440123Search in Google Scholar PubMed PubMed Central

[72] Suwanmanee T, Hu G, Gui T, Bartholomae CC, Kutschera I, von Kalle C, et al. Integration-deficient lentiviral vectors expressing codon-optimized R338L human FIX restore normal hemostasis in Hemophilia B mice. Mol Ther. 2014;22:567–74.10.1038/mt.2013.188Search in Google Scholar PubMed PubMed Central

[73] VandenDriessche T, Chuah MK. Moving forward toward a cure for hemophilia B. Mol Ther. 2015;23:809–11.10.1038/mt.2015.56Search in Google Scholar PubMed PubMed Central

[74] Monahan PE, Sun J, Gui T, Hu G, Hannah WB, Wichlan DG, et al. Employing a gain-of-function factor IX variant R338L to advance the efficacy and safety of hemophilia B human gene therapy: preclinical evaluation supporting an ongoing adeno-associated virus clinical trial. Hum Gene Ther. 2015;26:69–81.10.1089/hum.2014.106Search in Google Scholar PubMed PubMed Central

[75] Herzog RW. Hemophilia gene therapy: caught between a cure and an immune response. Mol Ther. 2015;23:1411–2.10.1038/mt.2015.135Search in Google Scholar PubMed PubMed Central

[76] Marsic D, Govindasamy L, Currlin S, Markusic DM, Tseng YS, Herzog RW, et al. Vector design tour de force: integrating combinatorial and rational approaches to derive novel adeno-associated virus variants. Mol Ther. 2014;22:1900–9.10.1038/mt.2014.139Search in Google Scholar PubMed PubMed Central

[77] Lisowski L, Dane AP, Chu K, Zhang Y, Cunningham SC, Wilson EM, et al. Selection and evaluation of clinically relevant AAV variants in a xenograft liver model. Nature. 2014;506:382–6.10.1038/nature12875Search in Google Scholar PubMed PubMed Central

[78] Schaffer DV. AAV shuffles to the liver: commentary on Lisowski et al. Mol Ther Methods Clin Dev. 2014;1:14006.10.1038/mtm.2014.6Search in Google Scholar PubMed PubMed Central

[79] Kotterman MA, Schaffer DV. Engineering adeno-associated viruses for clinical gene therapy. Nat Rev Genet. 2014;15:445–51.10.1038/nrg3742Search in Google Scholar PubMed PubMed Central

[80] Corti M, Elder M, Falk D, Lawson L, Smith B, Nayak S, et al. B-cell depletion is protective against anti-AAV capsid immune response: a human subject case study. Mol Ther Methods Clin Dev. 2014;1:14033.10.1038/mtm.2014.33Search in Google Scholar PubMed PubMed Central

[81] Sarkar D, Biswas M, Liao G, Seay HR, Perrin GQ, Markusic DM, et al. Ex vivo expanded autologous polyclonal regulatory T cells suppress inhibitor formation in hemophilia. Mol Ther Methods Clin Dev. 2014;1:14030.10.1038/mtm.2014.30Search in Google Scholar PubMed PubMed Central

[82] Sack BK, Herzog RW, Terhorst C, Markusic DM. Development of gene transfer for induction of antigen-specific tolerance. Mol Ther Methods Clin Dev. 2014;1:14013.10.1038/mtm.2014.13Search in Google Scholar PubMed PubMed Central

[83] Crudele JM, Finn JD, Siner JI, Martin NB, Niemeyer GP, Zhou S, et al. AAV liver expression of FIX-Padua prevents and eradicates FIX inhibitor without increasing thrombogenicity in hemophilia B dogs and mice. Blood. 2015;125:1553–61.10.1182/blood-2014-07-588194Search in Google Scholar PubMed PubMed Central

[84] Komor AC, Badran AH, Liu DR. CRISPR-based technologies for the manipulation of eukaryotic genomes. Cell. 2017;169(3):559.10.1016/j.cell.2017.04.005Search in Google Scholar PubMed

[85] Lander ES. The heroes of CRISPR. Cell. 2016;164(1–2):18–28.10.1016/j.cell.2015.12.041Search in Google Scholar PubMed

[86] Kevin D, John S. CRISPR: a powerful lab tool and a boon to gene therapy. Adv Gene Ther. 2019;39:4.Search in Google Scholar

[87] Mulvihill JJ, Capps B, Joly Y, Lysaght T, Zwart HAE, Chadwick R. International human genome organisation (HUGO) committee of ethics, law, and society (CELS). Ethical issues of CRISPR technology and gene editing through the lens of solidarity. Br Med Bull. 2017;122(1):17–29.10.1093/bmb/ldx002Search in Google Scholar PubMed

[88] Lavina ST, Nathan P, Rebecca P, Wei LC, Prashant M. Translating CRISPR-Cas therapeutics: approaches and challenges. CRISPR J. 2020;8:253–75.Search in Google Scholar

[89] Konstantinos S, Kiran M. Challenges and advances of CRISPR-Cas9 genome editing in therapeutics. Card Res. 2019;115:e12–4.10.1093/cvr/cvy300Search in Google Scholar PubMed

[90] Yang Y, Xu J, Ge S, Lai L. CRISPR/Cas: advances, limitations, and applications for precision cancer research. Front Med. 2021;8:649896.10.3389/fmed.2021.649896Search in Google Scholar PubMed PubMed Central

[91] Xu Y, Li Z. CRISPR-Cas systems: overview, innovations and applications in human disease research and gene therapy. Comput Struct Biotechnol J. 2020;18:2401–15.10.1016/j.csbj.2020.08.031Search in Google Scholar PubMed PubMed Central

Received: 2019-11-03
Revised: 2021-01-21
Accepted: 2021-01-24
Published Online: 2021-05-03

© 2021 Arome Solomon Odiba et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 19.4.2024 from https://www.degruyter.com/document/doi/10.1515/biol-2021-0033/html
Scroll to top button