Skip to main content

Nanoparticles-mediated CRISPR-Cas9 gene therapy in inherited retinal diseases: applications, challenges, and emerging opportunities

Abstract

Inherited Retinal Diseases (IRDs) are considered one of the leading causes of blindness worldwide. However, the majority of them still lack a safe and effective treatment due to their complexity and genetic heterogeneity. Recently, gene therapy is gaining importance as an efficient strategy to address IRDs which were previously considered incurable. The development of the clustered regularly-interspaced short palindromic repeats (CRISPR)-CRISPR-associated protein 9 (Cas9) system has strongly empowered the field of gene therapy. However, successful gene modifications rely on the efficient delivery of CRISPR-Cas9 components into the complex three-dimensional (3D) architecture of the human retinal tissue. Intriguing findings in the field of nanoparticles (NPs) meet all the criteria required for CRISPR-Cas9 delivery and have made a great contribution toward its therapeutic applications. In addition, exploiting induced pluripotent stem cell (iPSC) technology and in vitro 3D retinal organoids paved the way for prospective clinical trials of the CRISPR-Cas9 system in treating IRDs. This review highlights important advances in NP-based gene therapy, the CRISPR-Cas9 system, and iPSC-derived retinal organoids with a focus on IRDs. Collectively, these studies establish a multidisciplinary approach by integrating nanomedicine and stem cell technologies and demonstrate the utility of retina organoids in developing effective therapies for IRDs.

Introduction

Inherited retinal diseases (IRDs) are a diverse group of rare genetic disorders associated with more than 280 different genes [1]. IRDs manifest varying degrees of clinical severity and variable inheritance patterns [2], leading to blindness in infancy/early childhood [3, 4] or a gradual and progressive vision loss during adulthood [5,6,7,8]. The development of comprehensive and effective treatment proves to be a challenge for scientists, particularly due to the diverse number of genes involved in IRDs. In 2017, Luxturna® (voretigene neparvovec), a gene therapy drug developed by Spark Therapeutics Inc., was approved by the Food and Drug Administration (FDA). Luxturna® uses adeno-associated virus serotype 2 (AAV2) as a delivery vehicle to carry the wild-type Retinal Pigment Epithelium 65 (RPE65) gene into the retinal cells with RPE65 mutation for treating patients with Leber congenital amaurosis (LCA), a rare form of inherited vision loss [9]. AAV-derived vectors have several advantages, including high biosafety, low immunogenicity, stable expression, and high infectivity in several cell types. Although AAV-derived vectors are the safest and most effective viral vectors for gene replacement therapy in the retina, they cannot accommodate genes larger than 4.7 kb, and the generation of neutralizing antibodies against AAV may attenuate the efficacy of AAV-mediated gene therapy [10, 11]. Moreover, the treatment with Luxturna® requires vitrectomy of the retina, followed by the retinal detachment using the air tamponade [12, 13]. This multi-step surgical procedure is a huge burden to patients’ fragile retinas. Another concern is the repeated treatments with Luxturna®, as a single Luxturna® dosage only lasts for five years [14] and patients are required to undergo these invasive procedures routinely. Therefore, innovative topical delivery and highly permeable gene therapy are urgently needed for the therapy of IRDs [15].

The development of the clustered regularly-interspaced short palindromic repeats (CRISPR)-CRISPR-associated protein 9 (Cas9) gene editing technique revolutionized molecular biology and showed great potential for improved gene therapy [16, 17]. The first clinical trial of CRISPR-Cas9 gene editing treatment for IRDs, delivered by the AAV, was launched for the most common cause of inherited childhood blindness, LCA type 10 (LCA10) [18]. Homology-independent targeted integration (HITI) is an advanced CRISPR-Cas9 technique that enables targeted gene insertion in non-dividing cells and presents a new approach to treating genetic disorders [19]. However, one of the greatest challenges in the therapeutic application of CRISPR-Cas9 for retinal diseases is the delivery efficiency of the CRISPR-Cas9 components into the retinal pigment epithelium (RPE) and the neurosensory retinal environment in the posterior pole of human eyeballs [19]. The gene delivery using viral vectors is efficient but associated with several disadvantages, including random insertion, mutagenesis, and biohazard concerns [20]. Recent developments in nanomedicine overcome this difficulty by introducing a nontoxic delivery of CRISPR-Cas9 components that can significantly alleviate safety concerns raised by viral vectors [20]. Today, researchers aim to engineer nanoparticles (NPs) with specialized properties to go beyond viral limitations and create new opportunities towards the application of CRISPR-Cas9 in treating IRDs. However, the translation of such technology to the clinic is hampered by several obstacles. The complexity of the retinal structure poses a significant challenge to the standard measurement of visual performance after treatment and causes unreliable diagnosis [21, 22]. Moreover, in vitro, in vivo, and species variations among disease models limit the ability to fully recapitulate the structure and functions of the retina [23,24,25]. With induced pluripotent stem cells (iPSC) technology's burgeoning field, researchers are now generating three-dimensional (3D) retinal organoids (ROs) from human iPSCs [26,27,28,29,30,31], aiming at recapitulating disease phenotypes and developing a clinically relevant resource to investigate the molecular mechanisms of diseases. This review highlights current nanomedicine applications in gene therapy, focusing on an NP-mediated delivery of the CRISPR-Cas9 system with respect to IRDs. Meanwhile, we elaborate on the application of iPSC-derived retinal organoids as a drug-screening platform, which, in combination, provide new opportunities for drug development and personalized medicine for IRDs.

The complex three-dimensional (3D) architecture of the retina

During embryonic development, the retina is derived from the prosencephalon, the anterior portion of the brain [21]. The retina’s unique architecture can be classified into two distinct parts: the posterior RPE layer with the most apparent light absorption function and the anterior multilayered neuroretina (Fig. 1). This multi-layered structure of the retina comprises two synaptic layers (the outer and inner plexiform layers) and three specialized neuronal cell layers (Fig. 1). The first neuronal layer comprises rod and cone photoreceptor (PR) cells that convert the absorbed light with different intensities into electrical signals via phototransduction. PR cells then form synapses with the second layers of neuronal cells, horizontal and bipolar cells present in the inner nuclear layer (INL), through the outer plexiform layer (OPL). This eventually leads to the transmission of the signal from PRs to the third class of neuronal layer cells, the retinal ganglion cells (RGCs), through the inner plexiform layer (IPL) (Fig. 1). Finally, the axons of RGCs converge to form the optic nerve, which in turn leads to the transmission of the visual impulses from the eye to the brain [21, 22]. Any impairment in this signaling cascade can result in visual disorder. In addition, the photoreceptor dysfunction or loss can be associated with age, diabetes, and genetics [32,33,34]. The latter causes a specific category of disease described as IRDs, which is the focus of this review.

Fig. 1
figure 1

The complex architecture of the retina. The general layout of the retinal layers is shown on the left and cell types on the right. Photoreceptor cells, bipolar cells, and retinal ganglion cells constitute the signal transmission pathway that conveys vision signal to the brain. Horizontal and amacrine cells are interneurons modulating visual signal transmission. Müller glial cells perform the neuronal support functions similar to those of astrocytes in the brain

Inherited retinal diseases and the treatment obstacles

IRDs associated with several clinically and genetically heterogeneous defects belong to a group of progressive retinal degeneration diseases that may lead to vision loss. They exhibit a wide variation in genetic mutations, age of onset, and disease severity [35,36,37], with an estimated incidence of 1 in 2000 to 3000 individuals [38,39,40,41,42]. Due to recent advances toward pathogenesis and characterization of genes responsible for IRDs, with more than 270 genes identified so far, a significant progress has been made in the field of incurable IRDs [43,44,45,46,47,48]. It has led to the development of treatments aimed at restoring vision or delaying the vision loss progression; however thus far, treatment options are still limited. IRD gene variants can be transmitted in an autosomal dominant (e.g. LCA and retinitis pigmentosa (RP)) [49, 50], recessive (e.g. cone-rod dystrophy) [51], and X-linked manner (e.g. X-linked juvenile retinoschisis (XLRS)) [52,53,54,55]. As shown in Table 1, IRDs can be clinically classified into six categories based on the affected retinal regions/cell types and four categories depending on the genetic inheritance mode; other classifications are based on the monogenic and multifactorial nature of IRD and the disease progression [56]. The genetic heterogeneity observed in these diseases manifests in patients with very similar clinical phenotypes but different genetic diagnoses, demanding gene-specific therapies or gene editing treatments to develop treatment strategies [8, 57]. Nevertheless, there is no available cure for IRDs currently, and ophthalmology has been at the forefront of utilizing gene therapy to treat these disorders.

Table 1 Classification of IRDs

Overview of gene therapy techniques for IRDs

Gene therapy application for ophthalmic diseases is a blooming field of research and currently overcoming the barriers for translation to the clinic, which is extensively described in multiple reviews [58,59,60,61,62,63]. AAV vector-based gene therapy has obtained the marketing approval for treating RPE65-associated LCA [18, 64,65,66], Leber hereditary optic neuropathy (LHON) [67, 68], and choroideremia (CHM) [69,70,71]. Treatments for LHON and CHM have entered phase III clinical trials and raised hopes that these approaches might be of practical use to delay or halt disease progression in patients with these IRDs [72,73,74]. Currently, more than 30 gene therapy trials for IRDs are being conducted in the United States, and some have entered the phase III clinical trials. Positive outcomes in these clinical trials were mainly due to the advantages offered by the retina as the target organ and AAV as the carrier [75,76,77,78]. The eye is an ideal target for gene therapy approaches due to several features of its anatomy and microenvironment. First, the eye is an immune-privileged microenvironment in which the tight junctions of the blood-retina barrier (BRB) limit systemic dissemination of intraocularly injected antigens, making it suitable for the introduction of viral vectors. Besides, BRB plays a crucial role in regulating the retina’s microenvironment as its enclosed orbit environment allows a small number of vectors encoding the gene editing components to reach the optimal therapeutic concentration with the desired effect. Secondly, the accessibility of the eye enables the intravitreal and subretinal delivery of vectors and allows the non-invasive monitoring of the patient's response to therapy [79,80,81]. Lastly, an indefinite expression of delivered genes in non-dividing retinal cells can potentially be achieved only after a single injection, which is a tremendous advantage to non-integrating vectors [82].

As previously mentioned, gene therapy can be conducted using viral vectors to introduce the transgene into the target cells. The most broadly used vectors for ocular gene delivery are AAVs [83]. AAVs are favored for gene therapy as they are non-pathogenic helper-dependent viruses with little immunogenicity and high diffusion capacity. They are also capable of efficient transduction of quiescent retinal cells and allow the long-term expression of the transgene in the host cells. However, their restricted packaging size of approximately 4.7 kb limits their application [84, 85]. Despite many reports on the higher efficiency of viral versus non-viral vectors [86,87,88], recent findings on NP-based delivery for retinal gene therapy have proven to be more efficient than their viral counterparts [89, 90]. Current DNA-based retinal therapy strategies involve gene augmentation/replacement and CRISPR-Cas9-mediated gene editing, as discussed below.

Gene augmentation/replacement therapy

Current gene therapy strategies targeting IRDs rely on gene augmentation/supplementation in which the wild copy of the gene is delivered to affected cells resulting in a functional protein product. In most retinal gene therapies, the transgene is introduced using the AAV vector and delivered via a subretinal injection after the vitrectomy [56, 91, 92]. In 2017, the FDA approved the first gene therapy using Luxturna® for treating a monogenic disease—biallelic RPE65-related retinal dystrophy. Luxturna® treatment delivers a functional copy of the RPE65 gene to the retina of patients suffering from severe night blindness (nyctalopia). It can significantly improve patient’s light sensitivity and the navigation in dim light. Among all IRDs, slowly progressing IRDs are more favorable for gene therapy as it can only target viable cells. In patients with advanced disease when most of the photoreceptors are degenerated, photoreceptor replacement therapy appears to be more attractive approach to restore vision [61, 73, 74, 93]. Furthermore, disorders with monogenic X-linked or autosomal recessive mutations are usually the potential targets for gene therapy, as such therapy is ineffective for disorders with gain-of-function mutations in which the mutant protein products can interfere with the functions of wild type proteins [94, 95]. Therefore, advanced technologies are required to overcome the limitations associated with the gene augmentation and broaden the scope of retinal gene therapy.

CRISPR-Cas9 genome engineering

The efficient genome editing relies on the formation of DNA double-strand breaks (DSBs) at the desired genomic loci by an engineered nuclease and the subsequent repair of DSBs, exploiting two main mechanisms: the homology-directed repair (HDR), a precise gene editing strategy using DNA template with homology arms [96], and the non-homologous end-joining (NHEJ), an error-prone process which may lead to insertions or deletions (indels) at the break site. Nuclease-based platforms including zinc finger nucleases (ZFNs) and transcription activator-like effector nucleases (TALENs) have been extensively used for generating precise DSBs in cells over the past few decades. However, the technical shortcomings and laborious protein engineering impose certain limitations on their in vivo applicability.

In 2012, the advent of CRISPR-Cas9 technology with multiplexed gene editing ability and high efficiency opened up new horizons to a novel type treatment, genome surgery [97, 98]. The programmable RNA-guided Cas9 endonuclease targets and cleaves DNA of any size to insert or remove DNA fragments in a sequence-dependent manner without affecting the gene's regulatory sequences. Unlike gene augmentation, CRISPR-Cas9-mediated gene editing avoids the risk of toxicity caused by the transgene overexpression and overcomes the AAV packaging limits by using smaller Cas9 variants (e.g., Staphylococcus aureus Cas9 (SaCas9) [99] and Streptococcus pyogenes Cas9 [100]), making in vivo genome surgery an achievable goal. Furthermore, this strategy enables eliminating genes with dominant gain-of-function mutations, not amenable to gene augmentation therapy. Thus, CRISPR-Cas9 provides ophthalmologists with a novel therapeutic tool for treating IRDs unreachable by established treatments [101,102,103,104]. The CRISPR-Cas9 system can function via either the NHEJ or HDR pathways. The latter is preferable but inefficient and not readily accessible to post-mitotic cells. In contrast, Cas-induced NHEJ strategy is efficient and active in both dividing and non-dividing cells. Therefore, increasing CRISPR-Cas9 efficiency is a critical task to make CRISPR a broadly applicable gene therapy approach for treating IRDs.

Recently, several advanced strategies, such as obligate ligation-gated recombination (ObLiGaRe) [105, 106], homology-independent targeted integration (HITI) [107,108,109], precise integration into target chromosome (PITCH) [105, 110], base editing [111], prime editing [112], and CRISPR activation/interference [113], have been developed aiming at genome editing. However, the in vivo application of these strategies should be further validated before their clinical translation. HITI strategy was demonstrated to be able to bypass the low efficiency of HDR to a significant extent. Despite the development of CRISPR-Cas9, the targeted integration of transgenes in vivo remains mostly challenging, especially in most adult tissues that are non-dividing. HITI, designed by Suzuki et al., is known as the gene knock-in strategy in which the foreign DNA is directly ligated to DSBs, which can be achieved in both dividing and non-dividing cells via the NHEJ pathway [114,115,116]. Generally, following the HITI strategy, the CRISPR-Cas9-mediated gene knockin can be carried out as follows. First, Cas9 sgRNA is used to specifically recognize and process the targeted sequences in both genomic and donor DNA. Second, after generating the formation of three double-stranded breaks, DNA repair initiates the site-specific integration of transgenes via the NHEJ pathway. For instance, Suzuki’s study demonstrated the promising utility of HITI strategy that can promote efficient CRISPR-Cas9-mediated gene knock-in in the brain and eye in vivo [108]. The genomic safe harbors are the sites which can be safely manipulated, allowing the integrated transgene to function properly without affecting the genome of host cells [117]. Further studies showed that HITI strategy can be used to safely and efficiently integrate transgenes into zebrafish and mammalian cells [114,115,116].

Comparing to HDR and NHEJ, microhomology-mediated end-joining (MMEJ) that requires relatively much smaller regions for DSB repair is an alternative form of end-joining. Sakuma and Yamamoto et al. created an alternative method for gene knock-in, termed PITCH. Assisted by MMEJ, the PITCH system is able to promote precise gene knock-in with the requirements of much smaller homologous regions. Therefore, complicated cloning of homology arms was not needed, facilitating the entire process of PITCH vector construction [110]. Base editing, developed by Komor and Liu et al., is able to achieve programmable genome editing with direct and irreversible conversion of a specific DNA base into anthor through a mechanism that does not require the cleavage of dsDNA backbone or the donor template [111]. Base editors that can convert the target DNA base within an ~ 5-nucleotide window is capable of gene correction in several point mutations in human genetic diseases [111]. Prime editing was designed to achieve efficient gene correction with the minimal formation of byproducts. This technology uses a prime editing guide RNA (pegRNA) and a fusion protein consisting of a catalytically impaired Cas9 endonuclease and an engineered reverse transcriptase. Through the coordination of pegRNA and the fusion protein, prime editing can recognize the target sites and undergo the desired gene correction. Comparing to HDR, prime editing carries the minimal production of byproducts but shows similar or higher efficiency. Without the requirements of DSBs and the donor DNA templates, prime editing can mediate targeted insertion, deletion, and transversions which may largely expand the capabilities of gene editing [112]. Nuclease-deactivated Cas9 (dCas9) is designed for RNA-guided genomic transcription regulation. This technology can be employed in both CRISPR-Cas9 interference (CRISPRi) and CRISPR-Cas9 activation (CRISPRa) to achieve transcription regulation. In CRISPRi, once the dCas9 binds to the DNA target without cleaving it, either transcription initiation or elongation will be blocked, leading to sequence-specific repression of gene expression. In CRISPRa, the dCas9 interacts or is fused with transcription activators, eventually leading to the upregulation of specific genes [118]. Together, CRISPRi and CRISPRa can provide precision control of gene expression, instead of relying on genome editing.

Considering the potential of CRISPR-Cas9 system as a promising tool for precise genome manipulation, several studies have focused on its application in different IRD models. For example, we previously reported that HITI strategy could integrate the RS1 gene into the mouse retina, providing a potential therapeutic solution for treating X-linked juvenile retinoschisis (XLRS; Fig. 2) [119], an IRD with a typical retinoschisis phenotype [82, 120, 121]. Bakondi et al. used CRISPR-Cas9 system to ablate a mutation in Rho gene which causes progressive photoreceptor loss and restored the normal gene function [122]. Other two studies further showed that the Cas9-induced NHEJ strategy can potentially prevent the progression of dominant monogenic diseases such as rhodopsin-associated retinitis pigmentosa (RP) and Best disease [122, 123]. More recently, CRISPR-mediated HDR approach has been adopted to validate the preservation of visual functions by correcting Pde6b mutations in mouse photoreceptors [124]. Yang et al. used CRISPR-generated mutant keratinocytes to identify the essential role of EXOSC2 mutation in the pathogenesis of short stature, hearing loss, retinitis pigmentosa and distinctive facies (SHRF) syndrome [125]. To examine the utility of CRISPR-Cas9 gene editing on retinal dystrophies in vivo, the first phase I/II trial (NCT03872479) was conducted in March 2019 [126]. LCA type 10 is an IRD with IVS26 point mutation that creates a de novo splicing donor site and leads to a functional loss in the CEP290 protein [12]. The efficacy of the CRISPR medicine AGN-151587 (EDIT-101) to remove the IVS26 point mutation was evaluated and this trial was scheduled to be completed in 2024 [126]. EDIT-101 uses two gRNAs, taking advantage of NHEJ, creating a deletion to eliminate the IVS26 mutation [126]. For the aforementioned gene editing strategies, base editing has been used in effectively correct the C625T mutation and disease phenotypes in patient-derived retinal organoids [30] and to restore RPE65 expression and visual functions in a mouse model of IRD [127]. The in vivo efficiency of prime editing was found to be much lower than its efficiency [128, 129]. The large size of prime editing machinery also limited its in vivo delivery to the target. Liu et al. used dual adeno associated virus-mediated delivery of intein prime editing machinery to achieve in vivo delivery and gene editing in the mouse liver [129]. However, the dual vector systems and low efficiency remains concerns regarding the use of prime editing in vivo. It was reported that CRISPRi showed promising efficacy in the treatment of autosomal dominant IRDs [130], whereas the variable efficacy of CRISPRi-mediated gene knockdown, poor delivery efficiency in post-mitotic cells, and the unknown immune responses that may be caused by dCas9 may be the potential challenges for the clinical application of CRISPRi [130].

Fig. 2
figure 2

CRISPR-Cas9-mediated gene knock-in using the homology-independent targeted integration (HITI) strategy. A The CRISPR-Cas9-mediated gene editing initiates after the delivery of CRISPR-Cas9 machinery into target cells. The HITI strategy consists of two steps, including CRISPR-Cas9-mediated DSB formation and DNA repair via the NHEJ pathway. BD An example of RS1/GFP knock-in using the HITI strategy. B Representative bright-field and fluorescence images of RS1/GFP-knock-in B16 cells. C PCR analysis showing the presence of right-arm (R-arm) junction (617 bp) and left-arm (L-arm) junction (748 bp) after the integration of RS1/GFP into the ROSA26 sites. D Sanger sequencing of the genome-donor boundaries in the R-arm and L-arm junctions confirming the integration of RS1/GFP. E Quantitative PCR analysis showing the upregulation of RS1 gene after the RS1/GFP gene knock-in. F Immunofluorescence staining showing the expression of RS1 and GFP after the RS1/GFP gene knock-in. All data are reproduced from our previous work [119]

While the CRISPR technology has a great potential for improving current treatments for IRDs, an efficient and safe delivery system is a critical prerequisite for its success. In the next section, we discuss the significance of nanomaterials for the therapeutic application of CRISPR systems for potential gene therapy of IRDs.

Nanoparticles (NPs)

Successful gene editing requires the efficient delivery of CRISPR-Cas9 machinery to the desired cells. To date, AAV is the most widely adopted delivery system targeting the retina and the eye. Despite Luxturna® as the FDA-approved recombinant AAV gene therapy product [131], there are still some challenges in gene therapy using AAV-based vectors to efficiently deliver the transgenes. As mentioned earlier, the limited virus packaging size of AAV (4.7 kb) can hardly accommodate a typical Cas9 from Staphylococcus Pyrogens (approximately 4.2 kb in size) popularly used in CRISPR-Cas9 studies [10, 11]. In addition, the application of conventional AAV2 has been reported to raise immunogenicity concerns [132]. Several efforts and modifications in vector engineering have been made to improve the AAV-based gene delivery, including the improvement of AAV transduction efficiency, the tropism of AAV vectors, the reduction of the immunogenicity of AAV capsid and transgene, and the optimization of large-scale AAV manufacture [133]. To overcome the size limit of AAV and deliver a large gene expression cassette, scientists have attempted to split transgenes and deliver them via two or three individual AAV vectors [134]. However, the lower transduction efficiency using split AAV vectors than that of a single AAV vector is still a matter of concern [135]. Also, it costs 425,000 US dollars per eye to receive Luxturna® treatment, imposing a heavy economic burden on the society or individuals [1]. Another concern in treating IRDs is the structural peculiarity of the retina, which demands a specific administration route depending on the choice of drug vehicles. With the advent of nanoparticle (NP)-mediated delivery, the unmet needs of efficient genome editing associated with viral vectors are expected to be greatly fulfilled. In therapeutics, the application of NPs as delivery carriers for genes and drugs has been profoundly investigated [136,137,138,139,140,141]. The classification of NPs based on different sizes and structures is shown in Fig. 3. Their nanoscale size enables them to interact with biological systems at the molecular level. In addition, numerous reports have documented that NPs can ensure successful targeted delivery and be transported across biological barriers that can make them an indispensable tool for the drug delivery [142,143,144,145]. The NP-mediated delivery of high molecular weight CRISPR-Cas9 complexes is one of the most significant approaches being developed for genome editing and other evolving applications [146,147,148,149]. Here, we review the promising gene delivery carriers with the potential for IRD treatment based on properties like the nanocarriers and load capacity.

Fig. 3
figure 3

Classifications of NPs. NPs are classified into organic, inorganic, and other NPs. The average sizes of the particles are shown relative to each other and the structural features are shown as discussed in the article

To effectively deliver a therapeutic agent to the retina, the particle size and charge are important parameters when developing nanocarriers [150]. The inner limiting membrane (ILM) with a negative charge located between the vitreous and the retina is a physical and electrostatic barrier [151], which limits the diffusion of NPs [151, 152]. The diffusion of NPs or drugs into the retina varies due to the architecture differences among species, including mice, bovines, and humans [23,24,25]. For example, the pore size of human ILM is about 10 nm, with the variable thickness ranging from 100 nm in the fovea to 4 μm in the thickest area [23]. However, the thickness of the ILM in small animals (such as mice and rats) is less than 100 nm, so the pharmacokinetic distribution of drugs observed in animals often cannot be applicable in clinics [24]. In addition to ILM, the vitreous and retinal cell membranes are all negatively charged. Therefore, after the injection of NPs into the vitreous, the NP’s charge generates an electrostatic interaction which affects the diffusion rate of NPs within the vitreous. Furthermore, the charge of NPs also affects the permeability across the ILM. Therefore, excessive positive and negative charges are not conductive for the drug delivery to the retina [153]. Huang et al. compared lipid NPs with different charges (− 30mv ~  + 50mv) and found that + 35mv lipid NPs can achieve the highest distribution efficiency in the retina [154]. Here, we mainly focus on recent applications of NPs for delivering CRISPR-Cas9 to the retina while an extensive discussion of NP synthesis and design is beyond the scope of this article and has already been reviewed by others [155,156,157,158,159,160,161,162,163,164,165].

Organic NP

Polymer-based cationic NPs

One of the well-known crucial properties of cationic polymers is to form the particular polymer/DNA polyplex. Cationic polymers carry no hydrophobic moiety, making them completely soluble in an aqueous solution. They can also compress DNA molecules to a small size which is considered a crucial feature for improving transfection efficiency in gene transfer [166]. These cationic polymer-based NPs coat negatively charged nucleic acids via electrostatic interactions that secure the nucleic acid's structural stability [158, 167]. Considerable progress has been made over the last decade in polymer-mediated CRISPR-Cas9 plasmid delivery for genome editing applications [149, 168,169,170,171]. The most widely used cationic polymers for pharmaceutical applications include poly-L-lysine (PLL), polyethylenimine (PEI), polyamidoamine (PAMAM) dendrimers, polyethylene glycol (PEG), and chitosan. Despite some evidences revealing the toxicity of PLL[172], PEG [173], PEI [174], and PAMAM [175], remarkable efforts have been made to modify and optimize these cationic polymers and render them as ideal vectors for effective gene delivery with lower toxicity, especially for the delivery of CRISPR-Cas9 [171, 176,177,178]. To date, no study reported the delivery of CRISPR-Cas9 machinery to the retina using this cationic polymer-based delivery strategy. In the following sections, we discuss the update on the main polymer-based NPs that have shown potential in the in vivo and in vitro delivery of CRISPR-Cas9. The application of chitosan in gene delivery is significantly limited due to its low transfection efficiency [179]. Thus, we exclude it from the discussion here.

The cationic polymer PAMAM is a spherical dendrimer composed of repeating branched subunits of amide and amine functional groups. PAMAM is classified into different “generations” (G) based on the number of surface groups that determine its size [180]. The higher the generation number, the greater the number of branches and surface positive charges. This enables PAMAM to interact with nucleic acids by electrostatic interactions and form polyplex complexes [180]. These branched architectures are extensively used in supramolecular chemistry as will be briefly discussed later [149, 181,182,183,184,185,186,187]. PAMAM dendrimers exhibit "proton sponge effect" that causes endosomal swelling and DNA release into the cytoplasm [146]. In several studies, PAMAM dendrimer has been shown to serve as an effective delivery system. Due to its unique characteristics, it can enhance the loading capacity, protect the drugs from degradation, and lessen systematic toxicity [188]. Yavuz et al. conjugated drugs with PAMAM to increase the drug release and ocular residence time and reported that PAMAM was safe for the retina and could be metabolized within 3 h after administration [189, 190]. Kretzmann et al. constructed a highly controllable dendronized polymer that consists of PAMAM dendrons and a linear copolymer backbone to deliver small and large plasmid DNA. This dendronized polymer is also capable of transfecting genome editing tools such as zinc fingers, transcription activator-like effectors (TALEs), and CRISPR/dCas9 platforms [191]. As for the CRISPR/dCas9 platform, Kretzmann et al. used this dendronized polymer to deliver CRISPRa that contains dCa9 fused to VP64 transactivator domain to achieve transcriptional activation of MASPIN (mammary serine protease inhibitor) in the MCF-7 breast cancer cells [191]. These findings highlighted the high transfection efficiency and packing capacity of the dendronized polymer that can deliver larger constructs. However, further in vivo studies are required to confirm its application. PAMAM dendrimers have also been used in CRISPR-Cas9 gene editing in HCT-116 and HT-29 cells [192] and to promote a CRISPR-Cas9-mediated gene editing of programmed death protein-1 (PD-L1) to obtain tumor immunotherapy in melanoma B16F10 cells [193]. So far there was no published studies using PAMAM dendrimers to deliver CRISPR-Cas9 machinery into the retina.

PEG is a polymer of choice for drug delivery applications, and it is often used to modify different nanoparticles [194,195,196,197]. PEG modification that involves the covalent conjugation of PEG to NPs, also called PEGylation, has been shown to enhance structural stability, electrostatic binding, and hydrophobicity. The conjugated protein can be tuned to specifically meet the requirements of drug delivery, for example, by increasing the solubility and stability of the drug while also reducing immunogenicity. This prolongs the retention time of the drug and the conjugate in the blood, thereby reducing the frequency of administration. Charged side chain polypeptide-based NPs [e.g., poly(lactic-co-glycolic acid) (PLG), PLL] have been investigated as drug and gene delivery vehicles [198,199,200,201,202,203,204,205,206]. However, their applications are limited due to the low water solubility and processing complications of these structures.

Previous studies demonstrated the efficient delivery of nucleic acids (such as siRNA) into cells using α-helical polypeptides [198,199,200,201,202,203,204,205,206,207]. A cationic α-helical poly-glutamate-based polypeptide, poly(γ-4-((2-(piperidin-1-yl)ethyl)aminomethyl)-benzyl-l-glutamate (PPABLG) was demonstrated to condense siRNAs or plasmids and maintain the helical structure that ensured high membrane penetration capacity for the cell entrance and endosomal escape. Furthermore, PPABLG was shown to protect the helical structure against environmental stress such as proteases, denaturing conditions, and pH [208]. In 2018, Wang et al. modified PPABLG by incorporating PEG-polythymine 40 (PEG-T 40) to generate PEGylayted helical polypeptide NPs (P-HNPs), 100 nm NPs, for the delivery of the CRISPR-Cas9 gene editing components. These NPs carried Cas9 and gRNA expression plasmids and delivered these components into various cell type to achieve efficient gene editing in vitro. They also used this P-HNP delivery system to achieve a CRISPR-Cas9-mediated gene disruption and repression of tumor growth in vivo [208]. The intratumoral injection of P-HNPs loaded with Cas9 DNA and sgRNA into the xenograft-transplanted mice model could target the survival gene, Plk. Although the reduction of tumor growth required ten injections, multiple administrations did not cause any weight loss in mice, indicating that P-HNPs lacked cytotoxicity. In-depth sequencing analysis of the tumor confirmed that the final genome editing efficiency was 35%, suggesting that P-HNPs may have the potential for medical applications [141].

PEI is one of the most notable cationic polymers with a large amount of positive surface charge exploited in gene transfection in vitro and in vivo [209]. PEI-modified NPs bind with and condense DNA to form spherical structures, which were fused with the endosome and elicited the “proton sponge effect” to escape endosomal degradation [210]. Several studies have emphasized on the application of this polymer in gene therapy [211,212,213]. In 2015, PEI was utilized for delivering Cas9 and gRNA for the first time, leading to the knockout of the Ptch1 gene in the cerebellum of newborn mice [214]. Liang et al. encapsulated the CRISPR-Cas9 plasmid into an aptamer-functionalized PEG-PEI-Cholesterol nanocarrier to target and reduce vascular endothelial growth factor A (VEGF) gene expression both in vivo and in vitro [168]. However, the moderate toxicity of PEG-PEI-Cholesterol in subsequent clinical trials limited its application for drug delivery [215].

Liao et al. reported that the intravitreal injection of PEI/DNA polyplexes could deliver plasmids into retinal ganglion cells in mice. However, the efficiency of PEI-mediated gene editing was low [216]. β-cyclodextrin (β-CD), an FDA-approved drug [217], is a cyclic oligosaccharide with a diameter of 200 nm and high transfection efficiency for small plasmids [171, 218]. To increase the efficiency of gene transfection by PEI, this polymer was covalently linked with β-CD [219, 220] and showed no reported cytotoxicity in HEK293 cells [221]. This strategy increased the transfection efficiency of the luciferase gene to nearly four folds, compared with PEI alone [171]. Given that β-cyclodextrin-PEI (PEI/β-CD) can condense large plasmids at a high nitrogen-to-phosphorus ratio (N/P ratio; the ratio of positively-chargeable polymer amine (N) groups to negatively-charged nucleic acid phosphate (P) groups), Zhang et al. evaluated the efficiency of PEI/β-CD-mediated delivery of CRISPR-Cas9 system in HeLa cells in which the gene transfection efficiency was about 34%. Meanwhile, this delivery resulted in effient editing at hemoglobin subunit beat loci and rhomboid 5 homolog 1 loci of 19.1% and 7%, respectively [171]. Concerning ocular gene delivery, PEI is one of the most widely investigated polymers. Several studies have explored its potential as an alternative delivery vehicle for the eye [25, 222,223,224,225]. For example, one study showed that the intravitreally injected PEI/DNA could be successfully delivered to mouse RGCs [216]. Conceivably, these findings suggest PEI/β-CD as a potential, efficient, and safe nanocarrier for delivering the CRISPR-Cas9 gene editing system into the retina. However, more investigations and efforts are still needed to elucidate the in vivo utility of PEI/β-CD to deliver CRISPR-Cas9 machinery into the retina.

Inorganic nanoparticles

Nanodiamonds (NDs)

Nanodiamonds (NDs) are a novel class of nanomaterials that have garnered a great deal of attention for their clinical application potential due to their low cost, fluorescent capability, low cytotoxicity, and superior biocompatibility [226,227,228,229,230,231]. It has been reported that most of the NPs destroy the endosomal membrane and release the cargo via the proton sponge effect or chemical reaction [232]. Following the entry of NDs into the cells through endocytosis, the sharp structure of these nanocarriers destroys the endosomal membrane resulting in a quick escape. This unique mechanism of endosomal escape makes NDs more biologically safe and stable [233]. Notably, 2–10 nm NDs provided long-term stability without causing cell death and oxidative stress [231, 234]. In addition, DNA, protein, and drugs can be delivered through different surface modifications of NDs (e.g. carboxylation, hydroxylation, hydrogenation, amination, and halogenation) that improve their intracellular uptake and ability to target specific cells [235,236,237,238]. Despite the progress in ND technology, only one study demonstrated the potential therapeutic application of NDs in treating retinal diseases. In our laboratory, we utilized mCherry protein as a critical linker between 3 nm NDs and DNA in which the amide (–NH) and histidine groups (–His) on mCherry protein bind to the carboxylic groups (–COOH) and phosphate groups (–PO43−) on the ND and DNA molecules, respectively [239]. These chemical reactions formed a stable link between NDs and the components of the CRISPR-Cas9 genome editing system. The final size of this nanocarrier is about 5 nm, which facilitates its penetration and the delivery of CRISPR-Cas9 components to all layers of the retina, including photoreceptor and retinal pigmentation epithelium layers. The resulting NDs effectively promoted the delivery of the CRISPR-Cas9 components to initiate HDR to direct the c.625C > T mutation of RS1 gene in human iPSCs and mouse retina, generating an X-linked retinoschisis-like disease model characterized by severe perturbations of the retinal structure (Fig. 4) [239]. The regulatory approval of new drugs by FDA requires inorganic NPs to be cleared via the kidneys to minimize systemic toxicity and improve drug efficacy [240,241,242]. Since NDs are not biodegradable, using them as the delivery vehicle in therapeutics demands an articulate engineering of NPs to meet FDA standards [231, 243,244,245]. NDs with the size of 3 nm show a great promise in satisfying the requirements for the drug clearance by the kidneys and the successful transport and release of CRISPR-Cas9 components.

Fig. 4
figure 4

In vivo delivery of the CRISPR-Cas9 machinery using nanodiamond. A Functionalized nanodiamond can be covalently bound with mCherry protein and linear plasmid DNA encoding the CRISPR-Cas9 gene editing machinery. After mixing with bovine serum albumin (BSA) at defined conditions, functionalized nanodiamond can promote the delivery of CRISPR-Cas9 machinery in vitro or in vivo to cause precise gene editing. B Fluorescence microscopy showing the GFP and mCherry signals in the retinal sections from carboxylated nanodiamond-mCherry- CRISPR-Cas9-treated mouse retina. C ddPCR analysis of RS1 c.625 C > T copy number and D optical coherence tomography visualization of the mouse retina treated with control (Cas9 only) or carboxylated nanodiamond-mCherry-CRISPR-Cas9 NPs (Cas9 + sgRNA). E H&E staining of the retinal sections treated with control (Cas9 only) or carboxylated nanodiamond-mCherry-CRISPR-Cas9 NPs (Cas9 + sgRNA). Visualization methods (D and E) show the reduction and structural perturbation of photoreceptor inner segment/outer segment layer affected in retinoschisis. Data reproduced from our previous study [239]

Gold nanoparticles (AuNPs)

Gold NPs (AuNPs) are small gold particles with a diameter of 1 to 100 nm that are also known as colloidal gold once suspended in a fluid, usually water [246]. Due to the chemical inertness of gold, AuNPs are not considered as biodegradable materials. However, a previous study conducted more than 3 months of observation and found that AuNPs may be degraded by NADPH oxidase (NOX)-mediated reactive oxygen species (ROS) pathways in human fibroblasts [247]. The biomedical applications of AuNPs have been extensively explored [248,249,250,251,252]. Due to the excellent chemical stability, good biocompatibility, and large specific surface area of AuNPs, they can be a promising alternative for gene delivery, including the CRISPR-Cas9 system. To our knowledge, there are very few in vivo studies that have utilized AuNPs for the delivery of CRISPR-Cas9 system [253,254,255,256].

Wang et al. selected AuNPs with a size of approximately 2.4 nm and modified the TAT peptide by glutathione reduction to form the AuNP core. TAT peptide (GRKKRRQRRRPQ) is derived from the human immunodeficiency virus. It is a positively-charged cell-penetrating peptide (CPP) that can overcome the cell membrane's lipophilic barrier and deliver various types of molecules, such as proteins, DNA, antibodies, contrast agents, and toxins. Therefore, AuNP is an attractive approach to deliver the expression plasmids of CRISPR components [257] or the Cas9 protein/gRNA plasmid to target cells [258]. To increase the uptake and delivery efficiency, AuNPs are encapsulated with an anionic lipid shell (1,2-dioleoyl‐3‐trimethylammoniumpropane (DOTAP)/dioleoylphosphatidylethanolamine (DOPE)/cholesterol), which is critical for enhancing the therapeutic application at a low drug dose. The average diameter of the AuNP core is about 20 nm, and the final encapsulated product diameter is about 70 nm (Cas9/gRNA plasmid) or 101 nm (AuNP-CRISPR plasmids).

Although the robust application of AuNP-CRISPR has been found in cancer research, there are some doubts about its application in ophthalmic diseases that need to be further explored. Several in vitro studies reported that AuNP (5–30 nm) may induce cell apoptosis and oxidative stress in the retina [259, 260]. However, in vivo studies did not support the AuNP-induced inflammatory reactions or damage to the retinal structure [261, 262]. Based on the above results, the AuNPs have a potential for gene therapy applications by introducing CRISPR-Cas9 system to the retina. Nonetheless, the excretion, and toxicity of inorganic NPs are important concerns that must be addressed.

Graphene oxide (GO)

Graphene, synthesized from graphite, is a nanoscaled monolayer of carbon atoms, arranged in a two-dimensional (2D) crystal structure [263, 264]. Graphene and its family members, including graphene oxide (GO), have been extensively studied owing to their unique properties: high surface area, mechanical and chemical stability, and biocompatibility [264]. Previous studies have found that nano-grade GO, a single-layer graphene oxide sheet, may directly penetrate the cell membrane [265] or enter the cell through endocytosis [266].

Here, we briefly discuss the therapeutic potential of GO nanomaterials in retinal drug delivery. However, only a few studies have utilized GO to deliver CRISPR-Cas9 system [267,268,269,270,271]. Yue et al. constructed a stable and functional dual polymer-GO nanocarrier (GO-PEG-PEI) by conjugating the hydrophilic PEG polymer and the positively charged PEI polymer together with the GO via amide bonds. This dual nanocarrier was utilized to adsorb the negatively charged Cas9/gRNA. The diameter and length of GO-PEG-PEI were 1 nm and 165 nm, respectively, which were increased to 4 nm and 220 nm after loading of the Cas9/gRNA complex. Next, they evaluated the efficiency of GO-PEG-PEI nanocarrier for delivering Cas9/single-guide RNA (sgRNA) complex into the human gastric adenocarcinoma cell line. The results showed 36% of cells underwent NHEJ-based gene editing with 95% cell viability, indicating a successful in vitro gene editing using this nanocarrier [271]. Another in vitro study examined the influence of GO on human retinal pigment epithelium (RPE) cells and demonstrated its satisfactory biocompatibility. However, the cell viability and morphology were slightly affected following a prolnged exposure to GO. Besides, a small amount of GO aggregation was reported [272]. Surprisingly, the intravitreal injection of GO into rabbits’ eyes did not cause obvious ocular structural defects, vision loss, or inflammation [272]. Another study showed that PEG coating strategy promotes the clearance of GO from liver, lung, and spleen in mice, supporting its feasibility for biomedical applications [273]. These results suggested that GO has the potential to carry and deliver CRISPR-Cas9 into the retina; however, the two-dimensional element of GO poses some challenges for its application to treat IRDs.

Artificial virus NPs

As mentioned earlier, considerable progress has been made in optimizing CRISPR-Cas9 delivery in vitro. However, further modifications are necessary to improve the delivery efficiency of CRISPR-Cas9 components in vivo. The ligand-receptor interactions strategy is a promising approach for targeted drug and nucleic acid delivery, by which it reduces the drug toxicity and enhances its effectiveness [274]. Since NP’s infection capacity is significantly lower than that of viruses, the strategy of “artificial viruses” was proposed to improve this shortcoming. This strategy takes advantage of a virus-like core (composed of plasmid DNA, condensing agent, and functional peptides) and a hydrophilic shell that can expose specific targeting ligands [275]. Li et al. developed an artificial virus with enhanced endosomal escape and transfection efficiency to successfully carry and release the CRISPR-Cas9 system. In this study, the fluorinated branched PEI (PF33) was utilized as a condensing agent and combined with an expression plasmid containing Cas9 and gRNA directed against the MTH1 gene as the artificial virus core. The main chain of the shell was natural hyaluronic acid (HA) polymer, and the side chain was modified by PEG and R8-RGD tandem peptide called RGD-R8-PEG-HA (RRPH) multifunctional shell. The combination of PEG side chains with HA backbone improved the artificial virus stability and uptake efficiency. Besides, the terminal attachment of R8-RGD tandem peptide with PEG chain provided the artificial virus with increased target specificity and penetration ability. Notably, the transfection efficiency of the artificial virus was reported as high as > 90% in human ovarian cancer SKOV3 cells, and the resultant CRISPR-Cas9-mediated MTH1 gene disruption efficiency was 60%. Furthermore, the xenograft analysis of immunodeficient nude mice demonstrated the therapeutic effect of the artificial virus in inhibiting tumor metastasis [276]. So far, there has been no study using this artificial virus nanoparticle in the retina.

Supramolecular nanoparticles (SMNPs)

Another promising nanocarriers for drug delivery with high transfection efficiency are supramolecular nanoparticles (SMNPs). The main features of these NPs are their unique assembly through specific non-covalent interactions and molecular recognition properties that render distinct advantages, including controllable particle size, optimizable surface charge, and enhanced delivery efficiency [277]. The latest SMNP design mainly comprises PAMAM, PEI, and TAT. Among these components, PAMAM produces self-assembled nanoparticles, the cationic polymer PEI condenses DNA, and TAT can penetrate the membrane. In our previous study, we combined PEI/β-CD, adamantane-grafted PAMAM (Ad-PAMAM), adamantine-grafted PEG (Ad-PEG), and Ad-PEG-TAT to form SMNPs to deliver CRISPR components into the retina. To evaluate the efficiency and safety of the designed SMNPs for delivering CRISPR-Cas9 components in gene therapy, they were loaded with donor-RS1/GFP DNA and Cas9/gRNA expression plasmids, resulting in the final size of 110–127 nm. The donor-RS1/GFP was successfully inserted into the ROSA26 (safe harbor) locus using HITI strategy. The delivery efficiencies of donor-RS1/GFP-plasmid using SMNPs and Lipofectamine 3000 were comparable, however, the latter was associated with higher cytotoxicity. This study further showed that SMNPs could deliver the plasmids into the RGC layer via intravitreal injection (Fig. 5A–C) and successfully conducted the knock-in of donor RS1 into the ROSA26 locus (Fig. 5D–G) [119]. This study introduced a new supramolecular particle with promising features for treating IRDs.

Fig. 5
figure 5

In vivo delivery of the CRISPR-Cas9 machinery using SMNPs. A, B Schemes of the self-assembly of SMNPs loaded with plasmid DNA encoding Cas9/gRNA or donor-RS1/GFP. C A schematic presentation showing the intravitreal injection of SMNPs loaded with the indicated plasmids, resulting in the delivery of CRISPR-Cas9 machinery to express RS1 and GFP proteins in retinal layers. D Fundus photography (left) and optical coherence tomography (right) images showing GFP signals and retinal structure after SMNP-mediated gene delivery. E H&E and immunohistochemistry staining of GFP-positive cells in the retinal layers. F Electrophoretogram showing PCR amplification of the right-arm junction (617 bp) and left-arm junction (748 bp). G Sanger sequencing of the genome-donor boundaries showing the effective CRISPR-Cas9-mediated knock-in of RS1/GFP genes in vivo. Data reproduced from our previous study [119]

Lipid-based cationic nanoparticles (LNPs)

Lipid-based nanoparticles (LNPs) are a mixture of cationic lipids (such as DOTAP, MVL5) and neutral lipids (such as DOPE, cholesterol). These lipids are mixed with DNA under controlled microfluidic conditions and self-assemble into spherical LNPs with a final diameter of 100–200 nm [278, 279]. The cationic lipids are responsible for the electrostatic interaction of LNPs with DNA and their encapsulation [279]. LNPs can be further modified by PEGylation, to reduce the drug aggregation and degradation, which contributes to its therapeutic effect [158]. Additionally, the biodegradability of LNPs can be optimized by modifying the functional groups of ester chains [158].

The transfection efficiency of LNPs is positively correlated with the surface charge density, which is measured by the ratio of cationic lipids to neutral lipids [280]. The high surface charge density of LNPs enables them to effectively fuse with the endosomal membranes, thereby releasing DNA into the cytoplasm [280]. However, the excessive surface charge density hinders DNA dissociation from the liposome complex following endosomal escape [281]. Besides, the excessive positive charge of LNPs causes their poor diffusibility in the negatively charged vitreous as well as their limited passage to the ILM. Although several modifications such as PEG, DOPE, cholesteryl hemisuccinate (CHEMS) have been developed to reduce the positive charge of LNPs [282,283,284], their safety and transfection efficiency need to be evaluated to expand the utility of gene therapy for IRDs. Despite the challenges mentioned above, LNPs have emerged as a promising alternative for the drug delivery vehicle in clinics [179]. In 2018, Onpattro® (patisiran) was approved by FDA as the first LNP-based nanomedicine used for the treatment of polyneuropathy in patients with hereditary transthyretin-mediated amyloidosis (hATTR amyloidosis) [285, 286]. Later in 2020, LNPs were utilized in FDA-approved COVID-19 vaccines (mRNA-1273 from Moderna, and BNT162b2 from BioNTech/Pfizer) for mRNA delivery [287]. Moreover, D-Lin-MC3-DMA (MC3), an FDA-approved cationic ionizable lipid (CIL)-like LNP, could successfully deliver mCherry mRNA to the RPE layer of the retina in mice [279]. Several in vitro and in vivo studies have demonstrated the application of LNPs for the delivery of the CRISPR-Cas9 system [161,162,163,164,165]. For example, Zhang et al. established a phospholipid-modified cationic lipid nanoparticle (PLNP) delivery system modified with PEG for the delivery of the Cas9/sgRNA plasmid (DNA). Their study demonstrated the delivery efficiency of 47.4% into the A375 cell line (human malignant melanoma), while it was 3.09% for Lipofectamine. Furthermore, the intratumoral injection of PLNP/DNA to a mouse model with a melanoma tumor showed a significant tumor growth reduction [164]. As mentioned earlier, one major concern for the application of CRISPR-Cas9 system in clinics is the uncontrolled expression of Cas9 and gRNA in the cells of interests. Fin et al., created a biodegradable LNP-based delivery system (LNP-INT01) and injected this LNP containing Cas9 and gRNA into mice. The results clearly showed rapid clearance of the molecules 72 h post administration. In addition, they reported a high level (> 97%) knockdown of serum transthyretin levels in mice [161]. These findings are indicative of the efficacy and capacity of LNPs for CRISPR-Cas9 delivery in vivo.

Other nanoparticles

DNA nanoclews

Yarn-like DNA nanoparticles, known as DNA nanoclews, are delivery vehicles synthesized by the rolling circle amplification (RCA) method, producing long-chain ssDNAs with palindromic sequences required for their self-assembly [288]. Since DNA is intrinsically biocompatible and degradable, DNA nanoclews hold great promise in developing ideal drug delivery nanocarriers [289, 290]. However, the immune issues related to these nanocarriers are a matter of concern, therefore, their clinical applications should be further investigated [291]. To the best of our knowledge, no study explored the potential of DNA nanoclews for drug delivery into the eye and only one study reported the utility of DNA nanoclews for delivering CRISPR-Cas9 as briefly described below [292]. Since this type of NPs is made of ssDNA, a sequence complementary to gRNA can be designed to match the base-pair of the guide portion of Cas9-sgRNA [293]. Sun et al. coated the DNA nanoclews/Cas9 protein/gRNA mixture with PEI to improve its cellular uptake and the endosomal escape. The resulting DNA nanoclews were able to deliver the CRISPR-Cas9 components into the target cells in vitro and in vivo with the editing efficiencies of 36% and 25%, respectively [292].

Nanoscale zeolitic imidazole frameworks (ZIFs)

Nanoscale zeolitic imidazole frameworks (ZIFs), a subclass of metal–organic frameworks (MOFs), are composed of divalent metal cations and imidazolate bridging ligands with pH buffering capacity. These features enable ZIFs to facilitate endosomal escape [294]. ZIFs combine the advantages of the 3D network and porous structure of zeolite with traditional metal–organic clusters [295, 296] that have recently attracted more attention due to their great potential for delivering drugs, genes, and proteins [297,298,299]. Alsaiari et al. reported for the first time that ZIF-8 could encapsulate Cas9 protein and gRNA and subsequently undergo genome editing in Chinese hamster ovary (CHO) cells, with loading and editing efficiencies of 17% and 37%, respectively [294]. To the best of our knowledge, no report on the delivery of CRISPR-Cas9 system to the retina using such nanocarrier has been published.

A very recent study utilized a water-in-oil emulsion approach to fabricate a pH-responsive silica–metal–organic framework hybrid NP (SMOF NP) consisting of silica and ZIF. Subretinal injection of SMOF NPs induced efficient genome editing in mouse retinal pigment epithelium. Furthermore, both in vitro loading and delivery efficiencies of CRISPR-Cas9 components by SMOF NPs were high but varied depending on different cell lines [300]. These data introduced a promising nanoplatform that may improve gene therapy in the treatment of IRDs.

To summarize, most modern NPs use non-covalent bonding to carry plasmids expressing Cas9 and gRNA or Cas9 protein and gRNA expression plasmids. These NPs are about 100–200 nm in size with a slight positive charge, which may be suitable for intravitreal injection into the eyes of patients with IRDs. We have also described ND carriers as the only carriers smaller than 10 nm. The size may allow the NPs to pass through the human ILM barrier and effectively perform HDR-based CRISPR gene editing in all retinal layers. Also, SMNPs can adapt to various in vivo environments with their unique molecular recognition ability and therefore have the potential to be an alternative approach for the delivery of CRISPR system to the vitreous. Yet, the means of safe and efficient delivery remain to be fully investigated. The properties and advantages/disatvatages of different types of NPs are summarized in Table 2.

Table 2 Summary and comparison of properties of NPs mentioned in this review

The retinal organoids and precision medicine

The heterogeneity of IRDs hampers the development of an effective strategy to tackle a wide range of disorders. One of the major hurdles that hinders the translation of basic retinal research into clinical applications is mainly due to the poor relevance in existing preclinical models. For example, in the mouse model, more than 90% of photoreceptors are rod cells, whereas, in humans, the visual acuity is mostly dependent on cone photoreceptors [301]. Notably, many in vitro and in vivo findings could not be reproduced in humans. Drugs that proved to be safe and effective in animal studies failed to exert the same efficacy in clinical trials. In addition, the information obtained by studying two-dimensional cultures does not recapitulate the heterogeneous complexity and critical features of the microenvironment of cells in vivo [136,137,138]. This gap causes a noticeable lack of fidelity between the aforementioned experimental models and human outcomes. The advent of three-dimensional (3D) multicellular constructs, referred to as the organoid technology, offers a promising complementary model to pursue clinical translation and precision medicine applications. Human pluripotent stem cells (hPSC)-derived retinal organoids hold an excellent value for modeling the human retina features [302,303,304,305,306]. In particular, by using patient-derived cells combined with reprogramming strategies, this technique could represent an efficient pre-clinical approach toward the personalized therapeutic strategies adaptable to a broad number of IRDs and provide the link to disease-specific human drug screening models [307]. The iPSC-derived organoids strategy can provide a means for assessing the efficiency and efficacy of NP-mediated delivery of CRISPR-Cas9 system tailored to each patient's genetic makeup. In this part, we mainly focus on organoid as a technology platform for precision medicine in IRDs, especially with potential translational applications for evaluating new therapeutic drugs.

Organoids as a drug testing platform for translational research

Introducing a new pharmaceutical drug to the market is a complicated and costly process, especially when the in vivo testing result of drug candidates fail to reach the requirements initially fulfilled by the in vitro test. The gap between in vitro validation and clinical application is significant, mainly because the simplicity of the in vitro model cannot mimic the complex nature and heterogeneous characteristics of clinical patients [308]. These critical problems impose significant limitations on the translation of candidate drugs to the clinic and require advanced strategies to improve this shortage. Organoid systems show considerable reliability of recapitulating features and functions of the human system offering a great potential for testing drug efficiency in target organs. Patient-derived organoids (PDOs) can be generated particularly by reprogramming patient-derived cells to induced pluripotent stem cells (iPSCs), followed by the differentiation into the desired cell lineage and organoids [309]. Notably, several reports showed that in vitro PDOs could highly match and reproduce patients’ response to candidate drugs in most cases, highlighting the merit of this system in personalized medicine as a predictor of therapeutic outcome [310,311,312,313]. The organ-like structure technology offers a more efficient screening of candidate drugs prior to in vivo testing, which helps to reduce drug development costs. Organoids have been a powerful tool for functional drug testing, personalized therapy and disease modeling [314,315,316,317,318,319,320]. For iPSC-derived organoids, various organoid models have been generated using human iPSCs, including heart [321, 322], kidney [323], brain [324, 325], intestine [312, 313, 326], liver [327,328,329], lung [330] and retina [302,303,304,305]. Furthermore, organoids have been successfully applied to model human genetic diseases. For example, intestinal organoids derived from cystic fibrosis (CF) patients proved to be a reliable tool for effective drug treatment [314]. Of note, brain [315,316,317] and kidney [318,319,320] organoids generated from patient-derived iPSCs have also been established to model diseases. Xu et al. subjected brain organoids to Zika virus infection and used them as the platform for drug repurposing [331]. Bian et al. demonstrated that the neoplastic cerebral organoids are suitable for targeted drug testing [332]. Saengwimol et al. used retinoblastoma organoids for the evaluation of cellular response to chemotherapy drugs. [333]. However, the consistency and reducibility of this system at a scale consistent with clinically associated cell numbers is still a matter of concern [334]. For retinal organoids, Vergara et al. developed an iPSC-derived retinal organoid-based screening platform that allows the accurate quantification of fluorescent reporters [335]. Despite the progress of some organoid-based researches, the organoid technologies remains immature and not ready for the demands of high-throughput screening in drug screening [336]. It was attributed to the developmental variability and diversity of retinal organoids that may hinder the utility of retinal organoids in the evaluation of therapeutic effects and comparative analysis [337]. Nevertheless, considering that retinal organoids hold promising potential in new drug development, it would be still expected and encouraging to use retinal organoid technologies to augment the existing drug development pipelines. Collectively, these findings highlight the utility of organoids as a part of the drug testing pipeline, creating the opportunities for more effective therapies, especially for patients with rare genetic diseases in a cost-effective manner.

Retinal organoid applications in precision medicine

To establish a drug testing platform for IRD treatment based on organoid technologies, generating a representative disease model is a fundamental step. As highlighted below, several reports demonstrated the utility of organoids in eye disease modeling. Ohlemacher et al. developed a retinal organoid model using patient-specific iPSC-derived RGCs to study an inherited form of glaucoma [26]. Tucker et al. generated multi-layer optic cup-like structures representing photoreceptor precursor cells for investigating the pathogenesis of RP [27]. A separate study focused on different frameshift mutations in the RPGR gene, one of the most prevalent causes of autosomal recessive RP, and generated patient-specific retinal organoids with defects in morphology and functionality of photoreceptors accompanied with decreased cilia length as a disease model [28]. The constructed vectors for the CRISPR-Cas9 machinery were delivered into the patient-derived iPSCs via electroporation, and the mutation-corrected iPSCs were then differentiated into retinal organoids. Notably, the reversal of morphological and functional defects in retinal organoids with RPGR mutation was observed after the CRISPR-Cas9-mediated gene correction [28]. Using a similar approach, Buskin et al. demonstrated the severe RP defects observed in patient-specific retinal organoids harboring the CRISPR-Cas9-induced PRPF31 mutation [29]. This further proved the effectiveness of this combined strategy toward personalized and targeted gene therapy. Another example for coupling retinal organoids with the genome engineering technique is the application of CRISPR-Cas9 technology to correct RS1 mutation in retinal organoids derived from XLRS-patients [30]. Huang et al. successfully established the XLRS patient-derived retinal organoids that recapitulate the retinal splitting feature of the disease. Meanwhile, they delivered the CRISPR-Cas9 system using electroporation to correct the mutation and showed that CRISPR-Cas9-mediated correction of the disease-associated C625T mutation efficiently rescued the disease phenotype (Fig. 6) [30]. Parfitt et al. used LCA patient-derived iPSCs to generate 3D optic cups with the mutation of a cilia-related gene, CEP290 [31]. Introducing an antisense oligonucleotide to patient iPSC-derived organoids could effectively prevent the aberrant splicing, restore the expression of full-length CEP290 protein, and repaired the cilia defects [31]. Currently, CEP290 treatment is in phase III clinical trial (NCT03913143), paralleling the classic augmentation RPE65 trials initiated in 2007 [338, 339]. Overall, these studies demonstrated that 3D retinal organoids derived from patients with various retinal diseases are able to recapitulate the complex retinal architecture, rendering them an ideal platform for examining the safety and specificity of CRISPR-Cas9 system for the therapeutics applications.

Fig. 6
figure 6

Reproduced from our previous study [30]

Patient-derived retinal organoids recapitulate disease-specific features. A Bright-field images and B H&E staining of control and XLRS patient-derived retinal organoids exhibit schisis feature at day 150 of differentiation. C Quantification of splitting area in control and XLRS-patient-derived retinal organoids. D Bright-field images of control and XLRS patient-derived retinal organoids at days 90, 100, and 110 of differentiation. E A schematic presentation of the time course for the generation of control and XLRS patient-derived retinal organoids. Disease-specific features can be observed after applying defined differentiation stimuli and time course.

Conclusions and perspectives

The combination of NPs, CRISPR-Cas9, and retinal organoids as a promising therapeutic platform

IRDs have long been viewed as a class of disorders with no effective treatment. This maxim is now being reversed by tremendous efforts in nanomedicine and gene engineering, which results in promising clinical trials for blinding diseases. As therapeutic strategies for IRDs expand, the importance of molecular diagnosis is gaining momentum. Most IRD gene supplementation therapies are in phase I/II clinical trials, with LCA therapy approved by the FDA to treat patients carrying biallelic RPE65 mutations [59, 62, 340,341,342]. Although these efforts are still evolving, the importance of gene therapy for elevating the life quality of IRD patients has never been more apparent. In this review, we introduced a forward movement of therapy by combining the advances in CRISPR-mediated gene editing, NP-based delivery, and iPSC-derived retinal organoids technologies, to assess the potential safety and efficacy of designed CRISPR-Cas9 components and nanocarriers in a clinically relevant in vitro model. As mentioned earlier, NP-mediated delivery of high molecular weight CRISPR-Cas9 complexes combined with the advanced CRISPR-Cas9 technologies, applicable in non-dividing retinal cells (e.g. HITI), introduces a simple yet efficient approach for precise gene therapy in IRDs. In addition, patient-derived retinal organoids can mimic typical disease features, providing a reliable platform for disease modeling. Collectively, this integrated strategy is expected to facilitate the evaluation of the gene editing in preclinical tests and be a major driver towards advancing IRD’s personalized medicine (Fig. 7).

Fig. 7
figure 7

An integrative, multidisciplinary approach for future gene therapy in IRDs. IRD patient’s blood sample can be reprogrammed into induced pluripotent stem cells (iPSCs) followed by the differentiation into retinal organoids. These patient-derived retinal organoids exhibit disease-specific features and can be applied as a reliable platform for assessing disease progression and treatment outcome (e.g. XLRS-patient-derived retinal organoids exhibit severe retinoschisis-like features). Researchers can utilize optimized NPs loaded with plasmid DNA encoding CRISPR-Cas9 machinery to achieve efficient gene delivery and precise gene editing. This results in the rescue of the disease phenotypes associated with the specific IRDs (e.g. the splitting phenotype in XLRS-patient-derived retinal organoids can be rescued as shown above). Integrating patient-derived retinal organoids, CRISPR-Cas9 technologies, and NPs promotes precision gene therapy applications for IRDs

During the past few years, our preliminary research on combining CRISPR-Cas9 gene editing, patient-derived retinal organoids, and NPs has been conducted to investigate the potential of this integrated strategy in IRD therapeutics. For example, Yang et al. utilized ND-mediated delivery of CRISPR-Cas9 to introduce the mutated RS1 gene into human iPSCs and mouse retina, leading to the generation of XLRS-like disease model [239], however, the potential of this strategy for gene therapy is yet to be explored. In another study, Chou et al. successfully combined SMNP nanocarrier and CRISPR-Cas9-mediated HITI strategy to knock-in the RS1 gene in the mouse retina [119]. In the future, this strategy can be further applied in patient-derived retinal organoids model to assess its efficiency and efficacy in clinically relevant disease setting. In a separate study, Huang et al. successfully differentiated iPSCs from XLRS patients into retinal organoids presenting disease features. They further coupled this disease model with CRISPR-Cas9 technology and repaired the RS1 gene mutation with 50% efficiency [30]. However, the electroporation method used for the transfection of CRISPR-Cas9 machinery caused a massive cell death [30]. Therefore, future attempts to utilize NPs as an alternative method will be of great interest to resolve this drawback and further boost the therapeutic effects. We hope that this combined strategy will become a treatment modality for other IRDs and elevate the life quality of patients.

Technology hurdles

Although the rationale for this integrated approach is clear, several technology hurdles remain to be addressed. The development of advanced CRISPR-Cas9 systems with high specificity has armed researchers worldwide with a powerful tool to study human diseases. However, utilizing this technique for translational medicine research has inevitable concerns. Since patient-derived iPSCs not only carry the specific mutation intended to be repaired but also harbor the entire human genome, this makes CRISPR-Cas9-mediated gene editing more susceptible to undesired off-target effects. For example, gRNA may recognize sequences similar to the target loci and cause permanent sequence alterations resulting in abnormal gene function. Furthermore, no in vivo study examined how long the nuclease remains active before its degradation and what the possible adverse effect(s) might be. Anti-CRISPR proteins can be used to limit the off-target editing, however, the best time to shut off Cas activity requires optimizations [343]. Another concern is that CRISPR-Cas9-mediated knockout or overexpression of the gene of interest may be compensated by neighboring cells, which would interfere with the expected outcome [344]. As for the organoid technology platform, although the 3D retina organoid has equipped ophthalmology with a unique and relatively accurate representation of the human eye, it still lacks a high level of morphological and functional complexity demonstrated by the mammalian retina in vivo. For example, the 3D retinal organoid with retinal pigment epithelium (RPE) layer provides a more physiologically relevant disease model for photoreceptor-associated diseases. However, both simple and complex organoid models have their pros and cons; thus, the appropriate level of complexity should be designed according to the purpose of the study [345]. A more challenging issue is the disease modeling of late-onset retinal diseases by manipulating culture medium to induce ageing factors. It demands a profound knowledge of factors involved in each specific developmental stage which are yet to be investigated. More in-depth knowledge and assessment of the culture medium composition and distribution are required for modeling complex IRDs. Nevertheless, retinal organoid technologies have only been utilized for monogenic IRDs so far. Lastly, although NP-based delivery has been proposed and proven to be a promising drug delivery vehicle, improving the therapeutic application of NP-based gene therapy remains an important concern. It requires in-depth investigation on the cytotoxicity of the NPs under variable conditions and on the key factors determining the release rate of drugs from NPs to the retina. The successful drug delivery to the neuroretina, and even more specifically to photoreceptors, highly depends on the choice of the NPs most suitable for both the drug and the target tissue. Besides, different gene mutations may affect the complex retina structure and cause anatomical obstacles for nanomedicine drug delivery. Moreover, the choice of delivery route, immunoreactivity, and nucleic acid-based drug stability are critical factors that need to be addressed for a successful clinical application. Nevertheless, in this ever-evolving field, it is crucial to move scientific discoveries into clinics and new therapies for vision restoration in more patients than ever before. Ultimately, the application of such technologies in the clinic and industry should fulfill four criteria: reproducibility, standardization, validation, and quality assurance. Although NP-mediated delivery of the CRISPR-Cas9 system shows a great promise in repairing the IRDs, the time window for treatment is a critical determinant for the therapeutic outcome. For example, in XLRS, the differentiated retinal cells with the splitting phenotype are not responsive to gene therapy, indicating the delayed treatment by the time the disease is already progressed. However, more preclinical data will be required to prove this concept. Taken together, the advances and current progress of basic research hold the promises that laboratory findings can be translated into clinical applications in the near future and bring hope to patients who have blindness and other hereditary diseases. The integration of nanotechnology, CRISPR, and stem cell technologies present a novel platform and is expected to accelerate bridging the basic research and translational medicine, and further promote medical precision therapies.

References

  1. Daiger S, Sullivan L, Bowne S, Rossiter B. RetNet: retinal information network. Na+ Ca2. 2013;5.

  2. Farrar GJ, Carrigan M, Dockery A, Millington-Ward S, Palfi A, Chadderton N, Humphries M, Kiang AS, Kenna PF, Humphries P. Toward an elucidation of the molecular genetics of inherited retinal degenerations. Hum Mol Genet. 2017;26:R2–11.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  3. Koenekoop RK, Sui R, Sallum J, Van Den Born LI, Ajlan R, Khan A, Den Hollander AI, Cremers FP, Mendola JD, Bittner AK. Oral 9-cis retinoid for childhood blindness due to Leber congenital amaurosis caused by RPE65 or LRAT mutations: an open-label phase 1b trial. Lancet. 2014;384:1513–20.

    Article  PubMed  CAS  Google Scholar 

  4. Bennett J, Wellman J, Marshall KA, McCague S, Ashtari M, DiStefano-Pappas J, Elci OU, Chung DC, Sun J, Wright JF. Safety and durability of effect of contralateral-eye administration of AAV2 gene therapy in patients with childhood-onset blindness caused by RPE65 mutations: a follow-on phase 1 trial. Lancet. 2016;388:661–72.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  5. Finger RP, Fimmers R, Holz FG, Scholl HP. Prevalence and causes of registered blindness in the largest federal state of Germany. Br J Ophthalmol. 2011;95:1061–7.

    Article  PubMed  Google Scholar 

  6. Liew G, Michaelides M, Bunce C. A comparison of the causes of blindness certifications in England and Wales in working age adults (16–64 years), 1999–2000 with 2009–2010. BMJ Open. 2014;4:e004015.

    Article  PubMed  PubMed Central  Google Scholar 

  7. Khan NW, Falsini B, Kondo M, Robson AG. Inherited retinal degeneration: genetics, disease characterization, and outcome measures. J Ophthalmol. 2017. https://doi.org/10.1155/2017/2109014.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Sullivan LS, Daiger SP. Inherited retinal degeneration: exceptional genetic and clinical heterogeneity. Mol Med Today. 1996;2:380–6.

    Article  PubMed  CAS  Google Scholar 

  9. Gupta PR, Huckfeldt RM. Gene therapy for inherited retinal degenerations: initial successes and future challenges. J Neural Eng. 2017;14: 051002.

    Article  PubMed  Google Scholar 

  10. Flotte TR. Size does matter: overcoming the adeno-associated virus packaging limit. Respir Res. 2000;1:16–8.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  11. Peng R, Lin G, Li J. Potential pitfalls of CRISPR/Cas9-mediated genome editing. FEBS J. 2016;283:1218–31.

    Article  PubMed  CAS  Google Scholar 

  12. Grieger JC, Samulski RJ. Packaging capacity of adeno-associated virus serotypes: impact of larger genomes on infectivity and postentry steps. J Virol. 2005;79:9933–44.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  13. Selot R S, Hareendran S, Jayandharan G R. Developing immunologically inert adeno-associated virus (AAV) vectors for gene therapy: possibilities and limitations. Curr Pharm Biotechnol. 2013;14:1072–82.

    Article  Google Scholar 

  14. Weleber RG, Pennesi ME, Wilson DJ, Kaushal S, Erker LR, Jensen L, McBride MT, Flotte TR, Humphries M, Calcedo R. Results at 2 years after gene therapy for RPE65-deficient Leber congenital amaurosis and severe early-childhood–onset retinal dystrophy. Ophthalmology. 2016;123:1606–20.

    Article  PubMed  Google Scholar 

  15. Thompson DA, Iannaccone A, Ali RR, Arshavsky VY, Audo I, Bainbridge JW, Besirli CG, Birch DG, Branham KE, Cideciyan AV. Advancing clinical trials for inherited retinal diseases: recommendations from the Second Monaciano Symposium. Transl Vis Sci Technol. 2020;9:2–2.

    Article  PubMed  PubMed Central  Google Scholar 

  16. Ran FA, Hsu PD, Wright J, Agarwala V, Scott DA, Zhang F. Genome engineering using the CRISPR-Cas9 system. Nat Protoc. 2013;8:2281–308.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  17. Schwank G, Koo B-K, Sasselli V, Dekkers JF, Heo I, Demircan T, Sasaki N, Boymans S, Cuppen E, van der Ent CK. Functional repair of CFTR by CRISPR/Cas9 in intestinal stem cell organoids of cystic fibrosis patients. Cell Stem Cell. 2013;13:653–8.

    Article  PubMed  CAS  Google Scholar 

  18. Bainbridge JW, Smith AJ, Barker SS, Robbie S, Henderson R, Balaggan K, Viswanathan A, Holder GE, Stockman A, Tyler N. Effect of gene therapy on visual function in Leber’s congenital amaurosis. N Engl J Med. 2008;358:2231–9.

    Article  PubMed  CAS  Google Scholar 

  19. Moreno AM, Fu X, Zhu J, Katrekar D, Shih YRV, Marlett J, Cabotaje J, Tat J, Naughton J, Lisowski L. In situ gene therapy via AAV-CRISPR-Cas9-mediated targeted gene regulation. Mol Ther. 2018;26:1818–27.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  20. Thomas CE, Ehrhardt A, Kay MA. Progress and problems with the use of viral vectors for gene therapy. Nat Rev Genet. 2003;4:346–58.

    Article  PubMed  CAS  Google Scholar 

  21. Nakamura H, Matsui KA, Takagi S, Fujisawa H. Projection of the retinal ganglion cells to the tectum differentiated from the prosencephalon. Neurosci Res. 1991;11:189–97.

    Article  PubMed  CAS  Google Scholar 

  22. Polyak SL. The retina. Chicago: University of Chicago Press; 1941.

    Google Scholar 

  23. Henrich PB, Monnier CA, Halfter W, Haritoglou C, Strauss RW, Lim RY, Loparic M. Nanoscale topographic and biomechanical studies of the human internal limiting membrane. Invest Ophthalmol Vis Sci. 2012;53:2561–70.

    Article  PubMed  Google Scholar 

  24. Slijkerman RW, Song F, Astuti GD, Huynen MA, van Wijk E, Stieger K, Collin RW. The pros and cons of vertebrate animal models for functional and therapeutic research on inherited retinal dystrophies. Prog Retin Eye Res. 2015;48:137–59.

    Article  PubMed  Google Scholar 

  25. Pitkänen L, Pelkonen J, Ruponen M, Rönkkö S, Urtti A. Neural retina limits the nonviral gene transfer to retinal pigment epithelium in an in vitro bovine eye model. AAPS J. 2004;6:72–80.

    Article  PubMed Central  Google Scholar 

  26. Ohlemacher SK, Sridhar A, Xiao Y, Hochstetler AE, Sarfarazi M, Cummins TR, Meyer JS. Stepwise differentiation of retinal ganglion cells from human pluripotent stem cells enables analysis of glaucomatous neurodegeneration. Stem Cells. 2016;34:1553–62.

    Article  PubMed  CAS  Google Scholar 

  27. Tucker BA, Mullins RF, Streb LM, Anfinson K, Eyestone ME, Kaalberg E, Riker MJ, Drack AV, Braun TA, Stone EM. Patient-specific iPSC-derived photoreceptor precursor cells as a means to investigate retinitis pigmentosa. Elife. 2013;2: e00824.

    Article  PubMed  PubMed Central  Google Scholar 

  28. Deng W-L, Gao M-L, Lei X-L, Lv J-N, Zhao H, He K-W, Xia X-X, Li L-Y, Chen Y-C, Li Y-P, et al. Gene correction reverses ciliopathy and photoreceptor loss in iPSC-derived retinal organoids from retinitis pigmentosa patients. Stem Cell Rep. 2018;10:1267–81.

    Article  CAS  Google Scholar 

  29. Buskin A, Zhu L, Chichagova V, Basu B, Mozaffari-Jovin S, Dolan D, Droop A, Collin J, Bronstein R, Mehrotra S, et al. Disrupted alternative splicing for genes implicated in splicing and ciliogenesis causes PRPF31 retinitis pigmentosa. Nat Commun. 2018;9:4234.

    Article  PubMed  PubMed Central  Google Scholar 

  30. Huang K-C, Wang M-L, Chen S-J, Kuo J-C, Wang W-J, Nguyen PNN, Wahlin KJ, Lu J-F, Tran AA, Shi M. Morphological and molecular defects in human three-dimensional retinal organoid model of X-linked juvenile retinoschisis. Stem Cell Rep. 2019;13:906–23.

    Article  CAS  Google Scholar 

  31. Parfitt DA, Lane A, Ramsden CM, Carr AJ, Munro PM, Jovanovic K, Schwarz N, Kanuga N, Muthiah MN, Hull S, et al. Identification and correction of mechanisms underlying inherited blindness in human iPSC-derived optic cups. Cell Stem Cell. 2016;18:769–81.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  32. Bennett J, Tanabe T, Sun D, Zeng Y, Kjeldbye H, Gouras P, Maguire AM. Photoreceptor cell rescue in retinal degeneration (rd) mice by in vivo gene therapy. Nat Med. 1996;2:649–54.

    Article  PubMed  CAS  Google Scholar 

  33. Jackson GR, Barber AJ. Visual dysfunction associated with diabetic retinopathy. Curr DiabRep. 2010;10:380–4.

    Google Scholar 

  34. Jackson GR, Owsley C, Curcio CA. Photoreceptor degeneration and dysfunction in aging and age-related maculopathy. Ageing Res Rev. 2002;1:381–96.

    Article  PubMed  Google Scholar 

  35. Huang X-F, Huang F, Wu K-C, Wu J, Chen J, Pang C-P, Lu F, Qu J, Jin Z-B. Genotype–phenotype correlation and mutation spectrum in a large cohort of patients with inherited retinal dystrophy revealed by next-generation sequencing. Genet Med. 2015;17:271–8.

    Article  PubMed  CAS  Google Scholar 

  36. Glöckle N, Kohl S, Mohr J, Scheurenbrand T, Sprecher A, Weisschuh N, Bernd A, Rudolph G, Schubach M, Poloschek C. Panel-based next generation sequencing as a reliable and efficient technique to detect mutations in unselected patients with retinal dystrophies. Eur J Hum Genet. 2014;22:99–104.

    Article  PubMed  Google Scholar 

  37. Bernardis I, Chiesi L, Tenedini E, Artuso L, Percesepe A, Artusi V, Simone ML, Manfredini R, Camparini M, Rinaldi C. Unravelling the complexity of inherited retinal dystrophies molecular testing: added value of targeted next-generation sequencing. Biomed Res Int. 2016;2016:14.

    Article  Google Scholar 

  38. Novak-Lauš K, Kukulj S, Zorić-Geber M, Bastaić O. Primary tapetoretinal dystrophies as the cause of blindness and impaired vision in the republic of Croatia. Acta Clin Croat. 2002;41:23–7.

    Google Scholar 

  39. Haim M. Epidemiology of retinitis pigmentosa in Denmark. Acta Ophthalmol Scand. 2002;80:1.

    Article  Google Scholar 

  40. Grøndahl J. Estimation of prognosis and prevalence of retinitis pigmentosa and Usher syndrome in Norway. Clin Genet. 1987;31:255–64.

    Article  PubMed  Google Scholar 

  41. Bunker CH, Berson EL, Bromley WC, Hayes RP, Roderick TH. Prevalence of retinitis pigmentosa in Maine. Am J Ophthalmol. 1984;97:357–65.

    Article  PubMed  CAS  Google Scholar 

  42. Broadgate S, Yu J, Downes SM, Halford S. Unravelling the genetics of inherited retinal dystrophies: past, present and future. Prog Retin Eye Res. 2017;59:53–96.

    Article  PubMed  CAS  Google Scholar 

  43. Wright AF, Chakarova CF, Abd El-Aziz MM, Bhattacharya SS. Photoreceptor degeneration: genetic and mechanistic dissection of a complex trait. Nat Rev Genet. 2010;11:273–84.

    Article  PubMed  CAS  Google Scholar 

  44. Nash BM, Wright DC, Grigg JR, Bennetts B, Jamieson RV. Retinal dystrophies, genomic applications in diagnosis and prospects for therapy. Transl Pediatr. 2015;4:139–63.

    PubMed  PubMed Central  Google Scholar 

  45. Iannaccone A. The genetics of hereditary retinopathies and optic neuropathies. Compr Ophthalmol Update. 2005;6:39–62.

    Google Scholar 

  46. Hartong DT, Berson EL, Dryja TP. Retinitis pigmentosa. Lancet. 2006;368:1795–809.

    Article  PubMed  CAS  Google Scholar 

  47. Daiger S, Rossiter B, Greenberg J, Christoffels A, Hide W. Data services and software for identifying genes and mutations causing retinal degeneration. Invest Ophthalmol Vis Sci. 1998;39:S295.

    Google Scholar 

  48. Bramall AN, Wright AF, Jacobson SG, McInnes RR. The genomic, biochemical, and cellular responses of the retina in inherited photoreceptor degenerations and prospects for the treatment of these disorders. Annu Rev Neurosci. 2010;33:441–72.

    Article  PubMed  CAS  Google Scholar 

  49. Erkilic N, Sanjurjo-Soriano C, Manes G, Dubois G, Hamel CP, Meunier I, Kalatzis V. Generation of a human iPSC line, INMi004-A, with a point mutation in CRX associated with autosomal dominant Leber congenital amaurosis. Stem Cell Res. 2019;38: 101476.

    Article  PubMed  CAS  Google Scholar 

  50. Diakatou M, Manes G, Bocquet B, Meunier I, Kalatzis V. Genome editing as a treatment for the most prevalent causative genes of autosomal dominant retinitis pigmentosa. Int J Mol Sci. 2019;20:2542.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  51. Ahn J, Chiang J, Gorin MB. Novel mutation in SLC4A7 gene causing autosomal recessive progressive rod-cone dystrophy. Ophthalmic Genet. 2020;41:386–9.

    Article  PubMed  CAS  Google Scholar 

  52. Vijayasarathy C, Takada Y, Zeng Y, Bush RA, Sieving PA. Retinoschisin is a peripheral membrane protein with affinity for anionic phospholipids and affected by divalent cations. Invest Ophthalmol Vis Sci. 2007;48:991–1000.

    Article  PubMed  Google Scholar 

  53. Sauer CG, Gehrig A, Warneke-Wittstock R, Marquardt A, Ewing CC, Gibson A, Lorenz B, Jurklies B, Weber BH. Positional cloning of the gene associated with X-linked juvenile retinoschisis. Nat Genet. 1997;17:164–70.

    Article  PubMed  CAS  Google Scholar 

  54. Hiriyanna KT, Bingham EL, Yashar BM, Ayyagari R, Fishman G, Small KW, Weinberg DV, Weleber RG, Lewis RA, Andreasson S. Novel mutations in XLRS1 causing retinoschisis, including first evidence of putative leader sequence change. Hum Mutat. 1999;14:423–7.

    Article  PubMed  CAS  Google Scholar 

  55. Ali MH, Vajzovic L. X-Linked Juvenile Retinoschisis. In: Toth CA, Ong SS, editors. Handbook of pediatric retinal OCT and the eye-brain connection. Philadelphia: Elsevier; 2020. p. 119–23.

    Chapter  Google Scholar 

  56. Dalkara D, Sahel J-A. Gene therapy for inherited retinal degenerations. CR Biol. 2014;337:185–92.

    Article  Google Scholar 

  57. Dryja T, Li T. Molecular genetics of retinitis pigmentosa. Hum Mol Genet. 1995;4:1739–43.

    Article  PubMed  CAS  Google Scholar 

  58. Arbabi A, Liu A, Ameri H. Gene therapy for inherited retinal degeneration. J Ocul Pharmacol Ther. 2019;35:79–97.

    Article  PubMed  CAS  Google Scholar 

  59. Lee JH, Wang J-H, Chen J, Li F, Edwards TL, Hewitt AW, Liu G-S. Gene therapy for visual loss: opportunities and concerns. Prog Retin Eye Res. 2019;68:31–53.

    Article  PubMed  CAS  Google Scholar 

  60. Ramlogan-Steel CA, Murali A, Andrzejewski S, Dhungel B, Steel JC, Layton CJ. Gene therapy and the adeno-associated virus in the treatment of genetic and acquired ophthalmic diseases in humans: trials, future directions and safety considerations. Clin Exp Ophthalmol. 2019;47:521–36.

    Article  PubMed  Google Scholar 

  61. Soofiyani SR, Baradaran B, Lotfipour F, Kazemi T, Mohammadnejad L. Gene therapy, early promises, subsequent problems, and recent breakthroughs. Adv Pharm Bull. 2013;3:249.

    Google Scholar 

  62. Trapani I, Auricchio A. Seeing the light after 25 years of retinal gene therapy. Trends Mol Med. 2018;24:669–81.

    Article  PubMed  Google Scholar 

  63. Wert KJ, Davis RJ, Sancho-Pelluz J, Nishina PM, Tsang SH. Gene therapy provides long-term visual function in a pre-clinical model of retinitis pigmentosa. Hum Mol Genet. 2013;22:558–67.

    Article  PubMed  CAS  Google Scholar 

  64. Cideciyan AV, Aleman TS, Boye SL, Schwartz SB, Kaushal S, Roman AJ, Pang JJ, Sumaroka A, Windsor EA, Wilson JM. 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.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  65. Maguire AM, Simonelli F, Pierce EA, Pugh EN Jr, Mingozzi F, Bennicelli J, Banfi S, Marshall KA, Testa F, Surace EM. Safety and efficacy of gene transfer for Leber’s congenital amaurosis. N Engl J Med. 2008;358:2240–8.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  66. Smalley E. First AAV gene therapy poised for landmark approval. Nat Biotech. 2017. https://doi.org/10.1038/nbt1117-998.

    Article  Google Scholar 

  67. Feuer WJ, Schiffman JC, Davis JL, Porciatti V, Gonzalez P, Koilkonda RD, Yuan H, Lalwani A, Lam BL, Guy J. Gene therapy for Leber hereditary optic neuropathy: initial results. Ophthalmology. 2016;123:558–70.

    Article  PubMed  Google Scholar 

  68. Guy J, Feuer WJ, Davis JL, Porciatti V, Gonzalez PJ, Koilkonda RD, Yuan H, Hauswirth WW, Lam BL. Gene therapy for Leber hereditary optic neuropathy: low-and medium-dose visual results. Ophthalmology. 2017;124:1621–34.

    Article  PubMed  Google Scholar 

  69. Fischer MD, Ochakovski GA, Beier B, Seitz IP, Vaheb Y, Kortuem C, Reichel FF, Kuehlewein L, Kahle NA, Peters T. Efficacy and safety of retinal gene therapy using adeno-associated virus vector for patients with choroideremia: a randomized clinical trial. JAMA Ophthalmol. 2019;137:1247–54.

    Article  PubMed  PubMed Central  Google Scholar 

  70. Lam BL, Davis JL, Gregori NZ, MacLaren RE, Girach A, Verriotto JD, Rodriguez B, Rosa PR, Zhang X, Feuer WJ. Choroideremia gene therapy phase 2 clinical trial: 24-month results. Am J Ophthalmol. 2019;197:65–73.

    Article  PubMed  CAS  Google Scholar 

  71. MacLaren RE, Groppe M, Barnard AR, Cottriall CL, Tolmachova T, Seymour L, Clark KR, During MJ, Cremers FP, Black GC. Retinal gene therapy in patients with choroideremia: initial findings from a phase 1/2 clinical trial. Lancet. 2014;383:1129–37.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  72. McClements ME, MacLaren RE. Gene therapy for retinal disease. Transl Res. 2013;161:241–54.

    Article  PubMed  CAS  Google Scholar 

  73. Russell S, Bennett J, Wellman J, Chung D, High K, Tillman A. Phase 3 trial update of voretigene neparvovec in biallelic RPE65-mediated inherited retinal disease. Am Acad Ophthalmol AAO. 2017;2017:11–4.

    Google Scholar 

  74. Russell S, Bennett J, Wellman JA, Chung DC, Yu Z-F, Tillman A, Wittes J, Pappas J, Elci O, McCague S. Efficacy and safety of voretigene neparvovec (AAV2-hRPE65v2) in patients with RPE65-mediated inherited retinal dystrophy: a randomised, controlled, open-label, phase 3 trial. Lancet. 2017;390:849–60.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  75. Vandenberghe LH, Bell P, Maguire AM, Cearley CN, Xiao R, Calcedo R, Wang L, Castle MJ, Maguire AC, Grant R. Dosage thresholds for AAV2 and AAV8 photoreceptor gene therapy in monkey. Sci Transl Med. 2011;3:88ra54-88ra54.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  76. Koerber JT, Klimczak R, Jang J-H, Dalkara D, Flannery JG, Schaffer DV. Molecular evolution of adeno-associated virus for enhanced glial gene delivery. Mol Ther. 2009;17:2088–95.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  77. Dyka FM, Molday LL, Chiodo VA, Molday RS, Hauswirth WW. Dual ABCA4-AAV vector treatment reduces pathogenic retinal A2E accumulation in a mouse model of autosomal recessive stargardt disease. Hum Gene Ther. 2019;30:1361–70.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  78. Dyka FM, Boye SL, Chiodo VA, Hauswirth WW, Boye SE. Dual adeno-associated virus vectors result in efficient in vitro and in vivo expression of an oversized gene, MYO7A. Hum Gene Ther Methods. 2014;25:166–77.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  79. Zeng Y, Takada Y, Kjellstrom S, Hiriyanna K, Tanikawa A, Wawrousek E, Smaoui N, Caruso R, Bush RA, Sieving PA. RS-1 gene delivery to an adult Rs1h knockout mouse model restores ERG b-wave with reversal of the electronegative waveform of X-linked retinoschisis. Invest Ophthalmol Vis Sci. 2004;45:3279–85.

    Article  PubMed  Google Scholar 

  80. Park T, Wu Z, Kjellstrom S, Zeng Y, Bush RA, Sieving P, Colosi P. Intravitreal delivery of AAV8 retinoschisin results in cell type-specific gene expression and retinal rescue in the Rs1-KO mouse. Gene Ther. 2009;16:916–26.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  81. Byrne LC, Öztürk BE, Lee T, Fortuny C, Visel M, Dalkara D, Schaffer DV, Flannery JG. Retinoschisin gene therapy in photoreceptors, Müller glia or all retinal cells in the Rs1h−/− mouse. Gene Ther. 2014;21:585–92.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  82. Sengillo JD, Justus S, Tsai YT, Cabral T, Tsang SH. Gene and cell-based therapies for inherited retinal disorders: an update. In: Tan WH, Bird LM, editors. American journal of medical genetics part c: seminars in medical genetics. Toronto: Wiley; 2016. p. 349–66.

    Google Scholar 

  83. Lewin AS, Rossmiller B, Mao H. Gene augmentation for adRP mutations in RHO. Cold Spring Harb Perspect Med. 2014;4: a017400.

    Article  PubMed  PubMed Central  Google Scholar 

  84. Davis JL, Gregori NZ, MacLaren RE, Lam BL. Surgical technique for subretinal gene therapy in humans with inherited retinal degeneration. Retina. 2019;39:S2–8.

    Article  PubMed  Google Scholar 

  85. Davis JL. The blunt end: surgical challenges of gene therapy for inherited retinal diseases. Am J Ophthalmol. 2018;196:1–3.

    Article  Google Scholar 

  86. Andrieu-Soler C, Bejjani R-A, de Bizemont T, Normand N, BenEzra D, Behar-Cohen F. Ocular gene therapy: a review of nonviral strategies. Mol Vis. 2006;12:1334–47.

    PubMed  CAS  Google Scholar 

  87. Han Z, Conley SM, Naash MI. AAV and compacted DNA nanoparticles for the treatment of retinal disorders: challenges and future prospects. Invest Ophthalmol Vis Sci. 2011;52:3051–9.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  88. Koirala A, Conley SM, Naash MI. A review of therapeutic prospects of non-viral gene therapy in the retinal pigment epithelium. Biomaterials. 2013;34:7158–67.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  89. Cai X, Conley S, Naash M. Nanoparticle applications in ocular gene therapy. Vision Res. 2008;48:319–24.

    Article  PubMed  CAS  Google Scholar 

  90. Han Z, Conley SM, Makkia RS, Cooper MJ, Naash MI. DNA nanoparticle-mediated ABCA4 delivery rescues Stargardt dystrophy in mice. J Clin Investig. 2012;122:3221–6.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  91. Dalkara D, Byrne LC, Klimczak RR, Visel M, Yin L, Merigan WH, Flannery JG, Schaffer DV. In vivo—directed evolution of a new adeno-associated virus for therapeutic outer retinal gene delivery from the vitreous. Sci Transl Med. 2013;5:189ra176-189ra176.

    Article  Google Scholar 

  92. Petrs-Silva H, Dinculescu A, Li Q, Min S-H, Chiodo V, Pang J-J, Zhong L, Zolotukhin S, Srivastava A, Lewin AS. High-efficiency transduction of the mouse retina by tyrosine-mutant AAV serotype vectors. Mol Ther. 2009;17:463–71.

    Article  PubMed  CAS  Google Scholar 

  93. Leroy B, Pennesi M, Ohnsman C. Brave new world: gene therapy for inherited retinal disease. In: Leroy B, Pennesi M, Ohnsman C, editors. American academy of ophthalmology. San Francisco: EyeNet; 2018. p. 1–16.

    Google Scholar 

  94. Lipinski DM, Thake M, MacLaren RE. Clinical applications of retinal gene therapy. Prog Retin Eye Res. 2013;32:22–47.

    Article  PubMed  CAS  Google Scholar 

  95. Pennesi ME, Birch DG, Duncan JL, Bennett J, Girach A. Choroideremia: retinal degeneration with an unmet need. Retina. 2019;39:2059–69.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  96. Liu M, Rehman S, Tang X, Gu K, Fan Q, Chen D, Ma W. Methodologies for improving HDR efficiency. Front Genet. 2019;9:691.

    Article  PubMed  PubMed Central  Google Scholar 

  97. Yang H, Wang H, Shivalila CS, Cheng AW, Shi L, Jaenisch R. One-step generation of mice carrying reporter and conditional alleles by CRISPR/Cas-mediated genome engineering. Cell. 2013;154:1370–9.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  98. Wang H, Yang H, Shivalila CS, Dawlaty MM, Cheng AW, Zhang F, Jaenisch R. One-step generation of mice carrying mutations in multiple genes by CRISPR/Cas-mediated genome engineering. Cell. 2013;153:910–8.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  99. Ran F, Cong L, Yan WX, Scott DA, Gootenberg JS, Kriz AJ, Zetsche B, Shalem O, Wu X, Makarova KS. In vivo genome editing using Staphylococcus aureus Cas9. Nature. 2015;520:186–91.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  100. Le Rhun A, Escalera-Maurer A, Bratovic M, Charpentier E. CRISPR-Cas in Streptococcus pyogenes. RNA Biol. 2019;16:380–9.

    Article  PubMed  PubMed Central  Google Scholar 

  101. Aghaizu ND, Kruczek K, Gonzalez-Cordero A, Ali RR, Pearson RA. Pluripotent stem cells and their utility in treating photoreceptor degenerations. Prog Brain Res. 2017;231:191–223.

    Article  PubMed  Google Scholar 

  102. Hazim RA, Karumbayaram S, Jiang M, Dimashkie A, Lopes VS, Li D, Burgess BL, Vijayaraj P, Alva-Ornelas JA, Zack JA. Differentiation of RPE cells from integration-free iPS cells and their cell biological characterization. Stem Cell Res Ther. 2017;8:1–17.

    Article  Google Scholar 

  103. Jones MK, Lu B, Girman S, Wang S. Cell-based therapeutic strategies for replacement and preservation in retinal degenerative diseases. Prog Retin Eye Res. 2017;58:1–27.

    Article  PubMed  PubMed Central  Google Scholar 

  104. Reichman S, Terray A, Slembrouck A, Nanteau C, Orieux G, Habeler W, Nandrot EF, Sahel J-A, Monville C, Goureau O. From confluent human iPS cells to self-forming neural retina and retinal pigmented epithelium. Proc Natl Acad Sci. 2014;111:8518–23.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  105. Nami F, Basiri M, Satarian L, Curtiss C, Baharvand H, Verfaillie C. Strategies for in vivo genome editing in nondividing cells. Trends Biotechnol. 2018;36:770–86.

    Article  PubMed  CAS  Google Scholar 

  106. Yamamoto Y, Bliss J, Gerbi SA. Whole organism genome editing: Targeted large DNA insertion via ObLiGaRe nonhomologous end-joining in vivo capture. G3. 2015;5:1843–7.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  107. Ishizu T, Higo S, Masumura Y, Kohama Y, Shiba M, Higo T, Shibamoto M, Nakagawa A, Morimoto S, Takashima S. Targeted genome replacement via homology-directed repair in non-dividing cardiomyocytes. Sci Rep. 2017;7:1–11.

    Article  CAS  Google Scholar 

  108. Suzuki K, Tsunekawa Y, Hernandez-Benitez R, Wu J, Zhu J, Kim EJ, Hatanaka F, Yamamoto M, Araoka T, Li Z. In vivo genome editing via CRISPR/Cas9 mediated homology-independent targeted integration. Nature. 2016;540:144–9.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  109. Waldron D. In vivo gene editing in non-dividing cells. Nat Rev Genet. 2017;18:1–1.

    Article  PubMed  CAS  Google Scholar 

  110. Sakuma T, Nakade S, Sakane Y, Suzuki KT, Yamamoto T. MMEJ-assisted gene knock-in using TALENs and CRISPR-Cas9 with the PITCh systems. Nat Protoc. 2016;11:118–33.

    Article  PubMed  CAS  Google Scholar 

  111. Komor AC, Kim YB, Packer MS, Zuris JA, Liu DR. Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage. Nature. 2016;533:420–4.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  112. Anzalone AV, Randolph PB, Davis JR, Sousa AA, Koblan LW, Levy JM, Chen PJ, Wilson C, Newby GA, Raguram A, Liu DR. Search-and-replace genome editing without double-strand breaks or donor DNA. Nature. 2019;576:149–57.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  113. Dominguez AA, Lim WA, Qi LS. Beyond editing: repurposing CRISPR-Cas9 for precision genome regulation and interrogation. Nat Rev Mol Cell Biol. 2016;17:5–15.

    Article  PubMed  CAS  Google Scholar 

  114. Auer TO, Duroure K, De Cian A, Concordet JP, Del Bene F. Highly efficient CRISPR/Cas9-mediated knock-in in zebrafish by homology-independent DNA repair. Genome Res. 2014;24:142–53.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  115. Suzuki K, Izpisua Belmonte JC. In vivo genome editing via the HITI method as a tool for gene therapy. J Hum Genet. 2018;63:157–64.

    Article  PubMed  CAS  Google Scholar 

  116. He X, Tan C, Wang F, Wang Y, Zhou R, Cui D, You W, Zhao H, Ren J, Feng B. Knock-in of large reporter genes in human cells via CRISPR/Cas9-induced homology-dependent and independent DNA repair. Nucleic Acids Res. 2016;44: e85.

    Article  PubMed  PubMed Central  Google Scholar 

  117. Papapetrou EP, Schambach A. Gene insertion into genomic safe harbors for human gene therapy. Mol Ther. 2016;24:678–84.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  118. Kampmann M. CRISPRi and CRISPRa screens in mammalian cells for precision biology and medicine. ACS Chem Biol. 2018;13:406–16.

    Article  PubMed  CAS  Google Scholar 

  119. Chou SJ, Yang P, Ban Q, Yang YP, Wang ML, Chien CS, Chen SJ, Sun N, Zhu Y, Liu H, et al. Dual supramolecular nanoparticle vectors enable CRISPR/Cas9-mediated knockin of Retinoschisin 1 Gene-A potential nonviral therapeutic solution for X-linked Juvenile Retinoschisis. Adv Sci (Weinh). 2020;7:1903432.

    Article  PubMed  CAS  Google Scholar 

  120. Sikkink SK, Biswas S, Parry NR, Stanga PE, Trump D. X-linked retinoschisis: an update. J Med Genet. 2007;44:225–32.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  121. Tantri A, Vrabec TR, Cu-Unjieng A, Frost A, Annesley WH Jr, Donoso LA. X-linked retinoschisis: a clinical and molecular genetic review. Surv Ophthalmol. 2004;49:214–30.

    Article  PubMed  Google Scholar 

  122. Bakondi B, Lv W, Lu B, Jones MK, Tsai Y, Kim KJ, Levy R, Akhtar AA, Breunig JJ, Svendsen CN. In vivo CRISPR/Cas9 gene editing corrects retinal dystrophy in the S334ter-3 rat model of autosomal dominant retinitis pigmentosa. Mol Ther. 2016;24:556–63.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  123. Burnight ER, Gupta M, Wiley LA, Anfinson KR, Tran A, Triboulet R, Hoffmann JM, Klaahsen DL, Andorf JL, Jiao C. Using CRISPR-Cas9 to generate gene-corrected autologous iPSCs for the treatment of inherited retinal degeneration. Mol Ther. 2017;25:1999–2013.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  124. Vagni P, Perlini LE, Chenais N, Marchetti T, Parrini M, Contestabile A, Cancedda L, Ghezzi D. Gene editing preserves visual functions in a mouse model of retinal degeneration. Front Neurosci. 2019;13:945.

    Article  PubMed  PubMed Central  Google Scholar 

  125. Yang X, Bayat V, DiDonato N, Zhao Y, Zarnegar B, Siprashvili Z, Lopez-Pajares V, Sun T, Tao S, Li C. Genetic and genomic studies of pathogenic EXOSC2 mutations in the newly described disease SHRF implicate the autophagy pathway in disease pathogenesis. Hum Mol Genet. 2020;29:541–53.

    Article  PubMed  CAS  Google Scholar 

  126. Philippidis A. One small dose, one giant leap for CRISPR gene editing. Hum Gene Ther. 2020;31:402–4.

    Article  PubMed  CAS  Google Scholar 

  127. Suh S, Choi EH, Leinonen H, Foik AT, Newby GA, Yeh WH, Dong Z, Kiser PD, Lyon DC, Liu DR, Palczewski K. Restoration of visual function in adult mice with an inherited retinal disease via adenine base editing. Nat Biomed Eng. 2021;5:169–78.

    Article  PubMed  CAS  Google Scholar 

  128. Liu Y, Li X, He S, Huang S, Li C, Chen Y, Liu Z, Huang X, Wang X. Efficient generation of mouse models with the prime editing system. Cell Discov. 2020;6:27.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  129. Liu P, Liang SQ, Zheng C, Mintzer E, Zhao YG, Ponnienselvan K, Mir A, Sontheimer EJ, Gao G, Flotte TR, et al. Improved prime editors enable pathogenic allele correction and cancer modelling in adult mice. Nat Commun. 2021;12:2121.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  130. Peddle CF, Fry LE, McClements ME, MacLaren RE. CRISPR interference-potential application in retinal disease. Int J Mol Sci. 2020;21:2329.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  131. Keeler AM, Flotte TR. Recombinant adeno-associated virus gene therapy in light of Luxturna (and Zolgensma and Glybera): where are we, and how did we get here? Annu Rev Virol. 2019;6:601–21.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  132. Verdera HC, Kuranda K, Mingozzi F. AAV vector immunogenicity in humans: a long journey to successful gene transfer. Mol Ther. 2020;28:723–46.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  133. Li C, Samulski RJ. Engineering adeno-associated virus vectors for gene therapy. Nat Rev Genet. 2020;21:255–72.

    Article  PubMed  CAS  Google Scholar 

  134. Patel A, Zhao J, Duan D, Lai Y. Design of AAV vectors for delivery of large or multiple transgenes. Methods Mol Biol. 2019;1950:19–33.

    Article  PubMed  CAS  Google Scholar 

  135. Duan D, Yue Y, Engelhardt JF. Expanding AAV packaging capacity with trans-splicing or overlapping vectors: a quantitative comparison. Mol Ther. 2001;4:383–91.

    Article  PubMed  CAS  Google Scholar 

  136. Amreddy N, Babu A, Muralidharan R, Panneerselvam J, Srivastava A, Ahmed R, Mehta M, Munshi A, Ramesh R. Recent advances in nanoparticle-based cancer drug and gene delivery. Adv Cancer Res. 2018;137:115–70.

    Article  PubMed  CAS  Google Scholar 

  137. Kim HS, Sun X, Lee J-H, Kim H-W, Fu X, Leong KW. Advanced drug delivery systems and artificial skin grafts for skin wound healing. Adv Drug Deliv Rev. 2019;146:209–39.

    Article  PubMed  CAS  Google Scholar 

  138. Kong F-Y, Zhang J-W, Li R-F, Wang Z-X, Wang W-J, Wang W. Unique roles of gold nanoparticles in drug delivery, targeting and imaging applications. Molecules. 2017;22:1445.

    Article  PubMed  PubMed Central  Google Scholar 

  139. Matoba T, Koga JI, Nakano K, Egashira K, Tsutsui H. Nanoparticle-mediated drug delivery system for atherosclerotic cardiovascular disease. J Cardiol. 2017;70:206–11.

    Article  PubMed  Google Scholar 

  140. Mirza Z, Karim S. Nanoparticles-based drug delivery and gene therapy for breast cancer: recent advancements and future challenges. Semin Cancer Biol. 2019. https://doi.org/10.1016/j.semcancer.2019.10.020.

    Article  PubMed  Google Scholar 

  141. Zahin N, Anwar R, Tewari D, Kabir MT, Sajid A, Mathew B, Uddin MS, Aleya L, Abdel-Daim MM. Nanoparticles and its biomedical applications in health and diseases: special focus on drug delivery. Environ Sci Pollut Res. 2019. https://doi.org/10.1007/s11356-019-05211-0.

    Article  Google Scholar 

  142. Blanco E, Shen H, Ferrari M. Principles of nanoparticle design for overcoming biological barriers to drug delivery. Nat Biotechnol. 2015;33:941.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  143. Juillerat-Jeanneret L. The targeted delivery of cancer drugs across the blood–brain barrier: chemical modifications of drugs or drug-nanoparticles? Drug Discov Today. 2008;13:1099–106.

    Article  PubMed  CAS  Google Scholar 

  144. Kievit FM, Zhang M. Cancer therapy: cancer nanotheranostics: improving imaging and therapy by targeted delivery across biological barriers (Adv. Mater. 36/2011). Adv Mater. 2011;23:H209–H209.

    Article  Google Scholar 

  145. Steichen SD, Caldorera-Moore M, Peppas NA. A review of current nanoparticle and targeting moieties for the delivery of cancer therapeutics. Eur J Pharm Sci. 2013;48:416–27.

    Article  PubMed  CAS  Google Scholar 

  146. Givens BE, Naguib YW, Geary SM, Devor EJ, Salem AK. Nanoparticle-based delivery of CRISPR/Cas9 genome-editing therapeutics. AAPS J. 2018;20:108.

    Article  PubMed  Google Scholar 

  147. Nakade S, Yamamoto T, Sakuma T. Cas9, Cpf1 and C2c1/2/3-what’s next? Bioengineered. 2017;8:265–73.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  148. Xu Y, Liu R, Dai Z. Key considerations in designing CRISPR/Cas9-carrying nanoparticles for therapeutic genome editing. Nanoscale. 2020;12:21001–14.

    Article  PubMed  CAS  Google Scholar 

  149. Zhang S, Shen J, Li D, Cheng Y. Strategies in the delivery of Cas9 ribonucleoprotein for CRISPR/Cas9 genome editing. Theranostics. 2021;11:614.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  150. Huang X, Chau Y. Intravitreal nanoparticles for retinal delivery. Drug Discov Today. 2019;24:1510–23.

    Article  PubMed  CAS  Google Scholar 

  151. Jackson TL, Antcliff RJ, Hillenkamp J, Marshall J. Human retinal molecular weight exclusion limit and estimate of species variation. Invest Ophthalmol Vis Sci. 2003;44:2141–6.

    Article  PubMed  Google Scholar 

  152. Sebag J. Anatomy and pathology of the vitreo-retinal interface. Eye (Lond). 1992;6(Pt 6):541–52.

    Article  PubMed  Google Scholar 

  153. Tavakoli S, Peynshaert K, Lajunen T, Devoldere J, Del Amo EM, Ruponen M, De Smedt SC, Remaut K, Urtti A. Ocular barriers to retinal delivery of intravitreal liposomes: impact of vitreoretinal interface. J Control Release. 2020;328:952–61.

    Article  PubMed  CAS  Google Scholar 

  154. Huang X, Chau Y. Investigating impacts of surface charge on intraocular distribution of intravitreal lipid nanoparticles. Exp Eye Res. 2019;186: 107711.

    Article  PubMed  CAS  Google Scholar 

  155. Altınoglu S, Wang M, Xu Q. Combinatorial library strategies for synthesis of cationic lipid-like nanoparticles and their potential medical applications. Nanomedicine. 2015;10:643–57.

    Article  PubMed  Google Scholar 

  156. Freitag F, Wagner E. Optimizing synthetic nucleic acid and protein nanocarriers: the chemical evolution approach. Adv Drug Deliv Rev. 2021;168:30–54.

    Article  PubMed  CAS  Google Scholar 

  157. Fu A, Tang R, Hardie J, Farkas ME, Rotello VM. Promises and pitfalls of intracellular delivery of proteins. Bioconjug Chem. 2014;25:1602–8.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  158. Rahimi H, Salehiabar M, Charmi J, Barsbay M, Ghaffarlou M, Razlighi MR, Davaran S, Khalilov R, Sugiyama M, Nosrati H. Harnessing nanoparticles for the efficient delivery of the CRISPR/Cas9 system. Nano Today. 2020;34: 100895.

    Article  CAS  Google Scholar 

  159. Tang H, Zhao X, Jiang X. Synthetic multi-layer nanoparticles for CRISPR-Cas9 genome editing. Adv Drug Deliv Rev. 2020. https://doi.org/10.1016/j.addr.2020.03.001.

    Article  PubMed  Google Scholar 

  160. Wan T, Niu D, Wu C, Xu F-J, Church G, Ping Y. Material solutions for delivery of CRISPR/Cas-based genome editing tools: current status and future outlook. Mater Today. 2019;26:40–66.

    Article  CAS  Google Scholar 

  161. Finn JD, Smith AR, Patel MC, Shaw L, Youniss MR, van Heteren J, Dirstine T, Ciullo C, Lescarbeau R, Seitzer J. A single administration of CRISPR/Cas9 lipid nanoparticles achieves robust and persistent in vivo genome editing. Cell Rep. 2018;22:2227–35.

    Article  PubMed  CAS  Google Scholar 

  162. Liu J, Chang J, Jiang Y, Meng X, Sun T, Mao L, Xu Q, Wang M. Fast and efficient CRISPR/Cas9 genome editing in vivo enabled by bioreducible lipid and messenger RNA nanoparticles. Adv Mater. 2019;31:1902575.

    Article  Google Scholar 

  163. Miller JB, Zhang S, Kos P, Xiong H, Zhou K, Perelman SS, Zhu H, Siegwart DJ. Non-viral CRISPR/Cas gene editing in vitro and in vivo enabled by synthetic nanoparticle co-delivery of Cas9 mRNA and sgRNA. Angew Chem Int Ed. 2017;56:1059–63.

    Article  CAS  Google Scholar 

  164. Zhang L, Wang P, Feng Q, Wang N, Chen Z, Huang Y, Zheng W, Jiang X. Lipid nanoparticle-mediated efficient delivery of CRISPR/Cas9 for tumor therapy. NPG Asia Mater. 2017;9:e441–e441.

    Article  CAS  Google Scholar 

  165. Zhang X, Li B, Luo X, Zhao W, Jiang J, Zhang C, Gao M, Chen X, Dong Y. Biodegradable amino-ester nanomaterials for Cas9 mRNA delivery in vitro and in vivo. ACS Appl Mater Interfaces. 2017;9:25481–7.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  166. Ruponen M, Yla-Herttuala S, Urtti A. Interactions of polymeric and liposomal gene delivery systems with extracellular glycosaminoglycans: physicochemical and transfection studies. Biochim Biophys Acta. 1999;1415:331–41.

    Article  PubMed  CAS  Google Scholar 

  167. Anderson DG, Akinc A, Hossain N, Langer R. Structure/property studies of polymeric gene delivery using a library of poly(beta-amino esters). Mol Ther. 2005;11:426–34.

    Article  PubMed  CAS  Google Scholar 

  168. Liang C, Li F, Wang L, Zhang Z-K, Wang C, He B, Li J, Chen Z, Shaikh AB, Liu J. Tumor cell-targeted delivery of CRISPR/Cas9 by aptamer-functionalized lipopolymer for therapeutic genome editing of VEGFA in osteosarcoma. Biomaterials. 2017;147:68–85.

    Article  PubMed  CAS  Google Scholar 

  169. Salameh JW, Zhou L, Ward SM, Santa Chalarca CF, Emrick T, Figueiredo ML. Polymer-mediated gene therapy: recent advances and merging of delivery techniques. Wiley Interdiscip Rev Nanomed Nanobiotechnol. 2020;12: e1598.

    Article  PubMed  Google Scholar 

  170. Wan T, Chen Y, Pan Q, Xu X, Kang Y, Gao X, Huang F, Wu C, Ping Y. Genome editing of mutant KRAS through supramolecular polymer-mediated delivery of Cas9 ribonucleoprotein for colorectal cancer therapy. J Control Release. 2020;322:236–47.

    Article  PubMed  CAS  Google Scholar 

  171. Zhang Z, Wan T, Chen Y, Chen Y, Sun H, Cao T, Songyang Z, Tang G, Wu C, Ping Y. Cationic polymer-mediated CRISPR/Cas9 plasmid delivery for genome editing. Macromol Rapid Commun. 2019;40:1800068.

    Article  Google Scholar 

  172. Wolfert MA, Dash PR, Nazarova O, Oupicky D, Seymour LW, Smart S, Strohalm J, Ulbrich K. Polyelectrolyte vectors for gene delivery: influence of cationic polymer on biophysical properties of complexes formed with DNA. Bioconjug Chem. 1999;10:993–1004.

    Article  PubMed  CAS  Google Scholar 

  173. D’Souza AA, Shegokar R. Polyethylene glycol (PEG): a versatile polymer for pharmaceutical applications. Expert Opin Drug Deliv. 2016;13:1257–75.

    Article  PubMed  CAS  Google Scholar 

  174. Kim YH, Park JH, Lee M, Kim YH, Park TG, Kim SW. Polyethylenimine with acid-labile linkages as a biodegradable gene carrier. J Control Release. 2005;103:209–19.

    Article  PubMed  CAS  Google Scholar 

  175. Sadekar S, Ghandehari H. Transepithelial transport and toxicity of PAMAM dendrimers: implications for oral drug delivery. Adv Drug Deliv Rev. 2012;64:571–88.

    Article  PubMed  CAS  Google Scholar 

  176. Lv H, Zhang S, Wang B, Cui S, Yan J. Toxicity of cationic lipids and cationic polymers in gene delivery. J Control Release. 2006;114:100–9.

    Article  PubMed  CAS  Google Scholar 

  177. Chen K, Jiang S, Hong Y, Li Z, Wu Y-L, Wu C. Cationic polymeric nanoformulation: recent advances in material design for CRISPR/Cas9 gene therapy. Prog Nat Sci. 2019;29:617–27.

    Article  CAS  Google Scholar 

  178. Zhang H, Bahamondez-Canas TF, Zhang Y, Leal J, Smyth HD. PEGylated chitosan for nonviral aerosol and mucosal delivery of the CRISPR/Cas9 system in vitro. Mol Pharm. 2018;15:4814–26.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  179. Li L, He Z-Y, Wei X-W, Gao G-P, Wei Y-Q. Challenges in CRISPR/CAS9 delivery: potential roles of nonviral vectors. Hum Gene Ther. 2015;26:452–62.

    Article  PubMed  CAS  Google Scholar 

  180. Abedi-Gaballu F, Dehghan G, Ghaffari M, Yekta R, Abbaspour-Ravasjani S, Baradaran B, Dolatabadi JEN, Hamblin MR. PAMAM dendrimers as efficient drug and gene delivery nanosystems for cancer therapy. Appl Mater Today. 2018;12:177–90.

    Article  PubMed  PubMed Central  Google Scholar 

  181. Avila-Salas FN, González RI, Ríos PL, Araya-Durán I, Camarada MB. Effect of the generation of PAMAM dendrimers on the stabilization of gold nanoparticles. J Chem Inform Model. 2020;60:2966–76.

    Article  CAS  Google Scholar 

  182. Islam MT, Shi X, Balogh L, Baker JR. HPLC separation of different generations of poly (amidoamine) dendrimers modified with various terminal groups. Anal Chem. 2005;77:2063–70.

    Article  PubMed  CAS  Google Scholar 

  183. Kurbatov AO, Balabaev NK, Mazo MA, Kramarenko EY. Effects of generation number, spacer length and temperature on the structure and intramolecular dynamics of siloxane dendrimer melts: molecular dynamics simulations. Soft Matter. 2020;16:3792–805.

    Article  PubMed  CAS  Google Scholar 

  184. Maiti PK, Çaǧın T, Wang G, Goddard WA. Structure of PAMAM dendrimers: generations 1 through 11. Macromolecules. 2004;37:6236–54.

    Article  CAS  Google Scholar 

  185. Pavan GM, Albertazzi L, Danani A. Ability to adapt: different generations of PAMAM dendrimers show different behaviors in binding siRNA. J Phys Chem B. 2010;114:2667–75.

    Article  PubMed  CAS  Google Scholar 

  186. Sebby KB, Walter ED, Usselman RJ, Cloninger MJ, Singel DJ. End-group distributions of multiple generations of spin-labeled PAMAM dendrimers. J Phys Chem B. 2011;115:4613–20.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  187. Vinicius R, Ara D, Santos S, Ferreira EI, Giarolla J. New advances in general biomedical applications of PAMAM dendrimer. Molecules. 2018;23:2849.

    Article  Google Scholar 

  188. Thanh VM, Nguyen TH, Tran TV, Ngoc UT, Ho MN, Nguyen TT, Chau YN, Tran NQ, Nguyen CK, Nguyen DH. Low systemic toxicity nanocarriers fabricated from heparin-mPEG and PAMAM dendrimers for controlled drug release. Mater Sci Eng C. 2018;82:291–8.

    Article  CAS  Google Scholar 

  189. Yavuz B, Bozdağ Pehlivan S, Sümer Bolu B, Nomak Sanyal R, Vural İ, Ünlü N. Dexamethasone-PAMAM dendrimer conjugates for retinal delivery: preparation, characterization and in vivo evaluation. J Pharm Pharmacol. 2016;68:1010–20.

    Article  PubMed  CAS  Google Scholar 

  190. Yavuz B, Pehlivan SB, Vural İ, Ünlü N. In vitro/in vivo evaluation of dexamethasone-PAMAM dendrimer complexes for retinal drug delivery. J Pharm Sci. 2015;104:3814–23.

    Article  PubMed  CAS  Google Scholar 

  191. Kretzmann JA, Ho D, Evans CW, Plani-Lam JH, Garcia-Bloj B, Mohamed AE, O’Mara ML, Ford E, Tan DE, Lister R. Synthetically controlling dendrimer flexibility improves delivery of large plasmid DNA. Chem Sci. 2017;8:2923–30.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  192. Liu C, Wan T, Wang H, Zhang S, Ping Y, Cheng Y. A boronic acid-rich dendrimer with robust and unprecedented efficiency for cytosolic protein delivery and CRISPR-Cas9 gene editing. Sci Adv. 2019;5:eaaw8922.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  193. Wei S, Shao X, Liu Y, Xiong B, Cui P, Liu Z, Li Q. Genome editing of PD-L1 mediated by nucleobase-modified polyamidoamine for cancer immunotherapy. J Mater Chem B. 2022;10:1291–300.

    Article  PubMed  CAS  Google Scholar 

  194. Gref R, Lück M, Quellec P, Marchand M, Dellacherie E, Harnisch S, Blunk T, Müller R. ‘Stealth’corona-core nanoparticles surface modified by polyethylene glycol (PEG): influences of the corona (PEG chain length and surface density) and of the core composition on phagocytic uptake and plasma protein adsorption. Colloids Surf B. 2000;18:301–13.

    Article  CAS  Google Scholar 

  195. Nunes R, Araújo F, Tavares J, Sarmento B. Surface modification with polyethylene glycol enhances colorectal distribution and retention of nanoparticles. Eur J Pharm Biopharm. 2018. https://doi.org/10.1016/j.ejpb.2018.06.029.

    Article  PubMed  Google Scholar 

  196. Pelaz B, del Pino P, Maffre P, Hartmann R, Gallego M, Rivera-Fernandez S, de la Fuente JM, Nienhaus GU, Parak WJ. Surface functionalization of nanoparticles with polyethylene glycol: effects on protein adsorption and cellular uptake. ACS Nano. 2015;9:6996–7008.

    Article  PubMed  CAS  Google Scholar 

  197. Ruiz A, Hernandez Y, Cabal C, Gonzalez E, Veintemillas-Verdaguer S, Martinez E, Morales M. Biodistribution and pharmacokinetics of uniform magnetite nanoparticles chemically modified with polyethylene glycol. Nanoscale. 2013;5:11400–8.

    Article  PubMed  CAS  Google Scholar 

  198. Chen CL, Rosi NL. Peptide-based methods for the preparation of nanostructured inorganic materials. Angew Chem Int Ed. 2010;49:1924–42.

    Article  CAS  Google Scholar 

  199. Huang H, Li J, Liao L, Li J, Wu L, Dong C, Lai P, Liu D. Poly (l-glutamic acid)-based star-block copolymers as pH-responsive nanocarriers for cationic drugs. Eur Polymer J. 2012;48:696–704.

    Article  CAS  Google Scholar 

  200. Li Z, Chen Q, Qi Y, Liu Z, Hao T, Sun X, Qiao M, Ma X, Xu T, Zhao X. Rational design of multifunctional polymeric nanoparticles based on poly (L-histidine) and d-α-Vitamin E Succinate for reversing tumor multidrug resistance. Biomacromol. 2018;19:2595–609.

    Article  CAS  Google Scholar 

  201. Liu B, Gao GH, Liu P, Yi HQ, Wei W, Ge ZC, Cai LT. A tunable pH-responsive nanomaterials for cancer delivery. Adv Mater Res. 2013;750–752:1476–9.

    Article  Google Scholar 

  202. Lv S, Tang Z, Li M, Lin J, Song W, Liu H, Huang Y, Zhang Y, Chen X. Co-delivery of doxorubicin and paclitaxel by PEG-polypeptide nanovehicle for the treatment of non-small cell lung cancer. Biomaterials. 2014;35:6118–29.

    Article  PubMed  CAS  Google Scholar 

  203. Shi C, He Y, Feng X, Fu D. ε-Polylysine and next-generation dendrigraft poly-l-lysine: chemistry, activity, and applications in biopharmaceuticals. J Biomater Sci Polym Ed. 2015;26:1343–56.

    Article  PubMed  CAS  Google Scholar 

  204. Yi H, Liu P, Sheng N, Gong P, Ma Y, Cai L. In situ crosslinked smart polypeptide nanoparticles for multistage responsive tumor-targeted drug delivery. Nanoscale. 2016;8:5985–95.

    Article  PubMed  CAS  Google Scholar 

  205. Zhang R, Zheng N, Song Z, Yin L, Cheng J. The effect of side-chain functionality and hydrophobicity on the gene delivery capabilities of cationic helical polypeptides. Biomaterials. 2014;35:3443–54.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  206. Zhou H, Lv S, Zhang D, Deng M, Zhang X, Tang Z, Chen X. A polypeptide based podophyllotoxin conjugate for the treatment of multi drug resistant breast cancer with enhanced efficiency and minimal toxicity. Acta Biomater. 2018;73:388–99.

    Article  PubMed  CAS  Google Scholar 

  207. Yin L, Song Z, Qu Q, Kim KH, Zheng N, Yao C, Chaudhury I, Tang H, Gabrielson NP, Uckun FM, Cheng J. Supramolecular self-assembled nanoparticles mediate oral delivery of therapeutic TNF-α siRNA against systemic inflammation. Angew Chem Int Ed Engl. 2013;52:5757–61.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  208. Wang HX, Song Z, Lao YH, Xu X, Gong J, Cheng D, Chakraborty S, Park JS, Li M, Huang D, et al. Nonviral gene editing via CRISPR/Cas9 delivery by membrane-disruptive and endosomolytic helical polypeptide. Proc Natl Acad Sci USA. 2018;115:4903–8.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  209. Zakeri A, Kouhbanani MAJ, Beheshtkhoo N, Beigi V, Mousavi SM, Hashemi SAR, Karimi Zade A, Amani AM, Savardashtaki A, Mirzaei E, et al. Polyethylenimine-based nanocarriers in co-delivery of drug and gene: a developing horizon. Nano Rev Exp. 2018;9:1488497.

    Article  PubMed  PubMed Central  Google Scholar 

  210. Pishavar E, Shafiei M, Mehri S, Ramezani M, Abnous K. The effects of polyethylenimine/DNA nanoparticle on transcript levels of apoptosis-related genes. Drug Chem Toxicol. 2017;40:406–9.

    Article  PubMed  CAS  Google Scholar 

  211. Bahadur KR, Uludağ H. PEI and its derivatives for gene therapy. In: Narain R, editor. Polymers and nanomaterials for gene therapy. Amsterdam: Elsevier; 2016. p. 29–54.

    Chapter  Google Scholar 

  212. Ryu N, Kim MA, Park D, Lee B, Kim YR, Kim KH, Baek JI, Kim WJ, Lee KY, Kim UK. Effective PEI-mediated delivery of CRISPR-Cas9 complex for targeted gene therapy. Nanomed Nanotechnol Biol Med. 2018;14:2095–102.

    Article  CAS  Google Scholar 

  213. Wu Y, Wang W, Chen Y, Huang K, Shuai X, Chen Q, Li X, Lian G. The investigation of polymer-siRNA nanoparticle for gene therapy of gastric cancer in vitro. Int J Nanomed. 2010;5:129.

    Article  CAS  Google Scholar 

  214. Zuckermann M, Hovestadt V, Knobbe-Thomsen CB, Zapatka M, Northcott PA, Schramm K, Belic J, Jones DT, Tschida B, Moriarity B. Somatic CRISPR/Cas9-mediated tumour suppressor disruption enables versatile brain tumour modelling. Nat Commun. 2015;6:1–9.

    Article  Google Scholar 

  215. Thaker PH, Brady WE, Lankes HA, Odunsi K, Bradley WH, Moore KN, Muller CY, Anwer K, Schilder RJ, Alvarez RD, Fracasso PM. A phase I trial of intraperitoneal GEN-1, an IL-12 plasmid formulated with PEG-PEI-cholesterol lipopolymer, administered with pegylated liposomal doxorubicin in patients with recurrent or persistent epithelial ovarian, fallopian tube or primary peritoneal cancers: an NRG Oncology/Gynecologic Oncology Group study. Gynecol Oncol. 2017;147:283–90.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  216. Liao H-W, Yau K-W. In vivo gene delivery in the retina using polyethylenimine. Biotechniques. 2007;42:285–8.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  217. Mendelsohn AR, Larrick JW. Preclinical reversal of atherosclerosis by FDA-approved compound that transforms cholesterol into an anti-inflammatory “prodrug.” Rejuvenation Res. 2016;19:252–5.

    Article  PubMed  CAS  Google Scholar 

  218. Lai WF. Cyclodextrins in non-viral gene delivery. Biomaterials. 2014;35:401–11.

    Article  PubMed  CAS  Google Scholar 

  219. Ping Y, Liu C, Zhang Z, Liu KL, Chen J, Li J. Chitosan-graft-(PEI-β-cyclodextrin) copolymers and their supramolecular PEGylation for DNA and siRNA delivery. Biomaterials. 2011;32:8328–41.

    Article  PubMed  CAS  Google Scholar 

  220. Li J-M, Wang Y-Y, Zhang W, Su H, Ji L-N, Mao Z-W. Low-weight polyethylenimine cross-linked 2-hydroxypopyl-β-cyclodextrin and folic acid as an efficient and nontoxic siRNA carrier for gene silencing and tumor inhibition by VEGF siRNA. Int J Nanomed. 2013;8:2101.

    Article  Google Scholar 

  221. Forrest ML, Gabrielson N, Pack DW. Cyclodextrin–polyethylenimine conjugates for targeted in vitro gene delivery. Biotechnol Bioeng. 2005;89:416–23.

    Article  PubMed  CAS  Google Scholar 

  222. Borrás T. Recent developments in ocular gene therapy. Exp Eye Res. 2003;76:643–52.

    Article  PubMed  Google Scholar 

  223. Gomes dos Santos AL, Bochot A, Tsapis N, Artzner F, Bejjani RA, Thillaye-Goldenberg B, De Kozak Y, Fattal E, Behar-Cohen F. Oligonucleotide-polyethylenimine complexes targeting retinal cells: structural analysis and application to anti-TGFbeta-2 therapy. Pharm Res. 2006;23:770–81.

    Article  PubMed  CAS  Google Scholar 

  224. Pitkänen L, Ruponen M, Nieminen J, Urtti A. Vitreous is a barrier in nonviral gene transfer by cationic lipids and polymers. Pharm Res. 2003;20:576–83.

    Article  PubMed  Google Scholar 

  225. Reinisalo M, Urtti A, Honkakoski P. Freeze-drying of cationic polymer DNA complexes enables their long-term storage and reverse transfection of post-mitotic cells. J Control Release. 2006;110:437–43.

    Article  PubMed  CAS  Google Scholar 

  226. Huang H, Liu M, Jiang R, Chen J, Mao L, Wen Y, Tian J, Zhou N, Zhang X, Wei Y. Facile modification of nanodiamonds with hyperbranched polymers based on supramolecular chemistry and their potential for drug delivery. J Colloid Interface Sci. 2018;513:198–204.

    Article  PubMed  CAS  Google Scholar 

  227. Jung H-S, Cho K-J, Ryu S-J, Takagi Y, Roche PA, Neuman KC. Biocompatible fluorescent nanodiamonds as multifunctional optical probes for latent fingerprint detection. ACS Appl Mater Interfaces. 2020;12:6641–50.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  228. Li J, Zhu Y, Li W, Zhang X, Peng Y, Huang Q. Nanodiamonds as intracellular transporters of chemotherapeutic drug. Biomaterials. 2010;31:8410–8.

    Article  PubMed  CAS  Google Scholar 

  229. van der Laan K, Hasani M, Zheng T, Schirhagl R. Nanodiamonds for in vivo applications. Small. 2018;14:1703838.

    Article  Google Scholar 

  230. Woodhams B, Ansel-Bollepalli L, Surmacki J, Knowles H, Maggini L, De Volder M, Atatüre M, Bohndiek S. Graphitic and oxidised high pressure high temperature (HPHT) nanodiamonds induce differential biological responses in breast cancer cell lines. Nanoscale. 2018;10:12169–79.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  231. Zhu Y, Li J, Li W, Zhang Y, Yang X, Chen N, Sun Y, Zhao Y, Fan C, Huang Q. The biocompatibility of nanodiamonds and their application in drug delivery systems. Theranostics. 2012;2:302.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  232. Chu Z, Zhang S, Zhang B, Zhang C, Fang CY, Rehor I, Cigler P, Chang HC, Lin G, Liu R, Li Q. Unambiguous observation of shape effects on cellular fate of nanoparticles. Sci Rep. 2014;4:4495.

    Article  PubMed  PubMed Central  Google Scholar 

  233. Chu Z, Miu K, Lung P, Zhang S, Zhao S, Chang HC, Lin G, Li Q. Rapid endosomal escape of prickly nanodiamonds: implications for gene delivery. Sci Rep. 2015;5:11661.

    Article  PubMed  PubMed Central  Google Scholar 

  234. Schrand AM, Huang H, Carlson C, Schlager JJ, Omacr Sawa E, Hussain SM, Dai L. Are diamond nanoparticles cytotoxic? J Phys Chem B. 2007;111:2–7.

    Article  PubMed  CAS  Google Scholar 

  235. Fu CC, Lee HY, Chen K, Lim TS, Wu HY, Lin PK, Wei PK, Tsao PH, Chang HC, Fann W. Characterization and application of single fluorescent nanodiamonds as cellular biomarkers. Proc Natl Acad Sci USA. 2007;104:727–32.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  236. Chu HL, Chen HW, Tseng SH, Hsu MH, Ho LP, Chou FH, Li MP, Chang YC, Chen PH, Tsai LY, et al. Development of a growth-hormone-conjugated nanodiamond complex for cancer therapy. ChemMedChem. 2014;9:1023–9.

    Article  PubMed  CAS  Google Scholar 

  237. Mochalin VN, Pentecost A, Li XM, Neitzel I, Nelson M, Wei C, He T, Guo F, Gogotsi Y. Adsorption of drugs on nanodiamond: toward development of a drug delivery platform. Mol Pharm. 2013;10:3728–35.

    Article  PubMed  CAS  Google Scholar 

  238. Jung HS, Neuman KC. Surface modification of fluorescent nanodiamonds for biological applications. Nanomaterials (Basel). 2021;11:153.

    Article  PubMed  CAS  Google Scholar 

  239. Yang TC, Chang CY, Yarmishyn AA, Mao YS, Yang YP, Wang ML, Hsu CC, Yang HY, Hwang DK, Chen SJ, et al. Carboxylated nanodiamond-mediated CRISPR-Cas9 delivery of human retinoschisis mutation into human iPSCs and mouse retina. Acta Biomater. 2020;101:484–94.

    Article  PubMed  CAS  Google Scholar 

  240. Bobo D, Robinson KJ, Islam J, Thurecht KJ, Corrie SR. Nanoparticle-based medicines: a review of FDA-approved materials and clinical trials to date. Pharm Res. 2016;33:2373–87.

    Article  PubMed  CAS  Google Scholar 

  241. Mohammadpour R, Dobrovolskaia MA, Cheney DL, Greish KF, Ghandehari H. Subchronic and chronic toxicity evaluation of inorganic nanoparticles for delivery applications. Adv Drug Deliv Rev. 2019;144:112–32.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  242. Yang G, Phua SZF, Bindra AK, Zhao Y. Degradability and clearance of inorganic nanoparticles for biomedical applications. Adv Mater. 2019;31:1805730.

    Article  Google Scholar 

  243. Eidi H, David M-O, Crépeaux G, Henry L, Joshi V, Berger M-H, Sennour M, Cadusseau J, Gherardi RK, Curmi PA. Fluorescent nanodiamonds as a relevant tag for the assessment of alum adjuvant particle biodisposition. BMC Med. 2015;13:1–13.

    Article  CAS  Google Scholar 

  244. Mohan N, Chen C-S, Hsieh H-H, Wu Y-C, Chang H-C. In vivo imaging and toxicity assessments of fluorescent nanodiamonds in Caenorhabditis elegans. Nano Lett. 2010;10:3692–9.

    Article  PubMed  CAS  Google Scholar 

  245. Moore L, Yang J, Lan TTH, Osawa E, Lee DK, Johnson WD, Xi J, Chow EK, Ho D. Biocompatibility assessment of detonation nanodiamond in non-human primates and rats using histological, hematologic, and urine analysis. ACS Nano. 2016;10:7385–400.

    Article  PubMed  CAS  Google Scholar 

  246. Balfourier A, Luciani N, Wang G, Lelong G, Ersen O, Khelfa A, Alloyeau D, Gazeau F, Carn F. Unexpected intracellular biodegradation and recrystallization of gold nanoparticles. Proc Natl Acad Sci USA. 2020;117:103–13.

    Article  PubMed  CAS  Google Scholar 

  247. Bernard K, Thannickal VJ. NADPH oxidase inhibition in fibrotic pathologies. Antioxid Redox Signal. 2020;33:455–79.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  248. Lee K, Conboy M, Park HM, Jiang F, Kim HJ, Dewitt MA, Mackley VA, Chang K, Rao A, Skinner C, et al. Nanoparticle delivery of Cas9 ribonucleoprotein and donor DNA in vivo induces homology-directed DNA repair. Nat Biomed Eng. 2017;1:889–901.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  249. Castaneda MT, Merkoçi A, Pumera M, Alegret S. Electrochemical genosensors for biomedical applications based on gold nanoparticles. Biosens Bioelectron. 2007;22:1961–7.

    Article  PubMed  CAS  Google Scholar 

  250. da Silva AB, Rufato KB, de Oliveira AC, Souza PR, da Silva EP, Muniz EC, Vilsinski BH, Martins AF. Composite materials based on chitosan/gold nanoparticles: from synthesis to biomedical applications. Int J Biol Macromol. 2020. https://doi.org/10.1016/j.ijbiomac.2020.06.113.

    Article  PubMed  Google Scholar 

  251. Fan J, Cheng Y, Sun M. Functionalized gold nanoparticles: synthesis, properties and biomedical applications. Chem Rec. 2020;20:1474–504.

    Article  PubMed  CAS  Google Scholar 

  252. Tiwari PM, Vig K, Dennis VA, Singh SR. Functionalized gold nanoparticles and their biomedical applications. Nanomaterials. 2011;1:31–63.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  253. Chen F, Alphonse M, Liu Q. Strategies for nonviral nanoparticle-based delivery of CRISPR/Cas9 therapeutics. Wiley Interdiscip Rev Nanomed Nanobiotechnol. 2020;12: e1609.

    Article  PubMed  Google Scholar 

  254. Glass Z, Li Y, Xu Q. Nanoparticles for CRISPR–Cas9 delivery. Nat Biomed Eng. 2017;1:854–5.

    Article  PubMed  PubMed Central  Google Scholar 

  255. Shankar SS, Ahmad A, Pasricha R, Sastry M. Bioreduction of chloroaurate ions by geranium leaves and its endophytic fungus yields gold nanoparticles of different shapes. J Mater Chem. 2003;13:1822–6.

    Article  CAS  Google Scholar 

  256. Wang P, Zhang L, Zheng W, Cong L, Guo Z, Xie Y, Wang L, Tang R, Feng Q, Hamada Y. Thermo-triggered release of CRISPR-Cas9 system by lipid-encapsulated gold nanoparticles for tumor therapy. Angew Chem Int Ed. 2018;57:1491–6.

    Article  CAS  Google Scholar 

  257. Wang P, Zhang L, Zheng W, Cong L, Guo Z, Xie Y, Wang L, Tang R, Feng Q, Hamada Y, et al. Thermo-triggered release of CRISPR-Cas9 system by lipid-encapsulated gold nanoparticles for tumor therapy. Angew Chem Int Ed Engl. 2018;57:1491–6.

    Article  PubMed  CAS  Google Scholar 

  258. Wang P, Zhang L, Xie Y, Wang N, Tang R, Zheng W, Jiang X. Genome editing for cancer therapy: delivery of Cas9 Protein/sgRNA plasmid via a gold nanocluster/lipid core-shell nanocarrier. Adv Sci (Weinh). 2017;4:1700175.

    Article  PubMed  Google Scholar 

  259. Karakoçak BB, Raliya R, Davis JT, Chavalmane S, Wang WN, Ravi N, Biswas P. Biocompatibility of gold nanoparticles in retinal pigment epithelial cell line. Toxicol In Vitro. 2016;37:61–9.

    Article  PubMed  Google Scholar 

  260. Söderstjerna E, Bauer P, Cedervall T, Abdshill H, Johansson F, Johansson UE. Silver and gold nanoparticles exposure to in vitro cultured retina—studies on nanoparticle internalization, apoptosis, oxidative stress, glial- and microglial activity. PLoS ONE. 2014;9: e105359.

    Article  PubMed  PubMed Central  Google Scholar 

  261. Song HB, Wi JS, Jo DH, Kim JH, Lee SW, Lee TG, Kim JH. Intraocular application of gold nanodisks optically tuned for optical coherence tomography: inhibitory effect on retinal neovascularization without unbearable toxicity. Nanomedicine. 2017;13:1901–11.

    Article  PubMed  CAS  Google Scholar 

  262. Kim JH, Kim JH, Kim KW, Kim MH, Yu YS. Intravenously administered gold nanoparticles pass through the blood-retinal barrier depending on the particle size, and induce no retinal toxicity. Nanotechnology. 2009;20: 505101.

    Article  PubMed  Google Scholar 

  263. Dasari Shareena TP, McShan D, Dasmahapatra AK, Tchounwou PB. A review on graphene-based nanomaterials in biomedical applications and risks in environment and health. Nano Micro Lett. 2018;10:53.

    Article  Google Scholar 

  264. Liu J, Cui L, Losic D. Graphene and graphene oxide as new nanocarriers for drug delivery applications. Acta Biomater. 2013;9:9243–57.

    Article  PubMed  CAS  Google Scholar 

  265. Lammel T, Boisseaux P, Fernández-Cruz ML, Navas JM. Internalization and cytotoxicity of graphene oxide and carboxyl graphene nanoplatelets in the human hepatocellular carcinoma cell line Hep G2. Part Fibre Toxicol. 2013;10:27.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  266. Linares J, Matesanz MC, Vila M, Feito MJ, Gonçalves G, Vallet-Regí M, Marques PA, Portolés MT. Endocytic mechanisms of graphene oxide nanosheets in osteoblasts, hepatocytes and macrophages. ACS Appl Mater Interfaces. 2014;6:13697–706.

    Article  PubMed  CAS  Google Scholar 

  267. Burnett M, Abuetabh Y, Wronski A, Shen F, Persad S, Leng R, Eisenstat D, Sergi C. Graphene oxide nanoparticles induce apoptosis in wild-type and CRISPR/Cas9-IGF/IGFBP3 knocked-out osteosarcoma cells. J Cancer. 2020;11:5007.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  268. Carboni V, Maaliki C, Alyami M, Alsaiari S, Khashab N. Synthetic vehicles for encapsulation and delivery of CRISPR/Cas9 gene editing machinery. Adv Ther. 2019;2:1800085.

    Article  CAS  Google Scholar 

  269. Liu Z, Robinson JT, Sun X, Dai H. PEGylated nanographene oxide for delivery of water-insoluble cancer drugs. J Am Chem Soc. 2008;130:10876–7.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  270. Shen H, Liu M, He H, Zhang L, Huang J, Chong Y, Dai J, Zhang Z. PEGylated graphene oxide-mediated protein delivery for cell function regulation. ACS Appl Mater Interfaces. 2012;4:6317–23.

    Article  PubMed  CAS  Google Scholar 

  271. Yue H, Zhou X, Cheng M, Xing D. Graphene oxide-mediated Cas9/sgRNA delivery for efficient genome editing. Nanoscale. 2018;10:1063–71.

    Article  PubMed  CAS  Google Scholar 

  272. Yan L, Wang Y, Xu X, Zeng C, Hou J, Lin M, Xu J, Sun F, Huang X, Dai L, et al. Can graphene oxide cause damage to eyesight? Chem Res Toxicol. 2012;25:1265–70.

    Article  PubMed  CAS  Google Scholar 

  273. Li B, Zhang X-Y, Yang J-Z, Zhang Y-J, Li W-X, Fan C-H, Huang Q. Influence of polyethylene glycol coating on biodistribution and toxicity of nanoscale graphene oxide in mice after intravenous injection. Int J Nanomed. 2014;9:4697–707.

    Article  Google Scholar 

  274. Vázquez E, Villaverde A. Engineering building blocks for self-assembling protein nanoparticles. Microb Cell Fact. 2010;9:101.

    Article  PubMed  PubMed Central  Google Scholar 

  275. Mastrobattista E, van der Aa MA, Hennink WE, Crommelin DJ. Artificial viruses: a nanotechnological approach to gene delivery. Nat Rev Drug Discov. 2006;5:115–21.

    Article  PubMed  Google Scholar 

  276. Li L, Song L, Liu X, Yang X, Li X, He T, Wang N, Yang S, Yu C, Yin T, et al. Artificial virus delivers CRISPR-Cas9 system for genome editing of cells in mice. ACS Nano. 2017;11:95–111.

    Article  PubMed  CAS  Google Scholar 

  277. Mejia-Ariza R, Kronig GA, Huskens J. Size-controlled and redox-responsive supramolecular nanoparticles. Beilstein J Org Chem. 2015;11:2388–99.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  278. Rädler JO, Koltover I, Salditt T, Safinya CR. Structure of DNA-cationic liposome complexes: DNA intercalation in multilamellar membranes in distinct interhelical packing regimes. Science. 1997;275:810–4.

    Article  PubMed  Google Scholar 

  279. Patel S, Ryals RC, Weller KK, Pennesi ME, Sahay G. Lipid nanoparticles for delivery of messenger RNA to the back of the eye. J Control Release. 2019;303:91–100.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  280. Ewert KK, Zidovska A, Ahmad A, Bouxsein NF, Evans HM, McAllister CS, Samuel CE, Safinya CR. Cationic liposome-nucleic acid complexes for gene delivery and silencing: pathways and mechanisms for plasmid DNA and siRNA. Top Curr Chem. 2010;296:191–226.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  281. Ahmad A, Evans HM, Ewert K, George CX, Samuel CE, Safinya CR. New multivalent cationic lipids reveal bell curve for transfection efficiency versus membrane charge density: lipid-DNA complexes for gene delivery. J Gene Med. 2005;7:739–48.

    Article  PubMed  CAS  Google Scholar 

  282. Cheng X, Lee RJ. The role of helper lipids in lipid nanoparticles (LNPs) designed for oligonucleotide delivery. Adv Drug Deliv Rev. 2016;99:129–37.

    Article  PubMed  CAS  Google Scholar 

  283. Leung AK, Tam YY, Cullis PR. Lipid nanoparticles for short interfering RNA delivery. Adv Genet. 2014;88:71–110.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  284. Caracciolo G, Pozzi D, Capriotti AL, Cavaliere C, Laganà A. Effect of DOPE and cholesterol on the protein adsorption onto lipid nanoparticles. J Nanoparticle Res. 2013. https://doi.org/10.1007/s11051-013-1498-4.

    Article  Google Scholar 

  285. Adams D, Gonzalez-Duarte A, O’Riordan WD, Yang C-C, Ueda M, Kristen AV, Tournev I, Schmidt HH, Coelho T, Berk JL. Patisiran, an RNAi therapeutic, for hereditary transthyretin amyloidosis. N Engl J Med. 2018;379:11–21.

    Article  PubMed  CAS  Google Scholar 

  286. Akinc A, Maier MA, Manoharan M, Fitzgerald K, Jayaraman M, Barros S, Ansell S, Du X, Hope MJ, Madden TD. The Onpattro story and the clinical translation of nanomedicines containing nucleic acid-based drugs. Nat Nanotechnol. 2019;14:1084–7.

    Article  PubMed  CAS  Google Scholar 

  287. Pardi N, Hogan MJ, Porter FW, Weissman D. mRNA vaccines—a new era in vaccinology. Nat Rev Drug Discov. 2018;17:261–79.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  288. Ali MM, Li F, Zhang Z, Zhang K, Kang DK, Ankrum JA, Le XC, Zhao W. Rolling circle amplification: a versatile tool for chemical biology, materials science and medicine. Chem Soc Rev. 2014;43:3324–41.

    Article  PubMed  CAS  Google Scholar 

  289. Sun W, Jiang T, Lu Y, Reiff M, Mo R, Gu Z. Cocoon-like self-degradable DNA nanoclew for anticancer drug delivery. J Am Chem Soc. 2014;136:14722–5.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  290. Liu C, Zhang L, Liu H, Cheng K. Delivery strategies of the CRISPR-Cas9 gene-editing system for therapeutic applications. J Control Release. 2017;266:17–26.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  291. Tian J, Avalos AM, Mao S-Y, Chen B, Senthil K, Wu H, Parroche P, Drabic S, Golenbock D, Sirois C, et al. Toll-like receptor 9-dependent activation by DNA-containing immune complexes is mediated by HMGB1 and RAGE. Nat Immunol. 2007;8:487–96.

    Article  PubMed  CAS  Google Scholar 

  292. Sun W, Ji W, Hall JM, Hu Q, Wang C, Beisel CL, Gu Z. Self-assembled DNA nanoclews for the efficient delivery of CRISPR-Cas9 for genome editing. Angew Chem Int Ed Engl. 2015;54:12029–33.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  293. Gasiunas G, Barrangou R, Horvath P, Siksnys V. Cas9–crRNA ribonucleoprotein complex mediates specific DNA cleavage for adaptive immunity in bacteria. Proc Natl Acad Sci. 2012;109:E2579–86.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  294. Alsaiari SK, Patil S, Alyami M, Alamoudi KO, Aleisa FA, Merzaban JS, Li M, Khashab NM. Endosomal escape and delivery of CRISPR/Cas9 genome editing machinery enabled by Nanoscale Zeolitic Imidazolate Framework. J Am Chem Soc. 2018;140:143–6.

    Article  PubMed  CAS  Google Scholar 

  295. Zhang J, Tan Y, Song W-J. Zeolitic imidazolate frameworks for use in electrochemical and optical chemical sensing and biosensing: a review. Microchim Acta. 2020;187:234.

    Article  CAS  Google Scholar 

  296. Miensah ED, Khan MM, Chen JY, Zhang XM, Wang P, Zhang ZX, Jiao Y, Liu Y, Yang Y. Zeolitic imidazolate frameworks and their derived materials for sequestration of radionuclides in the environment: a review. Crit Rev Environ Sci Technol. 2020;50:1874–934.

    Article  CAS  Google Scholar 

  297. Chen TT, Yi JT, Zhao YY, Chu X. Biomineralized metal-organic framework nanoparticles enable intracellular delivery and endo-lysosomal release of native active proteins. J Am Chem Soc. 2018;140:9912–20.

    Article  PubMed  CAS  Google Scholar 

  298. Poddar A, Conesa JJ, Liang K, Dhakal S, Reineck P, Bryant G, Pereiro E, Ricco R, Amenitsch H, Doonan C, et al. Encapsulation, visualization and expression of genes with biomimetically mineralized Zeolitic Imidazolate Framework-8 (ZIF-8). Small. 2019;15: e1902268.

    Article  PubMed  Google Scholar 

  299. Yang X, Tang Q, Jiang Y, Zhang M, Wang M, Mao L. Nanoscale ATP-Responsive Zeolitic Imidazole Framework-90 as a general platform for cytosolic protein delivery and genome editing. J Am Chem Soc. 2019;141:3782–6.

    Article  PubMed  CAS  Google Scholar 

  300. Wang Y, Shahi PK, Xie R, Zhang H, Abdeen AA, Yodsanit N, Ma Z, Saha K, Pattnaik BR, Gong S. A pH-responsive silica-metal-organic framework hybrid nanoparticle for the delivery of hydrophilic drugs, nucleic acids, and CRISPR-Cas9 genome-editing machineries. J Control Release. 2020;324:194–203.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  301. Mustafi D, Engel AH, Palczewski K. Structure of cone photoreceptors. Prog Retin Eye Res. 2009;28:289–302.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  302. Chichagova V, Hilgen G, Ghareeb A, Georgiou M, Carter M, Sernagor E, Lako M, Armstrong L. Human iPSC differentiation to retinal organoids in response to IGF1 and BMP4 activation is line-and method-dependent. Stem Cells. 2020;38:195–201.

    Article  PubMed  CAS  Google Scholar 

  303. Lukovic D, Castro AA, Kaya KD, Munezero D, Gieser L, Davó-Martínez C, Corton M, Cuenca N, Swaroop A, Ramamurthy V. Retinal organoids derived from hiPSCs of an AIPL1-LCA patient maintain cytoarchitecture despite reduced levels of mutant AIPL1. Sci Rep. 2020;10:1–13.

    Article  Google Scholar 

  304. Regent F, Chen HY, Kelley RA, Qu Z, Swaroop A, Li T. A simple and efficient method for generating human retinal organoids. Mol Vis. 2020;26:97.

    PubMed  CAS  PubMed Central  Google Scholar 

  305. Sridhar A, Hoshino A, Finkbeiner CR, Chitsazan A, Dai L, Haugan AK, Eschenbacher KM, Jackson DL, Trapnell C, Bermingham-McDonogh O. Single-cell transcriptomic comparison of human fetal retina, hPSC-derived retinal organoids, and long-term retinal cultures. Cell Rep. 2020;30(1644–1659): e1644.

    Article  Google Scholar 

  306. Völkner M, Zschätzsch M, Rostovskaya M, Overall RW, Busskamp V, Anastassiadis K, Karl MO. Retinal organoids from pluripotent stem cells efficiently recapitulate retinogenesis. Stem Cell Rep. 2016;6:525–38.

    Article  Google Scholar 

  307. Wiegand C, Banerjee I. Recent advances in the applications of iPSC technology. Curr Opin Biotechnol. 2019;60:250–8.

    Article  PubMed  CAS  Google Scholar 

  308. Fatehullah A, Tan SH, Barker N. Organoids as an in vitro model of human development and disease. Nat Cell Biol. 2016;18:246–54.

    Article  PubMed  Google Scholar 

  309. Streeter I, Harrison PW, Faulconbridge A, Consortium H, Flicek P, Parkinson H, Clarke L. The human-induced pluripotent stem cell initiative—data resources for cellular genetics. Nucleic Acids Res. 2017;45:D691–7.

    Article  PubMed  CAS  Google Scholar 

  310. Kim M, Mun H, Sung CO, Cho EJ, Jeon H-J, Chun S-M, Shin TH, Jeong GS, Kim DK, Choi EK. Patient-derived lung cancer organoids as in vitro cancer models for therapeutic screening. Nat Commun. 2019;10:1–15.

    Article  Google Scholar 

  311. Lee SH, Hu W, Matulay JT, Silva MV, Owczarek TB, Kim K, Chua CW, Barlow LJ, Kandoth C, Williams AB. Tumor evolution and drug response in patient-derived organoid models of bladder cancer. Cell. 2018;173(515–528): e517.

    Google Scholar 

  312. Nadkarni RR, Abed S, Cox BJ, Bhatia S, Lau JT, Surette MG, Draper JS. Functional enterospheres derived in vitro from human pluripotent stem cells. Stem Cell Rep. 2017;9:897–912.

    Article  CAS  Google Scholar 

  313. Spence JR, Mayhew CN, Rankin SA, Kuhar MF, Vallance JE, Tolle K, Hoskins EE, Kalinichenko VV, Wells SI, Zorn AM. Directed differentiation of human pluripotent stem cells into intestinal tissue in vitro. Nature. 2011;470:105–9.

    Article  PubMed  Google Scholar 

  314. Dekkers JF, Wiegerinck CL, De Jonge HR, Bronsveld I, Janssens HM, De Winter-de Groot KM, Brandsma AM, De Jong NW, Bijvelds MJ, Scholte BJ. A functional CFTR assay using primary cystic fibrosis intestinal organoids. Nat Med. 2013;19:939–45.

    Article  PubMed  CAS  Google Scholar 

  315. Brawner AT, Xu R, Liu D, Jiang P. Generating CNS organoids from human induced pluripotent stem cells for modeling neurological disorders. Int J Physiol Pathophysiol Pharmacol. 2017;9:101.

    PubMed  CAS  PubMed Central  Google Scholar 

  316. Gabriel E, Gopalakrishnan J. Generation of iPSC-derived human brain organoids to model early neurodevelopmental disorders. J Vis Exper. 2017;14:e55372.

    Google Scholar 

  317. Sun G, Chiuppesi F, Chen X, Wang C, Tian E, Nguyen J, Kha M, Trinh D, Zhang H, Marchetto MC. Modeling human cytomegalovirus-induced microcephaly in human iPSC-derived brain organoids. Cell Rep Med. 2020;1:100002.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  318. Forbes TA, Howden SE, Lawlor K, Phipson B, Maksimovic J, Hale L, Wilson S, Quinlan C, Ho G, Holman K. Patient-iPSC-derived kidney organoids show functional validation of a ciliopathic renal phenotype and reveal underlying pathogenetic mechanisms. Am J Hum Genet. 2018;102:816–31.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  319. Freedman BS, Lam AQ, Sundsbak JL, Iatrino R, Su X, Koon SJ, Wu M, Daheron L, Harris PC, Zhou J. Reduced ciliary polycystin-2 in induced pluripotent stem cells from polycystic kidney disease patients with PKD1 mutations. J Am Soc Nephrol. 2013;24:1571–86.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  320. Shimizu T, Mae S-I, Araoka T, Okita K, Hotta A, Yamagata K, Osafune K. A novel ADPKD model using kidney organoids derived from disease-specific human iPSCs. Biochem Biophys Res Commun. 2020;529:1186–94.

    Article  PubMed  CAS  Google Scholar 

  321. Hulot JS. Modeling cardiac arrhythmias with organoids. J Am Coll Cardiol. 2019. https://doi.org/10.1016/j.jacc.2019.01.076.

    Article  PubMed  Google Scholar 

  322. Lee J, Sutani A, Kaneko R, Takeuchi J, Sasano T, Kohda T, Ihara K, Takahashi K, Yamazoe M, Morio T. In vitro generation of functional murine heart organoids via FGF4 and extracellular matrix. Nat Commun. 2020;11:1–18.

    Google Scholar 

  323. Takasato M, Pei XE, Chiu HS, Maier B, Baillie GJ, Ferguson C, Parton RG, Wolvetang EJ, Roost MS, de Sousa Lopes SMC. Kidney organoids from human iPS cells contain multiple lineages and model human nephrogenesis. Nature. 2015;526:564–8.

    Article  PubMed  CAS  Google Scholar 

  324. Lancaster MA, Renner M, Martin C-A, Wenzel D, Bicknell LS, Hurles ME, Homfray T, Penninger JM, Jackson AP, Knoblich JA. Cerebral organoids model human brain development and microcephaly. Nature. 2013;501:373–9.

    Article  PubMed  CAS  Google Scholar 

  325. Qian X, Jacob F, Song MM, Nguyen HN, Song H, Gl M. Generation of human brain region-specific organoids using a miniaturized spinning bioreactor. Nat Protoc. 2018;13:565.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  326. Forster R, Chiba K, Schaeffer L, Regalado SG, Lai CS, Gao Q, Kiani S, Farin HF, Clevers H, Cost GJ. Human intestinal tissue with adult stem cell properties derived from pluripotent stem cells. Stem Cell Rep. 2014;2:838–52.

    Article  CAS  Google Scholar 

  327. Ang LT, Tan AKY, Autio MI, Goh SH, Choo SH, Lee KL, Tan J, Pan B, Lee JJH, Lum JJ. A roadmap for human liver differentiation from pluripotent stem cells. Cell Rep. 2018;22:2190–205.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  328. Takebe T, Sekine K, Enomura M, Koike H, Kimura M, Ogaeri T, Zhang R-R, Ueno Y, Zheng Y-W, Koike N. Vascularized and functional human liver from an iPSC-derived organ bud transplant. Nature. 2013;499:481–4.

    Article  PubMed  CAS  Google Scholar 

  329. Takebe T, Sekine K, Kimura M, Yoshizawa E, Ayano S, Koido M, Funayama S, Nakanishi N, Hisai T, Kobayashi T. Massive and reproducible production of liver buds entirely from human pluripotent stem cells. Cell Rep. 2017;21:2661–70.

    Article  PubMed  CAS  Google Scholar 

  330. Dye BR, Dedhia PH, Miller AJ, Nagy MS, White ES, Shea LD, Spence JR. A bioengineered niche promotes in vivo engraftment and maturation of pluripotent stem cell derived human lung organoids. Elife. 2016;5: e19732.

    Article  PubMed  PubMed Central  Google Scholar 

  331. Xu M, Lee EM, Wen Z, Cheng Y, Huang WK, Qian X, Tcw J, Kouznetsova J, Ogden SC, Hammack C, et al. Identification of small-molecule inhibitors of Zika virus infection and induced neural cell death via a drug repurposing screen. Nat Med. 2016;22:1101–7.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  332. Bian S, Repic M, Guo Z, Kavirayani A, Burkard T, Bagley JA, Krauditsch C, Knoblich JA. Genetically engineered cerebral organoids model brain tumor formation. Nat Methods. 2018;15:631–9.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  333. Saengwimol D, Rojanaporn D, Chaitankar V, Chittavanich P, Aroonroch R, Boontawon T, Thammachote W, Jinawath N, Hongeng S, Kaewkhaw R. A three-dimensional organoid model recapitulates tumorigenic aspects and drug responses of advanced human retinoblastoma. Sci Rep. 2018;8:15664.

    Article  PubMed  PubMed Central  Google Scholar 

  334. Fernandes TG, Rodrigues CA, Diogo MM, Cabral JM. Stem cell bioprocessing for regenerative medicine. J Chem Technol Biotechnol. 2014;89:34–47.

    Article  CAS  Google Scholar 

  335. Vergara MN, Flores-Bellver M, Aparicio-Domingo S, McNally M, Wahlin KJ, Saxena MT, Mumm JS, Canto-Soler MV. Three-dimensional automated reporter quantification (3D-ARQ) technology enables quantitative screening in retinal organoids. Development. 2017;144:3698–705.

    PubMed  CAS  PubMed Central  Google Scholar 

  336. Aasen DM, Vergara MN. New drug discovery paradigms for retinal diseases: a focus on retinal organoids. J Ocul Pharmacol Ther. 2020;36:18–24.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  337. Kruczek K, Swaroop A. Pluripotent stem cell-derived retinal organoids for disease modeling and development of therapies. Stem Cells. 2020;38:1206–15.

    Article  PubMed  Google Scholar 

  338. Pierce EA, Bennett J. The status of RPE65 gene therapy trials: safety and efficacy. Cold Spring Harb Perspect Med. 2015;5: a017285.

    Article  PubMed  PubMed Central  Google Scholar 

  339. Simonelli F, Maguire AM, Testa F, Pierce EA, Mingozzi F, Bennicelli JL, Rossi S, Marshall K, Banfi S, Surace EM. Gene therapy for Leber’s congenital amaurosis is safe and effective through 1.5 years after vector administration. Mol Ther. 2010;18:643–50.

    Article  PubMed  CAS  Google Scholar 

  340. Benati D, Patrizi C, Recchia A. Gene editing prospects for treating inherited retinal diseases. J Med Genet. 2020;57:437–44.

    Article  PubMed  CAS  Google Scholar 

  341. Garafalo AV, Cideciyan AV, Héon E, Sheplock R, Pearson A, Yu CW, Sumaroka A, Aguirre GD, Jacobson SG. Progress in treating inherited retinal diseases: early subretinal gene therapy clinical trials and candidates for future initiatives. Prog Retin Eye Res. 2019;77:100827.

    Article  PubMed  PubMed Central  Google Scholar 

  342. Xu CL, Cho GY, Sengillo JD, Park KS, Mahajan VB, Tsang SH. Translation of CRISPR genome surgery to the bedside for retinal diseases. Front Cell Dev Biol. 2018;6:46.

    Article  PubMed  PubMed Central  Google Scholar 

  343. Bondy-Denomy J, Pawluk A, Maxwell KL, Davidson AR. Bacteriophage genes that inactivate the CRISPR/Cas bacterial immune system. Nature. 2013;493:429–32.

    Article  PubMed  CAS  Google Scholar 

  344. Ford K, McDonald D, Mali P. Functional genomics via CRISPR-Cas. J Mol Biol. 2019;431:48–65.

    Article  PubMed  CAS  Google Scholar 

  345. Kim J, Koo B-K, Knoblich JA. Human organoids: model systems for human biology and medicine. Nat Rev Mol Cell Biol. 2020;21:571–84.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

Download references

Acknowledgements

This study was funded by Ministry of Science and Technology (MOST) (MOST 108-2320-B-010 -019 -MY3; MOST 109-2327-B-010-007), Ministry of Health and Welfare (MOHW)(MOHW108-TDU-B-211-133001, MOHW109-TDU-B-211-114001), VGH, NTUH Joint Research Program (VN109-16), VGH, TSGH, NDMC, AS Joint Research Program (VTA107-V1-5-1, VTA108-V1-5-3, VTA109-V1-4-1), Academia Sinica (AS-TM-110-02-02), -AS Clinical Research Center (IBMS-CRC109-P04), the “Cancer Progression Research Center, National Yang-Ming University” from The Featured Areas Research Center Program within the framework of the Higher Education Sprout Project by the Ministry of Education (MOE) in Taiwan, and the Ministry of Education through the SPROUT Project- Center For Intelligent Drug Systems and Smart Bio-devices (IDS2B) of National Chiao Tung University and, Taiwan.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization-YC, YPY, SHC. Software-TYL, WYL, YC. Validation-MSL, YYL, TWL, DKH, TCL, SJC. writing—original draft preparation YC, YJH; writing—review and editing, YC, YJH, AAY; visualization, TYL, YC, YJH, AAY; supervision, YPY, SJC, YC, SHC. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Shih-Hwa Chiou or Yi-Ping Yang.

Ethics declarations

Competing interests

The authors have no Conflict of interest to declare.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chien, Y., Hsiao, YJ., Chou, SJ. et al. Nanoparticles-mediated CRISPR-Cas9 gene therapy in inherited retinal diseases: applications, challenges, and emerging opportunities. J Nanobiotechnol 20, 511 (2022). https://doi.org/10.1186/s12951-022-01717-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12951-022-01717-x

Keywords