Next Article in Journal
Increasing Angiogenesis Factors in Hypoxic Diabetic Wound Conditions by siRNA Delivery: Additive Effect of LbL-Gold Nanocarriers and Desloratadine-Induced Lysosomal Escape
Previous Article in Journal
Joint Degeneration in a Mouse Model of Pseudoachondroplasia: ER Stress, Inflammation, and Block of Autophagy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Understanding the Potential of Genome Editing in Parkinson’s Disease

1
Department of Biomedical Engineering, Universidad de los Andes, Bogotá 111711, Colombia
2
Grupo de Diseño de Productos y Procesos, Department of Chemical and Food Engineering, Universidad de los Andes, Bogotá 111711, Colombia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(17), 9241; https://doi.org/10.3390/ijms22179241
Submission received: 13 May 2021 / Revised: 7 July 2021 / Accepted: 8 July 2021 / Published: 26 August 2021
(This article belongs to the Section Molecular Neurobiology)

Abstract

:
CRISPR is a simple and cost-efficient gene-editing technique that has become increasingly popular over the last decades. Various CRISPR/Cas-based applications have been developed to introduce changes in the genome and alter gene expression in diverse systems and tissues. These novel gene-editing techniques are particularly promising for investigating and treating neurodegenerative diseases, including Parkinson’s disease, for which we currently lack efficient disease-modifying treatment options. Gene therapy could thus provide treatment alternatives, revolutionizing our ability to treat this disease. Here, we review our current knowledge on the genetic basis of Parkinson’s disease to highlight the main biological pathways that become disrupted in Parkinson’s disease and their potential as gene therapy targets. Next, we perform a comprehensive review of novel delivery vehicles available for gene-editing applications, critical for their successful application in both innovative research and potential therapies. Finally, we review the latest developments in CRISPR-based applications and gene therapies to understand and treat Parkinson’s disease. We carefully examine their advantages and shortcomings for diverse gene-editing applications in the brain, highlighting promising avenues for future research.

1. Introduction

The advent of simpler, accessible, and less expensive genome editing technologies such as CRISPR-Cas9 has opened up immense possibilities for research into the genetic underpinnings of complex neurodegenerative diseases, such as Parkinson’s disease (PD), and the development of genetic therapies to treat them. PD is a complex and heterogeneous neurological disorder characterized by Parkinsonian motor symptoms (muscle rigidity, hypokinesia, tremors, and impaired gait) that progress in time, leading to more severe movement disorders as the disease progresses [1]. Moreover, PD can alter cognition, sleep patterns, and autonomous function, generally before motor symptoms appear [2]. PD symptoms are typically caused by a specific cellular process disfunction, such as the early death of dopaminergic neurons in the brain’s Substantia nigra pars compacta (SNpC). In general, cellular dysfunction and the death of dopaminergic neurons result from α-synuclein (α-syn) aggregation into Lewy bodies. It is now well understood that PD pathological hallmarks result from a complex interplay between genetic and environmental factors (idiopathic origin). Only in a minority of PD cases (approximately 10%), PD can be associated with genetic factors and runs in families (familial PD) [1,3,4,5]. These familial PD cases have been attributed to mutations in genes such as SNCA (encoding for α-syn), PRKN, PARK2, PINK1, and PARK7 genes.
Genes identified in studies of familial PD have become candidates of interest for potential therapeutic intervention [2,5,6]. Additionally, these genes have been the focus of studies to better understand the biological routes disrupted in cases of idiopathic PD, likely caused by a combination of complex multigenic components and environmental triggers [5,7]. Understanding the genes involved in developing familial and idiopathic forms of the disease could elucidate the functional routes altered in PD and potentially guide novel treatments.
Based on the cellular processes implicated in PD, multiple treatment options have been investigated. Treatment targets and strategies can be classified as non-disease-modifying or disease-modifying depending on whether they aim to ease the disease’s symptoms or treating its underlying causes. Available treatment options for PD include only non-disease-modifying ones, targeting the dopaminergic function to control symptoms, and slowing down the disease [8,9]. These therapies either increase dopamine concentration or stimulate the brain’s dopamine receptors [1]. Treatment options remain limited and can lead to serious adverse side effects, such as dyskinesias, insomnia, and hallucinations, and can lead to additional complications over long-term use [10]. These limitations have motivated significant efforts to develop disease-modifying therapies that could slow down or stop the underlying disease’s progression. Efforts have focused on understanding new pharmacological targets for disease-modifying treatments [2], including targets for gene therapy [6] and the use of novel techniques such as cell transplantation (e.g., stem cell therapies) [11] and surgical interventions required for the administration of treatments to the target brain’s cell populations [2].
Gene editing has recently emerged as an innovative disease-modifying treatment alternative based on editing gene sequences or altering their expression profiles [2,5,6]. Gene therapies have the potential to permanently fix the underlying causes of the disease, which is not yet possible for most neurodegenerative diseases [12,13,14]. For more than two decades, site-directed genome editing has been possible using various molecules capable of recognizing and cleaving specific DNA sequences, such as zinc-finger nucleases and transcription activator-like effector nucleases (TALENs) [15]. Although these techniques represented the beginning of precise genome editing and its incursion into medical treatments, they are costly and challenging due to the complex nuclease design process [16]. A major revolution in genome engineering came with discovering CRISPR/Cas systems and their direct application as gene-editing platforms in mammalian cells [17].
Gene therapies based on CRISPR/Cas9 have already reached the clinical trial stage for many monogenic diseases, including sickle cell disease, b-thalassemia, and Leber congenital amaurosis [18,19,20]. Several of these therapies are currently in advanced preclinical testing stages, including the ones for Duchenne muscular dystrophy, hemoglobinopathies, and hereditary tyrosinemia type 1 [21]. Numerous gene therapies have also been tested to treat neurodegenerative diseases, including Alzheimer’s, metachromatic leukodystrophy, and spinal muscular atrophy [22,23,24,25]. The unique pathophysiology involving genetic, epigenetic, and idiopathic causes for each condition must be carefully evaluated to determine whether gene therapies are appropriate and which are the best strategies to implement them [26].
Before their widespread application in humans, several challenges and limitations must be overcome to ensure future therapies’ safety. Further research is required to address off-target modifications and rapid degradation and to develop solutions for addressing adverse reactions such as DNA damage [27] and immune responses [28]. The development of proper delivery vehicles could be a way to address some of the current challenges associated with CRISPR gene-editing applications [29]. Delivery vehicles based on nanostructured materials (both organic and inorganic, as addressed later) have emerged as powerful agents to internalize cells effectively and thus increase the efficiency of gene-editing technologies [30].
Here, we start by reviewing knowledge accumulated to date about the genes associated with PD, focusing on the different pathways known to become disrupted with the idiopathic form of the disease. We also examine how CRISPR-based gene therapies could be used as treatment alternatives for PD. In line with this, we review CRISPR techniques, their limitations, and the different types of vehicles that can be used to deliver the gene-editing components efficiently and safely. Finally, we review the current state of gene therapy-based clinical trials to treat PD [6].

2. The Genetic Basis of PD

PD is considered one of the most impactful neurodegenerative diseases, with an incidence of 17 per 100,000 persons per year [31]. Recently, significant research efforts have been devoted to developing disease-modifying alternatives based on gene therapies [32] targeting the genes and mutations that have been associated with familial Parkinson’s disease (Table 1). Consequently, understanding the genetic basis and the cellular pathways that become altered in PD is critical to identify gene-editing targets and ultimately develop effective gene therapies.

2.1. Genes at the Basis of PD

Familial PD has a more evident genetic basis than its idiopathic counterpart, facilitating the identification of genes underlying the disease and the ubiquitous pathways that might become altered in PD. Multiple genes involved in significant biological functions have been identified and studied in affected families (Table 1). Genome-wide association studies (GWAS) have been carried out to identify genes involved in PD and, with them, the biological pathways that become disrupted [33]. With this information, it has been possible to propose treatments targeting the affected biological routes and treat each PD patient according to their specific gene variants [34,35]. Most gene variants associated with familial PD are summarized in Table 1 [36,37,38].
Idiopathic PD is characterized by late-onset age, slow progress, and prominent cognitive impairment or dementia, mainly in processing and evoking information [39,40]. Despite its higher incidence, idiopathic PD is less understood than familial PD due to its symptom variability and complex combination of multigenic and environmental causes [41]. To date, research in diverse areas has identified three main cellular pathways that become altered in idiopathic PD: autophagy, lysosomal, and mitochondrial. Restoring the activity of these pathways through gene editing is a promising avenue for future research in our path to ameliorate the symptoms and progression of both idiopathic and familial PD. It is thus important to identify key gene candidates within these pathways as potential therapeutic targets. Below, we discuss these pathways and the genes studied in the context of PD. Although each pathway has particular alterations, they share common attributes, as shown in Figure 1.

2.2. Autophagic Pathway

Together with the proteasomal systems, the autophagic pathway is involved in removing damaged proteins and organelles in highly metabolic nondividing cells [42]. Lysosomes mediate this metabolic pathway, relying on enzymes responsible for protein degradation. The general process consists of forming a phagophore around the damaged elements to create an envelope known as an autophagosome [43]. Subsequently, this structure fuses with a primary lysosome (known as an autolysosome), in which the degradation process occurs, as shown in Figure 1a. A large number of genes encoding for lysosomal enzymes (also related to familial PD) and involved in transport to the lysosome, mitophagy, or other functions related to autophagy, have been associated with the detrimental alteration of this pathway [44].
Among these proteins, HSC70 is essential in the hydrolyzation cycle from ATP to ADP. Specifically, this is dependent on its conformational changes for nucleotide and protein binding [45]. Postmortem studies have provided evidence of decreased chaperone-mediated autophagy activity in PD, which has been associated with significantly decreased levels of HSC70 in the SNpC and amygdala. These findings demonstrate that chaperone-mediated autophagy can be reduced in PD’s brains, contributing to Lewy bodies’ formation and the pathogenesis of neuronal degeneration [46,47,48].
Autophagy disruptions in neurodegenerative diseases significantly impact protein misfolding and abnormal aggregation, impeding neuronal survival (Figure 1b) [49]. Neuronal death and Lewy bodies, protein inclusions of misfolded or non-functional α-synuclein (α-syn) proteins [4], characteristic of PD can be attributed to mutations in the SNCA gene degradation systems [50,51]. The SNCA protein is involved in several cellular processes, including autophagy, endocytosis, exocytosis, neurotransmitter vesicle cycling, and has recently been related to DNA repair mechanisms [52]. Along this pathway, mutations in the genes LRRK2 and UCH-L1 have been correlated with Parkinson’s cases by permanently blocking chaperone-mediated autophagy [53,54]. The mutated LRRK2 gene interacts with the chaperone-mediated autophagy (CMA) receptor, LAMP2A, which disrupts the receptor’s multimerization, resulting in substrates accumulation, including α-syn. Therefore, LRRK2 mutations are linked with defective autophagy and subsequent α-syn accumulation [44]. On the other hand, the UCH-L1 gene encodes a protein of the same name, present in Lewy bodies. UCH-L1 is an integral part of ubiquitin-dependent proteolysis because it allows the recycling of polymer chains from ubiquitin to monomeric ubiquitin. Studies in mesencephalic rat cultures found that inhibiting this gene causes dose-dependent degeneration of dopaminergic neurons and the formation of positive cytoplasmic inclusion of α-syn [55].
GWAS performed on familial PD have contributed significantly to identifying genes involved in the autophagic pathway damage, where 17 novel risk loci were reported, including CHMP2B, BAG3, ANK2, and KAT8 [56,57]. In the first place, CHMP2B is a gene encoding the charged multivesicular protein 2B with unclear function but thought to be related to the endosomal secretory complex required for autophagy [58,59,60]. Moreover, studies in murine PD models have demonstrated that excess protein 2B expression generates autophagy impairment, protein accumulation, and subsequent cellular death [61]. The BAG3 gene encodes for the BAG protein involved in chaperone-assisted selective macroautophagy. BAG is of particular interest because it binds to the protein HSPB8, facilitating the elimination of proteins prone to mutations, whose accumulation is expected in PD [62,63,64,65]. An in vivo study performed in an SNCAA53T transgenic murine model (expressing human A53T α-syn variant) and MG132-treated PC12 cells that overexpress wild-type α-syn confirmed that BAG3 plays a vital role in regulating the elimination of α-syn through macroautophagy [63].
Similarly, ANK2 is a gene encoding for a polypeptide called Ankyrin-B, which binds membrane proteins to the actin/myosin system in the cytoskeleton [66]. Studies have shown that the ANK2 gene’s phosphorylation events are involved in Parkinsonian neurodegeneration [67]. This gene interacts with PINK1/PARKIN target proteins such as MIRO1 or ATP1A2 and ANK2-derived peptides, which are known to be important autophagy inhibitors [68]. Therefore, phosphorylation of ANK2 inhibits autophagy of organelles, including mitochondria (i.e. mitophagy). Finally, KAT8 is a gene encoding for K(lysine) acetyltransferase 8, a protein involved in autophagic regulatory feedback loops because it is highly correlated to H4K16 acetylation [69]. Research has shown that the reduction in KAT8 acetylation is associated with the low regulation of autophagy genes resulting in cell death on mouse embryonic fibroblast [70]. K(lysine) also regulates the PINK1-mitophagy [71], as shown by research conducted to identify autophagic flow modulators on a high-content image-based ARNip. Moreover, this gene’s relationship with PD is verified by experiments where inhibiting KAT8 generates a decrease in the autophagic flow [72,73].

2.3. Mitochondrial Pathway

Mitochondria are essential for cellular function due to ATP production, calcium homeostasis, and apoptotic signaling, and mitochondrial damage is associated with many diseases such as Huntington’s, Alzheimer’s, and myalgic encephalomyelitis [74,75,76]. PD has also been associated with the accumulation of damaged mitochondria in the neuronal cytoplasm [77]. Due to their function, mitochondria are more vulnerable to stress and consequent dysfunction [78], thus need to be constantly renewed especially in cells that have lost their mitotic capacity. Mitochondrial autophagy disruptions, characteristic of PD, begin with a protein known as PINK1, which should be imported through the membrane into healthy mitochondria [77]. Deficient and depolarized mitochondria fail to carry out this process, thus accumulating PINK1 in their membranes, as shown in Figure 1b. Normally, PINK1 accumulation recruits PARKIN, a ubiquitin ligase that facilitates mitochondrial degradation by autophagic and proteasome mechanisms [77]. Mutations in these genes cause the failure of this process, leading to the accumulation of damaged mitochondria. Other mitochondrial proteins and pathways typically dysregulated in patients with PD include the molecular chaperone prohibitin, the OMM VDAC1 protein, the mitochondrial import protein Tom40, the serine protease HtrA2, and the mitochondrial complex I function in the SNpC [79,80].
Loss of mitochondrial integrity and functionality condemn the cell to apoptosis. Under oxidative stress conditions, increased mitochondrial fission (separation of one mitochondrion into two) contributes to mitochondrial damage and cellular energy metabolism failure (Figure 1b). Thus, oxidative stress contributes to well-documented PD dopaminergic neurodegeneration [81,82,83,84]. Environments rich in reactive oxygen species (ROS) trigger cell organelle dysfunction and post-translational modifications of α-syn, such as the phosphorylation of serine 129, nitration, and ubiquitination. These processes facilitate the formation of toxic oligomers [85]. Therefore, the release of cytochrome C into the cytoplasm induces an apoptotic pathway mediated by caspases. These enzymes are also involved in cell proliferation, cellular remodeling, cell fate determination, and immune responses [86]. Moreover, studies in PD patients and animal PD models revealed a positive correlation between neuronal loss and activated caspase-1, -3, -8, and -9 in SNpC [87,88,89].
Numerous studies have reported that the PINK1/PRKN pathway becomes disrupted in PD. This may also be induced in vitro after exogenously adding α-syn or chemical perturbation [90,91,92]. Specifically, in a PD patient’s brain, PARKIN is S-nitrosylated and sequestered into Lewy bodies, which leads to lower availability of soluble, functional PARKIN [92,93]. In contrast, the accumulation of PINK1 is blocked by the inactivation of PRKN. This is supported by the high levels of PARKIN substrates found in a patient’s midbrain tissue [94].
Recently, novel genomic regions associated with an increased risk of developing PD have been identified, including mitochondrial pathway genes such as MCCC1, ALAS1, ANK2, COQ7, CTSB, GALT, and ATP6V0A1 [95,96,97]. GWAS analyses worldwide have reported the presence of these risk loci in the Chinese [98,99], Taiwanese [100], European and American populations [101], thereby supporting the role of the genes mentioned above on PD susceptibility in multiple genetic backgrounds. For example, studies have identified a genetic association between MCCC1 (which encodes for methylcrotonoyl-CoA carboxylase 1 protein) genotypes and age at onset, motor progression, and the overall risk of developing PD [102]. ALAS1 (aminolevulinic acid synthase) encodes a protein with the same name whose function is related to biocatalysis of the aminolevulinic acid [103], a non-proteinogenic amino acid that plays an essential role in neuronal survival. Variation in its concentration can cause morphological and functional changes [104], which have been thought to be associated with PD.
Mitochondrial disruptions affecting the internal structure of mitochondria have been associated with PD symptomatology [105]. COQ7 is critical to maintaining mitochondrial integrity and CoQ synthesis [106]. CoQ is essential in the mitochondrial electron transport chain, and mutations in COQ7 have been linked to PD development [107]. Ebadi et al. also demonstrated that the action of neurotoxin 1-methyl-4-phenyl-1,2,3 6-tetrahydropyridine (MTPT) could induce Parkinson-like symptoms, which acts by inhibiting complex I [108]. GWAS findings in the Chinese population further support this locus’s association with PD susceptibility [106,109].

2.4. Lysosomal Pathway

Lysosomes allow the breakup of old and unnecessary structures, relying on digestive enzymes [110]. The lysosomal pathway and its role in the degradation of protein aggregates have emerged as a critical PD pathway because of its close relationship with the autophagic pathway. Lysosomes are essential for α-syn degradation, which forms the deposits responsible for the Lewy bodies and Lewy neurites observed in PD [111]. Traffic disruption occurs when lysosomal genes are altered, leading to intra-lysosomal buildup followed by persistence of α-syn in neurons, which ultimately results in apoptosis, as shown in Figure 1b [96]. For this reason, the lysosome pathway and its associated genes have been related to PD, including variants of GBA1, TMEM175, CTSB, and ATP6V0A1 genes [56].
The GBA1 gene encodes for lysosomal hydrolase β-glucocerebrosidase (GCasa), responsible for glucose hydrolysis. This ceramide glucocerebroside protein has shown a strong association with an increased risk of idiopathic PD with aging [112] and early-onset Parkinson’s with a higher incidence of neuropsychiatric symptoms [113]. Studies have shown that populations with both homozygote and heterozygote mutations in the GBA1 gene with no PD symptoms show deteriorated motor abilities, cognition, and olfaction [114]. Thus far, it has not been possible to define the exact mechanism by which GBA1 mutations mediate PD’s pathogenesis. However, it is known that not all carriers of the identified gene variants develop PD, providing further evidence for the complex multigenic nature of PD [115].
The TMEM175 gene, another PD risk locus [109], encodes for a transmembrane (TM) endolysosomal potassium channel. The TMEM175 p.M393T variant alters lysosomal and mitochondrial function, increasing the aggregation of α-syn, as well as the potassium conductance and luminal pH stability. Recent studies have shown a relationship between GCasa reduced activity and alterations in the TMEM175 gene due to loss of channel integrity. Nevertheless, other genes could also be associated with such reduction in activity since TMEM175 p.M393T and LRRK2 p.G2019S variations explain only 23% of the variance in the GCasa activity [109]. Variations of the TMEM175 gene can also cause a reduction in glucocerebrosidase activity, a deterioration in the autophagosome’s clearance by the lysosome, and a significant decrease in the mitochondrial respiration processes, which have also been strongly associated with PD [57,116].
The CTSB gene, belonging to the mitochondrial pathway, generates multiple protein products, including cathepsin B (CTSB). Ming Man et al. show that the CTSB protein ignites the MCOLN1/TRPML1 calcium channel in lysosomes, which maintains the suppression of the TFEB transcription factor and decreases protein expression related to lysosomes and autophagy [117]. Additionally, single-nucleus sequencing on postmortem brains shows that the CTSB gene is only expressed in neurons and microglia, suggesting that the variants observed in CTSB might have a significant impact on these cell types [97,117,118]. Together, these findings indicate that PD most likely involves a significant reduction in the lysosomal protease function [97].
Finally, the ATP6V0A1 gene is expressed in microglia and its cell precursors. It encodes for the α1 subunit of the H + vacuolar translocating ATPase, a heteromultimeric complex responsible for acidifying the compartment of the secreting pathway and the secretion of granules [119]. This is V-proton ATPase functions pumping protons into the luminal environment of the endolysosomal system. The loss of function generates an increase in lysosomal pH and the inhibition of lysosomes and phagosomes fusion in the endolysosomal pathway [120,121].

3. Advances in CRISPR/Cas Systems and Delivery Strategies

CRISPR/Cas are RNA-guided endonuclease systems first discovered as an adaptive defense mechanism of procaryotes and archaea against viral infections [122]. Compared to other gene-editing techniques, CRISPR/Cas9 technology is simple, flexible, and less costly, which has led to its increasing popularity [123]. DNA edits induced by CRISPR/Cas9 systems have contributed significantly to our understanding of diseases by allowing us to evaluate the role of gene candidates, creating relevant cell lines and animal models, and finally, it can be a powerful tool for efficient and safe gene therapies in the near future [21]. CRISPR/Cas9 systems have been further adapted to allow for many genome manipulations beyond site-directed gene editing and are now used to modify the expression of a gene by modifying a catalytically dead nuclease (dCas9) to perform precise base edits [124,125], known as CRISPR activation (CRISPRa) and CRISPR interference (CRISPRi) technologies [126,127]. Moreover, they have been further developed into tools for epigenetic research [128], gene location detection [129], or even modified RNA targeting [130].

3.1. CRISPR/Cas Component Design for Precise Gene Editing

The success and precision of CRISPR/Cas9 applications depend largely on proper sgRNA design for the intended gene target. Unspecified genome editing, also known as off-target editing, is a significant obstacle to overcome for an eventual translation of CRISPR technologies into clinical practice. Off-target editing occurs when the endonuclease binds and cleaves a genome region other than the expected one, thereby generating unintended mutations [131]. Off-target editing could occur by several mechanisms, including the unspecific coupling of sgRNA to unplanned regions of the genome, 5-base mismatch in the endonuclease-specific protospacer-adjacent motif (PAM)-distal sequence of sgRNA, and DNA methylation that impedes biding of Cas9 [132].
Multiple tools have been developed to design sgRNA and predict possible off-target effects in silico based on mismatches and bulge size [133,134,135]. RNA guides are designed according to the specific target organism, DNA-target sequence, PAM sequence, off-target effects, and the transcriptional start site (TSS) [136]. These tools use various algorithms to design sgRNA while examining the genome of the target organism to avoid the off-target caused by sequence similarity [137]. However, multiple sgRNA should be tested simultaneously for in vivo and in vitro studies due to possible unexpected outcomes [138]. Currently, further cutting-edge methodologies based on deep learning are under development to improve sgRNA design [139].
Off-target modifications remain possible even after careful sgRNA design. Zhang et al. [140] found that a sgRNA designed in silico with no predicted off-target activity still caused three silent mismatches in Arabidopsis, one of them in the proximal PAM sequence (8th nucleotide) and causing high off-target activity. Despite being located at a significant distance from the PAM sequence, the other two mismatches also had attributable off-target effects. Similarly, Kleinstiver et al. performed a study comparing wild-type SpCas9 and a SpCas9-HF1 variant and found that even one pair base mismatch in the proximal section of the sgRNA significantly diminished SpCas9-HF1 gene-editing efficiency, in comparison to the wild-type SpCas9 [141]. These findings are only a few examples highlighting the importance of optimizing the design of sgRNA and improving computational and experimental methods for off-target effects detection and quantification. This is important to mitigate the possible off-target effects of CRISPR-based editing technology.
In addition to sgRNA design, another critical feature for accurate and efficient genome editing is the intracellular bioavailability of the CRISPR elements. This refers to the proportion of gene-editing components able to bind to intracellular targets [142]. Rapid degradation of CRISPR elements due to changes in factors such as pH and electrochemical gradients impedes accurate genome editing. Changes in pH conditions may cause protein denaturation, affecting active sites of interest and inducing degradation reactions [143]. Similarly, reducing conditions, often found in diseased cells, may cleave dimer cysteine disulfide bonds that maintain tertiary and quaternary structures together [144]. Therefore, delivery carriers for CRISPR elements protecting them from rapid degradation can be an efficient solution for boosting the bioavailability in target cells [142,145,146,147].

3.2. Delivery Strategies for CRISPR Applications

CRISPR component delivery strategies for gene editing can be grouped into three general approaches. First, the sgRNA and nuclease mRNA directly delivered into the cell/tissue/embryo. Second, sgRNA is delivered along with the synthesized nuclease protein called a ribonucleoprotein complex (RNP) [148]. Finally, the third strategy focuses on combining both sequences of sgRNA and nuclease in a single plasmid, which the target cell translates, as shown in Figure 2a [149]. The choice of delivery strategy entirely depends upon the characteristics of the specific CRISPR application. The delivery strategy for CRISPR elements can significantly impact the gene-editing success and the duration of the nuclease activity [150]. Each delivery strategy has different limitations and advantages that should be considered when choosing the CRISPR system of interest for the intended application [151].
RNP delivery has shown a lower off-target activity and faster genome editing than plasmid-nuclease delivery [152]. When choosing to deliver RNP directly into the cell, the specificity of CRISPR/Cas gene editing is highly dependent on nuclease purity, as well as its source organism. The preferred nucleases are from Staphylococcus aureus and Streptococcus pyogenes [131,153]. Recombinant nuclease alternatives have emerged in recent years to enhance editing specificity. Notably, modifications of Cas9 nucleases alter the nuclease-specific PAM sequences and ultimately enhance specificity while maintaining suitable editing efficiency. Studies have compared the activity of different Cas9 variants to find one that is optimal for gene-editing applications [154,155]. Some Cas9 variants such as wild-type SpCas9 and Sniper-Cas9 have high activity but low specificity; meanwhile, some other variants such as evoCas9 show higher specificity and therefore are optimal for avoiding significant off-target effects, even if there are slight mismatches between the sgRNA and the target sequence [155,156].
In contrast to RNP, plasmid-nuclease can generate a constitutive expression that leads to a more extended genome edition [157] and is a more stable and cost-effective alternative to RNP delivery. However, plasmid-endonuclease delivery is significantly limited by plasmid size, where a plasmid over 7 kb can pose significant delivery challenges [149,158,159]. This is especially relevant for some CRISPR applications where a tissue-specific plasmid design is required to address possible variations in the promoters [160]. Since designing a single plasmid vector with all the CRISPR components can be challenging, promising results have been obtained using separate plasmids with the nuclease and sgRNA sequences [161,162]. Finally, even though the delivery of RNA sequences encoding for nuclease is an attractive strategy, it can be difficult in some scenarios due to their rapid degradation [163].
The selection of delivery carriers for different applications can also have an enormous impact on the success of CRISPR gene editing because they directly influence the bioavailability of CRISPR components and gene-editing specificity. Recently developed nanoscale carriers have been shown to significantly improve plasmid delivery efficiency by altering the possible intracellular trafficking routes for the vehicle [164]. For example, PEGylated nanoparticles can increase efficiency by 40% compared to free plasmids in HeLa and NIH-3T3 cells [164]. The general mechanism of trafficking for a delivery carrier transporting CRISPR components follows four major steps: (i) cell internalization via numerous pathways including clathrin-mediated, caveolin-mediated, or clathrin-caveolin independent; (ii) maturation of the early endosomes; (iii) endosome escape by protonation, charge modification or specific membrane interactions (e.g., changes in endosome osmotic pressure, electrostatic adsorption on delivery carriers and insertion of dendrimers on endosome membrane); and finally, (iv) transfection of CRISPR components for later transcription/translation, as shown in Figure 2a [149,165,166].

3.3. Delivery Carriers to Improve the Performance of CRISPR/Cas9-Based Therapies

Even if all the components necessary to perform CRISPR gene editing can be delivered directly into a cell or tissue, using vehicles to transport RNAs, plasmids, and/or nucleases is advantageous to increase editing efficiency [29,167] and to reduce possible immune responses that can occur in the host due to random off-target integration into its genome [168]. Viral vectors are often used as delivery carriers. Unfortunately, they have multiple disadvantages, including activating a cyclic GMP-AMP synthetase, which could ultimately trigger apoptosis in some cell lineages [168,169,170]. Consequently, independently of the exact CRISPR chosen, a carrier is highly recommended to bring stability, prevent degradation, reduce off-target effects, and increase the target cells’ transfection levels [171]. Moreover, delivery carriers can avoid sgRNA and nuclease degradation even after passing through organ and physiological barriers such as the BBB, the extracellular matrix, and the P-glycoprotein transport system [158,172]. Some delivery carriers might escape endosomes efficiently and, consequently, lead to higher transfection levels and ultimately achieve genome editing efficiencies of up to 70% in different human cells [173,174]. CRISPR delivery carriers include viral vectors, lipid micelles, and nanostructured materials, as shown in Figure 2b [29,171]. Furthermore, there are several delivery strategies where the genetic material is directly incorporated into the target tissues by physical means [175,176].
Viral vectors were the first vehicles to achieve the successful delivery of CRISPR genome editing components. The most successful viral vectors are lentivirus, retroviruses, adenoviruses, and adeno-associated viruses [177]. In all cases, these carriers can deliver sequences of 4.5 up to 5 kb, which usually include the sgRNA and, depending on the virus, a small regulatory portion with information for promoter and polyadenylation sequences [172,178]. Multiple studies have repeatedly shown how viral vectors can be successfully used in gene editing [179,180,181], with numerous clinical applications [182]. Among viral vectors, adenoviral vectors’ have unique features, such as increased carrying capacity and non-pathogenic properties, that led to their increasing popularity in gene editing and CRISPR implementation [183,184]. AAVs are the preferred choice among adenoviral vectors, even if inflammatory reactions have been reported as a possible complication [185]. The use of AAV in multiple clinical trials demonstrated their safety and biocompatibility, leading to their approval by the FDA [180]. Although AAV insertion into the host genome is possible, most studies have shown that it is uncommon and targets the mitochondria. AAV is thus considered a safe approach [186], and it is frequently used as a delivery vehicle in gene-editing applications [187,188,189].
Some alternative AAV versions have been successfully tested to improve viral capacity and specificity. In general, modifications focus on altering the capsid proteins. In particular, AAV concatemers can drive long-term transgene expression due to their stable life cycle [161,180]. Peptides added in the VP3 region of the AAV genome generate vector re-targeting and thus enhance specific-organ transduction [190]. Consequently, modifications on the VP2 region can affect viral delivery efficiency and the transduction capacity of the vector [190]. Capsid engineering is a suitable route to accomplish specific requirements of a particular gene-editing application and can efficiently target the nervous system and crossing the BBB due to the tropism inherent to viruses [191]. For CNS gene editing, this modification is crucial, as transduction in CNS can fail due to the BBB’s limited permeability [192]. The BBB is adapted to filter molecules larger than 400 Da, and thus it can be a significant challenge for viral vectors to cross this barrier [193].
Viral vectors can also enable proper transduction processes for both dividing and nondividing cells in a wide variety of lineages, including hematopoietic, hepatic, and T-cells [145,172,194]. Viral vectors can accomplish simultaneous co-delivery of endonuclease and single or multiple sgRNA sequences for applications where multiplex knock-out might be needed, e.g., autism spectrum disorder, post-mitotic neurons, and schizophrenia [172,195,196]. However, viral vectors exhibit limitations in their loading capacity, preventing complex sequence delivery containing markers such as GFP and td-Tomato [162]. Likewise, long-term insertional activation mechanisms using CRISPR are difficult to control by viral vectors, thereby generating toxicity in some organs in vivo [194,197]. Moreover, AVV vector-based treatments can be relatively expensive and therefore are prohibitive for some CRISPR and gene delivery applications [197,198].
Lipid-based carriers for biotechnology applications are formed by a self-assembly process with amphiphilic molecules (with hydrophilic heads and hydrophobic tails) dispersed in aqueous solutions. In this process, the molecules’ hydrophilic heads come together to form a layer facing the aqueous solution while the hydrophobic tails stack together in a core [199]. These unique supramolecular structures or micellar systems are well-suited for encapsulation and transport cargoes with different physicochemical properties ranging from small hydrophobic drugs to large biological molecules, as shown in Figure 2b [200]. It is possible to create various delivery systems such as liposomes, lipid nanoparticles, and nano- and microemulsions [201]. Each system has its advantages regarding physicochemical properties that can be adjusted for different applications. For example, nano-emulsions are typically <300 nm and are kinetically stable, but microemulsions are thermodynamically stable and can present different geometries (e.g., worm-like, hexagonal, and liquid crystalline) [202]. To facilitate RNA and DNA loading, the carriers can be formed by incorporating cationic lipids as they readily form complexes with the negatively charged nucleotide sequences [199]. Therefore, lipid carriers are well-suited to deliver the plasmids containing the sgRNA and endonuclease sequences of CRISPR editing systems. Moreover, lipid-based vehicles have proven beneficial for delivering RNP or even co-deliver nucleic acids and proteins [29].
Lipid-based carriers have been successfully used in various applications, showing reduced off-target effects [203]. For instance, a self-assembled micelle delivery system for plasmid-nuclease delivery to induce human papillomavirus (HPV) E7 oncogene disruption led to a significant reduction in HVP cancerous activity. Cellular uptake was quantified using fluorescence-activated cell sorting (FACS), which showed a peak of Cas9 at 24 h. Thus, a fast turnover rate can be associated with low off-target effects in comparison to previous studies. CAS9/GFP expression increased by nearly 35%, eight hours after delivery using this F127/PPO-NMe3/pCAs9 micellar system in HeLa cells compared to controls, with no detectable off-target editing [203]. Other examples have proved the potential of nano-liposomal particles to deliver CRISPR systems [204,205].
The ease of synthesis and absence of immune response makes lipid-based carriers perfectly adapted to gene delivery applications [206], leading to their approval by the FDA in 1995 [207]. However, low colloidal stability and poor performance have limited their full implementation in CRISPR applications in the long term. Moreover, some studies have shown that lysosome degradation still occurs despite internalization, resulting in low gene-editing efficiencies [29]. Several chemical modifications can be used to overcome these limitations, including peptide and protein attachments, folate conjugation, and PEGylation [208,209,210]. PEGylation is the most popular alternative because it efficiently prevents enzymatic degradation, increases systemic circulation stability, and reduces clearance and charge-based contact with proteins and small molecules [211]. Another modification of interest is the conjugation of lipid-based carriers to metallic nanoparticles (e.g., gold and silver) to form complexes that favor circulation time due to controlled adhesion of plasma proteins and phagocytosis [212]. Recent in vivo studies with lipid-based carriers have demonstrated promising genome editing results. For instance, Cheng et al. developed selective organ targeting (SORT) nanoparticles based on charged lipids 18PA and DOPA for intravenous delivery of Cre recombinase mRNA in a td-Tomato murine model. Their results indicate accurate, independent genome editing in the spleen, lungs, and liver [213]. Lipid-based carriers have been successfully employed in multiple gene-editing applications [206,214].
Inorganic nanostructured carriers have emerged as potent delivery vehicles due to the consolidation of nanotechnology as a mature field to produce materials with highly controlled physicochemical properties, including size, morphology, crystalline structure, surface charge and chemistry, and 3D topology [29,215,216]. Moreover, their synthesis schemes have improved over time, resulting in high replicability, low costs, and simple synthesis routes. These carriers are also highly biocompatible and have chemical versatility, enabling surface conjugation with many molecules, including polymers such as polyethylene glycol (PEG) and polyacrylic acid (PAA) that increase stability, biocompatibility, and solubility by avoiding aggregation. Their potential application in the delivery of CRISPR technologies is not the exception since inorganic nanostructured carriers can carry high loads, improve the stability of the transported cargoes, and facilitate cellular uptake and specific sub-cellular targeting [146,217,218]. Additionally, it is possible to conjugate inorganic nanostructured carriers with different types of biomolecules superficially, such as antibodies, polymers, vitamins and proteins, and peptides [146,217,218,219] to improve cell penetration and endosomal escape. Functionalization of nanomaterials with translocating proteins or peptides such as OmpA, Buf-II, INF7, and GALA helps to escape an important fraction of endosomal compartments upon cell internalization, which can enhance transfection efficiencies in gene delivery applications [146,173,220].
Thus far, the most explored inorganic nanomaterials for drug delivery and gene-editing applications include carbon nanotubes (CT), gold nanoparticles (AuNP), silica nanoparticles (SiNP), magnetite nanoparticles (MNP), and graphene oxide (GO) [173,206,221,222]. More recently, nanohybrids have gained significant attention due to the possibility of enabling inorganic-inorganic, organic-inorganic, and bio-inorganic conjugation of materials, as shown in Figure 2b. Naturally, improved benefits for drug and gene delivery can be achieved due to the unique properties that each material brings to the nanohybrid [223]. Recently explored nanohybrids include magnetoliposomes, CdTe/ZnTe, and gadolinium-doped hydroxyapatite nanoparticles [224,225,226]. The possibility of manipulating the surface chemistry facilitates different conjugation strategies for the nucleotide sequences depending on responsiveness and colocalization monitoring as the molecules penetrate cells. Additionally, it is relatively straightforward to co-immobilize recombinant Cas9 at different ratios to optimize the editing process as needed and as a function of target cell line and expected off-targets [227]. Cell internalization and the subsequent intracellular localization of nanostructured carriers are primarily dependent on physicochemical properties such as size, surface charge, steric hindrance, morphology, and hydrophobicity [165].
Besides inorganic nanoparticles, polymeric nanoparticles have gained significant attention over the past few years due to their high stability, ease of synthesis and functionalization, high permeability through biological barriers, and solubility at physiological conditions [228,229]. Examples of polymeric nanomaterials for gene and drug delivery include polyethyleneimine (PEI), cyclodextrins (CD), poly-L-lysine (PLL), and poly(lactide-co-glycolide) (PLGA) [230,231,232,233]. In vivo studies have shown that these nanoparticles can be completely degraded in the body after 48 h in murine models [234,235], making them a sound delivery system for gene-editing applications.
Dendrimers have emerged as reliable drug delivery vehicles due to unique attributes such as monodispersity and well-defined chemical structure [236]. Moreover, due to their versatile building blocks (also known as generations) and superficial functional groups, it is possible to tune specific interactions (either electrostatic or covalent) with an ample variety of biomacromolecules (e.g., therapeutic proteins and polynucleotides). In the case of gene therapies where nucleotide sequences need to form a stable complex with the delivery vehicle, cationic dendrimers have been exploited to develop dendriplexes that can protect the genetic material from degradation. This approach is also favorable to overcome endosomal entrapment as this type of dendrimers can escape them by the proton sponge mechanism. This capability has been attributed to the amine groups in their structure with pKa values at or below physiological pH [237]. Some of the most valuable dendrimers that can be tailored according to the final application include polyamidoamine (PANAM), polypropyleneimine (PPI), poly-l-lysine (PLL), phosphorus (PPH), and linear-dendritic block copolymers (LDBC).
Various dendrimers have been designed and tested for brain targeting, and much effort has been invested in their surface engineering to assure that they can successfully come across the BBB while maintaining high biocompatibility, superior drug-release kinetics, and specificity toward the CNS cells [237,238]. For instance, hydroxyl-terminated PANAM dendrimers have demonstrated targeting of astrocytes and microglia as they pass through the BBB largely intact. Another issue of some dendrimers (e.g., cationic) is their high cytotoxicity, which exacerbates as the number of generations increases. This major hurdle has been addressed by conjugating highly biocompatible neutral or negatively charged moieties such as PEG, carbohydrates, and acetyl groups [239]. For gene-editing purposes in the brain, Taharabaru et al. tested PANAM dendrimers modified with cyclodextrins to deliver f Cas9 RNP both in SH-SY5Y cells and in vivo in eight-week-old BALB/c mice. Their results showed higher genome editing activity than Lipofectamine and Lipofectamine CRISPRMAX [240]. Dendrimer engineering with angiopep-2 peptide (which is capable of targeting LRP1 on the BBB) has also been exploited for higher delivery efficacy as it exhibits high transcytosis capacity and parenchymal accumulation [241]. This has also been the case of the RVG29 peptide, a 29-amino-acid peptide that stemmed from the rabies virus glycoprotein (RVG29), which targets the nicotinic acetylcholine receptor on the BBB [242]. The potency of dendrimers has also been exploited to develop multimodal therapies where small pharmacological molecules are co-delivered with gene therapies in search of possible synergistic effects. In addition, they have been coupled with polymeric nanoparticles (e.g., PLA and gelatin), quantum dots, and bacterial magnetic nanoparticles to form hybrid systems exhibiting unique structural and functional properties unattainable with the independent components [236].
An essential aspect of nanocarriers’ delivery is their surface charge, as it can directly alter their possible interactions with proteins and cell membranes. This, in turn, might lead to different routes for nanoparticle degradation in vitro and in vivo. In this regard, endothelial (HUVEC) cells and macrophages (Kupffer) have shown a marked tendency to internalize gold nanoparticles, which subsequently compartmentalize into endosomes and lysosomes [235]. Phagocytic degradation appears avoided when nanoparticles have a slightly negative surface charge [243,244]. In general, a low absolute surface charge, known as zeta potential, has been reported to prevent nanoparticle degradation both in vitro and in vivo. For instance, a recent study demonstrated that PEG-oligocholic acid-based micelle nanoparticles exhibiting high surface charge (zeta potential > 15 mV) showed a rapid degradation by macrophages and higher liver accumulation [243]. In contrast, nanoparticles with a slightly negative surface charge (<−8.5 mV) appeared to induce significantly less hemolytic and cytotoxic effects and reduced undesirable clearance by the reticuloendothelial system (RES) [243]. Finally, in SORT nanoparticles, slight changes in surface charge can lead to different organ targeting [213].
Another critical characteristic of nanostructured carriers is size and distribution. Recent delivery reports suggest that the proper size distribution of nanocarriers for drug delivery applications needs to be between 20 to 1000 nm [245,246]. Nanostructured carriers’ size is crucial to minimize clearance mechanisms, prevent aggregation, and avoid mechanical retention within capillaries. For instance, nanoparticles with average diameters <100 nm avoid sequestration by sinusoids in the spleen, while those with >20 nm evade kidney filtration [247]. Cell internalization of nanoparticles <200 nm is enabled by clathrin-mediated endocytosis, which needs to be cleared by nanocarriers to assure high gene-editing efficiencies [248]. Internalization through this mechanism starts by membrane events where clathrin-coated pills are recruited and eventually cleaved to form clathrin-coated vesicles. Then, these vesicles fuse with early endosomes and mature to lysosomes, where nanoparticles are either degraded by enzymatic action or are sent back to the cell membrane [249].
The last important nanoparticle characteristic that needs consideration is hydrophobicity, which has been linked to the route of nanomaterial degradation under physiological conditions where interactions with immunoglobulins and other plasma proteins are highly favored. As a result, changes in hydrophobicity make it possible to enhance capturing by the reticuloendothelial system (RES) [248]. Therefore, conjugating nanomaterials with hydrophilic polymers such as PEG, PLGA, poloxamer, chitosan, and poly (ethylene oxide) is a common strategy to avoid RES capture while improving biocompatibility [250]. Moreover, PEG polymers adsorbed on nanoparticles form a hydrophilic steric barrier that prevents clearance by macrophages or interactions with plasma proteins that ultimately triggers carriers’ clearance [197]. In vitro studies have also shown that some polymers might have a stabilizing role as demonstrated by conjugating Poly lactic-co-glycolic acid (PLGA), poly(oxyethylene), poloxamer, and chitosan [248], allowing PLGA nanoparticles to be highly stable under physiological conditions [251]. In addition to enhancing nanoparticle stability and bioavailability, hydrophilic polymers can also improve biocompatibility (most likely due to lack of recognition by the immune system) and cell uptake [252].
Nanocarriers can also be modified with stimuli-responsive moieties such that their cargoes can be delivered when nanocarriers reach specific organs or tissues [253]. For instance, our research group developed a pH-responsive nanocarrier combining a pH-responsive polymer (pDMAEMA) to core/shell magnetite/silver nanoparticles. These nanocarriers load and deliver plasmids in response to changes in the pH [220]. Other pH-sensitive nanocarriers that release their cargoes under reducing conditions further support the potential of these nanocarriers for the delivery of the CRISPR components [254].
Thus far, little is known about the use of nanostructured carriers for the delivery of CRISPR editing systems. Only a few studies have been conducted in vivo to estimate the true potential of delivery vehicles in enhancing transfection and on-target efficiencies. Among these studies, Lee et al. show that gold nanoparticles (AuNP) were able to deliver a CRISPR RNP system to successfully edit the mGluR5 gene in the brain, rescuing exaggerated repetitive behaviors in a fragile X mouse model [29,95]. Wang et al. developed multi-AuNPs-lipid thermo-sensitive complexes to deliver CRISPR plasmids, achieving a release efficiency of 74% in mice [206]. Similarly, in vitro studies have used magnetic nanoparticles to introduce a complementary magnetically responsive adjuvant in a CRISPR application involving viral carriers, achieving a more precise genome editing [255]. This study proved that nanostructured vehicles could also be used for CRISPR editing systems as adjuvants to take advantage of their unique physicochemical properties. Although nanostructured vehicles have already enabled several gene delivery applications in recent years, issues regarding biosafety, wide particle size distribution, and low efficiency of surface chemical modifications are yet to be resolved prior to a full incursion pre-clinically and clinically [255,256].
Another important aspect to consider in CRISPR gene-editing applications is the delivery method for the vehicle carrying CRISPR components. Physical delivery is the most frequently used strategy for CRISPR systems in vitro and ex vivo delivery. This method dramatically increases the levels of readily available genetic material delivered [159]. Physical delivery standard techniques include microinjection, electroporation, and hydrodynamic delivery [29]. Although these techniques have not been specifically designed for in vivo studies due to cell structure damage, some reports have used physical delivery in zygotes to develop ex vivo transgenic animals [257]. Microinjection methodology was first used to generate transgenic animals by directly introducing DNA into the pronucleus [258]. On the other hand, electroporation is an alternative in which electrical currents alter the membrane potential. This creates transient nanopores to allow CRISPR elements to penetrate the cell by a concentration gradient [29]. Liang et al. achieved a 56% targeted integration efficiency for Cas9 RNP and ssDNA donors in HEK293 cells via electroporation, and precise genome editing rates of about 45% were reached in induced pluripotent stem cells (iPSCs) [259]. Compared with microinjection, electroporation has several advantages, including ease of implementation, its applicability in a broad range of cell types, simultaneous genome editing of different populations, and major genome editing in knock-out zygotes [260]. Lastly, hydrodynamic delivery takes advantage of the rise in permeability of specific cell types upon increasing the surrounding hydrodynamic pressure. This is performed by adding a large volume of solution to the target organs containing the CRISPR elements. Thus far, this approach has been the only physical delivery method used for in vivo testing of CRISPR systems [261].
Delivery carriers have recently emerged as excellent facilitators when using CRISPR systems for neurodegenerative disease research and treatment development. Encouraging results have been obtained using delivery carriers in Alzheimer’s, Huntington’s, and Parkinson’s disease research [262,263,264]. For instance, Park et al. developed nanocomplexes using the R7L10 amphiphilic peptide to deliver a silencing system of the Bace1 gene, related to the accumulation of Aβ peptides, one of the main hallmarks of Alzheimer’s disease [262]. In Huntington’s disease, CRISPR/Cas9 has been used as a method for an early diagnosis using an amplification-free long-read sequence of the HTT gene [263]. Lastly, CRISPR application in Parkinson’s includes targeting the LRRK2 gene, linked to familial PD. An in vitro study on hiPSC showed that an LRRK2 knock-out led to a reduction in TH-positive neurons [264]. For detailed information on gene therapies to treat neurodegenerative diseases, an excellent review by Karimian et al. can be consulted [265].

3.4. Perspectives on Delivery Carriers and Potential for Future Research

CRISPR is a constantly evolving genome editing technique, and therefore, there is still much room to continue improving and developing new multifunctional vehicles capable of passing through different biological barriers to reach the intended target cells. This could also contribute to minimizing off-target effects such that the process of reaching clinical implementation might be accelerated. In addition to designing proper delivery vehicles, the main CRISPR components need to be carefully planned. As mentioned, an appropriate design of the sgRNA is critical for accurate genome editing, as is the choice of nuclease delivery strategy. Depending on the intended CRISPR applications, each of the nuclease delivery alternatives may be preferable for implementation by considering that, for instance, protein delivery leads to faster results than mRNA, but mRNA delivery provides a longer-lasting effect.
Delivery carriers enhance transfection and provide protection and stability for CRISPR components. Therefore, the latest CRISPR studies employ a wide range of delivery carriers primarily defined by the physicochemical properties that impact gene-editing efficiency. Nanostructured vehicles are the most promising carriers in gene delivery applications due to their chemical versatility. It is possible to modify their surface charge, chemistry, and hydrophobicity to increase stability and eventually avoid rapid lysosomal degradation. In the same way, nanostructures can be combined with other vehicles, adjuvants, or physical delivery strategies to maximize internalization and transfection while minimizing possible off-target effects synergistically. Unfortunately, in vivo studies with CRISPR systems delivered by nanostructured materials remain scarce [30]. Tight international legislation controlling the use of carriers in vivo and the contradictory results regarding their biocompatibility pose challenges that must be overcome before delivery vehicles can move into clinical applications. For now, the pharmacokinetics of delivery carriers remains poorly understood, and we need more research to further understand the various aspects involved in using nanocarriers in gene editing and therapies.
As expected, only a few CRISPR applications have been developed and successfully applied for PD treatment. Although CRISPR has been used to study the genetic causes of several diseases, it also offers a promising application as a gene therapy tool. Therefore, as a more comprehensive arsenal of delivery platforms becomes available, chances are high for developing safe and effective gene therapies that offer treatment alternatives to PD patients. For instance, a more comprehensive range of options for gene therapy will be readily available, along with the use of vehicles for applications that require gene knock-out, knock-in, repression, or activation.

4. Potential of Genetic Therapies as Treatment Alternatives for Parkinson’s Disease

Gene editing offers unprecedented potential to understand the molecular basis of PD, identify new treatment targets, and ultimately develop new gene therapies. Correcting dysfunctions in the biological pathways associated with PD can be accomplished by selectively editing or up- and down-regulating gene expression in key genes known to become altered in PD, including GDNF, PINK1, PRKN, or AADC [6]. Even if idiopathic PD has a complex multigenic basis and could not be treated by correcting variants in a single gene, gene editing could still represent a promising strategy to restore the activity of the key biological pathways that can become disrupted and cause PD symptoms.
To date, gene-editing approaches for PD have been classified into four types according to the chosen therapeutic target [266]. An initial strategy aims to improve dopamine bioavailability in the brain. The second strategy focuses on neuronal regeneration by targeting neurotrophic factors. A third approach focuses on neuromodulation modifying genes in the subthalamic nucleus (STN). Finally, the fourth strategy is based on reducing α-syn production, thus ameliorating altered mitochondrial pathways [267,268,269]. All these approaches aim to modify the metabolic pathways involved in PD and neuron survival, mainly the aforementioned autophagic, mitochondrial, and lysosomal pathways (Figure 3).
The two pathways that have received the most attention in gene-editing studies related to PD are the dopamine pathway and neurotrophic factors. Manipulating neurotrophic factors can stop not only symptoms but also promote neuron survival. Alternatively, strategies involving the dopaminergic pathway mainly involve non-pulsatile stimulation of dopamine production that can drastically improve currently used treatments. The following sections will review gene editing and gene therapy advances in PD (summarized in Table 2).

4.1. Therapeutic Approaches Based on the Stimulation of Dopamine Production

Dopamine is a neurotransmitter involved in motivation, reward, pleasure, cognition, and motor control [285,286]. A dopamine deficiency detrimentally affects the neurons of the SNpC and is directly associated with PD symptoms (muscle stiffness, tremors, anxiety, and sleep disorders) [287]. Therefore, trying to recover this neurotransmitter’s normal levels is the aim of many currently available treatments. A typical strategy to normalize dopamine levels is via oral administration of levodopa (L-DOPA), which can cross the BBB, allowing an efficient treatment [288]. L-DOPA is a dopamine-precursor involved in the final step of the dopamine synthesis process. It is converted to dopamine through the decarboxylation performed by the aromatic L-amino acid decarboxylase (AADC) enzyme [275]. Even if these available treatments correct dopamine levels and may even offer some relief from PD symptoms, administered L-DOPA has low bioavailability (10%), and the fraction reaching the brain is only 1% [288]. In addition, using L-DOPA as a long-term oral treatment generate side effects such as dyskinesias and sleep disorders. [270,289,290]. Gene therapy offers an alternative route to increase dopamine production permanently by modifying the dopaminergic pathway and thus has received much attention in the last two decades.
Multiple studies have targeted the AADC (aromatic L-amino acid decarboxylase) [271] and TH (tyrosine hydroxylase) genes [270]. The TH gene encodes for the tyrosine hydroxylase responsible for converting tyrosine to L-DOPA during dopamine synthesis. Kirik et al. performed an in vivo study in a murine model where AAV was used to deliver the TH gene [270]. In this study, spontaneous and drug-induced behavior improvement was observed in rats with complete or partial 6-hydroxydopamine lesions of the nigrostriatal pathway, the bilateral dopaminergic pathway that connects the SNpC with the dorsal striatum [270]. The authors concluded that local intrastriatal TH delivery might be a viable therapeutic strategy in PD as it offers better control of orally administered L-DOPA’s adverse side effects. Further studies demonstrated the direct relationship between TH/GCH1 gene insertions and rats’ symptom improvement (behavioral recovery) [272]. The GCH1 gene, encoding the GTP cyclohydrolase 1 enzyme, is involved in the production of tetrahydrobiopterin (BPH4) and ultimately triggering dopamine production. However, the observed effect of editing the TH/GCH1 gene in murine models could not be replicated in other animal models [270,271], and the potential of modifying these pathways remains unclear.
Alternatively, it is possible to manipulate the AADC gene [291,292]. Gene therapy based on AADC modifications showed promising results in Phase I trials [275,293]. This therapy showed no complications in trial patients, was well tolerated, and no significant adverse effects related to the delivery vector (AAV) or the gene therapy treatment were observed [275]. Total and motor rating scales, quantified by Unified Parkinson’s Disease Rating Scale (UPDRS), improved over time for patients evaluated in the study. These results were corroborated by Muramatsu et al. [275,277].
Several Phase I studies have reported improvements in UPDRS for PD patients [276,291,292]. A comprehensive review on the use of ADDC in PD research was completed by Hitti et al. [266]. Despite all these promising results, Phase II trials are necessary to validate this approach’s safety and efficacy. Moreover, testing the use of delivery vehicles capable of crossing the BBB to avoid dangerous surgical procedures would be very useful, as these procedures can restrict clinical trials and in vivo studies [294]. Despite all these efforts, treatment alternatives focusing on dopamine stimulation remain a non-disease-modifying alternative [295].

4.2. Therapeutic Approaches Targeting Neurotrophic Genes

Current efforts to treat neuronal loss focus on glial cell line-derived neurotrophic factor (GFL) ligands, including glial cell line-derived neurotrophic factor (GDNF), neurturin (NRTN), artemin (ARTN), and persephin (PSPN). These neurotrophic factors are involved in the maintenance, survival, and differentiation of various neurons, including dopaminergic neurons [296]. As such, these genes represent important gene therapy targets to stop the progress of Parkinson’s disease and promote the generation of new neuronal tissue. Within this gene family, GDNF and NRTN have received the most attention. The therapeutic potential of GDNF for PD has been known for quite some time [297], as it has been associated with reduced neuronal development [298]. Attempts to develop therapeutic applications based on GDNF include the direct injection of the GNDF protein into the putamen, where results have shown a continued survival of lesioned nigral neurons for four months in rats [297] and monkeys [299]. These findings on animal models paved the way for Phase I clinical trials to evaluate the safety and efficacy of this type of therapy [300,301].
Contrary to expectations, Phase II trials yielded disappointing results. These studies concluded that gene editing on GDNF failed to confer a clinical benefit to PD patients [302]. It was uncertain whether technical differences between this trial and open-label studies contributed to this negative outcome [302]. Later research attributed this failure to inadequate diffusion of the GDNF proteins in the brain putamen [303]. New attempts to resolve this problem continue to emerge and have achieved protein diffusion throughout all the putamen [304]. Using nanostructured vehicles for delivery can be an alternative for boosting the bioavailability of proteins and improve clinical trial results [280].
Gene therapy presents a unique advantage to permanently correct PD abnormalities through modifications in the GDNF pathway, permanently increasing the production of GNDF. Chen et al. successfully evaluated this possibility showing that motor ability improved after gene editing in MitoPark mice, an animal model for PD [281]. In this novel study, the authors used macrophages that expressed GDNF for treatment and managed to increase the expression of GDNF up to three times compared to the control in plasma, SNpC, and striatum [278]. In addition, no evidence of side effects of macrophage-mediated GDNF therapy was detected in the study. This gene therapy evaluation was carried out parallel to the direct injection of the protein into the tissue, showing similar results to viral vectors [279,282]. Studies targeting GDNF for Parkinson’s disease treatment include various administration vehicles. For example, there are studies with encapsulated protein-producing cells [305], microspheres [283], cationic microbubbles [284], and BBB-penetrating nanoparticles (BPN) [306]. A more recent approach uses SINEUP-RNA delivered by AAV to increase the GNDF gene’s translation, demonstrating effective results in a PD mouse model by improving motor deficits and ceasing neurodegeneration [307].
The NRTN is another essential neurotrophic gene in PD, showing promising results as a therapy target. However, the therapy’s effectiveness was again disappointing in double-blind trials after initial validation of AAV-mediated NRTN delivery therapy in animals [308,309]. In Phase I trials [310], no differences were found between treatment and control groups [311]. Further analyses carried out postmortem 8–10 years after treatment revealed that NRTN had limited expression and coverage (~3%–12% of the putamen and ~9.8%–18.95% on the SNpC) in study patients. Moreover, it was determined that there was no difference in the degree of Lewy pathology between the treated group and untreated patients with Parkinson’s disease [312]. These findings are consistent with the results of previous clinical trials and currently halt further trials targeting the NRTN gene [310,311].

4.3. Other Gene Therapy Approaches

4.3.1. Approaches Targeting Mitochondrial Genes

Among genes in mitochondrial pathways, PRKN and PINK1 have been essential targets in PD research [313]. PRKN and PINK1 have been mainly associated with genetic PD [314] but have been considered targets to improve mitochondrial pathway function in all PD types [315]. Currently, strategies focusing on mitochondrial pathways are scarce, and there are no Phase I trials evaluating strategies targeting these genes. However, progress has been made in understanding these genes’ role in the disease and the possible therapeutic implications for addressing their alterations [316]. One example is the study of Yan et al., who used gene-editing techniques to understand the mitophagy process in detail. They corroborated the direct relationship between damage in PINK1 and PRKN with PD [317]. It has been shown that the natural replacement of damaged PINK1/PRKN genes generates protection against PD [267]. Koentjoro et al. demonstrated that Nix (Nip3-like protein X of the mitochondrial autophagy, a protein that induces autophagy) could function as a protective molecule, preventing a carrier of homozygous PRKN mutations from developing PD [267]. Using human fibroblasts, the study confirmed that Nix could facilitate mitochondrial clearance and, therefore, supports mitochondrial function despite the lack of the PINK1/PRKN pathway. Consequently, the Nix gene has therapeutic potential for PD treatment [267]. This approach can also provide a solution for idiopathic PD since proper regulation of these genes could favor a PD patient’s mitochondrial proliferation and health [315]. As shown by Chung et al., who developed PARKIN-permeable cells, its delivery increases mitophagy, mitochondrial biogenesis and suppresses α-syn accumulation in cells treated with mitochondrial toxins (sodium arsenite and rotenone) [317].

4.3.2. Approaches Focused on α-Synuclein

Parkinson’s disease is characterized by the formation of Lewy bodies in neurons (see Section 2, [318]). Therefore, interference with the sequence and expression of the SNCA gene offers a promising alternative to control Parkinson’s disease progression [268,269]. Studies focusing on SNCA as a gene-editing target have demonstrated that its silencing in the hippocampus of mice using an AAV-delivered microRNA can significantly reduce the behavioral deficits associated with PD [268,269]. In a study performed by Ye-Han et al., AAV gene silencing vectors were designed to determine SNCA silencing’s efficiency and specificity against human α-syn (hSNCA) [319]. Although neurotoxicity was observed [319,320] in vitro, data suggest that miRNA-embedded silencing vector may be ideal for SNCA silencing. However, there is a long list of issues and challenges to address before thinking about these studies’ clinical translation. Finally, in a recent study, gold nanoparticle composites were loaded with plasmid DNA (pDNA) to inhibit α-syn expression [321]. The authors observed that nanoparticles improved TH levels and decreased aggregation of α-syn in SNpC. The nanocomposites attenuated motor dysfunction and reversed the inhibition of long-term potentiation (LTP). LTP is one of the main cellular mechanisms that appear to underlie learning and memory. These results indicated that the nanocomposites had significantly neuroprotective effects in motor and non-motor dysfunction in PD mice [321].

4.4. Stem Cell-Derived Therapies and Stem Cell In Vitro Models

Our current knowledge about the pathophysiology and biological pathways of PD has directed research strategies using human stem cells to replace and regenerate dopaminergic (DA) and other cells. There are three main strategies based on the source and type of stem cells: (1) fetal mesencephalic tissue, (2) mesenchymal stem cells (MSCs), and (3) pluripotent stem cells (hPSC), including embryonic stem cells (ESC) and induced pluripotent stem cells (iPSC) [322]. A thorough review of the use of stem cells in PD can be found by Liu and Cheung [322].
Recent years have seen increased interest in the use of gene editing in stem cells to better understand the molecular basis of PD. Stem cells derived from PD patients and induced pluripotent stem (iPSCs) are relevant in vitro models to study the molecular mechanisms of the disease and develop therapeutic strategies [323]. IPSCs are cells that resemble embryonic stem cells by delivering OCT4, Sox2, Klf4, and c-Myc factors via retrovirus vehicles to somatic cells [324]. Furthermore, iPSCs derived from PD patients all for the control of genomic background, providing a personalized model to directly establish [325] and differentiate [326] iPSCs into relevant DA cells and subsequently assess the impact of genetic mutations on disease development and severity.
Integration of the CRISPR/Cas9 systems and the iPSC model offers the possibility of manipulating pathogenic genes (turn off/on), and eliminating phenotypic differences caused by individual inheritance, thereby providing a more direct understanding of the relationship between specific genes and PD. This platform can provide a vast iPSC PD library that allows analyzing the effects of single nucleotide polymorphism (SNPs) and drug response differences [264]. These studies have allowed significant advances in our understanding of the relevant molecular events and mechanisms associated with this disease [327,328,329,330]. Human iPSC technology allowed a better approach to preclinical studies and provided promising clinical trials results [331].
Work on the use of stem cells to study and treat PD paved the way to the 2015 global task force (called G-Force-PD) aimed at bringing the hPSC and iPSC work to clinical trials and to share the obtained information publicly [332]. The possibilities offered by these cells that can differentiate into multiple neuronal lineages as well as their ability to develop into three-dimensional aggregates, known as organoids, combined with gene-editing strategies such as CRISPR, provide a suitable platform in which to study the complexity of the disease much more accurately. Moreover, the development and progression of the disease and the association with genetic mutations can be targeted robustly with these platforms [323]. Studies where these approaches have been explored and studied in detail can be consulted elsewhere [329,333,334,335,336,337]

4.5. Future of Gene Therapy for Parkinson’s Disease

Parkinson’s disease has a complex pathological framework, including multiple metabolic pathways associated with the degradation of substrates and the accumulation of harmful material in neurons’ cytoplasm. The characteristic complexity of idiopathic Parkinson’s and the underlying cause of neuronal loss explain why the initial triggers for the disease remain largely unknown. Strategies based on gene editing aimed to evaluate how the underlying metabolic pathways become disrupted will contribute to our understanding of PD at the molecular level. This knowledge will be vital in identifying new genetic targets to enable more robust and comprehensive treatments of PD.
To date, a great diversity of approaches to treat PD have been evaluated due to the disease complexity. Strategies aimed at increasing dopamine levels, improving neuronal survival, preventing damage in the mitochondria, and preventing α-syn aggregation. These approaches have been extensively studied in cellular and animal models, revealing their potential as targets for future disease-modifying treatments. Clinical trials, however, have revealed challenges and limitations to these new treatment avenues that are yet to be resolved. Significant limitations in new PD treatment development regard delivery issues, including inadequate diffusion of proteins, rapid degradation of delivery vehicles, and viral vector drawbacks. The design and testing of new carriers that enhance transfection and limit immune response, such as multifunctional nanostructured vehicles, have immense potential to overcome these challenges. Furthermore, CRISPR technology now offers an efficient, accurate, and potentially safe toolkit for gene editing (i.e., knock-in and knock-out) or transcriptional modifications (i.e., CRISPRa and CRISPRi) [338]. Delivery of CRISPR components with efficient and safe carriers offers a unique opportunity to correct the biological pathways that become disrupted in Parkinson’s disease, offering a new disease-modifying treatment alternative for both familial and idiopathic PD. More research is still necessary to identify the best gene therapy targets and develop better gene-editing vehicles for superior tissue selectivity, increased safety, and higher on-target edition rates.

Author Contributions

N.I.B., C.M.-C., J.C.C. and L.H.R. conceived the review and were in charge of overall direction and planning. D.A., N.P.E., A.B. ad C.O. designed figures. All authors contributed equally to the literature searchand the writing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Minciencias, grant number 120484467666, and internal funding from the University of Los Andes.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data underlying this article are available in the article.

Acknowledgments

We thank the Department of Biomedical Engineering at the University of Los Andes for the support with laboratory materials, equipment, and graduate student funding. We also acknowledge the Department of Chemical and Food Engineering at the University of Los Andes for allowing us access to their laboratories.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kalia, L.V.; Lang, A.E. Parkinson’s disease. Lancet 2015, 386, 896–912. [Google Scholar] [CrossRef]
  2. Elkouzi, A.; Vedam-Mai, V.; Eisinger, R.S.; Okun, M.S. Emerging therapies in Parkinson disease—Repurposed drugs and new approaches. Nat. Rev. Neurol. 2019, 15, 204–223. [Google Scholar] [CrossRef]
  3. Correia Guedes, L.; Mestre, T.; Outeiro, T.F.; Ferreira, J.J. Are genetic and idiopathic forms of Parkinson’s disease the same disease? J. Neurochem. 2020, 152, 515–522. [Google Scholar] [CrossRef]
  4. Valdés, P.; Schneider, B.L. Gene Therapy: A Promising Approach for Neuroprotection in Parkinson’s Disease? Front. Neuroanat. 2016, 10, 123. [Google Scholar] [CrossRef] [Green Version]
  5. Lindholm, D.; Mäkelä, J.; Di Liberto, V.; Mudò, G.; Belluardo, N.; Eriksson, O.; Saarma, M. Current disease modifying approaches to treat Parkinson’s disease. Cell. Mol. Life Sci. 2016, 73, 1365–1379. [Google Scholar] [CrossRef]
  6. Axelsen, T.M.; Woldbye, D.P.D. Gene Therapy for Parkinson’s Disease, An Update. J. Park. Dis. 2018, 8, 195–215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Alberio, T.; Lopiano, L.; Fasano, M. Cellular models to investigate biochemical pathways in Parkinson’s disease. FEBS J. 2012, 279, 1146–1155. [Google Scholar] [CrossRef]
  8. Connolly, B.S.; Lang, A.E. Pharmacological treatment of Parkinson disease: A review. JAMA 2014, 311, 1670–1683. [Google Scholar] [CrossRef] [PubMed]
  9. Jamebozorgi, K.; Taghizadeh, E.; Rostami, D.; Pormasoumi, H.; Barreto, G.E.; Hayat, S.M.G.; Sahebkar, A. Cellular and Molecular Aspects of Parkinson Treatment: Future Therapeutic Perspectives. Mol. Neurobiol. 2019, 56, 4799–4811. [Google Scholar] [CrossRef]
  10. Foster, H.D.; Hoffer, A. The two faces of L-DOPA: Benefits and adverse side effects in the treatment of Encephalitis lethargica, Parkinson’s disease, multiple sclerosis and amyotrophic lateral sclerosis. Med. Hypotheses 2004, 62, 177–181. [Google Scholar] [CrossRef]
  11. Bjorklund, A.; Kordower, J.H. Cell Therapy for Parkinson’s Disease: What Next? Mov. Disord. 2013, 28, 110–115. [Google Scholar] [CrossRef]
  12. O’Connor, D.M.; Boulis, N.M. Gene therapy for neurodegenerative diseases. Trends Mol. Med. 2015, 21, 504–512. [Google Scholar] [CrossRef]
  13. Chen, W.; Hu, Y.; Ju, D. Gene therapy for neurodegenerative disorders: Advances, insights and prospects. Acta Pharm Sin. B 2020, 10, 1347–1359. [Google Scholar] [CrossRef]
  14. Fan, H.-C.; Chi, C.-S.; Lee, Y.-J.; Tsai, J.-D.; Lin, S.-Z.; Harn, H.-J. The Role of Gene Editing in Neurodegenerative Diseases. Cell Transplant. 2018, 27, 364–378. [Google Scholar] [CrossRef]
  15. Baker, M. Gene-editing nucleases. Nat. Methods 2012, 9, 23–26. [Google Scholar] [CrossRef]
  16. Zych, A.O.; Bajor, M.; Zagozdzon, R. Application of Genome Editing Techniques in Immunology. Arch. Immunol. Ther. Exp. 2018, 66, 289–298. [Google Scholar] [CrossRef] [Green Version]
  17. Doudna, J.A.; Charpentier, E. The new frontier of genome engineering with CRISPR-Cas9. Science 2014, 346, 125809. [Google Scholar] [CrossRef]
  18. Persaud, A.; Desine, S.; Blizinsky, K.; Bonham, V.L. A CRISPR focus on attitudes and beliefs toward somatic genome editing from stakeholders within the sickle cell disease community. Genet. Med. 2019, 21, 1726–1734. [Google Scholar] [CrossRef] [Green Version]
  19. Yang, Y.; Zhang, X.; Yi, L.; Hou, Z.; Chen, J.; Kou, X.; Zhao, Y.; Wang, H.; Sun, X.-F.; Jiang, C.; et al. Naïve Induced Pluripotent Stem Cells Generated From β-Thalassemia Fibroblasts Allow Efficient Gene Correction With CRISPR/Cas9. Stem Cells Transl. Med. 2016, 5, 8–19. [Google Scholar] [CrossRef] [Green Version]
  20. Chiu, W.; Lin, T.-Y.; Chang, Y.-C.; Lai, H.I.-A.M.; Lin, S.-C.; Ma, C.; Yarmishyn, A.A.; Lin, S.-C.; Chang, K.-J.; Chou, Y.-B.; et al. An Update on Gene Therapy for Inherited Retinal Dystrophy: Experience in Leber Congenital Amaurosis Clinical Trials. Int. J. Mol. Sci. 2021, 22, 4534. [Google Scholar] [CrossRef]
  21. Wu, S.-S.; Li, Q.-C.; Yin, C.-Q.; Xue, W.; Song, C.-Q. Advances in CRISPR/Cas-based Gene Therapy in Human Genetic Diseases. Theranostics 2020, 10, 4374–4382. [Google Scholar] [CrossRef]
  22. Hudry, E.; Vandenberghe, L.H. Therapeutic AAV Gene Transfer to the Nervous System: A Clinical Reality. Neuron 2019, 101, 839–862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Biffi, A.; Montini, E.; Lorioli, L.; Cesani, M.; Fumagalli, F.; Plati, T.; Baldoli, C.; Martino, S.; Calabria, A.; Canale, S.; et al. Lentiviral Hematopoietic Stem Cell Gene Therapy Benefits Metachromatic Leukodystrophy. Science 2013, 341, 1233158. [Google Scholar] [CrossRef] [Green Version]
  24. Eichler, F.; Duncan, C.; Musolino, P.L.; Orchard, P.J.; Oliveira, S.D.; Thrasher, A.J.; Armant, M.; Dansereau, C.; Lund, T.C.; Miller, W.P.; et al. Hematopoietic Stem-Cell Gene Therapy for Cerebral Adrenoleukodystrophy. N. Engl. J. Med. 2017, 377, 1630–1638. [Google Scholar] [CrossRef] [Green Version]
  25. Cartier, N.; Hacein-Bey-Abina, S.; Bartholomae, C.C.; Veres, G.; Schmidt, M.; Kutschera, I.; Vidaud, M.; Abel, U.; Dal-Cortivo, L.; Caccavelli, L.; et al. Hematopoietic Stem Cell Gene Therapy with a Lentiviral Vector in X-Linked Adrenoleukodystrophy. Science 2009, 326, 818–823. [Google Scholar] [CrossRef] [Green Version]
  26. Singh, A.; Kukreti, R.; Saso, L.; Kukreti, S. Oxidative Stress: A Key Modulator in Neurodegenerative Diseases. Molecules 2019, 24, 1583. [Google Scholar] [CrossRef] [Green Version]
  27. Haapaniemi, E.; Botla, S.; Persson, J.; Schmierer, B.; Taipale, J. CRISPR–Cas9 genome editing induces a p53-mediated DNA damage response. Nat. Med. 2018, 24, 927–930. [Google Scholar] [CrossRef] [Green Version]
  28. Charlesworth, C.T.; Deshpande, P.S.; Dever, D.P.; Camarena, J.; Lemgart, V.T.; Cromer, M.K.; Vakulskas, C.A.; Collingwood, M.A.; Zhang, L.; Bode, N.M.; et al. Identification of preexisting adaptive immunity to Cas9 proteins in humans. Nat. Med. 2019, 25, 249–254. [Google Scholar] [CrossRef]
  29. Lino, C.A.; Harper, J.C.; Carney, J.P.; Timlin, J.A. Delivering CRISPR: A review of the challenges and approaches. Drug Deliv. 2018, 25, 1234–1257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Tay, A.; Melosh, N. Nanostructured Materials for Intracellular Cargo Delivery. Acc. Chem. Res. 2019, 52, 2462–2471. [Google Scholar] [CrossRef] [PubMed]
  31. Twelves, D.; Perkins, K.S.M.; Counsell, C. Systematic review of incidence studies of Parkinson’s disease. Mov. Disord. 2003, 18, 19–31. [Google Scholar] [CrossRef]
  32. Niethammer, M.; Tang, C.C.; Vo, A.; Nguyen, N.; Spetsieris, P.; Dhawan, V.; Ma, Y.; Small, M.; Feigin, A.; During, M.J.; et al. Gene therapy reduces Parkinson’s disease symptoms by reorganizing functional brain connectivity. Sci. Transl. Med. 2018, 10, eaau0713. [Google Scholar] [CrossRef] [Green Version]
  33. Maraganore, D.M.; Andrade, M.D.; Lesnick, T.C.; Strain, K.J.; Farrer, M.J.; Rocca, W.A.; Pant, P.V.K.; Frazer, K.A.; Cox, D.R.; Ballinger, D.C. High-resolution whole-genome association study of Parkinson disease. Am. J. Hum. Genet. 2005, 77, 685–693. [Google Scholar] [CrossRef] [Green Version]
  34. Wider, C.; Lincoln, S.J.; Heckman, M.G.; Diehl, N.N.; Stone, J.T.; Haugarvoll, K.; Aasly, J.O.; Gibson, J.M.; Lynch, T.; Rajput, A.; et al. Phactr2 and Parkinson’s disease. Neurosci. Lett. 2009, 453, 9–11. [Google Scholar] [CrossRef] [Green Version]
  35. Galvin, J.E. Prevention of Alzheimer’s Disease: Lessons Learned and Applied. J. Am. Geriatr. Soc. 2017, 65, 2128–2133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. González-Casacuberta, I.; Juárez-Flores, D.L.; Morén, C.; Garrabou, G. Bioenergetics and Autophagic Imbalance in Patients-Derived Cell Models of Parkinson Disease Supports Systemic Dysfunction in Neurodegeneration. Front. Neurosci. 2019, 13, 1–20. [Google Scholar] [CrossRef] [PubMed]
  37. Daida, K.; Nishioka, K.; Li, Y.; Yoshino, H.; Kikuchi, A.; Hasegawa, T.; Funayama, M.; Hattori, N. Mutation analysis of LRP10 in Japanese patients with familial Parkinson’s disease, progressive supranuclear palsy, and frontotemporal dementia. Neurobiol. Aging 2019, 84, 235.e11–235.e16. [Google Scholar] [CrossRef]
  38. Barbeau, A.; Roy, M. Familial Subsets in Idiopathic Parkinson’s Disease. Can. J. Neurol. Sci. 1984, 11, 144–150. [Google Scholar] [CrossRef] [Green Version]
  39. Santangelo, G.; Vitale, C.; Trojano, L.; Errico, D.; Amboni, M.; Barbarulo, M.; Grossi, D.; Barone, P. Neurophysological correlates of theory of mind in patients with early Parkinson’s Disease. Mov. Disord. 2011, 27, 98–105. [Google Scholar] [CrossRef] [PubMed]
  40. Chartier-Harlin, M.-C.; Kachergus, J.; Roumier, C.; Mouroux, V.; Douay, X.; Lincoln, S.; Levecque, C.; Larvor, L.; Andrieux, J.; Hulihan, M.; et al. α-synuclein locus duplication as a cause of familial Parkinson’s disease. Lancet 2004, 364, 1167–1169. [Google Scholar] [CrossRef]
  41. Reetz, K.; Gaser, C.; Klein, C.; Hagenah, J.; Büchel, C.; Gottschalk, S.; Pramstaller, P.P.; Siebner, H.R.; Binkofski, F. Structural findings in the basal ganglia in genetically determined and idiopathic Parkinson’s disease. Mov. Disord. 2009, 24, 99–103. [Google Scholar] [CrossRef]
  42. Kwon, Y.H.; Kim, Y.A.; Yoo, Y.H. Loss of Pigment Epithelial Cells Is Prevented by Autophagy. In Autophagy: Cancer Other Pathologies, Inflammation, Immunity, Infection, and Aging; Elsevier: Amsterdam, The Netherlands, 2017; pp. 105–117. [Google Scholar]
  43. Hou, X.; Watzlawik, J.; Fiesel, F.; Springer, W. Autophagy in Parkinson’s Disease. J. Mol. Biol. 2020, 432, 2651–2672. [Google Scholar] [CrossRef]
  44. Gan-Or, Z.; Dion, P.A.; Rouleau, G.A. Genetic perspective on the role of the autophagy-lysosome pathway in Parkinson disease. Autophagy 2015, 11, 1443–1457. [Google Scholar] [CrossRef] [PubMed]
  45. Stricher, F.; Macri, C.; Ruff, M.; Muller, S. HSPA8/HSC70 chaperone protein. Autophagy 2013, 9, 1937–1954. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Pan, T.; Kondo, S.; Le, W.; Jankovic, J. The role of autophagy-lysosome pathway in neurodegeneration associated with Parkinson’s disease. Brain 2008, 131, 1969–1978. [Google Scholar] [CrossRef] [PubMed]
  47. Cook, C.; Stetler, C.; Petrucelli, L. Disruption of Protein Quality Control in Parkinson’s Disease. Cold Spring Harb. Perspect. Med. 2012, 2, a009423. [Google Scholar] [CrossRef] [Green Version]
  48. Alvarez-Erviti, L.; Rodriguez-Oroz, M.C.; Cooper, J.M.; Caballero, C.; Ferrer, I.; Obeso, J.A.; Schapira, A.H.V. Chaperone-Mediated Autophagy Markers in Parkinson Disease Brains. Arch. Neurol. 2010, 67, 1464–1472. [Google Scholar] [CrossRef] [Green Version]
  49. Lynch-Day, M.A.; Mao, K.; Wang, K.; Zhao, M.; Klionsky, D.J. The Role of Autophagy in Parkinson’s Disease. Cold Spring Harb. Perspect. Med. 2012, 2, a009357. [Google Scholar] [CrossRef] [Green Version]
  50. Cheung, Z.H.; Ip, N.Y. The emerging role of autophagy in Parkinson’s disease. Mol. Brain 2009, 2, 29. [Google Scholar] [CrossRef] [Green Version]
  51. Engelender, S. Ubiquitination of α-synuclein and autophagy in Parkinson’s disease. Autophagy 2008, 4, 372–374. [Google Scholar] [CrossRef] [Green Version]
  52. Schaser, A.J.; Osterberg, V.R.; Dent, S.E.; Stackhouse, T.L.; Wakeham, C.M.; Boutros, S.W.; Weston, L.J.; Owen, N.; Weissman, T.A.; Luna, E.; et al. Alpha-synuclein is a DNA binding protein that modulates DNA repair with implications for Lewy body disorders. Sci. Rep. 2019, 9, 1–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Alegre-Abarrategui, J.; Wade-Martins, R. Parkinson disease, LRRK2 and the endocytic-autophagic pathway. Autophagy 2009, 5, 1208–1210. [Google Scholar] [CrossRef] [Green Version]
  54. Volta, M.; Milnerwood, A.J.; Farrer, M.J. Insights from late-onset familial parkinsonism on the pathogenesis of idiopathic Parkinson’s disease. Lancet Neurol. 2015, 14, 1054–1064. [Google Scholar] [CrossRef]
  55. Maraganore, D.M.; Lesnick, T.G.; Elbaz, A.; Chartier-Harlin, M.-C.; Gasser, T.; Krüger, R.; Hattori, N.; Mellick, G.D.; Quattrone, A.; Satoh, J.-I.; et al. UCHL1 is a Parkinson’s disease susceptibility gene. Ann. Neurol. 2004, 55, 512–521. [Google Scholar] [CrossRef] [PubMed]
  56. Chang, D.; Nalls, M.A.; Hallgrímsdóttir, I.B.; Hunkapiller, J.; Brug, M.v.d.; Cai, F.; Kerchner, G.A.; Ayalon, G.; Bingol, B.; Sheng, M.; et al. A meta-analysis of genome-wide association studies identifies 17 new Parkinson’s disease risk loci. Nat. Genet. 2017, 49, 1511–1516. [Google Scholar] [CrossRef] [PubMed]
  57. Jinn, S.; Blauwendraat, C.; Toolan, D.; Gretzula, C.A.; Drolet, R.E.; Smith, S.; Nalls, M.A.; Marcus, J.; Singleton, A.B.; Stone, D.J. Functionalization of the TMEM175 p.M393T variant as a risk factor for Parkinson disease. Hum. Mol. Genet. 2019, 28, 3244–3254. [Google Scholar] [CrossRef] [Green Version]
  58. Cox, L.E.; Ferraiuolo, L.; Goodall, E.F.; Heath, P.R.; Higginbottom, A.; Mortiboys, H.; Hollinger, H.C.; Hartley, J.A.; Brockington, A.; Burness, C.E.; et al. Mutations in CHMP2B in Lower Motor Neuron Predominant Amyotrophic Lateral Sclerosis (ALS). PLoS ONE 2010, 5, e9872. [Google Scholar] [CrossRef]
  59. Rusten, T.E.; Filimonenko, M.; Rodahl, L.M.; Stenmark, H.; Simonsen, A. ESCRTing autophagic clearance of aggregating proteins. Autophagy 2008, 4, 233–236. [Google Scholar] [CrossRef] [Green Version]
  60. Tanikawa, S.; Mori, F.; Tanji, K.; Kakita, A.; Takahashi, H.; Wakabayashi, K. Endosomal sorting related protein CHMP2B is localized in Lewy bodies and glial cytoplasmic inclusions in α-synucleinopathy. Neurosci. Lett. 2012, 527, 16–21. [Google Scholar] [CrossRef]
  61. Zaglia, T.; Milan, G.; Ruhs, A.; Franzoso, M.; Bertaggia, E.; Pianca, N.; Carpi, A.; Carullo, P.; Pesce, P.; Sacerdoti, D.; et al. Atrogin-1 deficiency promotes cardiomyopathy and premature death via impaired autophagy. J. Clin. Investig. 2014, 124, 2410–2424. [Google Scholar] [CrossRef] [Green Version]
  62. Behl, C. Breaking BAG: The Co-Chaperone BAG3 in Health and Disease. Trends Pharmacol. Sci. 2016, 37, 672–688. [Google Scholar] [CrossRef]
  63. Cao, Y.-L.; Yang, Y.-P.; Mao, C.-J.; Zhang, X.-Q.; Wang, C.-T.; Yang, J.; Lv, D.-J.; Wang, F.; Hu, L.-F.; Liu, C.-F. A role of BAG3 in regulating SNCA/α-synuclein clearance via selective macroautophagy. Neurobiol. Aging 2017, 60, 104–115. [Google Scholar] [CrossRef]
  64. Gamerdinger, M.; Hajieva, P.; Kaya, A.M.; Wolfrum, U.; Hartl, F.U.; Behl, C. Protein quality control during aging involves recruitment of the macroautophagy pathway by BAG3. EMBO J. 2009, 28, 889–901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Seidel, K.; Vinet, J.; den Dunnen, W.F.A.; Brunt, E.R.; Meister, M.; Boncoraglio, A.; Zijlstra, M.P.; Boddeke, H.W.G.M.; Rüb, U.; Kampinga, H.H.; et al. The HSPB8-BAG3 chaperone complex is upregulated in astrocytes in the human brain affected by protein aggregation diseases. Neuropathol. Appl. Neurobiol. 2012, 38, 39–53. [Google Scholar] [CrossRef] [PubMed]
  66. Cunha, S.R.; Le Scouarnec, S.; Schott, J.-J.; Mohler, P.J. Exon organization and novel alternative splicing of the human ANK2 gene: Implications for cardiac function and human cardiac disease. J. Mol. Cell. Cardiol. 2008, 45, 724–734. [Google Scholar] [CrossRef] [Green Version]
  67. Weiner, A.T.; Seebold, D.Y.; Michael, N.L.; Guignet, M.; Feng, C.; Follick, B.; Yusko, B.A.; Wasilko, N.P.; Torres-Gutierrez, P.; Rolls, M.M. Identification of Proteins Required for Precise Positioning of Apc2 in Dendrites. G3 Genes Genomes Genet. 2018, 8, 1841–1853. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Auburger, G.; Gispert, S.; Torres-Odio, S.; Jendrach, M.; Brehm, N.; Canet-Pons, J.; Key, J.; Sen, N.-E. SerThr-PhosphoProteome of Brain from Aged PINK1-KO+A53T-SNCA Mice Reveals pT1928-MAP1B and pS3781-ANK2 Deficits as Hub between Autophagy and Synapse Changes. Int. J. Mol. Sci. 2019, 20, 3284. [Google Scholar] [CrossRef] [Green Version]
  69. Hale, C.M.; Cheng, Q.; Ortuno, D.; Huang, M.; Nojima, D.; Kassner, P.D.; Wang, S.; Ollmann, M.M.; Carlisle, H.J. Identification of modulators of autophagic flux in an image-based high content siRNA screen. Autophagy 2016, 12, 713–726. [Google Scholar] [CrossRef] [Green Version]
  70. Füllgrabe, J.; Lynch-Day, M.A.; Heldring, N.; Li, W.; Struijk, R.B.; Ma, Q.; Hermanson, O.; Rosenfeld, M.G.; Klionsky, D.J.; Joseph, B. The histone H4 lysine 16 acetyltransferase hMOF regulates the outcome of autophagy. Nature 2013, 500, 468–471. [Google Scholar] [CrossRef] [PubMed]
  71. Soutar, M.P.M.; Melandri, D.; Annuario, E.; Monaghan, A.E.; Welsh, N.J.; D’Sa, K.; Guelfi, S.; Zhang, D.; Pittman, A.; Trabzuni, D.; et al. Regulation of mitophagy by the NSL complex underlies genetic risk for Parkinson’s disease at Chr16q11.2 and on the MAPT H1 allele. boiRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  72. Füllgrabe, J.; Klionsky, D.J.; Joseph, B. Histone post-translational modifications regulate autophagy flux and outcome. Autophagy 2013, 9, 1621–1623. [Google Scholar] [CrossRef] [Green Version]
  73. Lapierre, L.R.; Kumsta, C.; Sandri, M.; Ballabio, A.; Hansen, M. Transcriptional and epigenetic regulation of autophagy in aging. Autophagy 2015, 11, 867–880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Quintanilla, R.A.; Tapia, C.; Pérez, M.J. Possible role of mitochondrial permeability transition pore in the pathogenesis of Huntington disease. Biochem. Biophys. Res. Commun. 2017, 483, 1078–1083. [Google Scholar] [CrossRef]
  75. Wang, W.; Zhao, F.; Ma, X.; Perry, G.; Zhu, X. Mitochondria dysfunction in the pathogenesis of Alzheimer’s disease: Recent advances. Mol. Neurodegener. 2020, 15, 1–22. [Google Scholar] [CrossRef] [PubMed]
  76. Holden, S.; Maksoud, R.; Eaton-Fitch, N.; Cabanas, H.; Staines, D.; Marshall-Gradisnik, S. A systematic review of mitochondrial abnormalities in myalgic encephalomyelitis/chronic fatigue syndrome/systemic exertion intolerance disease. J. Transl. Med. 2020, 18, 1–16. [Google Scholar]
  77. Ge, P.; Dawson, V.L.; Dawson, T.M. PINK1 and Parkin mitochondrial quality control: A source of regional vulnerability in Parkinson’s disease. Mol. Neurodegener. 2020, 15, 1–18. [Google Scholar] [CrossRef] [Green Version]
  78. Trinh, D.; Israwi, A.R.; Arathoon, L.R.; Gleave, J.A.; Nash, J.E. The multi-faceted role of mitochondria in the pathology of Parkinson’s disease. J. Neurochem. 2021, 156, 715–752. [Google Scholar] [CrossRef]
  79. Toulorge, D.; Schapira, A.H.V.; Hajj, R. Molecular changes in the postmortem parkinsonian brain. J. Neurochem. 2016, 139, 27–58. [Google Scholar] [CrossRef] [PubMed]
  80. Hardy, J. Genetic Analysis of Pathways to Parkinson Disease. Neuron 2010, 68, 201–206. [Google Scholar] [CrossRef] [Green Version]
  81. Anderson, J.; Walker, D.; Goldstein, J.; de Laat, R.; Banducci, K. Phosphorylation of Ser-129 is the dominant pathological modification of alpha-synuclein in familial and sporadic Lewy Body disease. J. Biol. Chem. 2006, 281, 29739–29752. [Google Scholar] [CrossRef] [Green Version]
  82. Giasson, B.; Duda, J.; Murray, I.; Chen, Q.; Souza, J.M.; Hurting, H.I.; Ischiropolous, H.; Trajanowski, J.Q.; Lee, V.M.-Y. Oxidative damage linked to neurodegeneration by selective alpha-synuclein nitration in synucleinopathy lesions. Science 2000, 290, 985–989. [Google Scholar] [CrossRef] [PubMed]
  83. Tofaris, G.K.; Razzaq, A.; Ghetti, B.; Lilley, K.S.; Spillantini, M.G. Ubiquitination of alpha-synuclein in Lewy bodies is a pathological event not associated with impairment of proteasome. J. Biol. Chem. 2003, 278, 44404–44411. [Google Scholar] [CrossRef] [Green Version]
  84. Neitemeier, S.; Jelinek, A.; Laino, V.; Hoffmann, L.; Eisenbach, I.; Eying, R.; Ganjam, G.K.; Dolga, A.M.; Oppermann, S.; Culmsee, C. BID links ferroptosis to mitochondrial cell death pathways. Redox Biol. 2017, 12, 558–570. [Google Scholar] [CrossRef]
  85. Ganjam, G.K.; Bolte, K.; Matschke, L.A.; Neitemeier, S.; Dolga, A.M.; Höllerhage, M.; Höglinger, G.U.; Adamczyk, A.; Decher, N.; Oertel, W.H.; et al. Mitochondrial damage by α-synuclein causes cell death in human dopaminergic neurons. Cell Death Dis. 2019, 10, 1–16. [Google Scholar] [CrossRef]
  86. Nakajima, Y.-I.; Kuranaga, E. Caspase-dependent non-apoptotic processes in development. Cell Death Differ. 2017, 24, 1422–1430. [Google Scholar] [CrossRef] [PubMed]
  87. Hartmann, A.; Hunot, S.; Michel, P.P.; Muriel, M.-P.; Vyas, S.; Faucheux, B.A.; Mouatt-Prigent, A.; Turmel, H.; Srinivasan, A.; Ruberg, M.; et al. Caspase-3. A vulnerability factor and final effector in apoptotic death of dopaminergic neurons in Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2000, 97, 2875–2880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Hartmann, A.; Troadec, J.-D.; Hunot, S.; Kikly, K.; Faucheux, B.A.; Mouatt-Prigent, A.; Ruberg, M.; Agid, Y.; Hirsch, E.C. Caspase-8 Is an Effector in Apoptotic Death of Dopaminergic Neurons in Parkinson’s Disease, But Pathway Inhibition Results in Neuronal Necrosis. J. Neurosci. 2001, 21, 2247–2255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Wang, W.; Nguyen, L.; Burlak, C.; Chegini, F.; Guo, F.; Chataway, T.; Ju, S.; Fisher, O.S.; Miller, D.W.; Datta, D.; et al. Caspase-1 causes truncation and aggregation of the Parkinson’s disease-associated protein α-synuclein. Proc. Natl. Acad. Sci. USA 2016, 113, 9587–9592. [Google Scholar] [CrossRef] [Green Version]
  90. Brahmachari, S.; Lee, S.; Kim, S.; Yuan, C.; Karuppagounder, S.S.; Ge, P.; Shi, R.; Kim, E.J.; Liu, A.; Kim, D.; et al. Parkin interacting substrate zinc finger protein 746 is a pathological mediator in Parkinson’s disease. Brain 2019, 142, 2380–2401. [Google Scholar] [CrossRef]
  91. LaVoie, M.J.; Ostaszewski, B.L.; Weihofen, A.; Schlossmacher, M.G.; Selkoe, D.J. Dopamine covalently modifies and functionally inactivates parkin. Nat. Med. 2005, 11, 1214–1221. [Google Scholar] [CrossRef]
  92. Sunico, C.R.; Nakamura, T.; Rockenstein, E.; Mante, M.; Adame, A.; Chan, S.; Newmeyer, T.; Masliah, E.; Nakanishi, N.; Lipton, S.A. S-Nitrosylation of parkin as a novel regulator of p53-mediated neuronal cell death in sporadic Parkinson’s disease. Mol. Neurodegener. 2013, 8, 1–16. [Google Scholar] [CrossRef] [Green Version]
  93. Schlossmacher, M.G.; Frosch, M.P.; Gai, W.P.; Medina, M.; Sharma, N.; Forno, L.; Ochiishi, T.; Shimura, H.; Sharon, R.; Hattori, N.; et al. Parkin Localizes to the Lewy Bodies of Parkinson Disease and Dementia with Lewy Bodies. Am. J. Pathol. 2002, 160, 1655–1667. [Google Scholar] [CrossRef] [Green Version]
  94. Ko, H.S.; von Coelln, R.; Sriram, S.R.; Kim, S.W.; Chung, K.K.K.; Pletnikova, O.; Troncoso, J.; Johnson, B.; Saffary, R.; Goh, E.L.; et al. Accumulation of the Authentic Parkin Substrate Aminoacyl-tRNA Synthetase Cofactor, p38/JTV-1, Leads to Catecholaminergic Cell Death. J. Neurosci. 2005, 25, 7968–7978. [Google Scholar] [CrossRef] [PubMed]
  95. Lee, B.; Lee, K.; Panda, S.; Gonzales-Rojas, R.; Chong, A.; Bugay, V.; Park, H.M.; Brenner, R.; Murthy, N.; Lee, H.Y. Nanoparticle delivery of CRISPR into the brain rescues a mouse model of fragile X syndrome from exaggerated repetitive behaviours. Nat. Biomed. Eng. 2018, 2, 497–507. [Google Scholar] [CrossRef]
  96. Klein, A.D.; Mazzulli, J.R. Is Parkinson’s disease a lysosomal disorder? Brain 2018, 141, 2255–2262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Blauwendraat, C.; Reed, X.; Krohn, L.; Heilbron, K.; Bandres-Ciga, S.; Tan, M.; Gibbs, J.R.; Hernandez, D.G.; Kumaran, R.; Langston, R.; et al. Genetic modifiers of risk and age at onset in GBA associated Parkinson’s disease and Lewy body dementia. Brain 2020, 143, 234–248. [Google Scholar] [CrossRef] [PubMed]
  98. Wang, L.; Cheng, L.; Lu, Z.-J.; Sun, X.-Y.; Li, J.-Y.; Peng, R. Association of three candidate genetic variants in RAB7L1/NUCKS1 MCCC1 and STK39 with sporadic Parkinson’s disease in Han Chinese. J. Neural Transm. 2016, 123, 425–430. [Google Scholar] [CrossRef] [PubMed]
  99. Zhao, A.; Li, Y.; Niu, M.; Li, G.; Luo, N.; Zhou, L.; Kang, W.; Liu, J. SNPs in SNCA, MCCC1, DLG2, GBF1 and MBNL2 are associated with Parkinson’s disease in southern Chinese population. J. Cell. Mol. Med. 2020, 24, 8744–8752. [Google Scholar] [CrossRef]
  100. Chang, K.-H.; Chen, C.-M.; Chen, Y.-C.; Fung, H.-C.; Wu, Y.-R. Polymorphisms of ACMSD—TMEM163, MCCC1, and BCKDK—STX1B Are Not Associated with Parkinson’s Disease in Taiwan. Parkinsons. Dis. 2019, 2019, 1–6. [Google Scholar] [CrossRef] [Green Version]
  101. Wang, Y.; Tang, B.; Yu, R.; Li, K.; Liu, Z.; Xu, Q.; Sun, Q.; Yan, X.; Guo, J. Association analysis of STK39 MCCC1/LAMP3 and sporadic PD in the Chinese Han population. Neurosci. Lett. 2014, 566, 206–209. [Google Scholar] [CrossRef]
  102. Brás, J.; Guerreiro, R.; Hardy, J. SnapShot: Genetics of Parkinson’s Disease. Cell 2015, 160, 570.e1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Horowitz, M.P.; Greenamyre, J.T. Mitochondrial Iron Metabolism and Its Role in Neurodegeneration. J. Alzheimers Dis. 2010, 20, S551–S568. [Google Scholar] [CrossRef] [Green Version]
  104. Smith, A.G.; Raven, E.L.; Chernova, T. The regulatory role of heme in neurons. Metallomics 2011, 3, 955–962. [Google Scholar] [CrossRef] [PubMed]
  105. Bose, A.; Beal, M.F. Mitochondrial dysfunction in Parkinson’s disease. J. Neurochem. 2016, 139, 216–231. [Google Scholar] [CrossRef] [PubMed]
  106. Chen, X.; Xiao, Y.; Guo, W.; Zhou, M.; Huang, S.; Mo, M.; Li, Z.; Li, G.; Liu, H.; Peng, G.; et al. Relationship between variants of 17 newly loci and Parkinson’s disease in a Chinese population. Neurobiol. Aging 2019, 73, 230.e1–230.e4. [Google Scholar] [CrossRef] [PubMed]
  107. Lohman, D.C.; Forouhar, F.; Beebe, E.T.; Stefely, M.S.; Minogue, C.E.; Ulbrich, A.; Stefely, J.A.; Sukumar, S.; Luna-Sanchez, M.; Jochem, A.; et al. Mitochondrial COQ9 is a lipid-binding protein that associates with COQ7 to enable coenzyme Q biosynthesis. Proc. Natl. Acad. Sci. USA 2014, 111, E4697–E4705. [Google Scholar] [CrossRef] [Green Version]
  108. Ebadi, M.; Govitrapong, P.; Sharma, S.; Muralikrishnan, D.; Shavali, S.; Pellett, L.; Schafer, R.; Albano, C.; Eken, J. Ubiquinone (coenzyme q10) and mitochondria in oxidative stress of Parkinson’s disease. Neurosignals 2001, 10, 224–253. [Google Scholar] [CrossRef]
  109. Krohn, L.; Öztürk, T.N.; Vanderperre, B.; Ouled Amar Bencheikh, B.; Ruskey, J.A.; Laurent, S.B.; Spiegelman, D.; Postuma, R.B.; Arnulf, I.; Hu, M.T.M.; et al. Genetic, Structural, and Functional Evidence Link TMEM175 to Synucleinopathies. Ann. Neurol. 2020, 87, 139–153. [Google Scholar] [CrossRef]
  110. Dehay, B.; Martinez-Vicente, M.; Caldwell, G.A.; Caldwell, K.A.; Yue, Z.; Cookson, M.R.; Bezard, E. Lysosomal impairment in Parkinson’s disease. Mov. Disord. 2013, 28, 725–732. [Google Scholar] [CrossRef]
  111. Fernández, E.; García-Moreno, J.-M.; Martín de Pablos, A.; Chacón, J. May the Evaluation of Nitrosative Stress Through Selective Increase of 3-Nitrotyrosine Proteins Other Than Nitroalbumin and Dominant Tyrosine-125/136 Nitrosylation of Serum α-Synuclein Serve for Diagnosis of Sporadic Parkinson’s Disease? Antioxid. Redox Signal. 2013, 19, 912–918. [Google Scholar] [CrossRef] [Green Version]
  112. Ambrosi, G.; Ghezzi, C.; Zangaglia, R.; Levandis, G.; Pacchetti, C.; Blandini, F. Ambroxol-induced rescue of defective glucocerebrosidase is associated with increased LIMP-2 and saposin C levels in GBA1 mutant Parkinson’s disease cells. Neurobiol. Dis. 2015, 82, 235–242. [Google Scholar] [CrossRef]
  113. Pitcairn, C.; Wani, W.Y.; Mazzulli, J.R. Dysregulation of the autophagic-lysosomal pathway in Gaucher and Parkinson’s disease. Neurobiol. Dis. 2019, 122, 72–82. [Google Scholar] [CrossRef]
  114. Migdalska-Richards, A.; Schapira, A.H.V. The relationship between glucocerebrosidase mutations and Parkinson disease. J. Neurochem. 2016, 139, 77–90. [Google Scholar] [CrossRef] [Green Version]
  115. Kilpatrick, B.S.; Magalhaes, J.; Beavan, M.S.; McNeill, A.; Gegg, M.E.; Cleeter, M.W.J.; Bloor-Young, D.; Churchill, G.C.; Duchen, M.R.; Schapira, A.H.; et al. Endoplasmic reticulum and lysosomal Ca2+ stores are remodelled in GBA1-linked Parkinson disease patient fibroblasts. Cell Calcium 2016, 59, 12–20. [Google Scholar] [CrossRef] [Green Version]
  116. Jinn, S.; Drolet, R.E.; Cramer, P.E.; Wong, A.H.-K.; Toolan, D.M.; Gretzula, C.A.; Voleti, B.; Vassileva, G.; Disa, J.; Tadin-Strapps, M.; et al. TMEM175 deficiency impairs lysosomal and mitochondrial function and increases α-synuclein aggregation. Proc. Natl. Acad. Sci. USA 2017, 114, 2389–2394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Man, S.M.; Kanneganti, T.-D. Regulation of lysosomal dynamics and autophagy by CTSB/cathepsin B. Autophagy 2016, 12, 2504–2505. [Google Scholar] [CrossRef] [Green Version]
  118. Alcalay, R.N.; Levy, O.A.; Waters, C.H.; Fahn, S.; Ford, B.; Kuo, S.-H.; Mazzoni, P.; Pauciulo, M.W.; Nichols, W.C.; Gan-Or, Z.; et al. Glucocerebrosidase activity in Parkinson’s disease with and without GBA mutations. Brain 2015, 138, 2648–2658. [Google Scholar] [CrossRef] [Green Version]
  119. Contu, R.; Condorelli, G. ATP6V0A1 Polymorphism and MicroRNA-637. Circ. Cardiovasc. Genet. 2011, 4, 337–338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Lin, G.; Wang, L.; Marcogliese, P.C.; Bellen, H.J. Sphingolipids in the Pathogenesis of Parkinson’s Disease and Parkinsonism. Trends Endocrinol. Metab. 2019, 30, 106–117. [Google Scholar] [CrossRef] [PubMed]
  121. Peri, F.; Nüsslein-Volhard, C. Live Imaging of Neuronal Degradation by Microglia Reveals a Role for v0-ATPase a1 in Phagosomal Fusion In Vivo. Cell 2008, 133, 916–927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Pickar-Oliver, A.; Gersbach, C.A. The next generation of CRISPR–Cas technologies and applications. Nat. Rev. Mol. Cell Biol. 2019, 20, 490–507. [Google Scholar] [CrossRef]
  123. Ran, F.A.; Hsu, P.D.; Wright, J.; Agarwala, V.; Scott, D.A.; Zhang, F. Genome engineering using the CRISPR-Cas9 system. Nat. Protoc. 2013, 8, 2281–2308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Qi, L.; Larson, M.; Gilbert, L.; Doudna, J.; Weissman, J.; Arkin, A.; Lim, W. Repurposing CRISPR as an RNA-Guided Platform for Sequence-Specific Control of Gene Expression. Cell 2013, 152, 1173–1183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Tanenbaum, M.; Gilbert, L.; Qi, L.; Weissman, J.; Vale, R. A Protein-Tagging System for Signal Amplification in Gene Expression and Fluorescence Imaging. Cell 2014, 159, 635–646. [Google Scholar] [CrossRef] [Green Version]
  126. Gaudelli, N.M.; Komor, A.C.; Rees, H.A.; Packer, M.S.; Badran, A.H.; Bryson, D.I.; Liu, D.R. Programmable base editing of A•T to G•C in genomic DNA without DNA cleavage. Nature 2017, 551, 464–471. [Google Scholar] [CrossRef]
  127. Kampmann, M. CRISPRi and CRISPRa Screens in Mammalian Cells for Precision Biology and Medicine. ACS Chem. Biol. 2018, 13, 406–416. [Google Scholar] [CrossRef]
  128. Kang, J.G.; Park, J.S.; Ko, J.-H.; Kim, Y.-S. Regulation of gene expression by altered promoter methylation using a CRISPR/Cas9-mediated epigenetic editing system. Sci. Rep. 2019, 9, 11960. [Google Scholar] [CrossRef] [Green Version]
  129. Deng, W.; Shi, X.; Tjian, R.; Lionnet, T.; Singer, R.H. CASFISH: CRISPR/Cas9-mediated in situ labeling of genomic loci in fixed cells. Proc. Natl. Acad. Sci. USA 2015, 112, 11870–11875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Ye, R.; Pi, M.; Cox, J.V.; Nishimoto, S.K.; Quarles, L.D. CRISPR/Cas9 targeting of GPRC6A suppresses prostate cancer tumorigenesis in a human xenograft model. J. Exp. Clin. Cancer Res. 2017, 36, 1–13. [Google Scholar] [CrossRef]
  131. Tycko, J.; Myer, V.; Hsu, P. Methods for Optimizing CRISPR-Cas9 Genome Editing Specificity. Mol. Cell 2016, 63, 355–370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Zhang, X.-H.; Tee, L.Y.; Wang, X.-G.; Huang, Q.-S.; Yang, S.-H. Off-target Effects in CRISPR/Cas9-mediated Genome Engineering. Mol. Ther. Nucleic Acids 2015, 4, e264. [Google Scholar] [CrossRef] [PubMed]
  133. Akcakaya, P.; Bobbin, M.L.; Guo, J.A.; Malagon-Lopez, J.; Clement, K.; Garcia, S.P.; Fellows, M.D.; Porritt, M.J.; Firth, M.A.; Carreras, A.; et al. In vivo CRISPR editing with no detectable genome-wide off-target mutations. Nature 2018, 561, 416–419. [Google Scholar] [CrossRef]
  134. Bae, S.; Park, J.; Kim, J.S. Cas-OFFinder: A fast and versatile algorithm that searches for potential off-target sites of Cas9 RNA-guided endonucleases. Bioinformatics 2014, 30, 1473–1475. [Google Scholar] [CrossRef] [Green Version]
  135. Haeussler, M.; Schönig, K.; Eckert, H.; Eschstruth, A.; Mianné, J.; Renaud, J.B.; Schneider-Maunoury, S.; Shkumatava, A.; Teboul, L.; Kent, J.; et al. Evaluation of off-target and on-target scoring algorithms and integration into the guide RNA selection tool CRISPOR. Genome Biol. 2016, 17, 148. [Google Scholar] [CrossRef]
  136. Hanna, R.E.; Doench, J.G. Design and analysis of CRISPR–Cas experiments. Nat. Biotechnol. 2020, 38, 813–823. [Google Scholar] [CrossRef] [PubMed]
  137. Fu, Y.; Sander, J.D.; Reyon, D.; Cascio, V.M.; Joung, J.K. Improving CRISPR-Cas nuclease specificity using truncated guide RNAs. Nat. Biotechnol. 2014, 32, 279–284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Mohr, S.E.; Hu, Y.; Ewen-Campen, B.; Housden, B.E.; Viswanatha, R.; Perrimon, N. CRISPR guide RNA design for research applications. FEBS J. 2016, 283, 3232–3238. [Google Scholar] [CrossRef]
  139. Wang, D.; Zhang, C.; Wang, B.; Li, B.; Wang, Q.; Liu, D.; Wang, H.; Zhou, Y.; Shi, L.; Lan, F.; et al. Optimized CRISPR guide RNA design for two high-fidelity Cas9 variants by deep learning. Nat. Commun. 2019, 10, 4284. [Google Scholar] [CrossRef] [PubMed]
  140. Zhang, Q.; Xing, H.L.; Wang, Z.P.; Zhang, H.Y.; Yang, F.; Wang, X.C.; Chen, Q.J. Potential high-frequency off-target mutagenesis induced by CRISPR/Cas9 in Arabidopsis and its prevention. Plant Mol. Biol. 2018, 96, 445–456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Kleinstiver, B.P.; Pattanayak, V.; Prew, M.S.; Tsai, S.Q.; Nguyen, N.T.; Joung, J.K. 731. High-Fidelity CRISPR-Cas9 Nucleases with No Detectable Genome-Wide Off-Target Effects. Mol. Ther. 2016, 24, S288. [Google Scholar] [CrossRef]
  142. Mateus, A.; Treyer, A.; Wegler, C.; Karlgren, M.; Matsson, P.; Artursson, P. Intracellular drug bioavailability: A new predictor of system dependent drug disposition. Sci. Rep. 2017, 7, 43047. [Google Scholar] [CrossRef] [PubMed]
  143. Singh, S.; Kumar, S.; Sharda, N.; Chakraborti, A.K. New Findings on Degradation of Famotidine under Basic Conditions: Identification of a Hitherto Unknown Degradation Product and the Condition for Obtaining the Propionamide Intermediate in Pure Form. J. Pharm. Sci. 2002, 91, 253–257. [Google Scholar] [CrossRef] [PubMed]
  144. Hogg, P.J. Disulfide bonds as switches for protein function. Trends Biochem. Sci. 2003, 28, 210–214. [Google Scholar] [CrossRef]
  145. Li, C.; Guan, X.; Du, T.; Jin, W.; Wu, B.; Liu, Y.; Wang, P.; Hu, B.; Griffin, G.E.; Shattock, R.J.; et al. Inhibition of HIV-1 infection of primary CD4+ T-cells by gene editing of CCR5 using adenovirus-delivered CRISPR/Cas9. J. Gen. Virol. 2015, 96, 2381–2393. [Google Scholar] [CrossRef]
  146. Perez, J.; Rueda, J.; Cuellar, M.; Suarez-Arnedo, A.; Cruz, J.C.; Muñoz-Camargo, C. Cell-Penetrating And Antibacterial BUF-II Nanobioconjugates: Enhanced Potency Via Immobilization On Polyetheramine-Modified Magnetite Nanoparticles. Int. J. Nanomed. 2019, 14, 8483–8497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Tong, S.; Moyo, B.; Lee, C.M.; Leong, K.; Bao, G. Engineered materials for in vivo delivery of genome-editing machinery. Nat. Rev. Mater. 2019, 4, 726–737. [Google Scholar] [CrossRef] [PubMed]
  148. DeWitt, M.A.; Corn, J.E.; Carroll, D. Genome editing via delivery of Cas9 ribonucleoprotein. Methods 2017, 121–122, 9–15. [Google Scholar] [CrossRef]
  149. Zhang, Z.; Wan, T.; Chen, Y.; Chen, Y.; Sun, H.; Cao, T.; Songyang, Z.; Tang, G.; Wu, C.; Ping, Y.; et al. Cationic Polymer-Mediated CRISPR/Cas9 Plasmid Delivery for Genome Editing. Macromol. Rapid Commun. 2019, 40, 1800068. [Google Scholar] [CrossRef]
  150. Hamilton, T.A.; Pellegrino, G.M.; Therrien, J.A.; Ham, D.T.; Bartlett, P.C.; Karas, B.J.; Gloor, G.B.; Edgell, D.R. Efficient inter-species conjugative transfer of a CRISPR nuclease for targeted bacterial killing. Nat. Commun. 2019, 10, 4544. [Google Scholar] [CrossRef]
  151. Wang, H.-X.; Li, M.; Lee, C.M.; Chakraborty, S.; Kim, H.-W.; Bao, G.; Leong, K.W. CRISPR/Cas9-Based Genome Editing for Disease Modeling and Therapy: Challenges and Opportunities for Nonviral Delivery. Chem. Rev. 2017, 117, 9874–9906. [Google Scholar] [CrossRef] [PubMed]
  152. Kang, S.; Jeon, S.; Kim, S.; Chang, Y.K.; Kim, Y.-C. Development of a pVEC peptide-based ribonucleoprotein (RNP) delivery system for genome editing using CRISPR/Cas9 in Chlamydomonas reinhardtii. Sci. Rep. 2020, 10, 22158. [Google Scholar] [CrossRef] [PubMed]
  153. Kleinstiver, B.P.; Pattanayak, V.; Prew, M.S.; Tsai, S.Q.; Nguyen, N.T.; Zheng, Z.; Joung, J.K. High-fidelity CRISPR-Cas9 nucleases with no detectable genome-wide off-target effects. Nature 2016, 529, 490–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Schimd-Burgk, J.L.; Gao, L.; Li, D.; Gardner, Z.; Strecker, J.; Lash, B.; Zhang, F. Highly parallel Profiling of Cas9 Variant Specificity. Mol. Cell 2020, 78, 794–800. [Google Scholar] [CrossRef]
  155. Kim, N.; Kim, H.K.; Lee, S.; Seo, J.H.; Choi, J.W.; Park, J.; Min, S.; Yoon, S.; Cho, S.-R.; Kim, H.H. Prediction of the sequence-specific cleavage activity of Cas9 variants. Nat. Biotechnol. 2020, 38, 1328–1336. [Google Scholar] [CrossRef]
  156. Casini, A.; Olivieri, M.; Petris, G.; Montagna, C.; Reginato, G.; Maule, G.; Cereseto, A. A highly specific SpCas9 variant is identified by in vivo screening in yeast. Nat. Biotechnol. 2018, 36, 265–271. [Google Scholar] [CrossRef]
  157. Zuris, J.A.; Thompson, D.B.; Shu, Y.; Guilinger, J.P.; Bessen, J.L.; Hu, J.H.; Maeder, M.L.; Joung, J.K.; Chen, Z.Y.; Liu, D.R. Cationic lipid-mediated delivery of proteins enables efficient protein-based genome editing in vitro and in vivo. Nat. Biotechnol. 2015, 33, 73–80. [Google Scholar] [CrossRef] [Green Version]
  158. 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. [Google Scholar] [CrossRef]
  159. Li, L.; Hu, S.; Chen, X. Non-viral delivery systems for CRISPR/Cas9-based genome editing: Challenges and opportunities. Biomaterials 2018, 171, 207–218. [Google Scholar] [CrossRef]
  160. Ablain, J.; Durand, E.; Yang, S.; Zhou, Y.; Zon, L. A CRISPR/Cas9 Vector System for Tissue-Specific Gene Disruption in Zebrafish. Dev. Cell 2015, 32, 756–764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Yu, W.; Mookherjee, S.; Chaitankar, V.; Hiriyanna, S.; Kim, J.W.; Brooks, M.; Ataeijannati, Y.; Sun, X.; Dong, L.; Li, T.; et al. Nrl knockdown by AAV-delivered CRISPR/Cas9 prevents retinal degeneration in mice. Nat. Commun. 2017, 8, 14716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Swiech, L.; Heidenreich, M.; Banerjee, A.; Habib, N.; Li, Y.; Trombetta, J.; Sur, M.; Zhang, F. In vivo interrogation of gene function in the mammalian brain using CRISPR-Cas9. Nat. Biotechnol. 2015, 33, 102–106. [Google Scholar] [CrossRef] [PubMed]
  163. Yue, H.; Zhou, X.; Chenga, M.; Xing, D. Graphene oxide-mediated Cas9/sgRNA delivery for efficient genome editing. Nanoscale 2018, 10, 1063–1071. [Google Scholar] [CrossRef]
  164. Wang, H.X.; Song, Z.; Lao, Y.H.; Xu, X.; Gong, J.; Cheng, D.; Chakraborty, S.; Park, J.S.; 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, 4903–4908. [Google Scholar] [CrossRef] [Green Version]
  165. Rueda-Gensini, L.; Cifuentes, J.; Castellanos, M.C.; Puentes, P.R.; Serna, J.A.; Muñoz-Camargo, C.; Cruz, J.C. Tailoring iron oxide nanoparticles for efficient cellular internalization and endosomal escape. Nanomaterials 2020, 10, 1816. [Google Scholar] [CrossRef] [PubMed]
  166. Tiana, W.; Ma, Y. Insights into the endosomal escape mechanism via investigation of dendrimer–membrane interactions. Soft Matter 2012, 8, 6378–6384. [Google Scholar] [CrossRef]
  167. Mout, R.; Ray, M.; Lee, Y.W.; Scaletti, F.; Rotello, V.M. In Vivo Delivery of CRISPR/Cas9 for Therapeutic Gene Editing: Progress and Challenges. Bioconjug Chem. 2017, 28, 880–884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Han, X.; Liu, Z.; Jo, M.C.; Zhang, K.; Li, Y.; Zeng, Z.; Li, N.; Zu, Y.; Qin, L. CRISPR-Cas9 delivery to hard-to-transfect cells via membrane deformation. Sci. Adv. 2015, 1, e1500454. [Google Scholar] [CrossRef] [Green Version]
  169. Sun, L.; Wu, J.; Du, F.; Chen, X.; Chen, Z.J. Cyclic GMP-AMP Synthase Is a Cytosolic DNA Sensor That Activates the Type I Interferon Pathway. Science 2013, 339, 786–791. [Google Scholar] [CrossRef] [Green Version]
  170. Kim, S.; Kim, D.; Cho, S.W.; Kim, J.; Kim, J.-S. Highly efficient RNA-guided genome editing in human cells via delivery of purified Cas9 ribonucleoproteins. Genome Res. 2014, 24, 1012–1019. [Google Scholar] [CrossRef] [Green Version]
  171. Imani, R.; Emami, S.H.; Faghihi, S. Synthesis and characterization of an octaarginine functionalized graphene oxide nano-carrier for gene delivery applications. Phys. Chem. Chem. Phys. 2015, 17, 6328–6339. [Google Scholar] [CrossRef]
  172. Friedland, A.E.; Baral, R.; Singhal, P.; Loveluck, K.; Shen, S.; Sanchez, M.; Marco, E.; Gotta, G.M.; Maeder, M.L.; Kennedy, E.M.; et al. Characterization of Staphylococcus aureus Cas9: A smaller Cas9 for all-in-one adeno-associated virus delivery and paired nickase applications. Genome Biol. 2015, 16, 257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Lopez-Barbosa, N.; Suárez-Arnedo, A.; Cifuentes, J.; Gonzalez Barrios, A.F.; Silvera Batista, C.A.; Osma, J.F.; Muñoz-Camargo, C.; Cruz, J.C. Magnetite–OmpA Nanobioconjugates as Cell-Penetrating Vehicles with Endosomal Escape Abilities. ACS Biomater. Sci. Eng. 2020, 6, 415–424. [Google Scholar] [CrossRef] [PubMed]
  174. Li, L.; Song, L.; Liu, X.; Yang, X.; Li, X.; He, T.; Wei, Y. Artificial Virus Delivers CRISPR-Cas9 System for Genome Editing of Cells in Mice. ACS Nano 2016, 11, 95–111. [Google Scholar] [CrossRef] [PubMed]
  175. 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, 452–462. [Google Scholar] [CrossRef] [PubMed]
  176. Hanlon, K.S.; Kleinstiver, B.P.; Garcia, S.P.; Zaborowski, M.P.; Volak, A.; Spirig, S.E.; Muller, A.; Sousa, A.A.; Tsai, S.Q.; Bengtsson, N.E.; et al. High levels of AAV vector integration into CRISPR-induced DNA breaks. Nat. Commun. 2019, 10, 4439. [Google Scholar] [CrossRef]
  177. Schmidt, F.; Grimm, D. CRISPR genome engineering and viral gene delivery: A case of mutual attraction. Biotechnol. J. 2015, 10, 258–272. [Google Scholar] [CrossRef] [Green Version]
  178. Wilbie, D.; Walther, J.; Mastrobattista, E. Delivery Aspects of CRISPR/Cas for in Vivo Genome Editing. Acc. Chem. Res. 2019, 52, 1555–1564. [Google Scholar] [CrossRef] [Green Version]
  179. Platt, R.J.; Chen, S.; Zhou, Y.; Yim, M.J.; Swiech, L.; Kempton, H.R.; Dahlman, J.E.; Parnas, O.; Eisenhaure, T.M.; Jovanovic, M.; et al. CRISPR-Cas9 knockin mice for genome editing and cancer modeling. Cell 2014, 159, 440–455. [Google Scholar] [CrossRef] [Green Version]
  180. Xu, C.L.; Ruan, M.Z.C.; Mahajan, V.B.; Tsang, S.H. Viral Delivery Systems for CRISPR. Viruses 2019, 11, 28. [Google Scholar] [CrossRef] [Green Version]
  181. György, B.; Lööv, C.; Zaborowski, M.P.; Takeda, S.; Kleinstiver, B.P.; Commins, C.; Kastanenka, K.; Mu, D.; Volak, A.; Giedraitis, V.; et al. CRISPR/Cas9 Mediated Disruption of the Swedish APP Allele as a Therapeutic Approach for Early-Onset Alzheimer’s Disease. Mol. Ther. Nucleic Acids 2018, 11, 429–440. [Google Scholar] [CrossRef] [Green Version]
  182. Wang, D.; Zhang, F.; Gao, G. CRISPR-Based Therapeutic Genome Editing: Strategies and In Vivo Delivery by AAV Vectors. Cell 2020, 181, 136–150. [Google Scholar] [CrossRef]
  183. Senís, E.; Fatouros, C.; Große, S.; Wiedtke, E.; Niopek, D.; Mueller, A.K.; Börner, K.; Grimm, D. CRISPR/Cas9-mediated genome engineering: An adeno-associated viral (AAV) vector toolbox. Biotechnol. J. 2014, 9, 1402–1412. [Google Scholar] [CrossRef]
  184. Choi, J.-H.; Yu, N.-K.; Baek, G.-C.; Bakes, J.; Seo, D.; Nam, H.; Baek, S.; Lim, C.-S.; Lee, Y.-S.; Kaang, B.-K. Optimization of AAV expression cassettes to improve packaging capacity and transgene expression in neurons. Mol. Brain 2014, 7, 1–10. [Google Scholar] [CrossRef] [Green Version]
  185. Daya, S.; Berns, K.I. Gene Therapy Using Adeno-Associated Virus Vectors. Clin. Microbiol. Rev. 2008, 21, 583–593. [Google Scholar] [CrossRef] [Green Version]
  186. Kaeppel, C.; Beattie, S.G.; Fronza, R.; van Logtenstein, R.; Salmon, F.; Schmidt, S.; Wolf, S.; Nowrouzi, A.; Glimm, H.; von Kalle, C.; et al. A largely random AAV integration profile after LPLD gene therapy. Nat. Med. 2013, 19, 889–891. [Google Scholar] [CrossRef] [PubMed]
  187. Newman, N.J.; Yu-Wai-Man, P.; Carelli, V.; Moster, M.L.; Biousse, V.; Vignal-Clermont, C.; Sergott, R.C.; Klopstock, T.; Sadun, A.A.; Barboni, P.; et al. Efficacy and Safety of Intravitreal Gene Therapy for Leber Hereditary Optic Neuropathy Treated within 6 Months of Disease Onset. Ophthalmology 2021, 128, 649–660. [Google Scholar] [CrossRef]
  188. Yang, L.; Slone, J.; Li, Z.; Lou, X.; Hu, Y.-C.; Queme, L.F.; Jankowski, M.P.; Huang, T. Systemic administration of AAV-Slc25a46 mitigates mitochondrial neuropathy in Slc25a46−/− mice. Hum. Mol. Genet. 2020, 29, 649–661. [Google Scholar] [CrossRef] [PubMed]
  189. Silva-Pinheiro, P.; Cerutti, R.; Luna-Sanchez, M.; Zeviani, M.; Viscomi, C. A Single Intravenous Injection of AAV-PHP.B-hNDUFS4 Ameliorates the Phenotype of Ndufs4 Mice. Mol. Ther. Methods Clin. Dev. 2020, 17, 1071–1078. [Google Scholar] [CrossRef] [PubMed]
  190. Burning, H.; Srivastava, A. Capsid modification for targeting and improving the efficacy of AAV vectors. Mol. Ther. Methods Clin. Dev. 2019, 12, 248–265. [Google Scholar] [CrossRef] [PubMed]
  191. Castle, M.J.; Turunen, H.T.; Vandenberghe, L.H.; Wolfe, J.H. Controlling AAV Tropism in the Nervous System with Natural and Engineered Capsids. In Methods in Molecular Biology; Humana Press: New York, NY, USA, 2016; pp. 133–149. [Google Scholar]
  192. Yang, B.; Li, S.; Wang, H.; Guo, Y.; Gessler, D.J.; Cao, C.; Su, Q.; Kramer, J.; Zhong, L.; Ahmed, S.S.; et al. Global CNS Transduction of Adult Mice by Intravenously Delivered rAAVrh.8 and rAAVrh.10 and Nonhuman Primates by rAAVrh.10. Mol. Ther. 2014, 22, 1299–1309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Fu, H.; McCarty, D.M. Crossing the blood–brain-barrier with viral vectors. Curr. Opin. Virol. 2016, 21, 87–92. [Google Scholar] [CrossRef] [PubMed]
  194. Hong, S.H.; Park, S.J.; Lee, S.; Cho, C.S.; Cho, M.H. Aerosol gene delivery using viral vectors and cationic carriers for in vivo lung cancer therapy. Expert Opin. Drug Deliv. 2015, 12, 977–991. [Google Scholar] [CrossRef] [PubMed]
  195. Matos, M.R.; Ho, S.-M.; Schrode, N.; Brennand, K.J. Integration of CRISPR-engineering and hiPSC-based models of psychiatric genomics. Mol. Cell. Neurosci. 2020, 107, 103532. [Google Scholar] [CrossRef] [PubMed]
  196. Williams, M.R.; Fricano-Kugler, C.J.; Getz, S.A.; Skelton, P.D.; Lee, J.; Rizzuto, C.P.; Geller, J.S.; Li, M.; Luikart, B.W. A Retroviral CRISPR-Cas9 System for Cellular Autism-Associated Phenotype Discovery in Developing Neurons. Sci. Rep. 2016, 6, 25611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Karimi, M.; Solati, N.; Ghasemi, A.; Estiar, M.A.; Hashemkhani, M.; Kiani, P.; Mohamed, E.; Saeidi, A.; Taheri, M.; Avci, P.; et al. Carbon nanotubes part II: A remarkable carrier for drug and gene delivery. Expert Opin. Drug Deliv. 2015, 12, 1089–1105. [Google Scholar] [CrossRef] [PubMed]
  198. Chen, S.; Sun, S.; Moonen, D.; Lee, C.; Lee, A.Y.; Schaffer, D.V.; He, L. CRISPR-READI: Efficient Generation of Knockin Mice by CRISPR RNP Electroporation and AAV Donor Infection. Cell Rep. 2019, 27, 3780–3789. [Google Scholar] [CrossRef] [Green Version]
  199. Glass, Z.; Lee, M.; Li, Y.; Xu, Q. Engineering the Delivery System for CRISPR-Based Genome Editing. Trends Biotechnol. 2018, 36, 173–185. [Google Scholar] [CrossRef]
  200. Naseri, N.; Valizadeh, H.; Zakeri-Milani, P. Solid Lipid Nanoparticles and Nanostructured Lipid Carriers: Structure, Preparation and Application. Adv. Pharm. Bull. 2015, 5, 305–313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Allen, T.M.; Cullis, P.R. Liposomal drug delivery systems: From concept to clinical applications. Adv. Drug Deliv. Rev. 2013, 65, 36–48. [Google Scholar] [CrossRef]
  202. Anton, N.; Vandamme, T.F. Nano-emulsions and Micro-emulsions: Clarifications of the Critical Differences. Pharm. Res. 2011, 28, 978–985. [Google Scholar] [CrossRef]
  203. Lao, Y.H.; Li, M.; Gao, M.A.; Shao, D.; Chi, C.W.; Huang, D.; Chakraborty, S.; Ho, T.C.; Jiang, W.; Wang, H.X.; et al. HPV Oncogene Manipulation Using Nonvirally Delivered CRISPR/Cas9 or Natronobacterium gregoryi Argonaute. Adv. Sci. 2018, 5, 1700540. [Google Scholar] [CrossRef]
  204. Röhrborn, D. DPP4 in diabetes. Front. Immunol. 2015, 6, 386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Sun, D.; Sun, Z.; Jiang, H.; Vaidya, A.M.; Xin, R.; Ayat, N.R.; Schilb, A.L.; Qiao, P.L.; Han, Z.; Naderi, A.; et al. Synthesis and Evaluation of pH-Sensitive Multifunctional Lipids for Efficient Delivery of CRISPR/Cas9 in Gene Editing. Bioconjug Chem. 2019, 30, 667–678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. 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. 2018, 57, 1491–1496. [Google Scholar] [CrossRef]
  207. Barenholz, Y. Doxil®—The first FDA-approved nano-drug: Lessons learned. J. Control. Release 2012, 160, 117–134. [Google Scholar] [CrossRef] [PubMed]
  208. Martins, S.; Sarmento, B.; Ferreira, D.C.; Souto, E.B. Lipid-based colloidal carriers for peptide and protein delivery—Liposomes versus lipid nanoparticles. Int. J. Nanomed. 2007, 2, 595–607. [Google Scholar]
  209. Chikh, G.G.; Li, W.M.; Schutze-Redelmeier, M.-P.; Meunier, J.-C.; Bally, M.B. Attaching histidine-tagged peptides and proteins to lipid-based carriers through use of metal-ion-chelating lipids. Biochim. Biophys. Acta Biomembr. 2002, 1567, 204–212. [Google Scholar] [CrossRef] [Green Version]
  210. Li, X.; Tian, X.; Zhang, J.; Zhao, X.; Chen, X.; Jiang, Y.; Wang, D.; Pan, W. In vitro and in vivo evaluation of folate receptor-targeting amphiphilic copolymer-modified liposomes loaded with docetaxel. Int. J. Nanomed. 2011, 6, 1167–1184. [Google Scholar] [CrossRef] [Green Version]
  211. Jokerst, J.V.; Lobovkina, T.; Zare, R.N.; Gambhir, S.S. Nanoparticle PEGylation for imaging and therapy. Nanomedicine 2011, 6, 715–728. [Google Scholar] [CrossRef] [Green Version]
  212. Xie, X.; Liao, J.; Shao, X.; Li, Q.; Lin, Y. The Effect of shape on Cellular Uptake of Gold Nanoparticles in the forms of Stars, Rods, and Triangles. Sci. Rep. 2017, 7, 3827. [Google Scholar] [CrossRef]
  213. Cheng, Q.; Wei, T.; Farbiak, L.; Johnson, L.T.; Dilliard, S.A.; Siegwart, D.J. Selective organ targeting (SORT) nanoparticles for tissue-specific mRNA delivery and CRISPR\Cas gene editing. Nat. Nanotechnol. 2020, 15, 313–320. [Google Scholar] [CrossRef]
  214. Wei, T.; Cheng, Q.; Min, Y.-L.; Olson, E.N.; Siegwart, D.J. Systemic nanoparticle delivery of CRISPR-Cas9 ribonucleoproteins for effective tissue specific genome editing. Nat. Commun. 2020, 11, 3232. [Google Scholar] [CrossRef]
  215. Podstawczyk, D.; Nizioł, M.; Szymczyk, P.; Wiśniewski, P.; Guiseppi-Elie, A. 3D printed stimuli-responsive magnetic nanoparticle embedded alginate-methylcellulose hydrogel actuators. Addit. Manuf. 2020, 34, 101275. [Google Scholar] [CrossRef]
  216. Zhuang, J.; Tan, J.; Wu, C.; Zhang, J.; Liu, T.; Fan, C.; Li, J.; Zhang, Y. Extracellular vesicles engineered with valency-controlled DNA nanostructures deliver CRISPR/Cas9 system for gene therapy. Nucleic Acids Res. 2020, 48, 8870–8882. [Google Scholar] [CrossRef]
  217. Xie, J.; Zhao, C.; Lin, Z.; Gu, P.; Zhang, Q. Nanostructured Conjugated Polymers for Energy-Related Applications beyond Solar Cells. Chem. Asian J. 2016, 11, 1489–1511. [Google Scholar] [CrossRef] [PubMed]
  218. Abdolahpour, S.; Toliyat, T.; Omidfar, K.; Modjtahedi, H.; Wong, A.J.; Rasaee, M.J.; Kashanian, S.; Paknejad, M. Targeted delivery of doxorubicin into tumor cells by nanostructured lipid carriers conjugated to anti-EGFRvIII monoclonal antibody. Artif. Cells Nanomed. Biotechnol. 2018, 46, 89–94. [Google Scholar] [CrossRef] [PubMed]
  219. Hameed, S.; Munawar, A.; Khan, W.S.; Mujahid, A.; Ihsan, A.; Rehman, A.; Ahmed, I.; Bajwa, S.Z. Assessing manganese nanostructures based carbon nanotubes composite for the highly sensitive determination of vitamin C in pharmaceutical formulation. Biosens. Bioelectron. 2017, 89, 822–828. [Google Scholar] [CrossRef]
  220. Ramírez-Acosta, C.M.; Cifuentes, J.; Castellanos, M.C.; Moreno, R.J.; Muñoz-Camargo, C.; Cruz, J.C.; Reyes, L.H. PH-Responsive, Cell-Penetrating, Core/Shell Magnetite/Silver Nanoparticles for the Delivery of Plasmids: Preparation, Characterization, and Preliminary In Vitro Evaluation. Pharmaceutics 2020, 12, 561. [Google Scholar] [CrossRef] [PubMed]
  221. He, Y.; Zhao, Y. Technological breakthroughs in generating transgene-free and genetically stable CRISPR-edited plants. aBIOTECH 2020, 1, 88–96. [Google Scholar] [CrossRef] [Green Version]
  222. Noureddine, A.; Maestas-Olguin, A.; Saada, E.A.; LaBauve, A.E.; Agola, J.O.; Baty, K.E.; Howard, T.; Sabo, J.K.; Espinoza, C.R.S.; Doudna, J.A.; et al. Engineering of monosized lipid-coated mesoporous silica nanoparticles for CRISPR delivery. Acta Biomater. 2020, 114, 358–368. [Google Scholar] [CrossRef] [PubMed]
  223. Choi, G.; Rejinold, N.S.; Piao, H.; Choy, J.-H. Inorganic–inorganic nanohybrids for drug delivery, imaging and photo-therapy: Recent developments and future scope. Chem. Sci. 2021. [Google Scholar] [CrossRef]
  224. Farcas, C.G.; Dehelean, C.; Pinzaru, I.A.; Mioc, M.; Socoliuc, V.; Moaca, E.-A.; Avram, S.; Ghiulai, R.; Coricovac, D.; Pavel, I.; et al. Thermosensitive Betulinic Acid-Loaded Magnetoliposomes: A Promising Antitumor Potential for Highly Aggressive Human Breast Adenocarcinoma Cells Under Hyperthermic Conditions. Int. J. Nanomed. 2020, 15, 8175–8200. [Google Scholar] [CrossRef]
  225. Shen, K.; Wang, X.; Zhang, Y.; Zhu, H.; Chen, Z.; Huang, C.; Mai, Y. Insights into the role of interface modification in performance enhancement of ZnTe:Cu contacted CdTe thin film solar cells. Sol. Energy 2020, 201, 55–62. [Google Scholar] [CrossRef]
  226. Paterlini, V.; Bettinelli, M.; Rizzi, R.; El Khouri, A.; Rossi, M.; Della Ventura, G.; Capitelli, F. Characterization and Luminescence of Eu3+- and Gd3+-Doped Hydroxyapatite Ca10(PO4)6(OH)2. Crystals 2020, 10, 806. [Google Scholar] [CrossRef]
  227. Mout, R.; Ray, M.; Yesilbag, T.G.; Lee, Y.W.; Tay, T.; Sasaki, K.; Rotello, V.M. Direct Cytosolic Delivery of CRISPR/Cas9-Ribonucleoprotein for Efficient Gene Editing. ACS Nano 2017, 11, 2452–2458. [Google Scholar] [CrossRef] [Green Version]
  228. Fonte, P.; Reis, S.; Sarmento, B. Facts and evidences on the lyophilization of polymeric nanoparticles for drug delivery. J. Control. Release 2016, 225, 75–86. [Google Scholar] [CrossRef] [PubMed]
  229. Masood, F. Polymeric nanoparticles for targeted drug delivery system for cancer therapy. Mater. Sci. Eng. C 2016, 60, 569–578. [Google Scholar] [CrossRef]
  230. Chertok, B.; David, A.E.; Yang, V.C. Polyethyleneimine-modified iron oxide nanoparticles for brain tumor drug delivery using magnetic targeting and intra-carotid administration. Biomaterials 2010, 31, 6317–6324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Díaz-Moscoso, A.; Guilloteau, N.; Bienvenu, C.; Méndez-Ardoy, A.; Jiménez Blanco, J.L.; Benito, J.M.; Le Gourriérec, L.; Di Giorgio, C.; Vierling, P.; Defaye, J.; et al. Mannosyl-coated nanocomplexes from amphiphilic cyclodextrins and pDNA for site-specific gene delivery. Biomaterials 2011, 32, 7263–7273. [Google Scholar] [CrossRef] [PubMed]
  232. Ayyappan, J.P.; Sami, H.; Rajalekshmi, D.C.; Sivakumar, S.; Abraham, A. Immunocompatibility and Toxicity Studies of Poly-L-Lysine Nanocapsules in Sprague–Dawley Rats for Drug-Delivery Applications. Chem. Biol. Drug Des. 2014, 84, 292–299. [Google Scholar] [CrossRef]
  233. Fornaguera, C.; Dols-Perez, A.; Calderó, G.; García-Celma, M.J.; Camarasa, J.; Solans, C. PLGA nanoparticles prepared by nano-emulsion templating using low-energy methods as efficient nanocarriers for drug delivery across the blood–brain barrier. J. Control. Release 2015, 211, 134–143. [Google Scholar] [CrossRef]
  234. Llop, J.; Jiang, P.; Marradi, M.; Gómez-Vallejo, V.; Echeverría, M.; Yu, S.; Puigivila, M.; Baz, Z.; Szczupak, B.; Pérez-Campaña, C.; et al. Visualisation of dual radiolabelled poly(lactide-co-glycolide) nanoparticle degradation in vivo using energy-discriminant SPECT. J. Mater. Chem. B 2015, 3, 6293–6300. [Google Scholar] [CrossRef] [Green Version]
  235. Kreyling, W.G.; Abdelmonem, A.M.; Ali, Z.; Alves, F.; Geiser, M.; Haberl, N.; Hartmann, R.; Hirn, S.; de Aberasturi, D.J.; Kantner, K.; et al. In vivo integrity of polymer-coated gold nanoparticles. Nat. Nanotechnol. 2015, 10, 619–623. [Google Scholar] [CrossRef] [Green Version]
  236. Zhu, Y.; Liu, C.; Pang, Z. Dendrimer-based drug delivery systems for brain targeting. Biomolecules 2019, 9, 790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Gauro, R.; Nandave, M.; Jain, V.K.; Jain, K. Advances in dendrimer-mediated targeted drug delivery to the brain. J. Nanopart. Res. 2021, 23, 1–20. [Google Scholar] [CrossRef]
  238. Moscariello, P.; Ng, D.Y.W.; Jansen, M.; Weil, T.; Luhmann, H.J.; Hedrich, J. Brain delivery of multifunctional dendrimer protein bioconjugates. Adv. Sci. 2018, 5, 1700897. [Google Scholar] [CrossRef] [PubMed]
  239. Janaszewska, A.; Lazniewska, J.; Trzepiński, P.; Marcinkowska, M.; Klajnert-Maculewicz, B. Cytotoxicity of dendrimers. Biomolecules 2019, 9, 330. [Google Scholar] [CrossRef] [Green Version]
  240. Taharabaru, T.; Yokoyama, R.; Higashi, T.; Mohammed, A.F.A.; Inoue, M.; Maeda, Y.; Niidome, T.; Onodera, R.; Motoyama, K. Genome editing in a wide area of the brain using dendrimer-based ternary polyplexes of Cas9 ribonucleoprotein. ACS Appl. Mater. Interfaces 2020, 12, 21386–21397. [Google Scholar] [CrossRef]
  241. Ke, W.; Shao, K.; Huang, R.; Han, L.; Liu, Y.; Li, J.; Kuang, Y.; Ye, L.; Lou, J.; Jiang, C. Gene delivery targeted to the brain using an Angiopep-conjugated polyethyleneglycol-modified polyamidoamine dendrimer. Biomaterials 2009, 30, 6976–6985. [Google Scholar] [CrossRef] [PubMed]
  242. Liu, Y.; Huang, R.; Han, L.; Ke, W.; Shao, K.; Ye, L.; Lou, J.; Jiang, C. Brain-targeting gene delivery and cellular internalization mechanisms for modified rabies virus glycoprotein RVG29 nanoparticles. Biomaterials 2009, 30, 4195–4202. [Google Scholar] [CrossRef]
  243. Xiao, K.; Li, Y.; Luo, J.; Lee, J.S.; Xiao, W.; Gonik, A.M.; Agarwal, R.G.; Lam, K.S. The effect of surface charge on in vivo biodistribution of PEG-oligocholic acid based micellar nanoparticles. Biomaterials 2011, 32, 3435–3446. [Google Scholar] [CrossRef] [Green Version]
  244. Gustafson, H.H.; Holt-Casper, D.; Grainger, D.W.; Ghandehari, H. Nanoparticle uptake: The phagocyte problem. Nano Today 2015, 10, 487–510. [Google Scholar] [CrossRef] [Green Version]
  245. Kulkarni, S.A.; Feng, S.-S. Effects of Particle Size and Surface Modification on Cellular Uptake and Biodistribution of Polymeric Nanoparticles for Drug Delivery. Pharm. Res. 2013, 30, 2512–2522. [Google Scholar] [CrossRef] [PubMed]
  246. Florence, A.T.; Hillery, A.M.; Hussain, N.; Jani, P.U. Nanoparticles as carriers for oral peptide absorption: Studies on particle uptake and fate. J. Control. Release 1995, 36, 39–46. [Google Scholar] [CrossRef]
  247. Davis, M.E.; Chen, Z.; Shin, D.M. Nanoparticle therapeutics: An emerging treatment modality for cancer. In Nanoscience and Technology; Co-Published with Macmillan Publishers Ltd.: London, UK, 2009; pp. 239–250. [Google Scholar]
  248. Duan, X.; Li, Y. Physicochemical characteristics of nanoparticles affect circulation, biodistribution, cellular internalization, and trafficking. Small 2013, 9, 1521–1532. [Google Scholar] [CrossRef]
  249. Zhao, Z.; Ukidve, A.; Krishnan, V.; Mitragotri, S. Effect of physicochemical and surface properties on in vivo fate of drug nanocarriers. Adv. Drug Deliv. Rev. 2019, 143, 3–21. [Google Scholar] [CrossRef] [PubMed]
  250. Makadia, H.K.; Siegel, S.J. Poly Lactic-co-Glycolic Acid (PLGA) as biodegradable controlled drug delivery carrier. Polymers 2011, 3, 1377–1397. [Google Scholar] [CrossRef]
  251. Rescignano, N.; Tarpani, L.; Romani, A.; Bicchi, I.; Mattioli, S.; Emiliani, C.; Torre, L.; Kenny, J.M.; Martino, S.; Latterini, L.; et al. In-vitro degradation of PLGA nanoparticles in aqueous medium and in stem cell cultures by monitoring the cargo fluorescence spectrum. Polym. Degrad. Stab. 2016, 134, 296–304. [Google Scholar] [CrossRef]
  252. Zhao, J.; Wang, Y.; Ma, Y.; Liu, Y.; Yan, B.; Wang, L. Smart nanocarrier based on PEGylated hyaluronic acid for deacetyl mycoepoxydience: High stability with enhanced bioavailability and efficiency. Carbohydr. Polym. 2019, 203, 356–368. [Google Scholar] [CrossRef]
  253. Adamo, G.; Campora, S.; Ghersi, G. Functionalization of nanoparticles in specific targeting and mechanism release. In Nanostructures for Novel Therapy; Elsevier: Amsterdam, The Netherlands, 2017; pp. 57–80. [Google Scholar]
  254. Adamo, G.; Grimaldi, N.; Campora, S.; Bulone, D.; Bondì, M.; Al-Sheikhly, M.; Sabatino, M.; Dispenza, C.; Ghersi, G. Multi-Functional Nanogels for Tumor Targeting and Redox-Sensitive Drug and siRNA Delivery. Molecules 2016, 21, 1594. [Google Scholar] [CrossRef]
  255. Zhu, H.; Zhang, L.; Tong, S.; Lee, C.M.; Deshmukh, H.; Bao, G. Spatial control of in vivo CRISPR–Cas9 genome editing via nanomagnets. Nat. Biomed. Eng. 2019, 3, 126–136. [Google Scholar] [CrossRef]
  256. Mondal, S.; Dorozhkin, S.V.; Pal, U. Recent progress on fabrication and drug delivery applications of nanostructured hydroxyapatite. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2018, 10, e1504. [Google Scholar] [CrossRef]
  257. Chuang, C.K.; Chen, C.H.; Huang, C.L.; Su, Y.H.; Peng, S.H.; Lin, T.Y.; Tai, H.C.; Yang, T.S.; Tu, C.F. Generation of GGTA1 Mutant Pigs by Direct Pronuclear Microinjection of CRISPR/Cas9 Plasmid Vectors. Anim. Biotechnol. 2017, 28, 174–181. [Google Scholar] [CrossRef]
  258. Love, J.; Gribbin, C.; Mather, C.; Sang, H. Transgenic Birds by DNA Microinjection. Bio/Technology 1994, 12, 60–63. [Google Scholar] [CrossRef]
  259. Liang, X.; Potter, J.; Kumar, S.; Ravinder, N.; Chesnut, J.D. Enhanced CRISPR/Cas9-mediated precise genome editing by improved design and delivery of gRNA, Cas9 nuclease, and donor DNA. J. Biotechnol. 2017, 241, 136–146. [Google Scholar] [CrossRef] [PubMed]
  260. Chen, S.; Lee, B.; Lee, A.Y.; Modzelewski, A.J.; He, L. Highly Efficient Mouse Genome Editing by CRISPR Ribonucleoprotein Electroporation of Zygotes. J. Biol. Chem. 2016, 291, 14457–14467. [Google Scholar] [CrossRef] [Green Version]
  261. Dong, C.; Qu, L.; Wang, H.; Wei, L.; Dong, Y.; Xiong, S. Targeting hepatitis B virus cccDNA by CRISPR/Cas9 nuclease efficiently inhibits viral replication. Antivir. Res. 2015, 118, 110–117. [Google Scholar] [CrossRef] [PubMed]
  262. Park, H.; Oh, J.; Shim, G.; Cho, B.; Chang, Y.; Kim, S.; Baek, S.; Kim, H.; Shin, J.; Choi, H.; et al. In vivo neuronal gene editing via CRISPR–Cas9 amphiphilic nanocomplexes alleviates deficits in mouse models of Alzheimer’s disease. Nat. Neurosci. 2019, 22, 524–528. [Google Scholar] [CrossRef] [PubMed]
  263. Höijer, I.; Tsai, Y.-C.; Clark, T.A.; Kotturi, P.; Dahl, N.; Stattin, E.-L.; Bondeson, M.-L.; Feuk, L.; Gyllensten, U.; Ameur, A. Detailed analysis of HTT repeat elements in human blood using targeted amplification-free long-read sequencing. Hum. Mutat. 2018, 39, 1262–1272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Qing, X.; Walter, J.; Jarazo, J.; Arias-Fuenzalida, J.; Hillje, A.-L.; Schwamborn, J.C. CRISPR/Cas9 and piggyBac-mediated footprint-free LRRK2-G2019S knock-in reveals neuronal complexity phenotypes and α-Synuclein modulation in dopaminergic neurons. Stem Cell Res. 2017, 24, 44–50. [Google Scholar] [CrossRef] [PubMed]
  265. Karimian, A.; Gorjizadeh, N.; Alemi, F.; Asemi, Z.; Azizian, K.; Soleimanpour, J.; Malakouti, F.; Targhazeh, N.; Majidinia, M.; Yousefi, B. CRISPR/Cas9 novel therapeutic road for the treatment of neurodegenerative diseases. Life Sci. 2020, 259, 118165. [Google Scholar] [CrossRef]
  266. Hitti, F.L.; Yang, A.I.; Gonzalez-Alegre, P.; Baltuch, G.H. Human gene therapy approaches for the treatment of Parkinson’s disease: An overview of current and completed clinical trials. Parkinsonism Relat. Disord. 2019, 66, 16–24. [Google Scholar] [CrossRef] [PubMed]
  267. Koentjoro, B.; Park, J.-S.; Sue, C.M. Nix restores mitophagy and mitochondrial function to protect against PINK1/Parkin-related Parkinson’s disease. Sci. Rep. 2017, 7, 44373. [Google Scholar] [CrossRef] [PubMed]
  268. Han, Y.; Khodr, C.E.; Sapru, M.K.; Pedapati, J.; Bohn, M.C. A microRNA embedded AAV alpha-synuclein gene silencing vector for dopaminergic neurons. Brain Res. 2011, 1386, 15–24. [Google Scholar] [CrossRef] [Green Version]
  269. Lewis, J.; Melrose, H.; Bumcrot, D.; Hope, A.; Zehr, C.; Lincoln, S.; Braithwaite, A.; He, Z.; Ogholikhan, S.; Hinkle, K.; et al. In vivo silencing of alpha-synuclein using naked siRNA. Mol. Neurodegener. 2008, 3, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  270. Kirik, D.; Georgievska, B.; Burger, C.; Winkler, C.; Muzyczka, N.; Mandel, R.J.; Bjorklund, A. Reversal of motor impairments in parkinsonian rats by continuous intrastriatal delivery of L-dopa using rAAV-mediated gene transfer. Proc. Natl. Acad. Sci. USA 2002, 99, 4708–4713. [Google Scholar] [CrossRef] [Green Version]
  271. Leriche, L.; Bjorklund, T.; Breysse, N.; Besret, L.; Gregoire, M.-C.; Carlsson, T.; Dolle, F.; Mandel, R.J.; Deglon, N.; Hantraye, P.; et al. Positron Emission Tomography Imaging Demonstrates Correlation between Behavioral Recovery and Correction of Dopamine Neurotransmission after Gene Therapy. J. Neurosci. 2009, 29, 1544–1553. [Google Scholar] [CrossRef] [PubMed]
  272. Cederfjäll, E.; Broom, L.; Kirik, D. Controlled Striatal DOPA Production From a Gene Delivery System in a Rodent Model of Parkinson’s Disease. Mol. Ther. 2015, 23, 896–906. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Cederfjäll, E.; Nilsson, N.; Sahin, G.; Chu, Y.; Nikitidou, E.; Björklund, T.; Kordower, J.H.; Kirik, D. Continuous DOPA synthesis from a single AAV: Dosing and efficacy in models of Parkinson’s disease. Sci. Rep. 2013, 3, 2157. [Google Scholar] [CrossRef] [Green Version]
  274. Christine, C.W.; Starr, P.A.; Larson, P.S.; Eberling, J.L.; Jagust, W.J.; Hawkins, R.A.; VanBrocklin, H.F.; Wright, J.F.; Bankiewicz, K.S.; Aminoff, M.J. Safety and tolerability of putaminal AADC gene therapy for Parkinson disease. Neurology 2009, 73, 1662–1669. [Google Scholar] [CrossRef] [Green Version]
  275. Muramatsu, S.-I.; Fujimoto, K.-I.; Kato, S.; Mizukami, H.; Asari, S.; Ikeguchi, K.; Kawakami, T.; Urabe, M.; Kume, A.; Sato, T.; et al. A Phase I Study of Aromatic L-Amino Acid Decarboxylase Gene Therapy for Parkinson’s Disease. Mol. Ther. 2010, 18, 1731–1735. [Google Scholar] [CrossRef] [Green Version]
  276. Ciesielska, A.; Samaranch, L.; San Sebastian, W.; Dickson, D.W.; Goldman, S.; Forsayeth, J.; Bankiewicz, K.S. Depletion of AADC activity in caudate nucleus and putamen of Parkinson’s disease patients; implications for ongoing AAV2-AADC gene therapy trial. PLoS ONE 2017, 12, e0169965. [Google Scholar] [CrossRef] [Green Version]
  277. Christine, C.W.; Bankiewicz, K.S.; Van Laar, A.D.; Richardson, R.M.; Ravina, B.; Kells, A.P.; Boot, B.; Martin, A.J.; Nutt, J.; Thompson, M.E.; et al. Magnetic resonance imaging–guided phase 1 trial of putaminal AADC gene therapy for Parkinson’s disease. Ann. Neurol. 2019, 85, 704–714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  278. Wang, L.; Muramatsu, S.; Lu, Y.; Ikeguchi, K.; Fujimoto, K.; Okada, T.; Mizukami, H.; Hanazono, Y.; Kume, A.; Urano, F.; et al. Delayed delivery of AAV-GDNF prevents nigral neurodegeneration and promotes functional recovery in a rat model of Parkinson’s disease. Gene Ther. 2002, 9, 381–389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  279. Kordower, J.H.; Emborg, M.E.; Bloch, J.; Ma, S.Y.; Chu, Y.; Leventhal, L.; McBride, J.; Chen, E.-Y.; Palfi, S.; Roitberg, B.Z.; et al. Neurodegeneration Prevented by Lentiviral Vector Delivery of GDNF in Primate Models of Parkinson’s Disease. Science 2000, 290, 767–773. [Google Scholar] [CrossRef] [PubMed]
  280. Chen, C.; Li, X.; Ge, G.; Liu, J.; Biju, K.C.; Laing, S.D.; Qian, Y.; Ballard, C.; He, Z.; Masliah, E.; et al. GDNF-expressing macrophages mitigate loss of dopamine neurons and improve Parkinsonian symptoms in MitoPark mice. Sci. Rep. 2018, 8, 5460. [Google Scholar] [CrossRef]
  281. Chen, C.; Guderyon, M.J.; Li, Y.; Ge, G.; Bhattacharjee, A.; Ballard, C.; He, Z.; Masliah, E.; Clark, R.A.; O’Connor, J.C.; et al. Non-toxic HSC transplantation-based macrophage/microglia-mediated GDNF delivery for Parkinson’s disease. Mol. Ther. Clin. Dev. 2020, 17, 83–98. [Google Scholar] [CrossRef] [Green Version]
  282. Lindvall, O.; Wahlberg, L.U. Encapsulated cell biodelivery of GDNF: A novel clinical strategy for neuroprotection and neuroregeneration in Parkinson’s disease? Exp. Neurol. 2008, 209, 82–88. [Google Scholar] [CrossRef]
  283. Fan, C.-H.; Ting, C.-Y.; Lin, C.; Chan, H.-L.; Chang, Y.-C.; Chen, Y.-Y.; Liu, H.-L.; Yeh, C.-K. Noninvasive, Targeted and Non-Viral Ultrasound-Mediated GDNF-Plasmid Delivery for Treatment of Parkinson’s Disease. Sci. Rep. 2016, 6, 19579. [Google Scholar] [CrossRef]
  284. Mead, B.P.; Kim, N.; Miller, G.W.; Hodges, D.; Mastorakos, P.; Klibanov, A.L.; Mandell, J.W.; Hirsh, J.; Suk, J.S.; Hanes, J.; et al. Novel Focused Ultrasound Gene Therapy Approach Noninvasively Restores Dopaminergic Neuron Function in a Rat Parkinson’s Disease Model. Nano Lett. 2017, 17, 3533–3542. [Google Scholar] [CrossRef]
  285. Berke, J.D. What does dopamine mean? Nat. Neurosci. 2018, 21, 787–793. [Google Scholar] [CrossRef] [PubMed]
  286. Eban-Rothschild, A.; Rothschild, G.; Giardino, W.J.; Jones, J.R.; de Lecea, L. VTA dopaminergic neurons regulate ethologically relevant sleep–wake behaviors. Nat. Neurosci. 2016, 19, 1356–1366. [Google Scholar] [CrossRef]
  287. Martinez-Martín, P.; Rodriguez-Blazquez, C.; Paz, S.; Forjaz, M.J.; Frades-Payo, B.; Cubo, E.; Pedro-Cuesta, J.d.; Lizán, L.; ELEP Group. Parkinson symptoms and health related quality of life as predictors of costs: A longitudinal observational study with linear mixed model analysis. PLoS ONE 2015, 10, e0145310. [Google Scholar] [CrossRef] [Green Version]
  288. Haddad, F.; Sawalha, M.; Khawaja, Y.; Najjar, A.; Karaman, R. Dopamine and Levodopa Prodrugs for the Treatment of Parkinson’s Disease. Molecules 2018, 23, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  289. Porras, G.; De Deurwaerdere, P.; Li, Q.; Marti, M.; Morgenstern, R.; Sohr, R.; Bezard, E.; Morari, M.; Meissner, W.G. L-dopa-induced dyskinesia: Beyond an excessive dopamine tone in the striatum. Sci. Rep. 2015, 4, 3730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  290. Eberling, J.L.; Jagust, W.J.; Christine, C.W.; Starr, P.; Larson, P.; Bankiewicz, K.S.; Aminoff, M.J. Results from a phase I safety trial of hAADC gene therapy for Parkinson disease. Neurology 2008, 70, 1980–1983. [Google Scholar] [CrossRef] [PubMed]
  291. Bankiewicz, K.S.; Forsayeth, J.; Eberling, J.L.; Sanchez-Pernaute, R.; Pivirotto, P.; Bringas, J.; Herscovitch, P.; Carson, R.E.; Eckelman, W.; Reutter, B.; et al. Long-term clinical improvement in MPTP-lesioned primates after gene therapy with AAV-hAADC. Mol. Ther. 2006, 14, 564–570. [Google Scholar] [CrossRef]
  292. Merkel, S.F.; Andrews, A.M.; Lutton, E.M.; Mu, D.; Hudry, E.; Hyman, B.T.; Maguire, C.A.; Ramirez, S.H. Trafficking of adeno-associated virus vectors across a model of the blood–brain barrier; a comparative study of transcytosis and transduction using primary human brain endothelial cells. J. Neurochem. 2017, 140, 21. [Google Scholar] [CrossRef] [Green Version]
  293. Mittermeyer, G.; Christine, C.W.; Rosenbluth, K.H.; Baker, S.L.; Starr, P.; Larson, P.; Kaplan, P.L.; Forsayeth, J.; Aminoff, M.J.; Bankiewicz, K.S. Long-term evaluation of a phase 1 study of AADC gene therapy for parkinson’s disease. Hum. Gene Ther. 2011, 23, 377–381. [Google Scholar] [CrossRef] [Green Version]
  294. Witt, J.; Marks, W.J. An Update on Gene Therapy in Parkinson’s Disease. Curr. Neurol. Neurosci. Rep. 2011, 11, 362–370. [Google Scholar] [CrossRef]
  295. Airaksinen, M.S.; Saarma, M. The GDNF family: Signalling, biological functions and therapeutic value. Nat. Rev. Neurosci. 2002, 3, 383–394. [Google Scholar] [CrossRef]
  296. Winkler, C.; Sauer, H.; Lee, C.S.; Björklund, A. Short-term GDNF treatment provides long-term rescue of lesioned nigral dopaminergic neurons in a rat model of Parkinson’s disease. J. Neurosci. 1996, 16, 72. [Google Scholar] [CrossRef] [Green Version]
  297. Wong, C.E.D.; Hua, K.; Monis, S.; Saxena, V.; Norazit, A.; Noor, S.M.; Ekker, M. Gdnf affects early diencephalic dopaminergic neuron development through regulation of differentiation-associated transcription factors in zebrafish. J. Neurochem. 2020, 156, 481–498. [Google Scholar] [CrossRef] [PubMed]
  298. Gash, D.M.; Zhang, Z.; Ovadia, A.; Cass, W.A.; Yi, A.; Simmerman, L.; Russell, D.; Martin, D.; Lapchak, P.A.; Collins, F.; et al. Functional recovery in parkinsonian monkeys treated with GDNF. Nature 1996, 380, 252–255. [Google Scholar] [CrossRef]
  299. Gill, S.S.; Patel, N.K.; Hotton, G.R.; O’Sullivan, K.; McCarter, R.; Bunnage, M.; Brooks, D.J.; Svendsen, C.N.; Heywood, P. Direct brain infusion of glial cell line–derived neurotrophic factor in Parkinson disease. Nat. Med. 2003, 9, 589–595. [Google Scholar] [CrossRef]
  300. Patel, N.K.; Bunnage, M.; Plaha, P.; Svendsen, C.N.; Heywood, P.; Gill, S.S. Intraputamenal infusion of glial cell line-derived neurotrophic factor in PD: A two-year outcome study. Ann. Neurol. 2005, 57, 298–302. [Google Scholar] [CrossRef]
  301. Lang, A.E.; Gill, S.; Patel, N.K.; Lozano, A.; Nutt, J.G.; Penn, R.; Brooks, D.; Hotton, G.; Moro, E.; Heywood, P.; et al. Randomized controlled trial of intraputamenal glial cell line–derived neurotrophic factor infusion in Parkinson disease. Ann. Neurol. 2006, 59, 459. [Google Scholar] [CrossRef] [PubMed]
  302. Salvatore, M.; Ai, Y.; Fischer, B.; Zhang, A.; Grondin, R.; Zhang, Z.; Gerhardt, G.; Gash, D. Point source concentration of GDNF may explain failure of phase II clinical trial. Exp. Neurol. 2006, 202, 497–505. [Google Scholar] [CrossRef] [PubMed]
  303. Whone, A.; Luz, M.; Boca, M.; Woolley, M.; Mooney, L.; Dharia, S.; Broadfoot, J.; Cronin, D.; Schroers, C.; Barua, N.U.; et al. Randomized trial of intermittent intraputamenal glial cell line-derived neurotrophic factor in Parkinson’s disease. Brain 2019, 142, 512–525. [Google Scholar] [CrossRef] [Green Version]
  304. Whone, A.L.; Boca, M.; Luz, M.; Woolley, M.; Mooney, L.; Dharia, S.; Broadfoot, J.; Cronin, D.; Schroers, C.; Barua, N.U.; et al. Extended Treatment with Glial Cell Line-Derived Neurotrophic Factor in Parkinson’s Disease. J. Parkinsons. Dis. 2019, 9, 301–313. [Google Scholar] [CrossRef] [Green Version]
  305. Garbayo, E.; Montero-Menei, C.N.; Ansorena, E.; Lanciego, J.L.; Aymerich, M.S.; Blanco-Prieto, M.J. Effective GDNF brain delivery using microspheres—a promising strategy for Parkinson’s disease. J. Control. Release 2009, 135, 11. [Google Scholar] [CrossRef] [PubMed]
  306. Kordower, J.H.; Herzog, C.D.; Dass, B.; Bakay, R.A.E.; Stansell, J.; Gasmi, M.; Bartus, R.T. Delivery of neurturin by AAV2 (CERE-120)-mediated gene transfer provides structural and functional neuroprotection and neurorestoration in MPTP-treated monkeys. Ann. Neurol. 2006, 60, 706–715. [Google Scholar] [CrossRef] [PubMed]
  307. Espinoza, S.; Scarpato, M.; Damiani, D.; Managò, F.; Mereu, M.; Contestabile, A.; Peruzzo, O.; Carninci, P.; Santoro, C.; Papaleo, F.; et al. SINEUP Non-coding RNA Targeting GDNF Rescues Motor Deficits and Neurodegeneration in a Mouse Model of Parkinson’s Disease. Mol. Ther. 2020, 28, 642–652. [Google Scholar] [CrossRef] [Green Version]
  308. Herzog, C.D.; Dass, B.; Holden, J.E.; Stansell, J.; Gasmi, M.; Tuszynski, M.H.; Bartus, R.T.; Kordower, J.H. Striatal delivery of CERE-120, an AAV2 vector encoding human neurturin, enhances activity of the dopaminergic nigrostriatal system in aged monkeys. Mov. Disord. 2007, 22, 1124–1132. [Google Scholar] [CrossRef]
  309. Marks, W.J.; Ostrem, J.L.; Verhagen, L.; Starr, P.A.; Larson, P.S.; Bakay, R.A.; Taylor, R.; Cahn-Weiner, D.A.; Stoessl, A.J.; Olanow, C.W.; et al. Safety and tolerability of intraputaminal delivery of CERE-120 (adeno-associated virus serotype 2–neurturin) to patients with idiopathic Parkinson’s disease: An open-label, phase I trial. Lancet Neurol. 2008, 7, 400–408. [Google Scholar] [CrossRef]
  310. Olanow, C.W.; Bartus, R.T.; Baumann, T.L.; Factor, S.; Boulis, N.; Stacy, M.; Turner, D.A.; Marks, W.; Larson, P.; Starr, P.A.; et al. Gene delivery of neurturin to putamen and Substantia nigra in P arkinson disease: A double-blind, randomized, controlled trial. Ann. Neurol. 2015, 78, 248. [Google Scholar] [CrossRef]
  311. Chu, Y.; Bartus, R.T.; Manfredsson, F.P.; Olanow, C.W.; Kordower, J.H. Long-term post-mortem studies following neurturin gene therapy in patients with advanced Parkinson’s disease. Brain 2020, 143, 960–975. [Google Scholar] [CrossRef]
  312. Pickrell, A.M.; Youle, R.J. The Roles of PINK1, Parkin, and Mitochondrial Fidelity in Parkinson’s Disease. Neuron 2015, 85, 257–273. [Google Scholar] [CrossRef] [Green Version]
  313. Hernandez, D.G.; Reed, X.; Singleton, A.B. Genetics in Parkinson disease: Mendelian versus non-Mendelian inheritance. J. Neurochem. 2016, 139, 59–74. [Google Scholar] [CrossRef]
  314. Miller, S.; Muqit, M.M.K. Therapeutic approaches to enhance PINK1/Parkin mediated mitophagy for the treatment of Parkinson’s disease. Neurosci. Lett. 2019, 705, 7–13. [Google Scholar] [CrossRef]
  315. Malpartida, A.B.; Williamson, M.; Narendra, D.P.; Wade-Martins, R.; Ryan, B.J. Mitochondrial Dysfunction and Mitophagy in Parkinson’s Disease: From Mechanism to Therapy. Trends Biochem. Sci. 2021, 46, 329–343. [Google Scholar] [CrossRef] [PubMed]
  316. Yan, C.; Gong, L.; Chen, L.; Xu, M.; Abou-Hamdan, H.; Tang, M.; Désaubry, L.; Song, Z. PHB2 (prohibitin 2) promotes PINK1-PRKN/Parkin-dependent mitophagy by the PARL-PGAM5-PINK1 axis. Autophagy 2020, 16, 419–434. [Google Scholar] [CrossRef] [PubMed]
  317. Chung, E.; Choi, Y.; Park, J.; Nah, W.; Park, J.; Jung, Y.; Lee, J.; Lee, H.; Park, S.; Hwang, S.; et al. Intracellular delivery of Parkin rescues neurons from accumulation of damaged mitochondria and pathological α-synuclein. Sci. Adv. 2020, 6, eaba1193. [Google Scholar] [CrossRef] [PubMed]
  318. Khodr, C.E.; Sapru, M.K.; Pedapati, J.; Han, Y.; West, N.C.; Kells, A.P.; Bankiewicz, K.S.; Bohn, M.C. An alpha-synuclein AAV gene silencing vector ameliorates a behavioral deficit in a rat model of Parkinson’s disease, but displays toxicity in dopamine neurons. Brain Res. 2011, 1395, 94–107. [Google Scholar] [CrossRef] [Green Version]
  319. Khodr, C.E.; Becerra, A.; Han, Y.; Bohn, M.C. Targeting alpha-synuclein with a microRNA-embedded silencing vector in the rat Substantia nigra: Positive and negative effects. Brain Res. 2014, 1550, 47–60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  320. Alarcón-Arís, D.; Recasens, A.; Galofré, M.; Carballo-Carbajal, I.; Zacchi, N.; Ruiz-Bronchal, E.; Pavia-Collado, R.; Chica, R.; Ferrés-Coy, A.; Santos, M.; et al. Selective α-Synuclein Knockdown in Monoamine Neurons by Intranasal Oligonucleotide Delivery: Potential Therapy for Parkinson’s Disease. Mol. Ther. 2018, 26, 550–567. [Google Scholar] [CrossRef] [Green Version]
  321. Perez-Pinera, P.; Kocak, D.D.; Vockley, C.M.; Adler, A.F.; Kabadi, A.M.; Polstein, L.R.; Thakore, P.I.; Glass, K.A.; Ousterout, D.G.; Leong, K.W.; et al. RNA-guided gene activation by CRISPR-Cas9–based transcription factors. Nat. Methods 2013, 10, 973–976. [Google Scholar] [CrossRef] [Green Version]
  322. Liu, Z.; Cheung, H. Stem cell-based therapies for Parkinson disease. Int. J. Mol. Sci. 2020, 21, 8060. [Google Scholar] [CrossRef]
  323. Coccia, E.; Ahfeldt, T. Towards physiologically relevant human pluripotent stem cell (hPSC) models of Parkinson’s disease. Stem Cell Res. Ther. 2021, 12, 253. [Google Scholar] [CrossRef]
  324. Takahashi, K.; Yamanaka, S. Induction of pluripotent stem cells from mouse embyonic and adult fibroblast cultures by defined factors. Cell 2006, 126, 663–676. [Google Scholar] [CrossRef] [Green Version]
  325. Park, H.; Arora, N.; Huo, H. Disease-specfic induced pluripotent stem cells. Cell 2008, 134, 877–886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  326. Soldner, F.; Hockemeyer, D.; Beard, C. Parkinson’s Disease Patient-Derived Induced Pluripotent Stem Cells Free of Viral Reprogramming Factors. Cell 2009, 135, 964–977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  327. Zhang, Q.; Chen, W.; Tan, S.; Lin, T. Stem cells for modeling and therapy of Parkinson’s disease. Hum. Gene Ther. 2017, 28, 85–98. [Google Scholar] [CrossRef]
  328. Safari, F.; Hatam, G.; Behbahani, A.B.; Rezaei, V.; Barekati-Mowahed, M.; Petramfar, P.; Khademi, F. CRISPR system: A high-throughput toolbox for research and treatment of Parkinson’s disease. Cell. Mol. Neurobiol. 2020, 40, 477–493. [Google Scholar] [CrossRef]
  329. Sison, S.L.; Vermilyea, S.C.; Emborg, M.E.; Ebert, A.D. Using patient-derived induced pluripotent stem cells to identify Parkinson’s disease-relevant phenotypes. Curr. Neurol. Neurosci. Rep. 2018, 18, 1–14. [Google Scholar] [CrossRef] [PubMed]
  330. Ke, M.; Chong, C.M.; Su, H. Using induced pluripotent stem cells for modeling Parkinson’s disease. World J. Stem Cells 2019, 11, 634. [Google Scholar] [CrossRef] [PubMed]
  331. Hu, X.; Mao, C.; Fan, L.; Luo, H.; Hu, Z.; Zhang, S.; Yang, Z.; Zheng, H.; Sun, H.; Fan, Y.; et al. Modeling Parkinson’s Disease Using Induced Pluripotent Stem Cells. Stem Cells Int. 2020, 2020, 1–15. [Google Scholar] [CrossRef]
  332. Barker, R.A.; Studer, L.; Cattaneo, E.; Takahashi, J. G-Force PD: A global initiative in coordinating stem cell-based dopamine treatments for Parkinson’s disease. Parkinsons. Dis. 2015, 1, 1–5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  333. Ahfeldt, T.; Ordureau, A.; Bell, C.; Sarrafha, L.; Sun, C.; Piccinotti, S.; Grass, T.; Parfitt, G.M.; Paulo, J.A.; Yanagawa, F.; et al. Pathogenic pathways in early-onset autosomal recessive Parkinson’s disease discovered using isogenic human dopaminergic neurons. Stem Cell Rep. 2020, 14, 75–90. [Google Scholar] [CrossRef] [Green Version]
  334. Calatayud, C.; Carola, G.; Consiglio, A.; Raya, A. Modeling the genetic complexity of Parkinson’s disease by targeted genome edition in iPS cells. Curr. Opin. Genet. Dev. 2017, 46, 123–131. [Google Scholar] [CrossRef]
  335. Di Domenico, A.; Carola, G.; Calatayud, C.; Pons-Espinal, M.; Muñoz, J.P.; Richaud-Patin, Y.; Consiglio, A. Patient-specific iPSC-derived astrocytes contribute to non-cell-autonomous neurodegeneration in Parkinson’s disease. Stem Cell Rep. 2019, 12, 213–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. Jing, C.C.; Luo, X.G.; Cui, H.G.; Li, F.R.; Li, P.; Jiang, E.Z.; Ren, Y.; Pang, H. Screening of polymorphisms located in the FGF20 and TMEM175 genes in North Chinese Parkinson’s disease patients. Genet. Mol. Res. 2015, 14, 13679–13687. [Google Scholar] [CrossRef] [PubMed]
  337. Harjuhaahto, S.; Rasila, T.S.; Molchanova, S.M.; Woldegebriel, R.; Kvist, J.; Konovalova, S.; Tyynismaa, H. ALS and Parkinson’s disease genes CHCHD10 and CHCHD2 modify synaptic transcriptomes in human iPSC-derived motor neurons. Neurobiol. Dis. 2020, 14, 104940. [Google Scholar] [CrossRef] [PubMed]
  338. Reyes, L.H.; Kao, K.C. Growth-Coupled Carotenoids Production Using Adaptive Laboratory Evolution. In Methods in Molecular Biology; Humana Press: New York, NY, USA, 2018; pp. 319–330. [Google Scholar]
Figure 1. (a) Schematic of the normal autophagic, mitochondrial, and lysosomal pathways. The tight relationship between these pathways that become altered in PD enables the recycling of deficient organelles and degradation of outer material. Black solid arrows show the normal biological pathway. (b) Schematic of PD disrupted pathways. Dotted arrows indicate an up-regulation of the biological pathway specified due to variants in the genes involved. Meanwhile, inhibiting dotted lines indicate the down-regulation/inhibition of the biological pathway specified due to gene variants. Colored solid arrows show biological pathways that become disrupted in PD. Accumulation of damaged mitochondria and Lewy bodies due to dysfunctional lysosome degradation generates the typical PD physiopathology. Designed with biorender.com.
Figure 1. (a) Schematic of the normal autophagic, mitochondrial, and lysosomal pathways. The tight relationship between these pathways that become altered in PD enables the recycling of deficient organelles and degradation of outer material. Black solid arrows show the normal biological pathway. (b) Schematic of PD disrupted pathways. Dotted arrows indicate an up-regulation of the biological pathway specified due to variants in the genes involved. Meanwhile, inhibiting dotted lines indicate the down-regulation/inhibition of the biological pathway specified due to gene variants. Colored solid arrows show biological pathways that become disrupted in PD. Accumulation of damaged mitochondria and Lewy bodies due to dysfunctional lysosome degradation generates the typical PD physiopathology. Designed with biorender.com.
Ijms 22 09241 g001
Figure 2. (a) Schematic of cell internalization pathways. Internalization can occur by three main mechanisms: clathrin-mediated, caveolin-mediated, and clathrin-caveolin independent internalization. The schematic shows two alternatives of CRISPR delivery: one using a viral vector for plasmid delivery (left) and liposomes for delivering RNP (right). The processes of maturation and endosome escape are also illustrated. Specific routes for endosome escape can vary for each delivery vehicle. (b) Schematic showing the nanostructured carriers currently used for delivery in CRISPR applications. Overall, nanostructured carriers can be divided into three subgroups: inorganic, organic, and polymeric-based vehicles. Similarly, nanohybrids are shown at the intersection of these families of nanostructured carriers. Nanohybrids of inorganic-inorganic materials were not included in this schematic. Alternative surface modifications can be incorporated to enhance the properties of each nanostructured vehicle shown, incorporating translocating proteins and pH-sensitive moieties. Depending on their unique characteristics, each delivery vehicle can also deliver either RNP or plasmid encoding for the CRISPR elements. Designed with biorender.com.
Figure 2. (a) Schematic of cell internalization pathways. Internalization can occur by three main mechanisms: clathrin-mediated, caveolin-mediated, and clathrin-caveolin independent internalization. The schematic shows two alternatives of CRISPR delivery: one using a viral vector for plasmid delivery (left) and liposomes for delivering RNP (right). The processes of maturation and endosome escape are also illustrated. Specific routes for endosome escape can vary for each delivery vehicle. (b) Schematic showing the nanostructured carriers currently used for delivery in CRISPR applications. Overall, nanostructured carriers can be divided into three subgroups: inorganic, organic, and polymeric-based vehicles. Similarly, nanohybrids are shown at the intersection of these families of nanostructured carriers. Nanohybrids of inorganic-inorganic materials were not included in this schematic. Alternative surface modifications can be incorporated to enhance the properties of each nanostructured vehicle shown, incorporating translocating proteins and pH-sensitive moieties. Depending on their unique characteristics, each delivery vehicle can also deliver either RNP or plasmid encoding for the CRISPR elements. Designed with biorender.com.
Ijms 22 09241 g002
Figure 3. Summary of Parkinson’s gene therapy strategies and targets. The first approach (AP1) focuses on stimulating dopamine synthesis by targeting AADC and TH in dopaminergic neurons of the Substantia nigra. The second approach (AP2) focuses on GLF ligands. A third approach (AP3) focuses on the stimulation of mitophagy. Finally, the fourth approach (AP4) aims to reduce the presence of α-synuclein that ultimately generates Lewy bodies. Designed with biorender.com.
Figure 3. Summary of Parkinson’s gene therapy strategies and targets. The first approach (AP1) focuses on stimulating dopamine synthesis by targeting AADC and TH in dopaminergic neurons of the Substantia nigra. The second approach (AP2) focuses on GLF ligands. A third approach (AP3) focuses on the stimulation of mitophagy. Finally, the fourth approach (AP4) aims to reduce the presence of α-synuclein that ultimately generates Lewy bodies. Designed with biorender.com.
Ijms 22 09241 g003
Table 1. Familial Parkinson’s genes identified until 2020. Details are presented on the proteins encoded by these genes, their function, the pathway they are involved in, and what type of Parkinson’s they are associated with.
Table 1. Familial Parkinson’s genes identified until 2020. Details are presented on the proteins encoded by these genes, their function, the pathway they are involved in, and what type of Parkinson’s they are associated with.
Gene Alternative Gene NamesGene LocusProteinProtein Function and Cell Pathway GovernedOnset of Familial PD
SNCAPARK 1 or PARK44q22.1α-synucleinSynaptic vesicles traffickingEarly
UnknownPARK32p13UnknownUnknownLate
UCHL1PARK54p13Ubiquitin C-terminal hydrolase L1Proteasome systemLate
LRRK2PARK812q12Leucine-rich repeat kinase 2Autophagy processingLate
HTRA2PARK132p13.1HtrA serine peptidase 2Mitophagy developmentUnknown
VPS35PARK1716q12Vacuolar protein sorting 35Endosome regulationLate
EIF4G1PARK183q27.1Eukaryotic translation initiation factor 4 gamma 1Protein translationLate
DNAJC13PARK213q22.1DnaJ heat shock protein family (Hsp40) member C13Endosome regulationLate
CHCHD2PARK227p11.2Coiled-coil-helix-coiled-coil-helix domain containing 2Mitochondria-mediated apoptosis and metabolismLate/Early
PRKNPARK26q26ParkinMitophagy developmentEarly
PINK1PARK61p36.12PTEN-induced putative kinase 1Mitophagy developmentEarly
DJ-1PARK71p36.23DJ-1Mitophagy developmentEarly
ATP13A2PARK91p36.13ATPase cation transporting 13A2Lysosomal functionEarly
GIGYF2PARK112q36-7GRB10 interacting GYF protein 2Insulin-like growth factors (IGFs) signalingEarly
PLA2G6PARK1422q13.1Phospholipase A2 group VILipids metabolismEarly
FBXO7PARK1522q12.3F-box protein 7Mitophagy developmentEarly
DNAJC6PARK191p31.3DnaJ heat shock protein family (Hsp40) member C6Endosome regulationEarly
SYNJ1PARK2021q22.11Synaptojanin 1Endosome regulationEarly
VPS13CPARK2315q22.2Vacuolar protein sorting 13 homolog CMitophagy developmentEarly
UnknownPARK101p32UnknownUnknownUnknown
UnknownPARK12Xq21–q22UnknownUnknownUnknown
UnknownPARK161q32UnknownUnknownUnknown
GBA-1q22Glucosylceramidase betaLysosomal functionLate
LRP10--LDLR-related protein 10Retinoid metabolism and transportLate
Table 2. Summary of main research studies on neurotrophic factors and increased dopamine expression in the context of PD. DP: brain dopamine. AVV: adeno-associated virus, TH: tyrosine hydroxylase, AADC: aromatic L-amino acid decarboxylase.
Table 2. Summary of main research studies on neurotrophic factors and increased dopamine expression in the context of PD. DP: brain dopamine. AVV: adeno-associated virus, TH: tyrosine hydroxylase, AADC: aromatic L-amino acid decarboxylase.
ApproachVectorPhase of Clinical StudyReference
DP activity (TH)AAV-THAnimal Model: murine[270,271,272,273]
DP activity (AADC)AAV-AADCAnimal Model: murine (Phase I)[274,275,276,277]
Neurotrophic genes (GDNF)AAV-GDNFAnimal Model: murine and primate (Phase I and II)[278,279]
Neurotrophic genes (GNDF)Hematopoietic stem cell macrophagesAnimal Model: murine[280,281]
Neurotrophic genes (GNDF)Encapsulated GDNF-secreting cellsAnimal Model: murine[282]
Neurotrophic genes (GNDF)Cationic microbubblesAnimal Model: murine[283]
Neurotrophic genes (GNDF)Brain penetrating nanoparticlesAnimal Model: murine[284]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Arango, D.; Bittar, A.; Esmeral, N.P.; Ocasión, C.; Muñoz-Camargo, C.; Cruz, J.C.; Reyes, L.H.; Bloch, N.I. Understanding the Potential of Genome Editing in Parkinson’s Disease. Int. J. Mol. Sci. 2021, 22, 9241. https://doi.org/10.3390/ijms22179241

AMA Style

Arango D, Bittar A, Esmeral NP, Ocasión C, Muñoz-Camargo C, Cruz JC, Reyes LH, Bloch NI. Understanding the Potential of Genome Editing in Parkinson’s Disease. International Journal of Molecular Sciences. 2021; 22(17):9241. https://doi.org/10.3390/ijms22179241

Chicago/Turabian Style

Arango, David, Amaury Bittar, Natalia P. Esmeral, Camila Ocasión, Carolina Muñoz-Camargo, Juan C. Cruz, Luis H. Reyes, and Natasha I. Bloch. 2021. "Understanding the Potential of Genome Editing in Parkinson’s Disease" International Journal of Molecular Sciences 22, no. 17: 9241. https://doi.org/10.3390/ijms22179241

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop