Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 Sep 22.
Published in final edited form as: Curr Atheroscler Rep. 2016 May;18(5):22. doi: 10.1007/s11883-016-0579-0

Recent Developments in Gene Therapy for Homozygous Familial Hypercholesterolemia

Ezim Ajufo 1, Marina Cuchel 1
PMCID: PMC5609810  NIHMSID: NIHMS774326  PMID: 26980316

Abstract

Homozygous familial hypercholesterolemia (HoFH) is a life-threatening Mendelian disorder with a mean life expectancy of 33 years despite maximally tolerated standard lipid-lowering therapies. This disease is an ideal candidate for gene therapy and in the last few years, a number of exciting developments have brought this approach closer to the clinic than ever before. In this review, we discuss in detail the most advanced of these developments, a recombinant AAV vector carrying a low-density lipoprotein receptor (LDLR) transgene which has recently entered phase 1/2a testing. We also review ongoing development of approaches to enhance transgene expression, improve the efficiency of hepatocyte transduction and minimise the AAV capsid-specific adaptive immune response. We include a summary of key gene therapy approaches for HoFH in preclinical development, including RNA silencing of HMG-CoA reductase and induced pluripotent stem cell transplant therapy.

Introduction

Homozygous familial hypercholesterolemia (HoFH) is a life-threatening autosomal co-dominant disorder with an estimated prevalence of up to 1 in 160 000–300 000 [1,2]. It is estimated that more than 90% of HoFH patients harbor loss-of-function mutations in both alleles of the low-density lipoprotein receptor gene (LDLR). These mutations substantially or fully impair LDLR function, resulting in low-density lipoprotein cholesterol (LDL-C) levels frequently in excess of 500mg/dL and development of cardiovascular disease (CVD), principally coronary artery and aortic valve disease, during adolescence and death in the 20s, if left untreated [2]. Interestingly, significant phenotypic variability has recently been reported in a cohort of molecularly defined HoFH subjects, with LDL-C levels that extensively overlap with levels clinically associated with heterozygous status [1]. A major determinant of the severity of the HoFH phenotype is the residual LDLR activity associated with a given mutation. Based on in vitro assays in cultured fibroblasts, HoFH subjects can be classified as receptor-negative (<2% residual LDLR activity) or receptor-defective (2–25% residual LDLR activity) [3]. Receptor-defective FH subjects have significantly lower mean LDL-C levels and develop symptomatic coronary artery disease almost a decade later than their receptor-negative counterparts [3,4]. Furthermore, they respond better to conventional and novel therapies. In the TESLA B study with the proprotein convertase subtilisin/kexin type 9 (PCSK9) monoclonal antibody evolocumab, receptor-defective HoFH subjects experienced a ~25% LDL-C reduction, in contrast with the lack of effect seen in receptor-negative subjects [5]. Taken together, these findings suggest that even a small amount of LDLR activity is beneficial compared to no receptor activity at all.

The treatment of HoFH has historically been very challenging due to the extremely elevated baseline LDL-C levels and poor responsiveness to standard lipid-lowering therapies (LLT) [2]. Statin treatment is associated with a 10–25% reduction in LDL-C in this population [6] and the addition of ezetimibe results in a further reduction of 10–15% [7]. Despite a 60% reduction in mortality due to these treatments [8], mean life expectancy in HoFH remains low at 33 years [8]. Substantial reduction in LDL-C is obtained with LDL apheresis. However its effect is transitory and availability is limited [9]. The recently approved lipid-lowering agents mipomersen [10], lomitapide [11] and evolocumab, [5] produce additional LDL-C reduction on top of standard therapy in HoFH patients; however, issues of tolerability or lack of efficacy in receptor-negative patients will limit the potential of these agents to fully address the serious unmet medical need in this population. Thus, despite significant therapeutic progress over the last few years, HoFH remains a challenging condition to treat for many of these patients.

HoFH is an ideal candidate for liver-directed gene therapy. Approximately 75% of the body’s LDL receptors are expressed within the liver making it the most important organ in LDL metabolism [12]. This role is underscored by the dramatic correction and even resolution of the HoFH phenotype with orthotopic liver transplantation [1214]. The disadvantages and risks of transplantation and long-term immunosuppression limit the viability of this approach in this disease. However, the delivery of a functional LDLR transgene to the liver has been of great interest. In the early nineties, Grossman et al. conducted the first ever gene therapy clinical trial in HoFH subjects, using an ex vivo approach to transfer replication-deficient human LDLR-expressing retroviruses into the livers of five HoFH patients [15]. To achieve this, the patients underwent resection of the hepatic left-lateral segment, hepatocytes were harvested, transduced ex vivo using the retrovirus vector, and autologous transduced cells were then infused back into their donors via the portal vein. In contrast with its effects in a natural rabbit model of HoFH [16], disappointing results were observed in human subjects with few transduced hepatocytes and variable (6–25% LDL-C reduction in three subjects) but mostly clinically unremarkable transient metabolic effects. These findings precluded further clinical development of the ex vivo approach but confirmed the feasibility and safety of LDLR gene therapy in humans. A number of different vectors and transgene constructs have since been tested in the pre-clinical setting, however none have achieved stable gene expression with durable metabolic effects [17,18]. In the last few years, however, several novel approaches have yielded promising results, bringing us to an exciting time in this field. In this review we discuss the most advanced of these developments, a recombinant adeno-associated virus (AAV) vector which has recently entered clinical testing. We also review approaches in pre-clinical development including gain-of-function LDLR transgene variants, RNA silencing of HMG CoA reductase and induced pluripotent stem cell transplant therapy.

Approaches in clinical development: recombinant AAV vectors

Development of AAV vectors for gene therapy

AAV is a replication-defective, non-pathogenic parvovirus with a linear 4.7kB single-stranded DNA [19]. In its latent phase, AAV DNA is retained in the host cell nucleus in a circular episomal form with negligible rates of genomic integration allowing for stable gene expression in post-mitotic tissues [20]. The first AAV vector to be developed, AAV2, achieved efficient gene transfer and stable hepatocyte transduction with low immunogenicity in pre-clinical models [20,21]; in human subjects, however, AAV2 had poor liver tropism and was associated with a limiting adaptive immune response [20,22]. A number of different AAV vectors have since been developed from other naturally occurring AAV serotypes. These differ from AAV2 in tropism and immunogenicity allowing for broader use of the AAV vector class as gene therapy platforms [2325]. In addition, bioengineering approaches have generated novel variants that, in pre-clinical testing, evaded cellular and humoral immune responses more effectively, and transduced a wide range of cell types more efficiently than their naturally occurring parent serotypes [26]. In 2006, the first bioengineered AAV vector, AAV2.5 (a chimeric serotype derived from AAV1 and AAV2), entered clinical development in a Duchenne’s muscular dystrophy cohort with mixed results [27].

AAV vectors are now widely used in gene therapy agents and have performed well clinically [26,25]. Most notable among these successes is that of the lipoprotein lipase (LPL) expressing AAV1 vector, AAV1-LPLS447X, which became the first gene therapy agent to gain regulatory approval in the Western world in 2012 [28]. In LPL deficient subjects, AAV1-LPLS447X improved post-prandial triglyceride metabolism with clinically significant reductions in the incidence and severity of pancreatic episodes for up to 2 years after administration of the vector with follow-up ongoing [28,29]. The factor IX (FIX) expressing liver-targeted AAV8 vector, scAAV2/8-LP1-hFIXco, also performed well in a phase 1/2 hemophilia B trial [30] achieving serum FIX levels of 1–7% of normal from baseline levels of <1% with clinically significant phenotypic improvement. This has persisted for up to 4 years in all of the subjects treated [31]. The results from this trial hold particular significance for the prospects of a liver-targeted recombinant AAV8 agent for HoFH.

Efficacy of recombinant AAV8 in mouse models of HoFH

AAV8 was identified for liver-directed gene therapy in HoFH due to its strong liver tropism and relatively low (38%) seroprevalence in Western populations [24,32]. The first AAV8-based HoFH gene therapy study compared the efficacy of a human LDLR-expressing recombinant AAV8 vector to AAV2 and AAV7 in Ldr−/− mice on a high-fat diet. Compared to AAV2, AAV8 was associated with more efficient gene transfer (2.1 gc/cell vs. 52 gc/cell), hepatocyte transduction (4.2% vs. 81.2% of all hepatocytes), and more complete lipid correction (total cholesterol −600mg/dL vs. −1300mg/dL; p<0.001) [33]. AAV8 was also superior to AAV7 in these endpoints. Moreover, crucially, in AAV8-treated mice metabolic effects were maintained for up to 20 weeks and the development of atherosclerosis was almost completely attenuated.

The differences in the preclinical and clinical results observed with the HoFH ex vivo gene therapy approach highlight important limitations with the use of animal models. This has led to substantial efforts in developing an animal model for HoFH that more accurately reflects the human pathophysiology. By deleting the apolipoprotein B mRNA editing catalytic polypeptide-1 (Apobec1) and LDLR genes (Ldlr−/−/Apobec1−/−), Kassim et al. designed a murine model of HoFH that, on a low-fat chow diet, develops a lipid and atherosclerotic phenotype comparable to that seen in human HoFH subjects [34,35]. In Ldlr−/−/Apobec1−/− mice maintained on a chow diet, increasing doses of a recombinant AAV8 vector containing a mouse LDLR transgene driven from a liver-specific thyroxine-binding globulin (TBG) promoter (AAV8.TBG.mLDLR) produced a significant dose-dependent reduction in total cholesterol and non-HDL-C within seven days of treatment [35]. Interestingly, the lowest dose that produced a significant lipid improvement (total cholesterol: −141mg/dL; non-HDL −130mg/dL at day 35) was associated with transduction of only 5–10% of hepatocytes. Complete metabolic correction was achieved with doses at least an order of magnitude higher than this dose and maintained for up to six months. Gene transfer and hepatocyte transduction were also dose-dependent and, at the highest dose tested, achieved 60–70% hepatocyte transduction at 35 days post vector treatment.

In a further study, a more advanced humanized model was designed using Ldlr−/−/Apobec1−/− mice engineered to express the human apoB100 transgene (Ldlr−/−/Apobec1−/−/human ApoB transgenic) [36]. In these mice, an AAV8 vector expressing the human LDLR transgene (AAV8.TBG.hLDLR) produced comparable metabolic effects to that seen in the previous study [36]. Of interest, experiments performed in heterozygous FH (HeFH) mice in this study showed that, although AAV8.TBG.hLDLR also produced a dose-dependent reduction in serum lipid levels in HeFH mice, the minimum effective dose in this cohort was an order of magnitude higher than that recorded in the setting of the LDLR knockout mice.

Atherosclerotic regression is an important therapeutic endpoint for lipid-lowering therapies used in HoFH patients. Atherosclerotic progression was assessed in a cohort of Ldlr−/−/Apobec1−/− mice maintained on a high-fat for two months [36]. At this point, these mice received a tail-vein injection of AAV8.TBG.mLDLR and were followed for an additional two months while remaining on the high-fat diet. Within seven days of receiving the vector, a significant reduction in mean total cholesterol was observed in these mice (1555mg/dL to 266mg/dL), and that normalized by the end of the follow-up period (67mg/dL). This was associated with an 87% reduction in the aortic plaque burden. Histological analysis of these lesions showed smaller lesions, fewer inflammatory cells and favorable plaque remodeling [36].

Taken together, these studies demonstrate that a recombinant LDLR-expressing AAV8 vector produces efficient gene transfer, stable hepatocyte transduction, significant lipid correction and atherosclerotic regression in humanized mouse models of HoFH. However, it has been shown that AAV vectors transduce human hepatocytes less efficiently than murine hepatocytes [24,37]. Recombinant FIX-expressing AAV8 produced substantially greater serum FIX elevations in mouse models of hemophilia B [38] than it did in human hemophilia B subjects [30]. Crucially, however, the levels achieved in the human hemophiliacs produced clinically significant effects that have persisted for more than 3 years [31]. This is of relevance to the use of AAV8 in HoFH because patients carrying LDLR mutations associated with >2% residual LDLR activity (receptor-defective) have a significantly better prognosis than patients with LDLR mutations associated with <2% residual LDLR activity (receptor-negative) [4].

Safety profile of recombinant AAV8 vectors

Published studies in mice have not provided evidence of toxicity associated with the LDLR-expressing AAV8 vector [33,35,36]. However, caution must be taken in applying these findings to humans because substantial differences exist between the murine and human immune systems [39,40]. AAVs are naturally present in the environment [41], so it is expected that a percentage of the population have been exposed to them and will have developed anti-AAV8 neutralizing antibodies (nAbs)[32,41] and possibly also AAV memory T cells [42]. Furthermore, as demonstrated by previous gene therapy trials in LPL deficient [28] and hemophilia patients [30,31,43], administration of AAV vectors induces both cellular and humoral adaptive immune responses in human subjects.

Humoral immunity poses a challenge to gene therapy with AAV vectors for a number of reasons. First, approximately 30% of patients from Western populations have pre-existing anti-AAV8 neutralizing antibodies [32], and would be precluded from treatment with these vectors. Pre-existing antibodies against other serotypes such as AAV2 are even more prevalent [32]. In non-Western populations, such as the Chinese where anti-AAV8 seroprevalence is >80% [44], many more could be ineligible for AAV treatment [41]. Second, immunity that develops as a result of AAV8 administration precludes re-administration of the transgene with the same vector should transgene expression falls below therapeutic levels.

The cellular immune response poses the challenge of limiting transgene expression through the destruction of transduced cells [42]. It has been shown that a capsid and, less commonly, a transgene-specific T cell response is elicited by recombinant AAV vectors targeted to skeletal muscle [27,28,45] and the liver [30,43]. Focusing on the liver data, in the first liver-directed hemophilia B trial with AAV2, serum FIX levels of 10–12% normal were achieved at the highest dose tested in one patient, however within 2–4 weeks of vector administration, levels declined and returned to baseline by 10 weeks [43]. This was accompanied by a capsid-specific T cell response and a rise in transaminases which was asymptomatic and reversible. In this study, the T cell response was attributed to expansion of AAV2-specific memory T cells [42,43], however, alternative theories exist [40,4648]. The capsid-specific T cell response was subsequently shown to be dose-dependent, induced by other AAVs and amenable to immunosuppression in the AAV8 hemophilia B trial [30,31]. In this study, capsid-specific T cells were detected in the intermediate and higher dosing groups and accompanied by a reversible transaminitis in 4 out of the 6 subjects in the high-dose cohort. As in the previous study, FIX levels declined in association with the cellular response however, in the highest dosing group, mean steady-state FIX levels of up to 7% of normal were achieved and maintained [31]. Importantly, a course of prednisolone was administered to subjects that experienced a rise in transaminases. The decline in serum FIX levels was more pronounced when the initiation of prednisolone was delayed for >2 days [31]. Of note, a dramatic 50–70% reduction in serum FIX levels was also observed in association with modest transaminase elevations (peak ALT 36–202 IU/l), highlighting the detrimental effect of the cellular immune response on gene expression, and the importance of early initiation of immunosuppression [31]. Crucially, the capsid-specific T cell responses in these trials were subclinical and no transgene-specific T cell responses were detected.

Based on this preliminary safety and efficacy data, a first-in-human phase 1/2a trial has been announced with the recombinant AAV8 vector AAV8.TBG.hLDLR (NCT02651675) [see Table]. Its primary outcome measure is the number of participants that experience vector-related adverse effects by 52 weeks following administration. The total follow-up of subjects in this study is 5 years.

Table.

Gene therapy agents in recent development for Homozygous Familial Hypercholesterolemia

Agent Stage of development Efficacy data Safety data
Recombinant LDLR-expressing AAV8 (AAV8.TBG.hLDLR) Phase I/2a first-in-human study (NCT02651675). 12 patients. Primary safety endpoint at 1 year. Total follow-up of 5 years.
  • In LdLr−/−/Apobec1−/− mice maintained on a high-fat diet for two months prior to the administration of AAV8.TBG.mLDLR and then followed and analysed after two further months on this diet, TC was reduced by ~1500mg/dL and atherosclerotic burden by 87% [35].

  • Lipid lowering effects were maintained for up to 6 months post-vector administration in a cohort of LdLr−/−/Apobec1−/− chow-fed mice.

  • Comparable findings in LdLr−/−/Apobec1−/−/hApoB-Tg mice treated with AAV8.TBG.hLDLR in [36]

  • In LdLr−/−/Apobec1−/− mice that received AAV8.TBG.mLDLR, hepatic lipid accumulation measured at day 35 and transaminases measured at day 60 post-vector administration were within normal limits [35].

  • In Ldlr−/−/Apobec1−/−/hApoB-Tg given the highest deliverable dose of AAV8.TBG.hLDLR (5×1013 GC/kg), no transgene or capsid-specific T cell responses were detected. There was also no significant rise in transaminases and no hepatic inflammation or lipid accumulation in AAV8.TBG.hLDLR-treated mice [36].

Gain-of-function LDLR variants (delivered within recombinant AAV8 vectors):
  • AAV8.TBG.hLDLR-L318D (confers PCSK9 resistance)*

  • AAV8.TBG.hLDLR -K809R\C818A (confers IDOL resistance)*

  • AAV8.TBG.hLDLR -L318D\K809R\C818A (confers resistance to PCSK9 and IDOL)*

Pre-clinical (in vivo)
  • Humanized HoFH mice that received AAV8.TBG.hLDLR-L318D, AAV8.TBG.hLDLR-K809R\C818A or AAV8.TBG.hLDLR-L318D\K809R\C818A had lower TC levels and higher hepatic LDLR expression in the context of IDOL or PCKS9 over-expression compared to those that received the wild-type LDLR transgene [53].

N/A
HMGCR-specific miRNA/siRNA Pre-clinical (in vivo)
  • Co-transfecting a plasmid containing a reporter gene driven by LDLR genomic regulatory elements with HMGCR-specific siRNA or miRNA increased LDLR promoter activity 300-fold and 12-fold, after 7 and 10 days, respectively, in vitro.

  • In vivo, co-transfecting 4 HMGCR-specific miRNA plasmids with an LDLR-expressing plasmid with LDLR regulatory elements lowered plasma LDL-C by ~1.5mmol/L (p =0.0036) compared to mice treated with the LDLR plasmid alone [55].

N/A
Genetically corrected iPSC Pre-clinical (in vitro)
  • Transfection of LDLR-expressing plasmids into iPSCs derived from patients with HoFH generated hepatocyte-like cells with restored LDL internalisation under feedback control from extracellular cholesterol, statin and sterols [58,59].

N/A
*

All three agents are recombinant AAV8 vectors expressing gain-of-function hLDLR variants resistant to PCSK9/IDOL or both. PCSK9 and IDOL are negative regulators of LDLR.

Key; Apobec, apolipoprotein B mRNA editing catalytic polypeptide-1; gc, gene copies; hLDLR, human LDLR; HMGCR, HMG CoA reductase.; IDOL, inducible degrader of LDLR; iPSC, induced pluripotent stem cell; miRNA, microRNA; TC, total cholesterol; siRNA, small interfering RNA oligonucloetides

Recombinant AAV8 in HoFH: future directions

The development of strategies to improve transgene expression and minimise the immune response is ongoing. Bioengineering approaches are being used to develop AAV variants with attenuated immunogenicity and improved transgene transfer and hepatocyte transduction efficacies [40,49]. One AAV2 variant developed with this strategy was almost 100-fold more resistant to neutralization by nAbs specific for its parent AAV2 capsid in vitro with consistent findings in vivo [50]. AAV2.5 is thus far the only engineered AAV2 variant to undergo clinical testing. In the first-in-human study with this agent, six boys with Duchenne’s muscular dystrophy received an intramuscular injection of one of two doses of a recombinant AAV2.5 vector carrying a minidystophin transgene (AAV2.5-minidystrophin) under corticosteroid cover [27]. Vector genome transfer was confirmed in all of the subjects treated however gene expression was poor overall. Both humoral and adaptive immune responses to the AAV capsid were detected in this study although pre-existing seropositivity for AAV2.5 was shown to be ~25% lower than for AAV2. These clinical findings support the safety of engineered AAV serotypes in humans; however, the mixed efficacy and immunological results suggest further work is needed to develop variants with clear superiority to native AAV vectors.

Based on the successes of the LPL [28] and FIX [51,52] gain-of-function variants, a further approach in development are gain-of-function LDLR transgene variants. Using site-directed mutagenesis of hLDLR cDNA, Somanathan et al. developed three gain-of-function LDLR transgene variants that expressed LDLR proteins resistant to PCSK9 and/or inducible degrader of LDLR (IDOL)-mediated LDLR degradation. These variants (carried by AAV8 vectors) produced significantly greater cholesterol corrections in vivo and in vitro compared to the wild-type LDLR transgene in the setting of PCSK9 or IDOL overexpression (see Table) [53]. In vivo studies with physiological levels of PCSK9 and IDOL are needed to verify these findings. Moreover, it is possible that these variants could reduce responsiveness to the PCSK9 inhibitors so this should also be addressed in future studies.

Approaches in pre-clinical development

RNA silencing of 3-hydroxy-3-methyl-glutaryl-CoA reductase (HMGCR)

In a 2010 paper, Hibbitt et al. showed that combining statin therapy with a LDLR transgene construct containing LDLR genomic regulatory elements increased transgene expression up to 5-fold [54]. Building on this work, they piloted an approach that combined RNA silencing of HMGCR, which encodes the enzyme target of statins, with a plasmid that contained the LDLR transgene driven by LDLR genomic regulatory elements (pLDLR-LDLR) [55]. Using a luciferase reporter gene driven by LDLR regulatory elements, they showed that small interfering RNA oligonucleotides and microRNA (miRNA) specific for murine HMGCR transcripts increased LDLR promoter activity by 300-fold and 12-fold, respectively [55]. Furthermore, when they co-transfected four HMGCR-specific miRNA plasmids with pLDLR-LDLR into Ldr/− mice, they observed significantly lower LDL-C levels (−~1.5mmo/L, p=0.0036) in these mice as compared with controls that only received the pLDLR-LDLR plasmid. Thus, RNA-mediated knockdown of HMGCR provides an effective complementary strategy for enhancing LDLR transgene expression. How this approach would compare to combining LDLR gene transfer with regular statin therapy, which is the standard of care in HoFH, is of importance and remains to be established.

Induced pluripotent stem cell-based cell (iPSC) transplant therapy

iPSCs are somatic cells generated by overexpression of specific transcription factors. As orthotopic liver transplantation is curative in HoFH [2], autologous transplantation of genetically corrected cells derived from iPSCs has been of interest as a way of harnessing the benefits of transplantation while avoiding its undesirable effects. It has been shown that iPSCs derived from patients with familial hypercholesterolemia (FH) can be differentiated into hepatocyte-like cells that closely recapitulate the FH phenotype in vitro [56,57]. Building on this, Fattahi et al. showed that transfecting iPSCs derived from a HoFH patient with a plasmid containing the LDLR transgene generated hepatocyte-like cells that expressed LDLR mRNA and were able to internalize LDL under feedback regulation from extracellular LDL-C [58]. In a subsequent study, Ramakrishnan et al., showed that LDL internalization into such cells was also responsive to extracellular lovastatin and sterol, consistent with the restoration of physiological LDL metabolism [59].

In vivo data in HoFH preclinical models are not yet available, however, corrected iPSCs that were transplanted into the liver of fumarylacetoacetate hydrolase deficient (FAH) mice, another inherited metabolic disease of the liver, reversed the disease phenotype [60]. Of note, however, genetically corrected hepatocytes in FAH mice possess a unique survival advantage to native hepatocytes and expand robustly to repopulate the liver. Conversely, engraftment and repopulation of the liver is known to be less effective with iPSCs than with mature human hepatocytes [61]. Translation of iPSC technology to the clinic will also require significant refinement of the iPSC technology itself to ensure the genetic stability of derived cells, production of high-quality iPSCs, and efficient and reproducible differentiation to fully functional hepatocyte-like cells [61]. In addition, the risk of tumorigenesis associated with retained stem cells and inadvertently activated oncogenes during reprogramming must be mitigated in cells that are to be used therapeutically [62]. Thus, although appealing, the iPSC technology remains very much experimental and is unlikely to be a viable therapeutic for HoFH in the near future.

Conclusion

After more than two decades, gene therapy for HoFH has returned to clinical testing. The AAV vector class which has been widely successful in clinical trials in other Mendelian diseases is the backbone on which this agent has been developed. The recently announced first-in-human clinical trial is expected to recruit a dozen HoFH subjects, who will receive a recombinant LDLR-expressing AAV8 vector with follow-up of up to five years. It is worth noting that none of the AAV gene therapy agents that have been clinically successful in the past decade have produced complete reversal of the disease phenotype. As such, the recombinant LDLR-expressing AAV in testing is unlikely to prove curative but might produce clinically significant LDL-C reduction and improved response to lipid-lowering agents such as PCSK9 inhibitors. Nevertheless, a ‘cure’ remains the ultimate goal of any gene therapy approach. Achieving this will depend on achieving sufficient and sustained levels of stable LDLR expression. Bioengineering strategies have been employed to develop gain-of-function LDLR transgenes and AAV vectors adapted to enhance transgene transfer, transduction efficacy and minimize immunogenicity compared to their natural occurring AAV counterparts. Combining LDLR gene therapy with RNA silencing of HMGCR is a further strategy intended to enhance LDLR gene expression. These approaches are still largely in pre-clinical development however, with promising early results, it is hoped that they will advance us further towards a genetic cure for HoFH in the future.

References

  • 1.Sjouke B, Kusters DM, Kindt I, Besseling J, Defesche JC, Sijbrands EJG, et al. Homozygous autosomal dominant hypercholesterolaemia in the Netherlands: Prevalence, genotype-phenotype relationship, and clinical outcome. Eur Heart J. 2015;36:560–5. doi: 10.1093/eurheartj/ehu058. [DOI] [PubMed] [Google Scholar]
  • 2.Cuchel M, Bruckert E, Ginsberg HN, Raal FJ, Santos RD, Hegele RA, et al. Homozygous familial hypercholesterolaemia: new insights and guidance for clinicians to improve detection and clinical management. A position paper from the Consensus Panel on Familial Hypercholesterolaemia of the European Atherosclerosis Society. Eur Heart J England. 2014;35:2146–57. doi: 10.1093/eurheartj/ehu274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Goldstein JL, Brown MS. The LDL receptor defect in familial hypercholesterolemia. Implications for pathogenesis and therapy. Med Clin North Am UNITED STATES. 1982;66:335–62. doi: 10.1016/s0025-7125(16)31424-9. [DOI] [PubMed] [Google Scholar]
  • 4.Kolansky DM, Cuchel M, Clark BJ, Paridon S, McCrindle BW, Wiegers SE, et al. Longitudinal evaluation and assessment of cardiovascular disease in patients with homozygous familial hypercholesterolemia. Am J Cardiol. 2008;102:1438–43. doi: 10.1016/j.amjcard.2008.07.035. [DOI] [PubMed] [Google Scholar]
  • 5.Raal FJ, Honarpour N, Blom DJ, Hovingh GK, Xu F, Scott R, et al. Inhibition of PCSK9 with evolocumab in homozygous familial hypercholesterolaemia (TESLA Part B): a randomised, double-blind, placebo-controlled trial. Lancet. 2015;385:341–50. doi: 10.1016/S0140-6736(14)61374-X. 2014/10/06 ed. Carbohydrate and Lipid Metabolism Research Unit, Faculty of Health Sciences, University of Witwatersrand, Johannesburg, South AfricaAmgen Inc, One Amgen Center Drive, Thousand Oaks, CA, USA Division of Lipidology, Department of Medicine, University of. [DOI] [PubMed] [Google Scholar]
  • 6.Raal FJ, Pappu AS, Illingworth DR, Pilcher GJ, Marais AD, Firth JC, et al. Inhibition of cholesterol synthesis by atorvastatin in homozygous familial hypercholesterolaemia. Atherosclerosis. 2000;150:421–8. doi: 10.1016/s0021-9150(99)00435-9. [DOI] [PubMed] [Google Scholar]
  • 7.Gagné C, Gaudet D, Bruckert E. Efficacy and safety of ezetimibe coadministered with atorvastatin or simvastatin in patients with homozygous familial hypercholesterolemia. Circulation. 2002;105:2469–75. doi: 10.1161/01.cir.0000018744.58460.62. [DOI] [PubMed] [Google Scholar]
  • 8.Raal FJ, Pilcher GJ, Panz VR, Van Deventer HE, Brice BC, Blom DJ, et al. Reduction in mortality in subjects with homozygous familial hypercholesterolemia associated with advances in lipid-lowering therapy. Circulation. 2011;124:2202–7. doi: 10.1161/CIRCULATIONAHA.111.042523. [DOI] [PubMed] [Google Scholar]
  • 9.Thompson GR. LDL apheresis. Vol. 167. Atherosclerosis; Ireland: 2003. pp. 1–13. [DOI] [PubMed] [Google Scholar]
  • 10.Raal FJ, Santos RD, Blom DJ, Marais a D, Charng M-J, Cromwell WC, et al. Mipomersen, an apolipoprotein B synthesis inhibitor, for lowering of LDL cholesterol concentrations in patients with homozygous familial hypercholesterolaemia: a randomised, double-blind, placebo-controlled trial. Lancet. 2010;375:998–1006. doi: 10.1016/S0140-6736(10)60284-X. [DOI] [PubMed] [Google Scholar]
  • 11.Cuchel M, Meagher EA, du Toit Theron H, Blom DJ, Marais AD, Hegele RA, et al. Efficacy and safety of a microsomal triglyceride transfer protein inhibitor in patients with homozygous familial hypercholesterolaemia: a single-arm, open-label, phase 3 study. Lancet. 2013;381:40–6. doi: 10.1016/S0140-6736(12)61731-0. 2012/11/06 ed. Institute for Translational Medicine and Therapeutics, Cardiovascular Institute, University of Pennsylvania, Philadelphia, PA 19104, USA. mcuchel@mail.med.upenn.edu. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Bilheimer DW, Goldstein JL, Grundy SM, Starzl TE, Brown MS. Liver Transplantation to Provide Low-Density-Lipoprotein Receptors and Lower Plasma Cholesterol in a Child with Homozygous Familial Hypercholesterolemia. N Engl J Med. 1984;311:1658–64. doi: 10.1056/NEJM198412273112603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Ibrahim M, El-Hamamsy I, Barbir M, Yacoub MH. Translational lessons from a case of combined heart and liver transplantation for familial hypercholesterolemia 20 years post-operatively. J Cardiovasc Transl Res United States. 2012;5:351–8. doi: 10.1007/s12265-011-9311-1. [DOI] [PubMed] [Google Scholar]
  • 14.Kucukkartallar T, Yankol Y, Kanmaz T, Topaloglu S, Acarli K, Kalayoglu M. Liver transplantation as a treatment option for three siblings with homozygous familial hypercholesterolemia. Pediatr Transplant Denmark. 2011;15:281–4. doi: 10.1111/j.1399-3046.2010.01469.x. [DOI] [PubMed] [Google Scholar]
  • 15.Grossman M, Rader DJ, Muller DWM, Kolansky DM, Kozarsky K, Clark BJ, et al. A pilot study of ex vivo gene therapy for homozygous familial hypercholesterolaemia. Nat Med. 1995;1:1148–54. doi: 10.1038/nm1195-1148. [DOI] [PubMed] [Google Scholar]
  • 16.Chowdhury JR, Grossman M, Gupta S, Chowdhury NR, Baker JR, Wilson JM. Long-term improvement of hypercholesterolemia after ex vivo gene therapy in LDLR-deficient rabbits. Science. 1991;254:1802–5. doi: 10.1126/science.1722351. [DOI] [PubMed] [Google Scholar]
  • 17.Van Craeyveld E, Jacobs F, Gordts SC, De Geest B. Gene therapy for familial hypercholesterolemia. Curr Pharm Des. 2011;17:2575–91. doi: 10.2174/138161211797247550. [DOI] [PubMed] [Google Scholar]
  • 18.Al-Allaf FA, Coutelle C, Waddington SN, David AL, Harbottle R, Themis M. LDLR-Gene therapy for familial hypercholesterolaemia: problems, progress, and perspectives. Int Arch Med. 2010;3:36. doi: 10.1186/1755-7682-3-36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Buning H, Huber A, Zhang L, Meumann N, Hacker U. Engineering the AAV capsid to optimize vector-host-interactions. Curr Opin Pharmacol. 2015;24:94–104. doi: 10.1016/j.coph.2015.08.002. [DOI] [PubMed] [Google Scholar]
  • 20.Hastie E, Samulski RJ. AAV at 50: A golden anniversary of discovery, research, and gene therapy success, a personal perspective. Hum Gene Ther. 2015:1–24. doi: 10.1089/hum.2015.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Hermonat PL, Muzyczka N. Use of adeno-associated virus as a mammalian DNA cloning vector: transduction of neomycin resistance into mammalian tissue culture cells. Proc Natl Acad Sci USA UNITED STATES. 1984;81:6466–70. doi: 10.1073/pnas.81.20.6466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Mingozzi F, High KA. Therapeutic in vivo gene transfer for genetic disease using AAV: progress and challenges. Nat Rev Genet. 2011;12:341–55. doi: 10.1038/nrg2988. Nature Publishing Group, a division of Macmillan Publishers Limited All Rights Reserved. [DOI] [PubMed] [Google Scholar]
  • 23.Gao G-P, Alvira MR, Wang L, Calcedo R, Johnston J, Wilson JM. Novel adeno-associated viruses from rhesus monkeys as vectors for human gene therapy. Proc Natl Acad Sci USA United States. 2002;99:11854–9. doi: 10.1073/pnas.182412299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Lisowski L, Tay SS, Alexander IE. Adeno-associated virus serotypes for gene therapeutics. Curr Opin Pharmacol. 2015;24:59–67. doi: 10.1016/j.coph.2015.07.006. [DOI] [PubMed] [Google Scholar]
  • 25.Journal of Gene Medicine. Gene Therapy Clinical Trials Worldwide [Internet] Wiley; 2014. [cited 2015 Dec 12]. p. Charts & Tables, Vectors. Available from: http://www.abedia.com/wiley/ [Google Scholar]
  • 26•.Kotterman MA, Schaffer DV. Engineering adeno-associated viruses for clinical gene therapy. Nat Rev Genet. 2014;15:445–51. doi: 10.1038/nrg3742. This review provides a good overview of key bioengineering approaches being employed to improve AAV vectors. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Bowles DE, McPhee SWJ, Li C, Gray SJ, Samulski JJ, Camp AS, et al. Phase 1 gene therapy for Duchenne muscular dystrophy using a translational optimized AAV vector. Mol Ther United States. 2012;20:443–55. doi: 10.1038/mt.2011.237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Gaudet D, Méthot J, Déry S, Brisson D, Essiembre C, Tremblay G, et al. Efficacy and long-term safety of alipogene tiparvovec (AAV1-LPLS447X) gene therapy for lipoprotein lipase deficiency: an open-label trial. Gene Ther. 2013;20:361–9. doi: 10.1038/gt.2012.43. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Gaudet D, Méthot J, Kastelein J. Gene therapy for lipoprotein lipase deficiency. Curr Opin Lipidol. 2012;23:310–20. doi: 10.1097/MOL.0b013e3283555a7e. [DOI] [PubMed] [Google Scholar]
  • 30.Nathwani AC, Tuddenham EGD, Rangarajan S, Rosales C, McIntosh J, Linch DC, et al. Adenovirus-associated virus vector-mediated gene transfer in hemophilia B. N Engl J Med. 2011;365:2357–65. doi: 10.1056/NEJMoa1108046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31••.Nathwani AC, Reiss UM, Tuddenham EGD, 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. doi: 10.1056/NEJMoa1407309. This paper presents three-year follow-up data from the first liver-directed gene therapy clinical trial with AAV8 in hemophilia B. All of the subjects in this trial achieved steady-state serum FIX levels between 1–7% of normal from baseline levels of <1%. Capsid-specific T cells were associated with a reversible rise in transaminase in 4/6 patients that receieved the highest dose of the vector which was associated with a decline in serum FIX levels. This trial is of particular significance for the HoFH clinical trial underway due to key parallels between the two diseases being treated and use of the same vector and target. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Boutin S, Monteilhet V, Veron P, Leborgne C, Benveniste O, Montus MF, et al. Prevalence of serum IgG and neutralizing factors against adeno-associated virus (AAV) types 1, 2, 5, 6, 8, and 9 in the healthy population: implications for gene therapy using AAV vectors. Hum Gene Ther United States. 2010;21:704–12. doi: 10.1089/hum.2009.182. [DOI] [PubMed] [Google Scholar]
  • 33.Lebherz C, Gao G, Louboutin J-P, Millar J, Rader D, Wilson JM. Gene therapy with novel adeno-associated virus vectors substantially diminishes atherosclerosis in a murine model of familial hypercholesterolemia. J Gene Med. 2004;6:663–72. doi: 10.1002/jgm.554. [DOI] [PubMed] [Google Scholar]
  • 34.Powell-Braxton L, Véniant M, Latvala RD, Hirano KI, Won WB, Ross J, et al. A mouse model of human familial hypercholesterolemia: markedly elevated low density lipoprotein cholesterol levels and severe atherosclerosis on a low-fat chow diet. Nat Med. 1998;4:934–8. doi: 10.1038/nm0898-934. [DOI] [PubMed] [Google Scholar]
  • 35.Kassim SH, Li H, Vandenberghe LH, Hinderer C, Bell P, Marchadier D, et al. Gene therapy in a humanized mouse model of familial hypercholesterolemia leads to marked regression of atherosclerosis. PLoS One. 2010;5:e13424. doi: 10.1371/journal.pone.0013424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36••.Kassim SH, Li H, Bell P, Somanathan S, Lagor W, Jacobs F, et al. Adeno-associated virus serotype 8 gene therapy leads to significant lowering of plasma cholesterol levels in humanized mouse models of homozygous and heterozygous familial hypercholesterolemia. Hum Gene Ther. 2013;24:19–26. doi: 10.1089/hum.2012.108. In this follow-up to the 2010 Kassim et al. manuscript referenced above, a more advanced humanised mouse model was used to evaluate the efficacy and safety of a human LDLR transgene in an AAV8 vector. This agent is now in clinicial testing. Key safety analyses reported by this paper include transgene and capsid-specific T cell responses, liver histopathlogy, liver lipid accumulation and serum transaminases. These were found to be within normal limits. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.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 England. 2014;506:382–6. doi: 10.1038/nature12875. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Nathwani AC, Gray JT, Ng CYC, 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;107:2653–61. doi: 10.1182/blood-2005-10-4035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Herzog RW. Immune Responses to AAV Capsid: Are Mice Not Humans After All [quest] Mol Ther. 2007;15:649–50. doi: 10.1038/sj.mt.6300123. [DOI] [PubMed] [Google Scholar]
  • 40.Martino AT, Basner-Tschakarjan E, Markusic DM, Finn JD, Hinderer C, Zhou S, et al. Engineered AAV vector minimizes in vivo targeting of transduced hepatocytes by capsid-specific CD8+ T cells. Blood. 2013;121:2224–33. doi: 10.1182/blood-2012-10-460733. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Calcedo R, Vandenberghe LH, Gao G, Lin J, Wilson JM. Worldwide epidemiology of neutralizing antibodies to adeno-associated viruses. J Infect Dis. 2009;199:381–90. doi: 10.1086/595830. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Mingozzi F, Maus MV, Hui DJ, Sabatino DE, Murphy SL, Rasko JEJ, et al. CD8(+) T-cell responses to adeno-associated virus capsid in humans. Nat Med. 2007;13:419–22. doi: 10.1038/nm1549. [DOI] [PubMed] [Google Scholar]
  • 43.Manno CS, Pierce GF, Arruda VR, Glader B, Ragni M, Rasko JJ, et al. Successful transduction of liver in hemophilia by AAV-Factor IX and limitations imposed by the host immune response. Nat Med United States. 2006;12:342–7. doi: 10.1038/nm1358. [DOI] [PubMed] [Google Scholar]
  • 44.Liu Q, Huang W, Zhang H, Wang Y, Zhao J, Song A, et al. Neutralizing antibodies against AAV2, AAV5 and AAV8 in healthy and HIV-1-infected subjects in China: implications for gene therapy using AAV vectors. Gene Ther. 2014;21:732–8. doi: 10.1038/gt.2014.47. [DOI] [PubMed] [Google Scholar]
  • 45.Mendell JR, Campbell K, Rodino-Klapac L, Sahenk Z, Shilling C, Lewis S, et al. Dystrophin immunity in Duchenne’s muscular dystrophy. N Engl J Med. 2010;363:1429–37. doi: 10.1056/NEJMoa1000228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Hauck B, Murphy SL, Smith PH, Qu G, Liu X, Zelenaia O, et al. Undetectable transcription of cap in a clinical AAV vector: implications for preformed capsid in immune responses. Mol Ther United States. 2009;17:144–52. doi: 10.1038/mt.2008.227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Li C, Goudy K, Hirsch M, Asokan A, Fan Y, Alexander J, et al. Cellular immune response to cryptic epitopes during therapeutic gene transfer. Proc Natl Acad Sci USA. 2009;106:10770–4. doi: 10.1073/pnas.0902269106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Vandenberghe LH, Wang L, Somanathan S, Zhi Y, Figueredo J, Calcedo R, et al. Heparin binding directs activation of T cells against adeno-associated virus serotype 2 capsid. Nat Med. 2006;12:967–71. doi: 10.1038/nm1445. [DOI] [PubMed] [Google Scholar]
  • 49.Hui DJ, Basner-Tschakarjan E, Chen Y, Davidson RJ, Buchlis G, Yazicioglu M, et al. Modulation of CD8+ T cell responses to AAV vectors with IgG-derived MHC class II epitopes. Mol Ther. 2013;21:1727–37. doi: 10.1038/mt.2013.166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Maheshri N, Koerber JT, Kaspar BK, Schaffer DV. Directed evolution of adeno-associated virus yields enhanced gene delivery vectors. Nat Biotechnol. 2006;24:198–204. doi: 10.1038/nbt1182. [DOI] [PubMed] [Google Scholar]
  • 51.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. doi: 10.1182/blood-2012-06-440123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Baxalta Reports Continued Progress on Phase 1/2 Clinical Trial of BAX335, Investigational Gene Therapy Treatment for Hemophilia B [Internet] [cited 2015 Dec 7]. Available from: http://www.businesswire.com/news/home/20150624005677/en/Baxalta-Reports-Continued-Progress-Phase-12-Clinical.
  • 53•.Somanathan S, Jacobs F, Wang Q, Hanlon AL, Wilson JM, Rader DJ. AAV vectors expressing LDLR gain-of-function variants demonstrate increased efficacy in mouse models of familial hypercholesterolemia. Circ Res. 2014;115:591–9. doi: 10.1161/CIRCRESAHA.115.304008. This interesting paper presents in vitro and in vivo efficacy data from three engineered gain-of-function LDLR transgenes which improved LDLR expression and cholesterol reduction compared to the wild-type LDLR transgene. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Hibbitt OC, McNeil E, Lufino MM, Seymour L, Channon K, Wade-Martins R. Long-term physiologically regulated expression of the low-density lipoprotein receptor in vivo using genomic DNA mini-gene constructs. Mol Ther Nature Publishing Group. 2010;18:317–26. doi: 10.1038/mt.2009.249. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55•.Hibbitt O, Agkatsev S, Owen C, Cioroch M, Seymour L, Channon K, et al. RNAi-mediated knockdown of HMG CoA reductase enhances gene expression from physiologically regulated low-density lipoprotein receptor therapeutic vectors in vivo. Gene Ther. 2012;19:463–7. doi: 10.1038/gt.2011.103. This study showed that combining RNA-mediated knockdown of HMGCR with LDLR gene transfer increased LDLR gene expression up to 300-fold in vitro with significant metabolic effects in LDLR knockout mice. [DOI] [PubMed] [Google Scholar]
  • 56.Rashid ST, Corbineau S, Hannan N, Marciniak SJ, Miranda E, Alexander G, et al. Modeling inherited metabolic disorders of the liver using human induced pluripotent stem cells. J Clin Invest. 2010;120:3127–36. doi: 10.1172/JCI43122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Cayo MA, Cai J, DeLaForest A, Noto FK, Nagaoka M, Clark BS, et al. JD induced pluripotent stem cell-derived hepatocytes faithfully recapitulate the pathophysiology of familial hypercholesterolemia. Hepatology. 2012;56:2163–71. doi: 10.1002/hep.25871. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Fattahi F, Asgari S, Pournasr B, Seifinejad A, Totonchi M, Taei A, et al. Disease-corrected hepatocyte-like cells from familial hypercholesterolemia-induced pluripotent stem cells. Mol Biotechnol. 2013;54:863–73. doi: 10.1007/s12033-012-9635-3. [DOI] [PubMed] [Google Scholar]
  • 59•.Ramakrishnan VM, Yang J-Y, Tien KT, McKinley TR, Bocard BR, Maijub JG, et al. Restoration of Physiologically Responsive Low-Density Lipoprotein Receptor-Mediated Endocytosis in Genetically Deficient Induced Pluripotent Stem Cells. Sci Rep England. 2015;5:13231. doi: 10.1038/srep13231. This in vitro study, building on the Fattahi et al. study referenced above, showed that autologous iPSCs from HoFH patients could be genetically corrected using an LDLR expressing plasmid and differentiated into hepatocyte-like cells with restored LDL internalisation responsive to extracellular sterol and lovostatin. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Espejel S, Roll GR, Mclaughlin KJ, Lee AY, Zhang JY, Laird DJ, et al. Induced pluripotent stem cell-derived hepatocytes have the functional and proliferative capabilities needed for liver regeneration in mice. J Clin Invest. 2010;120:3120–6. doi: 10.1172/JCI43267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Cantz T, Sharma AD, Ott M. Concise review: cell therapies for hereditary metabolic liver diseases-concepts, clinical results, and future developments. Stem Cells. 2015;33:1055–62. doi: 10.1002/stem.1920. [DOI] [PubMed] [Google Scholar]
  • 62.Ordonez MP, Goldstein LSB. Using human-induced pluripotent stem cells to model monogenic metabolic disorders of the liver. Semin Liver Dis. 2012;32:298–306. doi: 10.1055/s-0032-1329898. [DOI] [PubMed] [Google Scholar]

RESOURCES