Next Article in Journal
Ovule Transcriptome Analysis Discloses Deregulation of Genes and Pathways in Sexual and Apomictic Limonium Species (Plumbaginaceae)
Next Article in Special Issue
Targeted Modulation of Chicken Genes In Vitro Using CRISPRa and CRISPRi Toolkit
Previous Article in Journal
Outlining the Clinical Profile of TCIRG1 14 Variants including 5 Novels with Overview of ARO Phenotype and Ethnic Impact in 20 Egyptian Families
Previous Article in Special Issue
Current Approaches to and the Application of Intracytoplasmic Sperm Injection (ICSI) for Avian Genome Editing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Strategies for the Generation of Gene Modified Avian Models: Advancement in Avian Germline Transmission, Genome Editing, and Applications

1
Avinnogen Co., Ltd., Seoul 08826, Republic of Korea
2
Department of Agricultural Biotechnology, Research Institute of Agriculture and Life Sciences, College of Agriculture and Life Sciences, Seoul National University, Seoul 08826, Republic of Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Genes 2023, 14(4), 899; https://doi.org/10.3390/genes14040899
Submission received: 24 February 2023 / Revised: 2 April 2023 / Accepted: 10 April 2023 / Published: 12 April 2023
(This article belongs to the Special Issue Avian Genomics and Transgenesis)

Abstract

:
Avian models are valuable for studies of development and reproduction and have important implications for food production. Rapid advances in genome-editing technologies have enabled the establishment of avian species as unique agricultural, industrial, disease-resistant, and pharmaceutical models. The direct introduction of genome-editing tools, such as the clustered regularly interspaced short palindromic repeats (CRISPR) system, into early embryos has been achieved in various animal taxa. However, in birds, the introduction of the CRISPR system into primordial germ cells (PGCs), a germline-competent stem cell, is considered a much more reliable approach for the development of genome-edited models. After genome editing, PGCs are transplanted into the embryo to establish germline chimera, which are crossed to produce genome-edited birds. In addition, various methods, including delivery by liposomal and viral vectors, have been employed for gene editing in vivo. Genome-edited birds have wide applications in bio-pharmaceutical production and as models for disease resistance and biological research. In conclusion, the application of the CRISPR system to avian PGCs is an efficient approach for the production of genome-edited birds and transgenic avian models.

1. Introduction

Although the chicken was the first organism to have a sequenced genome after the Human Genome Project [1], it is not a well-established model for genome editing. Avian model species, such as chickens, have unique developmental characteristics, in which the fertilized embryo grows in the eggshell until hatching, thus enabling easy access and manipulation of specific embryonic stages [2]. The avian-specific ex vivo research system can be used to directly monitor cell differentiation, transformation, and organogenesis, and contributes to basic and clinical research. Recent advances in genome modification technologies, such as the clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated (Cas) system and transgenic systems, have substantially expanded the applications of genome-edited avian models in various fields, including agriculture, healthcare, and disease control [3,4,5].
The chicken is a particularly advantageous model for recombinant protein production because chicken eggs contain about 3.5 g of egg white protein with similar glycosylation residues to those of human and target proteins are easily purified compared with other animal bioreactor systems [6,7]. Theoretically, if the protein composition of proteins is altered by genome-editing technologies, chickens could be a promising animal bioreactor system. Moreover, the risk of exposure to exogenous contaminants and the cost of feeding a hen are relatively low, enabling large-scale recombinant protein production. Therefore, in the poultry industry, there is substantial interest in the use of genome-editing technologies to efficiently produce recombinant proteins with applications in various fields, such as medicine. Furthermore, there is a focus on the development of disease-resistant lines to minimize economic losses in the poultry industry [8]. There are about 23 billion chickens, which is about 5 times more than the estimate 50 years ago, reflecting the dramatic increase in the poultry industry and market over the past few decades [9]. This trend is expected to continue, with corresponding increases in demand for eggs and meat. In this regard, disease control in the poultry industry is needed. In particular, avian influenza viruses have severely threatened poultry farms and caused enormous economic losses [10]. Therefore, the demand for genome-edited avian models with resistance to various poultry diseases is increasing.
Germline genome modification is essential to produce genome-edited avian model. In general, modification of totipotent embryo stem cells followed by transplantation into recipient embryos can produce transgenic mammals [11]. In addition, somatic cell nuclear transfer (SCNT) was adopted for transgenesis in mammalian species; however, physiological differences (discussed below) limit the application of these approaches to avian species. Although long-term inducible gene silencing by small hairpin RNA (shRNA) was demonstrated in chicken embryos [12], avian genome modification substantially expanded the scope of research on gene function and enabled the development of genome-edited avian models with multiple applications. In this regard, studies of germline-competent stem cells of avian species have focused on model development by genetic modification and genome editing [13]. This review provides an overview of genome editing in avian taxa and discusses the use of germline-competent cells to develop genome-edited avian models.

2. Historical Overview of Genome Modification Strategies for Avian Model Development

2.1. Overview of Avian Transgenesis

In initial techniques for mouse transgenesis, foreign genes from viral DNA were successfully introduced in the host genome by microinjection into the pronuclei of fertilized oocytes [14]. Putative genome-modified embryos were implanted in the oviducts of surrogate females, and the resulting progeny contained the foreign gene [14]. This approach was subsequently attempted in rabbit, livestock (sheep and pigs) [15], sea urchins [16], frogs [17], and flies [18,19]. This approach has been the most common transgenic and genome-edited animal production method for decades [20]. Although introducing foreign genes into fertilized eggs from various species is a well-established routine method for transgenesis, its application to birds has been limited. Unlike mammalian eggs, it is hard to obtain single chicken embryos and to manipulate the fertilized chicken embryos located within the eggshell and vitelline membrane, which is not conducive to microscopy. It is also difficult to introduce foreign genes, since the pronuclei of chicken eggs are surrounded by the opaque yolk-filled cytoplasm, and numerous sperm nuclei exist on the membrane [21,22,23]. Moreover, the Eyal–Giladi and Kochav (EGK) stage X blastoderms [24,25,26] of freshly laid eggs already consist of numerous cells (approximately 40,000–60,000 cells), and there are around 50–100 primordial germ cells (PGCs) until oviposition (EGK V-EGK X) [27,28]. This indicates that embryonic development progressed considerably before oviposition, distinct from fertilized eggs of mammals. Despite the production of transgenic birds by the microinjection of linearized DNA constructs in chicken zygotes, the practical application or feasibility of this method is limited by the complexity and low production rate [29].

2.2. Virus-Mediated Gene Insertion for Avian Transgenesis

An alternative to conventional foreign DNA transfer systems for bird transgenesis, virus-mediated gene transfer, has been established. The first transgenic chicken was established by transduction of EGK stage X blastoderms with a replication-competent retroviral system based on the avian leucosis virus (ALV) [30,31]. However, the germline transmission efficiency was too low to merit its application to further transgenic line development. In 1989, Bosselman et al. developed a replication-defective reticuloendotheliosis virus (REV) vector-mediated germline transgenic chicken system, including genes encoding thymidine kinase and neomycin phosphotransferase [32]. The replication-defective system is more widely and commercially available than the previous replication-competent system [30]; however, the germline transmission efficiency is still very low (approximately 8%) [32]. A replication-deficient retrovirus based on ALV was further applied to chickens to develop a transgenic bioreactor system. Harvey and colleagues have demonstrated β-lactamase production in chicken egg whites and serum using the ubiquitous cytomegalovirus (CMV) promoter [33]. Around this time, more effective viral vector systems were developed for transgenic chicken production. The lentiviral system uses a genus of retroviruses that can infect dividing cells, thus enabling transgene introduction into somatic and germ cells with a higher efficiency compared with that of the previous retroviral system [34]. The germline transmission efficiency varies from 4% to 45%; however, the frequency is comparatively higher than those of other retroviral systems. Recently, adenovirus-mediated CRISPR/Cas9 genome editing in chicken blastoderm PGCs produced genome-edited offspring with 11% germline transmission efficiency [35].
Similarly, virus-mediated gene transfer has been established in other avian species. For example, transgenic quails have been developed using a retroviral vector based on Moloney murine leukemia virus (MoMLV) pseudotyped with vesicular stomatitis virus, with high germline transmission efficiencies (around 80%) [36]. A lentivirus system has also been adopted to generate transgenic quails with Rous sarcoma virus (RSV) promoter-driven EGFP. Direct transgene introduction into gonadal PGCs achieves germline transmission, although the transmission rate is low (1.6–1.9%) [37]. The lentivirus system has also been introduced to generate transgenic zebra finches (Taeniopygia guttata), a representative animal model of vocal learning. After the introduction of a lentivirus containing a human ubiquitin-C promoter-driven GFP transgene into very early stage embryos, 13% (3/23) of founders were germline transgenics [38]. Similarly, a transgenic zebra finch model in which cAMP response element binding protein (CREB) is regulated or expressed in the human mutant huntingtin (mHTT) gene was also developed using the lentiviral system to monitor vocal learning behavior and to model Huntington’s disease [39,40]. Recently, embryonic zebra finch PGCs were successfully expanded in vitro for 15 days and the GFP gene was inserted by lentiviral transfer. The genome-modified cultured PGCs were used to generate transgenic zebra finch [41]. In addition, an adenovirus containing CRISPR/Cas9 targeting for melanophilin (MLPH) gene was injected into the quail blastoderm to produce transgenic quail with 45% germline chimera production efficiency and 2.4% to 10% germline transmission efficiency [42]. Similarly, adenovirus containing CRISPR/Cas9 was injected into duck blastoderms, producing genome-edited progenies with 2% (duck) germline efficiency [35]. Despite these successful cases, viral systems are difficult to regard as an effective method owing to the variation in transgenic production efficiency and the limited precision of gene editing.

2.3. Development of an Avian PGC Culture System and Transgenic Models

In most animal species, including the fly, fish, mouse, and chicken, germ cells are specified in the early embryo and migrate to the genital ridge [43,44,45,46,47]. However, the detailed processes, including the origin and migratory route of cells, vary between taxa. In the mouse system, PGCs, the precursor cells of gametes, are first detectable as cell clusters in the proximal epiblast region, then migrate to the extra-embryonic ectoderm (ExE) [48,49], followed by movement through the developing endodermal hindgut into the genital ridges [43]. In the avian system, different from mammals, PGCs are initially detected in a scattered pattern in the area pellucida, the central region of the blastoderm at EGK stage X of freshly laid fertilized eggs [24,26,50]. PGCs localized in EGK stage X move to the germinal crescent at Hamburger and Hamilton (HH) stage 4 [51,52,53]. PGCs enter embryonic blood vessels and circulate through the bloodstream between HH stages 9 and 12 [54,55] and finally settle in the genital ridge [55,56]. The distinct patterns of avian PGC development and migration [57] allow the easy isolation of PGCs at several developmental stages, including the germinal crescent of the embryo, blood vessels, and genital ridge [58]. In particular, avian PGCs from the chicken, pheasant, quail, turkey, duck, and guinea fowl have been isolated by density gradient centrifugation and size-dependent isolation methods, without requiring specific antibodies [59,60,61,62,63,64]. Antibody-mediated methods, such as fluorescence-activated cell sorting and magnetic-activated cell sorting, have also been developed to isolate avian PGCs [63,65,66,67,68].
PGCs derived from embryos have been directly introduced into recipient embryos to restore wild or endangered birds [69,70,71,72,73,74]; however, these previous studies are largely dependent on the chicken embryo as a recipient due to the requirement for the long-term cultivation of PGCs. Among vertebrates, the chicken is the only species in which a long-term PGC culture system has been successfully established and used to produce transgenic animals [75]. For the proliferation and survival of chicken PGCs, basic fibroblast growth factor (bFGF) is required [76,77] and stem cell factor (SCF) or leukemia inhibitory factor (LIF) [75]. When bFGF is added to the culture medium, PGCs can be propagated by the activation of the mitogen-activated protein kinase (MEK)/extracellular signal-regulated kinase (ERK) signaling pathway, maintaining their key characteristics, such as migratory activity and germline transmission, even after long-term cultivation [76,77]. Subsequent studies have suggested that the MEK1, AKT (also known as protein kinase B, PKB), SMAD family member 3 (SMAD3), and Wnt/β-catenin signaling promote PGC proliferation [78,79]. The long-term cultivation of PGCs has also been demonstrated for other avian species. Transforming growth factor β (TGF-β) activated by activin or inducer of definitive endoderm 1 (IDE1) for quail PGC proliferation is effective for in vitro culture for 40–50 days [80]. Gonadal PGCs in the zebra finch have been cultured for 30 days in vitro [81]. Circulating PGCs and gonadal PGCs of the Muscovy duck (Cairina moschata), the Pekin duck (Anas platyrhynchos) and hybrid mule duck (C. moschata × A. platyrhynchos) have been cultured for days in vitro in chemically defined medium [64]. These findings provide a basis for the production of transgenic models by manipulating PGCs in several avian species, such as quail [37] and zebra finch [41]; however, reliable long-term PGC culture systems are lacking for species other than chickens.
Since the establishment of long-term cultivation methods, a PGC-mediated transgenic chicken production system has been introduced [75]. The transposon system is an efficient gene delivery system for chicken PGCs [82,83,84]. Incorporation of the EGFP gene containing the piggyBac transposon into the genome of chicken PGCs after long-term culture results in a high transmission rate in offspring (over 90%). This represents a stable transgenic chicken production strategy, different from previous systems [82]. Transposon systems are easy to produce and inexpensive to purify, with non-immunogenic characteristics compared with virus-mediated systems [85,86]; accordingly, they are considered far more practical and reliable for the PGC-mediated production of transgenic chickens. This technique has been adopted to produce recombinant bioactive protein, human epidermal growth factor (EGF) [87], and human anti-CD20 monoclonal antibody in transgenic egg whites as animal bioreactors [88].

2.4. Development of Efficient Germline Chimeras by Depleting Endogenous PGCs for Avian Transgenesis

Cultured PGCs could be gene-edited by a programmable gene editing system (discussed in Section 3) and microinjected into the blood vessels of recipient embryos. The recipient embryos contain both endogenous gametes and donor-derived gametes and are referred to as “germline chimeras”. In 1976, turkey PGCs were injected into the blood vessels of chicken embryo and functional gametes derived from turkey PGCs were produced [89]. Subsequently, quail PGCs were successfully transferred to recipient embryos to produce quail germline chimeras [90]. Using this strategy, the first transgenic bird was produced via PGCs isolated from the germinal crescent of HH stage 5 chicken embryos [91]. In addition, isolated PGCs obtained by density gradient centrifugation and magnetic-activated cell sorting were transferred into a recipient embryo for the efficient production of germline chimeras [66,92]. Germline chimeras were also produced using the cryopreserved PGCs [93,94]. In summary, avian germline chimeras can be produced by microinjecting PGCs isolated from the blood of HH stage 14–16 embryos and gonads of HH stage 26–28 embryos in chicken and quail.
However, a major challenge to increasing germline transmission efficiency is competing endogenous PGCs of recipient embryos. One strategy to enhance germline chimerism is transplanting genome-edited PGCs into the adult testis to obtain functional sperm cells [95]. Another solution is depleting endogenous PGCs of recipient embryos by exposure to γ rays [96] and the elimination of blood from HH stage 14–15 recipient embryos [93]. In 2010, Nakamura et al. showed that busulfan treatment in recipient embryos depleted endogenous PGCs and resulted in a germline chimera efficiency of approximately 99%, compared with an efficiency of only 6% in an untreated group [97]. Although the effect of busulfan treatment could depend on the administration route, time point, and dosage [98], the depletion of endogenous PGCs using busulfan can promote the establishment of efficient germline chimeras and genome-edited birds. Recently, Kim et al. developed an in vivo selection model to increase the efficiency of transgenic chicken production by introducing microsomal glutathione-S-transferase II (MGSTII) into the PGCs to confer resistance to busulfan. The MGSTII-expressing PGCs were dominantly localized in the recipient testes after busulfan treatment compared to non-treated group. The rate of donor PGC-derived progeny production was 94.68% and the rate of transgenic chicken production was 80.33–95.23%, compared with 51.18% in the group not treated with busulfan [99]. A similar system was applied to zebra finch; MGSTII-expressing PGCs were enriched and underwent spermatogenesis in the recipient zebra finch testis [100]. In addition, the chicken DEAD-Box Helicase 4 (DDX4) gene, essential for formation of germ cell lineage formation, was disrupted by transcription activator-like effector nucleases (TALENs), resulting in the loss of PGCs and infertile hens [101]. Furthermore, the inducible caspase-9 (iCaspase9), conditionally activated in response to the chemical compound AP20187 (B/B), was combined with Deleted In Azoospermia Like (DAZL) for germ cell-specific expression; injecting AP20187 (B/B) conditionally inhibited the growth of endogenous iCaspase9-expressing PGCs, providing an alternative strategy to deplete endogenous PGCs for efficient germline chimera production [102].

2.5. Transgenic Systems Using Other Germline-Competent Stem Cells in Avian Species

Pluripotent stem cells, such as embryonic stem cells (ESCs), have been used to produce transgenic mammals [103,104,105,106]. In this regard, the establishment of pluripotent stem cells from undifferentiated blastodermal cells at EGK stage X and of germline chimeras by injecting the stem cells into the subgerminal cavity of recipient embryos has been described in avian species [96,107,108,109]. Culture conditions for chicken ESCs include LIF, bFGF, SCF, and insulin growth factor 1 (IGF-1), which are similar to the chemokines required for mammalian pluripotent stem cells [109]. These cells contribute to all three germ layers and somatic tissues, whereas their germline contribution is lacking or very limited [109,110]. Although the ectopic overexpression of chicken vasa homolog (CVH) on chicken ESCs increases germline markers, the contribution to the germline is still unknown [111]. A method for the direct reprogramming of somatic cells into a pluripotent state, or induced pluripotent stem cells (iPSCs), using several transcription factors, has been introduced [112]. This system has been rapidly applied to several species, including humans, monkeys, and pigs [113,114,115]. In avian species, the induction of pluripotent stem cells from embryonic fibroblast cells and feather follicle cells have been demonstrated in the chicken, quail, and zebra finch [116,117,118]. Recently, chicken embryonic fibroblasts (CEFs) were reprogrammed into iPSCs and further induced into PGCs, producing somatic cell-derived offspring [119]. However, more reproducible results for iPSC-mediated viable germline-competent stem cell development (e.g., induced PGCs) are needed in the future.
Germline stem cells or spermatogonial stem cells (SSCs) are also frequently used for transgenic research [120,121,122]. SSCs are reliable germline-competent stem cells able to deliver genetic information to successive generations. In rats, SSC transfection via lentiviral vectors has been used to generate genome-modified progeny after xenografting into heterologous testes [120,121,122]. Recently, this strategy has been successfully applied to other species, such as the tree shrew [123]. In addition, using SSCs, transposable element-mediated gene knockout and genome editing by homologous recombination have been successfully demonstrated [124,125]. However, SSCs only account for approximately 0.03% of germs cells in the mammalian testis. Therefore, SSC enrichment is important for the production of transgenic animals by transgenic cell transplantation [126]. Owing to the scarce availability of SSCs in the testes, efficient methods for their isolation and subsequent enrichment are needed. SSCs have been successfully isolated and cultured from many species, including humans, mice, cattle, and pigs [127,128,129,130]. Chicken and quail SSCs have been isolated and maintained over short durations in vitro [131,132]. More recently, quail SSCs were efficiently enriched by density gradient centrifugation using Ficoll-Paque PLUS (Ficoll) and transplanted into the recipient. The rate of germline transmission by SSC transplantation in germline chimera was 0–13.2% [59]. However, studies of SSCs in poultry and their utilization are still limited. Therefore, SSC-mediated genome-edited avian production requires more research on cell proliferation and gene regulatory systems. Collectively, the production of avian germline chimeras and transgenic animals using pluripotent stem cells or germline stem cells is not yet a reliable strategy compared with the PGC-mediated system.

3. Progression of Precise Genome-Editing Systems for Avian Taxa

3.1. Programmable Genome Editing in the Avian System

Targeted gene modifications at specific loci have broad implications, including therapeutic applications, implications in the areas of immunity, disease control, neuroscience, developmental biology, and agriculture [133,134,135,136,137]. In mammalian species, genome-edited production systems for the loss- and gain-of-function of loci that direct gene transfer and genome-edited ESCs into fertile eggs have been used for a long time [103,104,105,106]. Unlike in mammalian species, the absence of a reliable transgenesis system has prevented specific gene-targeting in birds until the development of PGC-mediated transgenesis. The first targeted gene knockout in chickens was achieved by homologous recombination in the immunoglobulin gene [138]. The homozygous knockout of Ig heavy chain (IgH) in chickens resulted in a lack of an antibody response after immunization. Despite the successful knockout, the homologous recombination efficiency in the PGC genome was extremely low (approximately 0.1%) because the method depends only on the homology of the template gene, limiting its practical implementation [139]. Thus, developing an efficient programmable genome editing system is a prerequisite for the efficient production of genome-edited birds.
In the last few decades, improvements in our understanding of the nuclease-based bacterial immune system have enabled the development of programmable genome editing, including targeted gene deletions and insertions, and the modification of host genomes [140]. Zinc-finger nucleases (ZFNs) and TALENs are chimeric nucleases comprising site-specific DNA-binding modules with DNA nucleases. These nuclease systems induce genome modifications via double-strand DNA breaks, resulting in error-prone non-homologous end joining (NHEJ) or homology-directed repair (HDR) at the target genome sequence [141]. In 2014, the first TALEN-mediated targeted deletion in chickens, targeting ovalbumin (OV), was reported [142]. The genome deletion efficiency for PGCs was up to 33.3%, which is a dramatic increase over that of homologous recombination [138]. The germline efficiency ranged from 22.3% to 53.2%, and the rate of OV mutant offspring was around 8% [142]. Since the development of TALEN-mediated genome editing in chickens, the TALEN system has been used for targeted GFP gene insertion with HDR repair in the chicken DDX4 gene, a representative germ cell marker [101]. The development of these programmable gene-editing technologies has made it possible to produce sustainable and reliable genome-edited chicken models.

3.2. CRISPR-Mediated Genome Editing in the Avian System

The recent development of the CRISPR/Cas9 system for gene editing is a major advancement; the system consists of two sets of RNA, 20 bp CRISPR RNA (crRNA) and universal trans-activating crRNA (tracrRNA), and a nuclease from Streptococcus pyogenes type II, Cas9 protein (Cas9), which induces the cleavage of target loci [143]. This CRISPR/Cas9 system can also edit targeted sequences via HDR or NHEJ, similar to other programmable gene-targeting systems, ZFN and TALEN [144]. Since its development, the CRISPR/Cas9 system has been applied to a wide range of taxa and has quickly become accepted as the most advanced and simple gene-editing system to date [145,146,147,148,149].
The PGC-mediated germline transmission system led to the development of genome-edited chickens via the CRISPR/Cas9 system, similar to the previously described method for transgenic bird production and the CRISPR/Cas9-mediated genome-editing system has been effectively applied to avian species for broad applications. [150,151]. Oishi et al. specifically knocked out chicken OV and ovomucoid (OVM), which are major egg allergens, in the genome of PGCs using the CRISPR/Cas9 system. CRISPR/Cas9-mediated targeted gene insertion in the chicken IgH variable region (V region) was reported at about the same time [150]. In particular, double-strand breaks were induced at the target site using the CRISPR/Cas9 system, followed by target gene insertion by homologous recombination. Surprisingly, the recombination efficiency was dramatically improved (to 9%) [151] compared with the efficiency of conventional homologous recombination methods for chicken PGCs [138]. In addition, genome-edited chicken with an increased muscle mass was generated by targeting the myostatin (MSTN) gene, which controls tissue growth and development, and the G0/G1 switch gene (G0S2), which regulates fat deposition [152,153]. In a similar study, recombinant adenovirus containing CRISPR/Cas9 was injected into the quail blastoderm, resulting in significant increases in body weight and muscle mass in homozygous mutants with the deletion of cysteine 42 in the MSTN gene [154]. Similarly, the targeted deletion of quail melanophilin (MLPH) was demonstrated.
CRISPR/Cas9-mediated in vivo genome editing in avian species has also been introduced. Cas9-expressing chickens showed successful in vivo gene disruption in lymphocytes and embryonic brains [155]. In addition, Challagulla et al. produced interferon α and β receptor subunit (IFAR1) knockout chicken by directly injecting transposon vector encoding guide RNAs (gRNAs) targeting chicken IFAR1 and high-fidelity Cas9 [156]. These results indicate that in vivo genome editing could be used to produce genome edited birds which lack the long term PGC culture method. Collectively, the CRISPR/Cas9-mediated genome editing system considerably increased the efficiency of gene-edited birds’ production; additional applications will be discussed in Section 4.
Although genome editing in animals represents a major advance, its low efficiency (<10%) needs to be resolved before it is considered a practical method [42]. Strategies aimed at increasing gene editing efficiency and accuracy are needed. Unlike the CRISPR/Cas9 system, the base editing (BE) system uses catalytically deactivated Cas9 (dCas9) or nickase Cas9 (nCas9) and cytidine deaminase, which can induce a C-to-T (or G-to-A) substitution, and thereby edit specific nucleotide sequences without double-strand breaks in DNA [157]. The BE system was successfully applied in chicken to induce mutations in MSTN and ovotransferrin (TF) and 35.7% and 55.5% of genome-edited progenies harbored the desired base substitution in TF and MSTN, respectively [158]. More recently, a prime editing guide RNA (pegRNA) was developed and Cas9 was fused with a reverse transcriptase (RT) domain to form a programmable nuclease, termed prime editing, able to remove, replace, and insert target sequences in the genome [159]. Prime editing was recently applied to the PGC genome. PEmax and ZsGreen1 were integrated into the PGCs genome by the transposon system and cells were transfected with several candidate pegRNAs inducing a stop codon in DDX4. Approximately 8.3% of prime edited PGCs showed the desired substitution in DDX4 [160]. Based on these studies, advanced genome editing systems are expected to be useful for the production of genome-edited birds in the near future.

4. Potential Applications of Genome Editing for Avian Model Development

The advancement of avian germline transmission and genome editing technology enabled researchers to develop various genome edited avian models, including a disease resistant model, efficient bioreactor, and academic model for scientific use (Figure 1). In the following sections, we will discuss various examples of genome modified avian models.

4.1. Genome Editing for the Development of Disease-Resistant Avian Models

Disease control and prevention in birds is an essential prerequisite for the sustainable poultry industry. In this regard, genome-editing technologies have been used to control avian diseases, such as avian influenza, Marek’s disease, and avian leukosis [161,162,163]. In some cases, the effective prevention of avian viral diseases has been achieved by the regulation of virus-specific receptors. For example, a subgroup of ALV can lead to cancer in chickens, and this could be prevented by the regulation of host specific ALV receptors [164]. The precise genome editing of chicken NHE1, TVA, TVB, and TVC (specific receptors of ALV subgroups ALV-J, ALV-A, ALV-B, and ALV-C) has been achieved in the chicken DF1 cell line [165,166,167,168]. Such gene editing effectively reduces viral infection, leading to the development of genome-edited chickens. Beyond the cell level, tryptophan 38 (W38) of NHE1, a critical residue for ALV-J entry, was precisely deleted in a commercial chicken line to produce ALV-J-resistant chicken [169].
In relation to avian influenza virus (AIV), resulting in high mortality rates in birds, the host factor ANP32A, which supports the vPol activity of influenza A virus (IAV) in a species-specific manner, is critical [170,171]. The 99 nucleotides of chicken ANP32A encoding the additional 33 amino acids in birds have been deleted in chicken DF1 cells and viral polymerase (vPol) activity was significantly reduced [162,170,171]. Recently, the critical residues (aspartate 149 and 152) for interactions with viral protein in the additional 33 amino acids of chicken ANP32A were revealed [172]. In addition, chickens lost Retinoic inducible gene 1 (RIG-I), a major IAV RNA sensor in mammals; they only harbor Melanoma differentiation associated protein 5 (MDA5), a member of RIG-I like receptor (RLR) family [173]. To gain resistance against IAV, duck RIG-I was introduced into the chicken DF-1 cells [173,174]. Furthermore, the chicken MDA5 C terminal domain (CTD) was replaced with that of RIG-I, resulting in greater inhibition of viral proliferation than that of wild-type chicken MDA5 [175]. Genome editing at these loci could be an effective strategy for the control of avian influenza in a host-specific manner.
Another model of avian disease control has also been proposed for Marek’s disease viruses (MDVs), a lymphotropic α-herpesvirus associated with T-cell lymphoma that induces asymmetric paralysis of the limbs, depression, and death. The targeted disruption of genes essential for MDV replication suggests that effective disease control is possible [163]. Recently, transgenic chickens expressing both Cas9 and gRNA specific to the immediate early infected-cell polypeptide-4 (ICP4) of MDV were produced. The chicken embryonic fibroblasts from transgenic chickens inhibited MDV infection with no effect on herpesvirus of turkeys (HVT) infection [176]. Based on these results, an in-depth understanding of the infection route, replication pathways, and host–virus interactions will provide a basis for the development of an effective genome editing-based disease control model in avian species.

4.2. Practical Bioreactor System for Recombinant Protein Production in Avian Systems

Transgenic avian bioreactors have great potential for recombinant protein production, including proteins with pharmaceutical and industrial applications in eggs [177,178,179]. Several transgenic avian bioreactors have been reported [88,180,181,182]. However, the conventional system is limited with respect to the quantity of recombinant protein produced. Still, the CRISPR/Cas9 system has made it possible to develop an effective chicken bioreactor system for large-scale production in eggs. Oishi et al. inserted human interferon-β (IFN-β) into the chicken ovalbumin gene via the CRISPR/Cas9 system to produce high levels of recombinant protein (3.5 mg/mL) in chicken egg whites [183]. Human interferon α 2a and porcine colony stimulating factor 1 (CSF1) fused with the Fc region was produced in transgenic chicken eggs and recombinant proteins produced from the chicken bioreactor were easily purified and showed comparable biological functions to those of recombinant proteins produced by other systems [6]. Kim et al. produced anti-cancer monoclonal antibodies against the CD20 with greater Fc effector function in chicken egg whites compared to a commercial counterpart [88]. In addition, gene encoding adiponectin (ADPN), a hormone derived from adipose tissue that can be used to treat insulin resistance, was precisely integrated into the Ovalbumin (OVA) by CRISPR/Cas9 system. The gene-edited chicken expressed high amount of high-molecular-weight (HMW) ADPN, considered to be a more active form [184]. Recently, the GFP gene was inserted into chicken ovalbumin (OVA) gene and system for evaluating protein production in a chicken bioreactor using young chicks was established. This system measured GFP expression in the oviduct of 3-week-old chicks after treatment with an estrogen agonist, diethylstilbestrol (DES) [185]. These results provide a basis for the development of an ideal animal bioreactor that overcomes issues related to yield.
Recombinant proteins from the chicken oviduct derived from egg bioreactor systems with unique post-translational modifications related to N-glycan species terminated with high mannose with a core afucosylated form have been reported [180]. Based on these characteristics, an enzyme produced by transgenic chickens was developed for enzyme replacement therapy for Gaucher disease [186]. As enzymes for the treatment of lysosomal storage diseases, including Gaucher disease, Pompe disease, and Fabry disease, are taken up by mannose receptors, the terminal mannosylation of these recombinant enzymes is critical for efficacy [187,188,189,190]. In this respect, recombinant proteins derived from chicken fallopian tubes are suitable bioreactors to produce these enzymes. Alternatively, the afucosylation of recombinant proteins produced in transgenic chickens can be an effective system for the production of anti-cancer antibodies. N-glycan afucosylation of the Fc domain affects the antibody-dependent cellular cytotoxicity of therapeutic antibodies [191]. Accordingly, levels of antibody-dependent cellular cytotoxicity of antibodies derived from transgenic chickens are significantly higher than those of control chickens [88,180]. Overall, an effective genome-edited avian bioreactor model obtained by CRISPR/Cas9-mediated target gene insertion could be an innovative approach to recombinant protein production for various purposes.

4.3. Genome-Edited Birds as Scientific Models

The modification of the avian genome provided a basis for the identification of specific gene functions and the development of genome-edited birds as scientific models. For example, recombination activating gene 1 (RAG1) gene was precisely disrupted by CRISPR/Cas9 to obtain chickens lacking mature B and T cells. These chickens could be used to study various lymphocytes in the absence of B and T cells and to study a wide range of diseases, such as cancer and viral infection [192]. In particular, chickens can spontaneously develop ovarian cancer [193]. Thus, RAG1-deficient chickens could be an effective model for studying ovarian cancer. In addition, the GFP gene was precisely inserted into chicken DAZL, a germ cell-specific marker in chicken, to trace germ cells from E2.5 to 1-week post-hatching. Using this model, sex-specific developmental stages and trajectories of chicken germ cells were identified and evolutionary conserved or species-specific genes involved in germ cell development were analyzed [194]. Furthermore, the PR domain zinc finger protein 14 (PRDM14) gene, a critical factor for PGC development in mice, was disrupted in PGCs by inserting eGFP gene via the CRISPR/Cas9 system to evaluate the important roles of PRDM14 in early chicken development [195]. Similarly, double sex and mab-3-related transcription factor 1 (DMRT1) was precisely deleted in chicken, revealing that this gene is one of the most important factors for testis development, while other factors, including sex hormones and DMRT1 gene networks, are key factors for sex determination [196,197].
The zebra finch is a promising bird for neurobiological studies. The Forkhead box protein 2 (FOXP2) mutations in humans lead to developmental verbal dyspraxia (DVD) and lentivirus mediated FOXP2 gene knockdown in zebra finch results in abnormal speech production [198,199]. Transgenic zebra finches carrying the GFP gene under the control of the human ubiquitin-C promoter were generated and GFP-expressing cells located in the forebrain could be traced and analyzed [38,200]. Gonadal PGCs of zebra finch are heterogenous, and the signaling pathways contributing to their development differ from those of chickens [201,202]. In addition, a retrovirus-mediated immortalized zebra finch fibroblast cell line was established and the Sox9 gene was knocked out using CRISPR/Cas9 [203]. Collectively, the generation of transgenic zebra finch models would be facilitated by PGC studies, and the utilization of immortalized cell line; such models are expected to contribute substantially to our understanding the mechanism underlying vocal learning in the near future.

5. Conclusions

In avian species, gene transfer to early embryos is a potential strategy for genome editing (Figure 2). To date, PGCs are the only avian germline-competent cells that can be cultured in vitro over long time periods and are a reliable means of genome modification. Using viral systems and transplantation methods for other types of cells (e.g., ESCs, SSCs, and iPSCs), the stability and efficiency of germline genome editing are major limitations. For genome editing in a wider variety of bird species, more in-depth studies of these systems are needed, with a focus on efficiency and sustainability. Recently developed genome-editing systems, such as TALEN and CRISPR/Cas, have made it possible to produce genome-edited avian models for various purposes [101,142,150,151]. In particular, the development of avian genetic resources for the control of a wide range of avian diseases, such as avian influenza, should be a focus of future research. The application of genome-editing techniques combined with germline-competent cell-based strategies to a variety of valuable avian species has extensive research applications, and practical applications in the poultry industry. These genome-edited birds will contribute to sustainable agriculture in an eco-friendly manner, providing a basis for reducing the massive culling of virus-infected flocks, improving animal welfare, and increasing production efficiency [204].

Author Contributions

Y.-M.K. and S.-J.W. participated in the design of the manuscript, reviewed the literature, interpreted the data, and wrote the draft of the manuscript. Y.-M.K., S.-J.W. and J.-Y.H. participated in the conceptualization and revision of the manuscript. J.-Y.H. participated in the writing of the final version of the manuscript and overall coordination. All authors reviewed the manuscript and approved the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIP) [NRF-2015R1A3A2033826], and Cooperative Research Program for Agriculture Science and Technology Development (RS-2023-00259807) from the Korean Rural Development Administration.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. International Chicken Genome Sequencing Consortium. Sequencing and comparative analysis of the chicken genome provide unique perspectives on vertebrate evolution. Nature 2004, 432, 695–716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Stern, C.D. The chick; A great model system becomes even greater. Dev. Cell 2005, 8, 9–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Tizard, M.L.; Jenkins, K.A.; Cooper, C.A.; Woodcock, M.E.; Challagulla, A.; Doran, T.J. Potential benefits of gene editing for the future of poultry farming. Transgenic Res. 2019, 28, 87–92. [Google Scholar] [CrossRef] [PubMed]
  4. Looi, F.Y.; Baker, M.L.; Townson, T.; Richard, M.; Novak, B.; Doran, T.J.; Short, K.R. Creating Disease Resistant Chickens: A Viable Solution to Avian Influenza? Viruses 2018, 10, 561. [Google Scholar] [CrossRef] [Green Version]
  5. Han, J.Y.; Park, Y.H. Primordial germ cell-mediated transgenesis and genome editing in birds. J. Anim. Sci. Biotechnol. 2018, 9, 19. [Google Scholar] [CrossRef] [Green Version]
  6. Herron, L.R.; Pridans, C.; Turnbull, M.L.; Smith, N.; Lillico, S.; Sherman, A.; Gilhooley, H.J.; Wear, M.; Kurian, D.; Papadakos, G.; et al. A chicken bioreactor for efficient production of functional cytokines. BMC Biotechnol. 2018, 18, 82. [Google Scholar] [CrossRef]
  7. Woodfint, R.M.; Hamlin, E.; Lee, K. Avian Bioreactor Systems: A Review. Mol. Biotechnol. 2018, 60, 975–983. [Google Scholar] [CrossRef]
  8. Mozdziak, P.E.; Petitte, J.N. Status of transgenic chicken models for developmental biology. Dev. Dyn. 2004, 229, 414–421. [Google Scholar] [CrossRef]
  9. Mottet, A.; Tempio, G. Global poultry production: Current state and future outlook and challenges. World’s Poult. Sci. J. 2017, 73, 245–256. [Google Scholar] [CrossRef] [Green Version]
  10. Peacock, T.H.P.; James, J.; Sealy, J.E.; Iqbal, M. A Global Perspective on H9N2 Avian Influenza Virus. Viruses 2019, 11, 620. [Google Scholar] [CrossRef] [Green Version]
  11. Gordon, J.W. Transgenic Technology and Laboratory Animal Science. ILAR J. 1997, 38, 32–41. [Google Scholar] [CrossRef]
  12. Serralbo, O.; Picard, C.A.; Marcelle, C. Long-term, inducible gene loss-of-function in the chicken embryo. Genesis 2013, 51, 372–380. [Google Scholar] [CrossRef]
  13. Han, J.Y.; Lee, H.C.; Park, T.S. Germline-competent stem cell in avian species and its application. Asian J. Androl. 2015, 17, 421–426. [Google Scholar] [CrossRef]
  14. Gordon, J.W.; Scangos, G.A.; Plotkin, D.J.; Barbosa, J.A.; Ruddle, F.H. Genetic transformation of mouse embryos by microinjection of purified DNA. Proc. Natl. Acad. Sci. USA 1980, 77, 7380–7384. [Google Scholar] [CrossRef] [Green Version]
  15. Hammer, R.E.; Pursel, V.G.; Rexroad, C.E., Jr.; Wall, R.J.; Bolt, D.J.; Ebert, K.M.; Palmiter, R.D.; Brinster, R.L. Production of transgenic rabbits, sheep and pigs by microinjection. Nature 1985, 315, 680–683. [Google Scholar] [CrossRef] [Green Version]
  16. McMahon, A.P.; Flytzanis, C.N.; Hough-Evans, B.R.; Katula, K.S.; Britten, R.J.; Davidson, E.H. Introduction of cloned DNA into sea urchin egg cytoplasm: Replication and persistence during embryogenesis. Dev. Biol. 1985, 108, 420–430. [Google Scholar] [CrossRef]
  17. Rusconi, S.; Schaffner, W. Transformation of frog embryos with a rabbit beta-globin gene. Proc. Natl. Acad. Sci. USA 1981, 78, 5051–5055. [Google Scholar] [CrossRef] [Green Version]
  18. Spradling, A.C.; Rubin, G.M. Transposition of cloned P elements into Drosophila germ line chromosomes. Science 1982, 218, 341–347. [Google Scholar] [CrossRef] [Green Version]
  19. Rubin, G.M.; Spradling, A.C. Genetic transformation of Drosophila with transposable element vectors. Science 1982, 218, 348–353. [Google Scholar] [CrossRef] [Green Version]
  20. Lanigan, T.M.; Kopera, H.C.; Saunders, T.L. Principles of Genetic Engineering. Genes 2020, 11, 291. [Google Scholar] [CrossRef] [Green Version]
  21. Fofanova, K.A. Morphologic Data on Polyspermy in Chickens. Fed. Proc. Transl. Suppl. 1965, 24, 239–247. [Google Scholar] [PubMed]
  22. Lee, H.C.; Choi, H.J.; Park, T.S.; Lee, S.I.; Kim, Y.M.; Rengaraj, D.; Nagai, H.; Sheng, G.; Lim, J.M.; Han, J.Y. Cleavage events and sperm dynamics in chick intrauterine embryos. PLoS ONE 2013, 8, e80631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Waddington, D.; Gribbin, C.; Sterling, R.J.; Sang, H.M.; Perry, M.M. Chronology of events in the first cell cycle of the polyspermic egg of the domestic fowl (Gallus domesticus). Int. J. Dev. Biol. 1998, 42, 625–628. [Google Scholar] [PubMed]
  24. Eyal-Giladi, H.; Kochav, S. From cleavage to primitive streak formation: A complementary normal table and a new look at the first stages of the development of the chick. I. General morphology. Dev. Biol. 1976, 49, 321–337. [Google Scholar] [CrossRef]
  25. Khaner, O.; Mitrani, E.; Eyal-Giladi, H. Developmental potencies of area opaca and marginal zone areas of early chick blastoderms. J. Embryol. Exp. Morphol. 1985, 89, 235–241. [Google Scholar] [CrossRef]
  26. Tsunekawa, N.; Naito, M.; Sakai, Y.; Nishida, T.; Noce, T. Isolation of chicken vasa homolog gene and tracing the origin of primordial germ cells. Development 2000, 127, 2741–2750. [Google Scholar] [CrossRef]
  27. Nakamura, Y.; Yamamoto, Y.; Usui, F.; Mushika, T.; Ono, T.; Setioko, A.R.; Takeda, K.; Nirasawa, K.; Kagami, H.; Tagami, T. Migration and proliferation of primordial germ cells in the early chicken embryo. Poult. Sci. 2007, 86, 2182–2193. [Google Scholar] [CrossRef]
  28. Lee, H.C.; Choi, H.J.; Lee, H.G.; Lim, J.M.; Ono, T.; Han, J.Y. DAZL Expression Explains Origin and Central Formation of Primordial Germ Cells in Chickens. Stem Cells Dev. 2016, 25, 68–79. [Google Scholar] [CrossRef]
  29. Love, J.; Gribbin, C.; Mather, C.; Sang, H. Transgenic birds by DNA microinjection. Biotechnology 1994, 12, 60–63. [Google Scholar] [CrossRef]
  30. Salter, D.W.; Smith, E.J.; Hughes, S.H.; Wright, S.E.; Fadly, A.M.; Witter, R.L.; Crittenden, L.B. Gene insertion into the chicken germ line by retroviruses. Poult. Sci. 1986, 65, 1445–1458. [Google Scholar] [CrossRef]
  31. Salter, D.W.; Smith, E.J.; Hughes, S.H.; Wright, S.E.; Crittenden, L.B. Transgenic chickens: Insertion of retroviral genes into the chicken germ line. Virology 1987, 157, 236–240. [Google Scholar] [CrossRef]
  32. Bosselman, R.A.; Hsu, R.Y.; Boggs, T.; Hu, S.; Bruszewski, J.; Ou, S.; Kozar, L.; Martin, F.; Green, C.; Jacobsen, F.; et al. Germline transmission of exogenous genes in the chicken. Science 1989, 243, 533–535. [Google Scholar] [CrossRef]
  33. Harvey, A.J.; Speksnijder, G.; Baugh, L.R.; Morris, J.A.; Ivarie, R. Expression of exogenous protein in the egg white of transgenic chickens. Nat. Biotechnol. 2002, 20, 396–399. [Google Scholar] [CrossRef]
  34. McGrew, M.J.; Sherman, A.; Ellard, F.M.; Lillico, S.G.; Gilhooley, H.J.; Kingsman, A.J.; Mitrophanous, K.A.; Sang, H. Efficient production of germline transgenic chickens using lentiviral vectors. EMBO Rep. 2004, 5, 728–733. [Google Scholar] [CrossRef] [Green Version]
  35. Lee, J.; Kim, D.-H.; Karolak, M.C.; Shin, S.; Lee, K. Generation of genome-edited chicken and duck lines by adenovirus-mediated in vivo genome editing. Proc. Natl. Acad. Sci. USA 2022, 119, e2214344119. [Google Scholar] [CrossRef]
  36. Mizuarai, S.; Ono, K.; Yamaguchi, K.; Nishijima, K.; Kamihira, M.; Iijima, S. Production of transgenic quails with high frequency of germ-line transmission using VSV-G pseudotyped retroviral vector. Biochem. Biophys. Res. Commun. 2001, 286, 456–463. [Google Scholar] [CrossRef]
  37. Shin, S.S.; Kim, T.M.; Kim, S.Y.; Kim, T.W.; Seo, H.W.; Lee, S.K.; Kwon, S.C.; Lee, G.S.; Kim, H.; Lim, J.M.; et al. Generation of transgenic quail through germ cell-mediated germline transmission. FASEB J. 2008, 22, 2435–2444. [Google Scholar] [CrossRef]
  38. Agate, R.J.; Scott, B.B.; Haripal, B.; Lois, C.; Nottebohm, F. Transgenic songbirds offer an opportunity to develop a genetic model for vocal learning. Proc. Natl. Acad. Sci. USA 2009, 106, 17963–17967. [Google Scholar] [CrossRef] [Green Version]
  39. Abe, K.; Matsui, S.; Watanabe, D. Transgenic songbirds with suppressed or enhanced activity of CREB transcription factor. Proc. Natl. Acad. Sci. USA 2015, 112, 7599–7604. [Google Scholar] [CrossRef] [Green Version]
  40. Liu, W.C.; Kohn, J.; Szwed, S.K.; Pariser, E.; Sepe, S.; Haripal, B.; Oshimori, N.; Marsala, M.; Miyanohara, A.; Lee, R. Human mutant huntingtin disrupts vocal learning in transgenic songbirds. Nat. Neurosci. 2015, 18, 1617–1622. [Google Scholar] [CrossRef]
  41. Gessara, I.; Dittrich, F.; Hertel, M.; Hildebrand, S.; Pfeifer, A.; Frankl-Vilches, C.; McGrew, M.; Gahr, M. Highly Efficient Genome Modification of Cultured Primordial Germ Cells with Lentiviral Vectors to Generate Transgenic Songbirds. Stem Cell Rep. 2021, 16, 784–796. [Google Scholar] [CrossRef] [PubMed]
  42. Lee, J.; Ma, J.; Lee, K. Direct delivery of adenoviral CRISPR/Cas9 vector into the blastoderm for generation of targeted gene knockout in quail. Proc. Natl. Acad. Sci. USA 2019, 116, 13288–13292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Richardson, B.E.; Lehmann, R. Mechanisms guiding primordial germ cell migration: Strategies from different organisms. Nat. Rev. Mol. Cell Biol. 2010, 11, 37–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Nakamura, Y.; Kagami, H.; Tagami, T. Development, differentiation and manipulation of chicken germ cells. Dev. Growth Differ. 2013, 55, 20–40. [Google Scholar] [CrossRef]
  45. Kang, K.S.; Lee, H.C.; Kim, H.J.; Lee, H.G.; Kim, Y.M.; Lee, H.J.; Park, Y.H.; Yang, S.Y.; Rengaraj, D.; Park, T.S.; et al. Spatial and temporal action of chicken primordial germ cells during initial migration. Reproduction 2015, 149, 179–187. [Google Scholar] [CrossRef] [Green Version]
  46. Paksa, A.; Raz, E. Zebrafish germ cells: Motility and guided migration. Curr. Opin. Cell Biol. 2015, 36, 80–85. [Google Scholar] [CrossRef]
  47. Sasado, T.; Morinaga, C.; Niwa, K.; Shinomiya, A.; Yasuoka, A.; Suwa, H.; Hirose, Y.; Yoda, H.; Henrich, T.; Deguchi, T.; et al. Mutations affecting early distribution of primordial germ cells in Medaka (Oryzias latipes) embryo. Mech. Dev. 2004, 121, 817–828. [Google Scholar] [CrossRef]
  48. Hayashi, K.; de Sousa Lopes, S.M.; Surani, M.A. Germ cell specification in mice. Science 2007, 316, 394–396. [Google Scholar] [CrossRef]
  49. Lawson, K.A.; Hage, W.J. Clonal analysis of the origin of primordial germ cells in the mouse. In Ciba Foundation Symposium 182-Germline Development: Germline Development: Ciba Foundation Symposium 182; John Willey & Sons: Chichester, UK, 1994; Volume 182, pp. 68–84, discussion 84–91. [Google Scholar]
  50. Ginsburg, M.; Eyal-Giladi, H. Primordial germ cells of the young chick blastoderm originate from the central zone of the area pellucida irrespective of the embryo-forming process. Development 1987, 101, 209–219. [Google Scholar] [CrossRef]
  51. Hamburger, V.; Hamilton, H.L. A series of normal stages in the development of the chick embryo. J. Morphol. 1951, 88, 49–92. [Google Scholar] [CrossRef]
  52. Swift, C.H. Origin and early history of the primordial germ cells of the chick. Am. J. Anat. 1914, 15, 483–516. [Google Scholar] [CrossRef] [Green Version]
  53. Eyal-Giladi, H.; Ginsburg, M.; Farbarov, A. Avian primordial germ cells are of epiblastic origin. J. Embryol. Exp. Morphol. 1981, 65, 139–147. [Google Scholar] [CrossRef]
  54. Ukeshima, A.; Yoshinaga, K.; Fujimoto, T. Scanning and transmission electron microscopic observations of chick primordial germ cells with special reference to the extravasation in their migration course. J. Electron. Microsc. 1991, 40, 124–128. [Google Scholar]
  55. Fujimoto, T.; Ukeshima, A.; Kiyofuji, R. The origin, migration and morphology of the primordial germ cells in the chick embryo. Anat. Rec. 1976, 185, 139–145. [Google Scholar] [CrossRef]
  56. Meyer, D.B. The Migration of Primordial Germ Cells in the Chick Embryo. Dev. Biol. 1964, 10, 154–190. [Google Scholar] [CrossRef]
  57. Niewkoop, P.; Sutasurya, L. Primordial Germ Cells in the Chordates; Cambridge University Press: Cambridge, UK, 1979; pp. 118–123. [Google Scholar]
  58. Kim, Y.M.; Han, J.Y. The early development of germ cells in chicken. Int. J. Dev. Biol. 2018, 62, 145–152. [Google Scholar] [CrossRef] [Green Version]
  59. Han, J.Y.; Cho, H.Y.; Kim, Y.M.; Park, K.J.; Jung, K.M.; Park, J.S. Production of quail (Coturnix japonica) germline chimeras by transfer of Ficoll-enriched spermatogonial stem cells. Theriogenology 2020, 154, 223–231. [Google Scholar] [CrossRef]
  60. Yamamoto, Y.; Usui, F.; Nakamura, Y.; Ito, Y.; Tagami, T.; Nirasawa, K.; Matsubara, Y.; Ono, T.; Kagami, H. A Novel Method to Isolate Primordial Germ Cells and Its Use for the Generation of Germline Chimeras in Chicken1. Biol. Reprod. 2007, 77, 115–119. [Google Scholar] [CrossRef] [Green Version]
  61. Chaipipat, S.; Prukudom, S.; Sritabtim, K.; Kuwana, T.; Piyasanti, Y.; Sinsiri, R.; Piantham, C.; Sangkalerd, S.; Boonsanong, S.; Pitiwong, K.; et al. Primordial germ cells isolated from individual embryos of red junglefowl and indigenous pheasants of Thailand. Theriogenology 2021, 165, 59–68. [Google Scholar] [CrossRef]
  62. Wade, A.J.; French, N.A.; Ireland, G.W. The potential for archiving and reconstituting valuable strains of turkey (Meleagris gallopavo) using primordial germ cells. Poult. Sci. 2014, 93, 799–809. [Google Scholar] [CrossRef]
  63. Jung, K.; Kim, Y.; Ono, T.; Han, J. Size-dependent isolation of primordial germ cells from avian species. Mol. Reprod. Dev. 2017, 84, 508–516. [Google Scholar] [CrossRef] [PubMed]
  64. Chen, Y.C.; Lin, S.P.; Chang, Y.Y.; Chang, W.P.; Wei, L.Y.; Liu, H.C.; Huang, J.F.; Pain, B.; Wu, S.C. In vitro culture and characterization of duck primordial germ cells. Poult. Sci. 2019, 98, 1820–1832. [Google Scholar] [CrossRef] [PubMed]
  65. Mozdziak, P.E.; Angerman-Stewart, J.; Rushton, B.; Pardue, S.L.; Petitte, J.N. Isolation of chicken primordial germ cells using fluorescence-activated cell sorting. Poult. Sci. 2005, 84, 594–600. [Google Scholar] [CrossRef] [PubMed]
  66. Ono, T.; Machida, Y. Immunomagnetic purification of viable primordial germ cells of Japanese quail (Coturnix japonica). Comp. Biochem. Physiol. Part A Mol. Integr. Physiol. 1999, 122, 255–259. [Google Scholar] [CrossRef] [PubMed]
  67. Zhao, D.F.; Kuwana, T. Purification of avian circulating primordial germ cells by Nycodenz density gradient centrifugation. Br. Poult. Sci. 2003, 44, 30–35. [Google Scholar] [CrossRef]
  68. Chang, I.; Tajima, A.; Yasuda, Y.; Chikamune, T.; Ohno, T. Simple method for isolation of primordial germ cells from chick embryos. Cell Biol. Int. Rep. 1992, 16, 853–857. [Google Scholar] [CrossRef]
  69. Ono, T.; Yokoi, R.; Aoyama, H. Transfer of male or female primordial germ cells of quail into chick embryonic gonads. Exp. Anim. 1996, 45, 347–352. [Google Scholar] [CrossRef] [Green Version]
  70. Reynaud, G. The transfer of turkey primordial germ cells to chick embryos by intravascular injection. J. Embryol. Exp. Morphol. 1969, 21, 485–507. [Google Scholar]
  71. Liu, C.; Khazanehdari, K.A.; Baskar, V.; Saleem, S.; Kinne, J.; Wernery, U.; Chang, I.K. Production of chicken progeny (Gallus gallus domesticus) from interspecies germline chimeric duck (Anas domesticus) by primordial germ cell transfer. Biol. Reprod. 2012, 86, 101. [Google Scholar] [CrossRef] [Green Version]
  72. Van de Lavoir, M.C.; Collarini, E.J.; Leighton, P.A.; Fesler, J.; Lu, D.R.; Harriman, W.D.; Thiyagasundaram, T.S.; Etches, R.J. Interspecific germline transmission of cultured primordial germ cells. PLoS ONE 2012, 7, e35664. [Google Scholar] [CrossRef]
  73. Kang, S.J.; Choi, J.W.; Kim, S.Y.; Park, K.J.; Kim, T.M.; Lee, Y.M.; Kim, H.; Lim, J.M.; Han, J.Y. Reproduction of wild birds via interspecies germ cell transplantation. Biol. Reprod. 2008, 79, 931–937. [Google Scholar] [CrossRef] [Green Version]
  74. Wernery, U.; Liu, C.; Baskar, V.; Guerineche, Z.; Khazanehdari, K.A.; Saleem, S.; Kinne, J.; Wernery, R.; Griffin, D.K.; Chang, I.K. Primordial germ cell-mediated chimera technology produces viable pure-line Houbara bustard offspring: Potential for repopulating an endangered species. PLoS ONE 2010, 5, e15824. [Google Scholar] [CrossRef] [Green Version]
  75. Van de Lavoir, M.C.; Diamond, J.H.; Leighton, P.A.; Mather-Love, C.; Heyer, B.S.; Bradshaw, R.; Kerchner, A.; Hooi, L.T.; Gessaro, T.M.; Swanberg, S.E.; et al. Germline transmission of genetically modified primordial germ cells. Nature 2006, 441, 766–769. [Google Scholar] [CrossRef]
  76. Choi, J.W.; Kim, S.; Kim, T.M.; Kim, Y.M.; Seo, H.W.; Park, T.S.; Jeong, J.W.; Song, G.; Han, J.Y. Basic fibroblast growth factor activates MEK/ERK cell signaling pathway and stimulates the proliferation of chicken primordial germ cells. PLoS ONE 2010, 5, e12968. [Google Scholar] [CrossRef]
  77. Macdonald, J.; Glover, J.D.; Taylor, L.; Sang, H.M.; McGrew, M.J. Characterisation and germline transmission of cultured avian primordial germ cells. PLoS ONE 2010, 5, e15518. [Google Scholar] [CrossRef] [Green Version]
  78. Lee, H.C.; Lim, S.; Han, J.Y. Wnt/beta-catenin signaling pathway activation is required for proliferation of chicken primordial germ cells in vitro. Sci. Rep. 2016, 6, 34510. [Google Scholar] [CrossRef] [Green Version]
  79. Whyte, J.; Glover, J.D.; Woodcock, M.; Brzeszczynska, J.; Taylor, L.; Sherman, A.; Kaiser, P.; McGrew, M.J. FGF, Insulin, and SMAD Signaling Cooperate for Avian Primordial Germ Cell Self-Renewal. Stem Cell Rep. 2015, 5, 1171–1182. [Google Scholar] [CrossRef] [Green Version]
  80. Yakhkeshi, S.; Rahimi, S.; Sharafi, M.; Hassani, S.N.; Taleahmad, S.; Shahverdi, A.; Baharvand, H. In vitro improvement of quail primordial germ cell expansion through activation of TGF-beta signaling pathway. J. Cell. Biochem. 2018, 119, 4309–4319. [Google Scholar] [CrossRef]
  81. Jung, K.M.; Kim, Y.M.; Keyte, A.L.; Biegler, M.T.; Rengaraj, D.; Lee, H.J.; Mello, C.V.; Velho, T.A.F.; Fedrigo, O.; Haase, B.; et al. Identification and characterization of primordial germ cells in a vocal learning Neoaves species, the zebra finch. FASEB J. 2019, 33, 13825–13836. [Google Scholar] [CrossRef] [Green Version]
  82. Park, T.S.; Han, J.Y. piggyBac transposition into primordial germ cells is an efficient tool for transgenesis in chickens. Proc. Natl. Acad. Sci. USA 2012, 109, 9337–9341. [Google Scholar] [CrossRef] [Green Version]
  83. Macdonald, J.; Taylor, L.; Sherman, A.; Kawakami, K.; Takahashi, Y.; Sang, H.M.; McGrew, M.J. Efficient genetic modification and germ-line transmission of primordial germ cells using piggyBac and Tol2 transposons. Proc. Natl. Acad. Sci. USA 2012, 109, E1466–E1472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Lee, H.J.; Lee, H.C.; Kim, Y.M.; Hwang, Y.S.; Park, Y.H.; Park, T.S.; Han, J.Y. Site-specific recombination in the chicken genome using Flipase recombinase-mediated cassette exchange. FASEB J. 2016, 30, 555–563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Ivics, Z.; Li, M.A.; Mates, L.; Boeke, J.D.; Nagy, A.; Bradley, A.; Izsvak, Z. Transposon-mediated genome manipulation in vertebrates. Nat. Methods 2009, 6, 415–422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Aronovich, E.L.; Bell, J.B.; Belur, L.R.; Gunther, R.; Koniar, B.; Erickson, D.C.; Schachern, P.A.; Matise, I.; McIvor, R.S.; Whitley, C.B.; et al. Prolonged expression of a lysosomal enzyme in mouse liver after Sleeping Beauty transposon-mediated gene delivery: Implications for non-viral gene therapy of mucopolysaccharidoses. J. Gene Med. 2007, 9, 403–415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Park, T.S.; Lee, H.G.; Moon, J.K.; Lee, H.J.; Yoon, J.W.; Yun, B.N.; Kang, S.C.; Kim, J.; Kim, H.; Han, J.Y.; et al. Deposition of bioactive human epidermal growth factor in the egg white of transgenic hens using an oviduct-specific minisynthetic promoter. FASEB J. 2015, 29, 2386–2396. [Google Scholar] [CrossRef] [Green Version]
  88. Kim, Y.M.; Park, J.S.; Kim, S.K.; Jung, K.M.; Hwang, Y.S.; Han, M.; Lee, H.J.; Seo, H.W.; Suh, J.Y.; Han, B.K.; et al. The transgenic chicken derived anti-CD20 monoclonal antibodies exhibits greater anti-cancer therapeutic potential with enhanced Fc effector functions. Biomaterials 2018, 167, 58–68. [Google Scholar] [CrossRef]
  89. Reynaud, G. Reproductive capacity and offspring of chickens submitted to a transfer of primordial germ cells during embryonic life. Wilhelm Roux’s Arch. Dev. Biol. 1976, 179, 85–110. [Google Scholar] [CrossRef]
  90. Wentworth, B.C.; Tsai, H.; Hallett, J.H.; Gonzales, D.S.; Rajcic-Spasojevic, G. Manipulation of avian primordial germ cells and gonadal differentiation. Poult. Sci. 1989, 68, 999–1010. [Google Scholar] [CrossRef]
  91. Vick, L.; Li, Y.; Simkiss, K. Transgenic birds from transformed primordial germ cells. Proc. Biol. Sci. 1993, 251, 179–182. [Google Scholar] [CrossRef]
  92. Yasuda, Y.; Tajima, A.; Fujimoto, T.; Kuwana, T. A method to obtain avian germ-line chimaeras using isolated primordial germ cells. J. Reprod. Fertil. 1992, 96, 521–528. [Google Scholar] [CrossRef] [Green Version]
  93. Naito, M.; Tajima, A.; Tagami, T.; Yasuda, Y.; Kuwana, T. Preservation of chick primordial germ cells in liquid nitrogen and subsequent production of viable offspring. J. Reprod. Fertil. 1994, 102, 321–325. [Google Scholar] [CrossRef] [Green Version]
  94. Tajima, A.; Naito, M.; Yasuda, Y.; Kuwana, T. Production of germ-line chimeras by transfer of cryopreserved gonadal primordial germ cells (gPGCs) in chicken. J. Exp. Zool. 1998, 280, 265–267. [Google Scholar] [CrossRef]
  95. Trefil, P.; Aumann, D.; Kosolová, A.; Mucksová, J.; Benešová, B.; Kallina, J.; Wurmser, C.; Fries, R.; Elleder, D.; Schsser, B.; et al. Male fertility restored by transplanting primordial germ cells into testes: A new way towards efficient transgenesis in chicken. Sci. Rep. 2017, 7, 14246. [Google Scholar] [CrossRef]
  96. Carsience, R.S.; Clark, M.E.; Verrinder Gibbins, A.M.; Etches, R.J. Germline chimeric chickens from dispersed donor blastodermal cells and compromised recipient embryos. Development 1993, 117, 669–675. [Google Scholar] [CrossRef]
  97. Nakamura, Y.; Usui, F.; Ono, T.; Takeda, K.; Nirasawa, K.; Kagami, H.; Tagami, T. Germline replacement by transfer of primordial germ cells into partially sterilized embryos in the chicken. Biol. Reprod. 2010, 83, 130–137. [Google Scholar] [CrossRef] [Green Version]
  98. Nakamura, Y. Poultry genetic resource conservation using primordial germ cells. J. Reprod. Dev. 2016, 62, 431–437. [Google Scholar] [CrossRef] [Green Version]
  99. Kim, Y.M.; Park, K.J.; Park, J.S.; Jung, K.M.; Han, J.Y. In vivo enrichment of busulfan-resistant germ cells for efficient production of transgenic avian models. Sci. Rep. 2021, 11, 9127. [Google Scholar] [CrossRef]
  100. Jung, K.M.; Kim, Y.M.; Han, J.Y. Transplantation and enrichment of busulfan-resistant primordial germ cells into adult testes for efficient production of germline chimeras in songbirds. Biol. Reprod. 2022, 108, 316–323. [Google Scholar] [CrossRef]
  101. Taylor, L.; Carlson, D.F.; Nandi, S.; Sherman, A.; Fahrenkrug, S.C.; McGrew, M.J. Efficient TALEN-mediated gene targeting of chicken primordial germ cells. Development 2017, 144, 928–934. [Google Scholar] [CrossRef] [Green Version]
  102. Ballantyne, M.; Woodcock, M.; Doddamani, D.; Hu, T.; Taylor, L.; Hawken, R.J.; McGrew, M.J. Direct allele introgression into pure chicken breeds using Sire Dam Surrogate (SDS) mating. Nat. Commun. 2021, 12, 659. [Google Scholar] [CrossRef]
  103. Rogers, C.S. Engineering Large Animal Species to Model Human Diseases. Curr. Protoc. Hum. Genet. 2016, 90, 15.9.1–15.9.14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Thomas, K.R.; Capecchi, M.R. Site-directed mutagenesis by gene targeting in mouse embryo-derived stem cells. Cell 1987, 51, 503–512. [Google Scholar] [CrossRef] [PubMed]
  105. Lai, L.; Kolber-Simonds, D.; Park, K.W.; Cheong, H.T.; Greenstein, J.L.; Im, G.S.; Samuel, M.; Bonk, A.; Rieke, A.; Day, B.N.; et al. Production of alpha-1,3-galactosyltransferase knockout pigs by nuclear transfer cloning. Science 2002, 295, 1089–1092. [Google Scholar] [CrossRef] [PubMed]
  106. Nottle, M.B.; Beebe, L.F.; Harrison, S.J.; McIlfatrick, S.M.; Ashman, R.J.; O’Connell, P.J.; Salvaris, E.J.; Fisicaro, N.; Pommey, S.; Cowan, P.J.; et al. Production of homozygous alpha-1,3-galactosyltransferase knockout pigs by breeding and somatic cell nuclear transfer. Xenotransplantation 2007, 14, 339–344. [Google Scholar] [CrossRef]
  107. Petitte, J.N.; Liu, G.; Yang, Z. Avian pluripotent stem cells. Mech. Dev. 2004, 121, 1159–1168. [Google Scholar] [CrossRef]
  108. Petitte, J.N.; Clark, M.E.; Liu, G.; Verrinder Gibbins, A.M.; Etches, R.J. Production of somatic and germline chimeras in the chicken by transfer of early blastodermal cells. Development 1990, 108, 185–189. [Google Scholar] [CrossRef]
  109. Pain, B.; Clark, M.E.; Shen, M.; Nakazawa, H.; Sakurai, M.; Samarut, J.; Etches, R.J. Long-term in vitro culture and characterisation of avian embryonic stem cells with multiple morphogenetic potentialities. Development 1996, 122, 2339–2348. [Google Scholar] [CrossRef]
  110. Van de Lavoir, M.C.; Mather-Love, C.; Leighton, P.; Diamond, J.H.; Heyer, B.S.; Roberts, R.; Zhu, L.; Winters-Digiacinto, P.; Kerchner, A.; Gessaro, T.; et al. High-grade transgenic somatic chimeras from chicken embryonic stem cells. Mech. Dev. 2006, 123, 31–41. [Google Scholar] [CrossRef]
  111. Lavial, F.; Acloque, H.; Bachelard, E.; Nieto, M.A.; Samarut, J.; Pain, B. Ectopic expression of Cvh (Chicken Vasa homologue) mediates the reprogramming of chicken embryonic stem cells to a germ cell fate. Dev. Biol. 2009, 330, 73–82. [Google Scholar] [CrossRef]
  112. Takahashi, K.; Yamanaka, S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 2006, 126, 663–676. [Google Scholar] [CrossRef] [Green Version]
  113. Liu, K.; Ji, G.; Mao, J.; Liu, M.; Wang, L.; Chen, C.; Liu, L. Generation of porcine-induced pluripotent stem cells by using OCT4 and KLF4 porcine factors. Cell. Reprogram. 2012, 14, 505–513. [Google Scholar] [CrossRef] [Green Version]
  114. Yu, J.; Vodyanik, M.A.; Smuga-Otto, K.; Antosiewicz-Bourget, J.; Frane, J.L.; Tian, S.; Nie, J.; Jonsdottir, G.A.; Ruotti, V.; Stewart, R.; et al. Induced pluripotent stem cell lines derived from human somatic cells. Science 2007, 318, 1917–1920. [Google Scholar] [CrossRef]
  115. Zhong, B.; Trobridge, G.D.; Zhang, X.; Watts, K.L.; Ramakrishnan, A.; Wohlfahrt, M.; Adair, J.E.; Kiem, H.P. Efficient generation of nonhuman primate induced pluripotent stem cells. Stem Cells Dev. 2011, 20, 795–807. [Google Scholar] [CrossRef] [Green Version]
  116. Rossello, R.A.; Chen, C.C.; Dai, R.; Howard, J.T.; Hochgeschwender, U.; Jarvis, E.D. Mammalian genes induce partially reprogrammed pluripotent stem cells in non-mammalian vertebrate and invertebrate species. Elife 2013, 2, e00036. [Google Scholar] [CrossRef] [Green Version]
  117. Lu, Y.; West, F.D.; Jordan, B.J.; Mumaw, J.L.; Jordan, E.T.; Gallegos-Cardenas, A.; Beckstead, R.B.; Stice, S.L. Avian-induced pluripotent stem cells derived using human reprogramming factors. Stem Cells Dev. 2012, 21, 394–403. [Google Scholar] [CrossRef]
  118. Kim, Y.M.; Park, Y.H.; Lim, J.M.; Jung, H.; Han, J.Y. Technical note: Induction of pluripotent stem cell-like cells from chicken feather follicle cells. J. Anim. Sci. 2017, 95, 3479–3486. [Google Scholar] [CrossRef]
  119. Zhao, R.; Zuo, Q.; Yuan, X.; Jin, K.; Jin, J.; Ding, Y.; Zhang, C.; Li, T.; Jiang, J.; Li, J.; et al. Production of viable chicken by allogeneic transplantation of primordial germ cells induced from somatic cells. Nat. Commun. 2021, 12, 2989. [Google Scholar] [CrossRef]
  120. Ryu, B.Y.; Orwig, K.E.; Oatley, J.M.; Lin, C.C.; Chang, L.J.; Avarbock, M.R.; Brinster, R.L. Efficient generation of transgenic rats through the male germline using lentiviral transduction and transplantation of spermatogonial stem cells. J. Androl. 2007, 28, 353–360. [Google Scholar] [CrossRef]
  121. Hamra, F.K.; Gatlin, J.; Chapman, K.M.; Grellhesl, D.M.; Garcia, J.V.; Hammer, R.E.; Garbers, D.L. Production of transgenic rats by lentiviral transduction of male germ-line stem cells. Proc. Natl. Acad. Sci. USA 2002, 99, 14931–14936. [Google Scholar] [CrossRef] [Green Version]
  122. Kanatsu-Shinohara, M.; Kato, M.; Takehashi, M.; Morimoto, H.; Takashima, S.; Chuma, S.; Nakatsuji, N.; Hirabayashi, M.; Shinohara, T. Production of transgenic rats via lentiviral transduction and xenogeneic transplantation of spermatogonial stem cells. Biol. Reprod. 2008, 79, 1121–1128. [Google Scholar] [CrossRef]
  123. Li, C.H.; Yan, L.Z.; Ban, W.Z.; Tu, Q.; Wu, Y.; Wang, L.; Bi, R.; Ji, S.; Ma, Y.H.; Nie, W.H.; et al. Long-term propagation of tree shrew spermatogonial stem cells in culture and successful generation of transgenic offspring. Cell Res. 2017, 27, 241–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Kanatsu-Shinohara, M.; Kato-Itoh, M.; Ikawa, M.; Takehashi, M.; Sanbo, M.; Morioka, Y.; Tanaka, T.; Morimoto, H.; Hirabayashi, M.; Shinohara, T. Homologous recombination in rat germline stem cells. Biol. Reprod. 2011, 85, 208–217. [Google Scholar] [CrossRef] [PubMed]
  125. Izsvak, Z.; Frohlich, J.; Grabundzija, I.; Shirley, J.R.; Powell, H.M.; Chapman, K.M.; Ivics, Z.; Hamra, F.K. Generating knockout rats by transposon mutagenesis in spermatogonial stem cells. Nat. Methods 2010, 7, 443–445. [Google Scholar] [CrossRef] [PubMed]
  126. Tegelenbosch, R.A.; de Rooij, D.G. A quantitative study of spermatogonial multiplication and stem cell renewal in the C3H/101 F1 hybrid mouse. Mutat. Res./Fundam. Mol. Mech. Mutagen. 1993, 290, 193–200. [Google Scholar] [CrossRef] [PubMed]
  127. Kossack, N.; Meneses, J.; Shefi, S.; Nguyen, H.N.; Chavez, S.; Nicholas, C.; Gromoll, J.; Turek, P.J.; Reijo-Pera, R.A. Isolation and characterization of pluripotent human spermatogonial stem cell-derived cells. Stem Cells 2009, 27, 138–149. [Google Scholar] [CrossRef] [Green Version]
  128. Nagano, M.; Ryu, B.Y.; Brinster, C.J.; Avarbock, M.R.; Brinster, R.L. Maintenance of mouse male germ line stem cells in vitro. Biol. Reprod. 2003, 68, 2207–2214. [Google Scholar] [CrossRef]
  129. Nasiri, Z.; Hosseini, S.M.; Hajian, M.; Abedi, P.; Bahadorani, M.; Baharvand, H.; Nasr-Esfahani, M.H. Effects of different feeder layers on short-term culture of prepubertal bovine testicular germ cells in-vitro. Theriogenology 2012, 77, 1519–1528. [Google Scholar] [CrossRef]
  130. Kuijk, E.W.; Colenbrander, B.; Roelen, B.A. The effects of growth factors on in vitro-cultured porcine testicular cells. Reproduction 2009, 138, 721–731. [Google Scholar] [CrossRef] [Green Version]
  131. Jung, J.G.; Lee, Y.M.; Park, T.S.; Park, S.H.; Lim, J.M.; Han, J.Y. Identification, culture, and characterization of germline stem cell-like cells in chicken testes. Biol. Reprod. 2007, 76, 173–182. [Google Scholar] [CrossRef]
  132. Pramod, R.K.; Lee, B.R.; Kim, Y.M.; Lee, H.J.; Park, Y.H.; Ono, T.; Lim, J.M.; Han, J.Y. Isolation, Characterization, and In Vitro Culturing of Spermatogonial Stem Cells in Japanese Quail (Coturnix japonica). Stem Cells Dev. 2017, 26, 60–70. [Google Scholar] [CrossRef]
  133. Zhang, Y.; Sastre, D.; Wang, F. CRISPR/Cas9 Genome Editing: A Promising Tool for Therapeutic Applications of Induced Pluripotent Stem Cells. Curr. Stem Cell Res. Ther. 2018, 13, 243–251. [Google Scholar] [CrossRef]
  134. De Silva Feelixge, H.S.; Stone, D.; Roychoudhury, P.; Aubert, M.; Jerome, K.R. CRISPR/Cas9 and Genome Editing for Viral Disease-Is Resistance Futile? ACS Infect. Dis. 2018, 4, 871–880. [Google Scholar] [CrossRef]
  135. Ingusci, S.; Verlengia, G.; Soukupova, M.; Zucchini, S.; Simonato, M. Gene Therapy Tools for Brain Diseases. Front. Pharmacol. 2019, 10, 724. [Google Scholar] [CrossRef]
  136. Chen, K.; Wang, Y.; Zhang, R.; Zhang, H.; Gao, C. CRISPR/Cas Genome Editing and Precision Plant Breeding in Agriculture. Annu. Rev. Plant Biol. 2019, 70, 667–697. [Google Scholar] [CrossRef]
  137. Zhu, H.; Li, C.; Gao, C. Applications of CRISPR-Cas in agriculture and plant biotechnology. Nat. Rev. Mol. Cell Biol. 2020, 21, 661–677. [Google Scholar] [CrossRef]
  138. Schusser, B.; Collarini, E.J.; Yi, H.; Izquierdo, S.M.; Fesler, J.; Pedersen, D.; Klasing, K.C.; Kaspers, B.; Harriman, W.D.; van de Lavoir, M.C.; et al. Immunoglobulin knockout chickens via efficient homologous recombination in primordial germ cells. Proc. Natl. Acad. Sci. USA 2013, 110, 20170–20175. [Google Scholar] [CrossRef] [Green Version]
  139. Kim, H.; Kim, J.S. A guide to genome engineering with programmable nucleases. Nat. Rev. Genet. 2014, 15, 321–334. [Google Scholar] [CrossRef]
  140. Li, H.; Yang, Y.; Hong, W.; Huang, M.; Wu, M.; Zhao, X. Applications of genome editing technology in the targeted therapy of human diseases: Mechanisms, advances and prospects. Signal Transduct. Target. Ther. 2020, 5, 1. [Google Scholar] [CrossRef] [Green Version]
  141. Gaj, T.; Gersbach, C.A.; Barbas, C.F., 3rd. ZFN, TALEN, and CRISPR/Cas-based methods for genome engineering. Trends Biotechnol. 2013, 31, 397–405. [Google Scholar] [CrossRef] [Green Version]
  142. Park, T.S.; Lee, H.J.; Kim, K.H.; Kim, J.S.; Han, J.Y. Targeted gene knockout in chickens mediated by TALENs. Proc. Natl. Acad. Sci. USA 2014, 111, 12716–12721. [Google Scholar] [CrossRef] [Green Version]
  143. Jinek, M.; Chylinski, K.; Fonfara, I.; Hauer, M.; Doudna, J.A.; Charpentier, E. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 2012, 337, 816–821. [Google Scholar] [CrossRef] [PubMed]
  144. 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]
  145. Cho, S.W.; Kim, S.; Kim, J.M.; Kim, J.S. Targeted genome engineering in human cells with the Cas9 RNA-guided endonuclease. Nat. Biotechnol. 2013, 31, 230–232. [Google Scholar] [CrossRef] [PubMed]
  146. Sung, Y.H.; Kim, J.M.; Kim, H.T.; Lee, J.; Jeon, J.; Jin, Y.; Choi, J.H.; Ban, Y.H.; Ha, S.J.; Kim, C.H.; et al. Highly efficient gene knockout in mice and zebrafish with RNA-guided endonucleases. Genome Res. 2014, 24, 125–131. [Google Scholar] [CrossRef] [Green Version]
  147. Shan, Q.; Wang, Y.; Li, J.; Zhang, Y.; Chen, K.; Liang, Z.; Zhang, K.; Liu, J.; Xi, J.J.; Qiu, J.L.; et al. Targeted genome modification of crop plants using a CRISPR-Cas system. Nat. Biotechnol. 2013, 31, 686–688. [Google Scholar] [CrossRef]
  148. Jiang, W.; Zhou, H.; Bi, H.; Fromm, M.; Yang, B.; Weeks, D.P. Demonstration of CRISPR/Cas9/sgRNA-mediated targeted gene modification in Arabidopsis, tobacco, sorghum and rice. Nucleic Acids Res. 2013, 41, e188. [Google Scholar] [CrossRef]
  149. Hwang, W.Y.; Fu, Y.; Reyon, D.; Maeder, M.L.; Tsai, S.Q.; Sander, J.D.; Peterson, R.T.; Yeh, J.R.; Joung, J.K. Efficient genome editing in zebrafish using a CRISPR-Cas system. Nat. Biotechnol. 2013, 31, 227–229. [Google Scholar] [CrossRef] [Green Version]
  150. Oishi, I.; Yoshii, K.; Miyahara, D.; Kagami, H.; Tagami, T. Targeted mutagenesis in chicken using CRISPR/Cas9 system. Sci. Rep. 2016, 6, 23980. [Google Scholar] [CrossRef] [Green Version]
  151. Dimitrov, L.; Pedersen, D.; Ching, K.H.; Yi, H.; Collarini, E.J.; Izquierdo, S.; van de Lavoir, M.C.; Leighton, P.A. Germline Gene Editing in Chickens by Efficient CRISPR-Mediated Homologous Recombination in Primordial Germ Cells. PLoS ONE 2016, 11, e0154303. [Google Scholar] [CrossRef]
  152. Park, T.S.; Park, J.; Lee, J.H.; Park, J.W.; Park, B.C. Disruption of G0/G1 switch gene 2 (G0S2) reduced abdominal fat deposition and altered fatty acid composition in chicken. FASEB J. 2019, 33, 1188–1198. [Google Scholar] [CrossRef]
  153. Kim, G.D.; Lee, J.H.; Song, S.; Kim, S.W.; Han, J.S.; Shin, S.P.; Park, B.C.; Park, T.S. Generation of myostatin-knockout chickens mediated by D10A-Cas9 nickase. FASEB J. 2020, 34, 5688–5696. [Google Scholar] [CrossRef] [Green Version]
  154. Lee, J.; Kim, D.H.; Lee, K. Muscle Hyperplasia in Japanese Quail by Single Amino Acid Deletion in MSTN Propeptide. Int. J. Mol. Sci. 2020, 21, 1504. [Google Scholar] [CrossRef] [Green Version]
  155. Rieblinger, B.; Sid, H.; Duda, D.; Bozoglu, T.; Klinger, R.; Schlickenrieder, A.; Lengyel, K.; Flisikowski, K.; Flisikowska, T.; Simm, N.; et al. Cas9-expressing chickens and pigs as resources for genome editing in livestock. Proc. Natl. Acad. Sci. USA 2021, 118, e2022562118. [Google Scholar] [CrossRef]
  156. Challagulla, A.; Jenkins, K.A.; O’Neil, T.E.; Morris, K.R.; Wise, T.G.; Tizard, M.L.; Bean, A.G.D.; Schat, K.A.; Doran, T.J. Germline engineering of the chicken genome using CRISPR/Cas9 by in vitro transfection of PGCs. Anim. Biotechnol. 2020, 1–10. [Google Scholar] [CrossRef]
  157. Komor, A.C.; Badran, A.H.; Liu, D.R. Editing the Genome Without Double-Stranded DNA Breaks. ACS Chem. Biol. 2018, 13, 383–388. [Google Scholar] [CrossRef]
  158. Lee, K.Y.; Lee, H.J.; Choi, H.J.; Han, S.T.; Lee, K.H.; Park, K.J.; Park, J.S.; Jung, K.M.; Kim, Y.M.; Han, H.J.; et al. Highly elevated base excision repair pathway in primordial germ cells causes low base editing activity in chickens. FASEB J. 2020, 34, 15907–15921. [Google Scholar] [CrossRef]
  159. Anzalone, A.V.; Randolph, P.B.; Davis, J.R.; Sousa, A.A.; Koblan, L.W.; Levy, J.M.; Chen, P.J.; Wilson, C.; Newby, G.A.; Raguram, A.; et al. Search-and-replace genome editing without double-strand breaks or donor DNA. Nature 2019, 576, 149–157. [Google Scholar] [CrossRef]
  160. Atsuta, Y.; Suzuki, K.; Iikawa, H.; Yaguchi, H.; Saito, D. Prime editing in chicken fibroblasts and primordial germ cells. Dev. Growth Differ. 2022, 64, 548–557. [Google Scholar] [CrossRef]
  161. Koslová, A.; Trefil, P.; Mucksová, J.; Reinišová, M.; Plachý, J.; Kalina, J.; Kučerová, D.; Geryk, J.; Krchlíková, V.; Lejčková, B.; et al. Precise CRISPR/Cas9 editing of the NHE1 gene renders chickens resistant to the J subgroup of avian leukosis virus. Proc. Natl. Acad. Sci. USA 2020, 117, 2108–2112. [Google Scholar] [CrossRef]
  162. Park, Y.H.; Chungu, K.; Lee, S.B.; Woo, S.J.; Cho, H.Y.; Lee, H.J.; Rengaraj, D.; Lee, J.H.; Song, C.S.; Lim, J.M.; et al. Host-Specific Restriction of Avian Influenza Virus Caused by Differential Dynamics of ANP32 Family Members. J. Infect. Dis. 2020, 221, 71–80. [Google Scholar] [CrossRef]
  163. Hagag, I.T.; Wight, D.J.; Bartsch, D.; Sid, H.; Jordan, I.; Bertzbach, L.D.; Schusser, B.; Kaufer, B.B. Abrogation of Marek’s disease virus replication using CRISPR/Cas9. Sci. Rep. 2020, 10, 10919. [Google Scholar] [CrossRef] [PubMed]
  164. Fadly, A.M. Isolation and identification of avian leukosis viruses: A review. Avian Pathol. 2000, 29, 529–535. [Google Scholar] [CrossRef] [PubMed]
  165. Lee, H.J.; Lee, K.Y.; Jung, K.M.; Park, K.J.; Lee, K.O.; Suh, J.Y.; Yao, Y.; Nair, V.; Han, J.Y. Precise gene editing of chicken Na+/H+ exchange type 1 (chNHE1) confers resistance to avian leukosis virus subgroup J (ALV-J). Dev. Comp. Immunol. 2017, 77, 340–349. [Google Scholar] [CrossRef] [PubMed]
  166. Lee, H.J.; Lee, K.Y.; Park, Y.H.; Choi, H.J.; Yao, Y.; Nair, V.; Han, J.Y. Acquisition of resistance to avian leukosis virus subgroup B through mutations on tvb cysteine-rich domains in DF-1 chicken fibroblasts. Vet. Res. 2017, 48, 48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Lee, H.J.; Park, K.J.; Lee, K.Y.; Yao, Y.; Nair, V.; Han, J.Y. Sequential disruption of ALV host receptor genes reveals no sharing of receptors between ALV subgroups A, B, and J. J. Anim. Sci. Biotechnol. 2019, 10, 23. [Google Scholar] [CrossRef] [Green Version]
  168. Koslova, A.; Kucerova, D.; Reinisova, M.; Geryk, J.; Trefil, P.; Hejnar, J. Genetic Resistance to Avian Leukosis Viruses Induced by CRISPR/Cas9 Editing of Specific Receptor Genes in Chicken Cells. Viruses 2018, 10, 605. [Google Scholar] [CrossRef] [Green Version]
  169. Hellmich, R.; Sid, H.; Lengyel, K.; Flisikowski, K.; Schlickenrieder, A.; Bartsch, D.; Thoma, T.; Bertzbach, L.D.; Kaufer, B.B.; Nair, V.; et al. Acquiring Resistance Against a Retroviral Infection via CRISPR/Cas9 Targeted Genome Editing in a Commercial Chicken Line. Front. Genome. Ed. 2020, 2, 3. [Google Scholar] [CrossRef]
  170. Long, J.S.; Giotis, E.S.; Moncorge, O.; Frise, R.; Mistry, B.; James, J.; Morisson, M.; Iqbal, M.; Vignal, A.; Skinner, M.A.; et al. Species difference in ANP32A underlies influenza A virus polymerase host restriction. Nature 2016, 529, 101–104. [Google Scholar] [CrossRef] [Green Version]
  171. Long, J.S.; Idoko-Akoh, A.; Mistry, B.; Goldhill, D.; Staller, E.; Schreyer, J.; Ross, C.; Goodbourn, S.; Shelton, H.; Skinner, M.A.; et al. Species specific differences in use of ANP32 proteins by influenza A virus. Elife 2019, 8, e45066. [Google Scholar] [CrossRef]
  172. Park, Y.H.; Woo, S.J.; Chungu, K.; Lee, S.B.; Shim, J.H.; Lee, H.J.; Kim, I.; Rengaraj, D.; Song, C.S.; Suh, J.Y.; et al. Asp149 and Asp152 in chicken and human ANP32A play an essential role in the interaction with influenza viral polymerase. FASEB J. 2021, 35, e21630. [Google Scholar] [CrossRef]
  173. Barber, M.R.W.; Aldridge, J.R., Jr.; Webster, R.G.; Magor, K.E. Association of RIG-I with innate immunity of ducks to influenza. Proc. Natl. Acad. Sci. USA 2010, 107, 5913–5918. [Google Scholar] [CrossRef] [Green Version]
  174. Ichikawa, K.; Motoe, Y.; Ezaki, R.; Matsuzaki, M.; Horiuchi, H. Knock-in of the duck retinoic acid-inducible gene I (RIG-I) into the Mx gene in DF-1 cells enables both stable and immune response-dependent RIG-I expression. Biochem. Biophys. Rep. 2021, 27, 101084. [Google Scholar] [CrossRef]
  175. Woo, S.J.; Choi, H.J.; Park, Y.H.; Rengaraj, D.; Kim, J.K.; Han, J.Y. Amplification of immunity by engineering chicken MDA5 combined with the C terminal domain (CTD) of RIG-I. Appl. Microbiol. Biotechnol. 2022, 106, 1599–1613. [Google Scholar] [CrossRef]
  176. Challagulla, A.; Jenkins, K.A.; O’Neil, T.E.; Shi, S.; Morris, K.R.; Wise, T.G.; Paradkar, P.N.; Tizard, M.L.; Doran, T.J.; Schat, K.A. In Vivo Inhibition of Marke’s Disease Virus in Transgenic Chickens Expressing Cas9 and gRNA against ICP4. Microorganisms 2021, 9, 164. [Google Scholar] [CrossRef]
  177. Ivarie, R. Avian transgenesis: Progress towards the promise. Trends Biotechnol. 2003, 21, 14–19. [Google Scholar] [CrossRef]
  178. Han, J.Y. Germ cells and transgenesis in chickens. Comp. Immunol. Microbiol. Infect. Dis. 2009, 32, 61–80. [Google Scholar] [CrossRef]
  179. Sang, H.M. Transgenics, chickens and therapeutic proteins. Vox Sang. 2004, 87 (Suppl. 2), 164–166. [Google Scholar] [CrossRef]
  180. Zhu, L.; van de Lavoir, M.C.; Albanese, J.; Beenhouwer, D.O.; Cardarelli, P.M.; Cuison, S.; Deng, D.F.; Deshpande, S.; Diamond, J.H.; Green, L.; et al. Production of human monoclonal antibody in eggs of chimeric chickens. Nat. Biotechnol. 2005, 23, 1159–1169. [Google Scholar] [CrossRef]
  181. Lillico, S.; Sherman, A.; McGrew, M.; Robertson, C.; Smith, J.; Haslam, C.; Barnard, P.; Radcliffe, P.; Mitrophanous, K.; Elliot, E.; et al. Oviduct-specific expression of two therapeutic proteins in transgenic hens. Proc. Natl. Acad. Sci. USA 2007, 104, 1771–1776. [Google Scholar] [CrossRef] [Green Version]
  182. Cao, D.; Wu, H.; Li, Q.; Sun, Y.; Liu, T.; Fei, J.; Zhao, Y.; Wu, S.; Hu, X.; Li, N. Expression of recombinant human lysozyme in egg whites of transgenic hens. PLoS ONE 2015, 10, e0118626. [Google Scholar] [CrossRef]
  183. Oishi, I.; Yoshii, K.; Miyahara, D.; Tagami, T. Efficient production of human interferon beta in the white of eggs from ovalbumin gene-targeted hens. Sci. Rep. 2018, 8, 10203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Kim, Y.M.; Park, J.S.; Choi, H.J.; Jung, K.M.; Lee, K.Y.; Shim, J.H.; Park, K.J.; Han, J.Y. Efficient production of recombinant human adiponectin in egg white using genome edited chickens. Front. Nutr. 2022, 9, 1068558. [Google Scholar] [CrossRef] [PubMed]
  185. Kim, Y.M.; Shim, J.H.; Park, J.S.; Choi, H.J.; Jung, K.M.; Lee, K.Y.; Park, K.J.; Han, J.Y. Sequential verification of exogenous protein production in OVA gene-targeted chicken bioreactors. Poult. Sci. 2023, 102, 102247. [Google Scholar] [CrossRef] [PubMed]
  186. Sheridan, C. FDA approves ‘farmaceutical’ drug from transgenic chickens. Nat. Biotechnol. 2016, 34, 117–119. [Google Scholar] [CrossRef]
  187. Sly, W.S.; Vogler, C.; Grubb, J.H.; Levy, B.; Galvin, N.; Tan, Y.; Nishioka, T.; Tomatsu, S. Enzyme therapy in mannose receptor-null mucopolysaccharidosis VII mice defines roles for the mannose 6-phosphate and mannose receptors. Proc. Natl. Acad. Sci. USA 2006, 103, 15172–15177. [Google Scholar] [CrossRef] [Green Version]
  188. Neufeld, E.F. Enzyme replacement therapy-a brief history. In Fabry Disease: Perspectives from 5 Years of FOS; Mehta, A., Beck, M., Sunder-Plassmann, G., Eds.; Oxford PharmaGenesis: Oxford, UK, 2006. [Google Scholar]
  189. Hintze, S.; Limmer, S.; Dabrowska-Schlepp, P.; Berg, B.; Krieghoff, N.; Busch, A.; Schaaf, A.; Meinke, P.; Schoser, B. Moss-Derived Human Recombinant GAA Provides an Optimized Enzyme Uptake in Differentiated Human Muscle Cells of Pompe Disease. Int. J. Mol. Sci. 2020, 21, 2642. [Google Scholar] [CrossRef] [Green Version]
  190. Jung, J.W.; Huy, N.X.; Kim, H.B.; Kim, N.S.; van Giap, D.; Yang, M.S. Production of recombinant human acid alpha-glucosidase with high-mannose glycans in gnt1 rice for the treatment of Pompe disease. J. Biotechnol. 2017, 249, 42–50. [Google Scholar] [CrossRef]
  191. Shields, R.L.; Lai, J.; Keck, R.; O’Connell, L.Y.; Hong, K.; Meng, Y.G.; Weikert, S.H.; Presta, L.G. Lack of fucose on human IgG1 N-linked oligosaccharide improves binding to human Fcgamma RIII and antibody-dependent cellular toxicity. J. Biol. Chem. 2002, 277, 26733–26740. [Google Scholar] [CrossRef] [Green Version]
  192. Lee, K.Y.; Choi, H.J.; Park, K.J.; Woo, S.J.; Kim, Y.M.; Han, J.Y. Development and characterization of a CRISPR/Cas9-mediated RAG1 knockout chicken model lacking mature B and T cells. Front. Immunol. 2022, 13, 892476. [Google Scholar] [CrossRef]
  193. Johnson, P.A.; Giles, J.R. The hen as a model of ovarian cancer. Nat. Rev. Cancer 2013, 13, 432–436. [Google Scholar] [CrossRef]
  194. Rengaraj, D.; Cha, D.G.; Lee, H.J.; Lee, K.Y.; Choi, Y.H.; Jung, K.M.; Kim, Y.M.; Choi, H.J.; Choi, H.J.; Yoo, E.; et al. Dissecting chicken germ cell dynamics by combining a germ cell tracing transgenic chicken model with single-cell RNA sequencing. Comput. Struct. Biotechnol. J. 2022, 20, 1654–1669. [Google Scholar] [CrossRef]
  195. Hagihara, Y.; Qkuzaki, Y.; Matsubayashi, K.; Kaneoka, H.; Suzuki, T.; Iijima, S.; Nishijima, K.-I. Primordial germ cell-specific expression of eGFP in transgenic chickens. Genesis 2020, 58, e23388. [Google Scholar] [CrossRef]
  196. Lee, H.J.; Seo, M.; Choi, H.J.; Rengaraj, D.; Jung, K.M.; Park, J.S.; Lee, K.Y.; Kim, Y.M.; Park, K.J.; Han, S.T.; et al. DMRT1 gene disruption alone induces incomplete gonad feminization in chicken. FASEB J. 2021, 35, e21876. [Google Scholar] [CrossRef]
  197. Ioannidis, J.; Taylor, G.; Zhao, D.; Liu, L.; Idoko-Akoh, A.; Gong, D.; Lovell-Badge, R.; Guioli, S.; McGrew, M.J.; Clinton, M. Primary sex determination in birds depends on DMRT1 dosage, but gonadal sex does not determine adult secondary sex characteristics. Proc. Natl. Acad. Sci. USA 2021, 118, e2020909118. [Google Scholar] [CrossRef]
  198. Lai, C.S.; Fisher, S.E.; Hurst, J.A.; Vargha-Khadem, F.; Monaco, A.P. A forkhead-domain gene is mutated in a severe speech and language disorder. Nature 2001, 413, 519–523. [Google Scholar] [CrossRef]
  199. Haesler, S.; Rochefort, C.; Georgi, B.; Licznerski, P.; Osten, P.; Scharff, C. Incomplete and inaccurate vocal imitation after knockdown of FoxP2 in songbird basal ganglia nucleus Area X. PLoS Biol. 2007, 5, e321. [Google Scholar] [CrossRef] [Green Version]
  200. Scott, B.B.; Velho, T.A.; Sim, S.; Lois, C. Applications of avian transgenesis. ILAR J. 2010, 51, 353–361. [Google Scholar] [CrossRef] [Green Version]
  201. Jung, K.M.; Seo, M.; Kim, Y.M.; Kim, J.L.; Han, J.Y. Single-Cell RNA Sequencing Revealed the Heterogeneity of Gonadal Primordial Germ Cells in Zebra Finch (Taeniopygia guttata). Front. Cell Dev. Biol. 2021, 9, 791335. [Google Scholar] [CrossRef]
  202. Jung, K.M.; Seo, M.; Han, J.Y. Comparative single-cell transcriptomic analysis reveals differences in signaling pathways in gonadal primordial germ cells between chicken (Gallus gallus) and zebra finch (Taeniopygia guttata). FASEB J. 2023, 37, e22706. [Google Scholar] [CrossRef]
  203. Jung, K.M.; Kim, Y.M.; Yoo, E.; Han, J.Y. Generation and characterization of genome-modified chondrocyte-like cells from the zebra finch cell line immortalized by c-MYC expression. Front. Zool. 2022, 19, 18. [Google Scholar] [CrossRef]
  204. Wray-Cahen, D.; Bodnar, A.; Rexroad, C.; Siewerdt, F.; Kovich, D. Advancing genome editing to improve the sustainability and resiliency of animal agriculture. CABI Agric. Biosci. 2022, 3, 21. [Google Scholar] [CrossRef]
Figure 1. Representative applications of genome edited chickens. The representative examples of genome edited birds introduced in this article are shown. The viruses must use host proteins to successfully complete their life cycle and the host gene edited birds are resistant to various types of viruses, including avian influenza viruses, avian leukosis viruses and Marek’s disease viruses. The target gene encoding therapeutic proteins are inserted into the chicken genome, such as the ovalbumin gene, to produce avian bioreactor. The egg white contains a high amount of target proteins that can be used for medical purpose. The specific gene knockout or gene tagging in the chicken provided suitable model for researchers to study the gene function.
Figure 1. Representative applications of genome edited chickens. The representative examples of genome edited birds introduced in this article are shown. The viruses must use host proteins to successfully complete their life cycle and the host gene edited birds are resistant to various types of viruses, including avian influenza viruses, avian leukosis viruses and Marek’s disease viruses. The target gene encoding therapeutic proteins are inserted into the chicken genome, such as the ovalbumin gene, to produce avian bioreactor. The egg white contains a high amount of target proteins that can be used for medical purpose. The specific gene knockout or gene tagging in the chicken provided suitable model for researchers to study the gene function.
Genes 14 00899 g001
Figure 2. Schematic representation of various strategies for genome-edited bird production. Avian germplasm is found in the zygotic stage as RNA granules and proteins, and a number of germ cells are found in the blastoderm at EGK stage X. In this stage, genome-editing tools, such as CRISPR/Cas9, can be inserted into the subgerminal cavity of the blastoderm with a virus system. The genomes of a small number of germ cells at this stage can be modified, and germline chimera can be formed. Blastoderm cells can be cultured in vitro, and chimera can be established by injecting cultured embryonic stem cells (ESCs) into the subgerminal cavity of recipients, similar to the virus injection system. Primordial germ cells (PGCs) can be isolated from blood vessels of HH stage 13–16 embryos and embryonic gonads of HH stage 28 embryos. The isolated PGCs can be propagated in vitro, and the genome of cultured PGC could be edited using established tools and subsequently transplanted into recipient blood vessels, ultimately forming a germline chimera. From the testes of sexually mature birds, spermatogonial stem cells (SSCs) can be isolated, edited, and implanted into recipient testes to form a germline chimera. The genome-edited germ cells formed from the germline chimera result in a genome-edited bird. The germline cells are shown in green, and cultured/genome-edited cells are shown in blue.
Figure 2. Schematic representation of various strategies for genome-edited bird production. Avian germplasm is found in the zygotic stage as RNA granules and proteins, and a number of germ cells are found in the blastoderm at EGK stage X. In this stage, genome-editing tools, such as CRISPR/Cas9, can be inserted into the subgerminal cavity of the blastoderm with a virus system. The genomes of a small number of germ cells at this stage can be modified, and germline chimera can be formed. Blastoderm cells can be cultured in vitro, and chimera can be established by injecting cultured embryonic stem cells (ESCs) into the subgerminal cavity of recipients, similar to the virus injection system. Primordial germ cells (PGCs) can be isolated from blood vessels of HH stage 13–16 embryos and embryonic gonads of HH stage 28 embryos. The isolated PGCs can be propagated in vitro, and the genome of cultured PGC could be edited using established tools and subsequently transplanted into recipient blood vessels, ultimately forming a germline chimera. From the testes of sexually mature birds, spermatogonial stem cells (SSCs) can be isolated, edited, and implanted into recipient testes to form a germline chimera. The genome-edited germ cells formed from the germline chimera result in a genome-edited bird. The germline cells are shown in green, and cultured/genome-edited cells are shown in blue.
Genes 14 00899 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kim, Y.-M.; Woo, S.-J.; Han, J.-Y. Strategies for the Generation of Gene Modified Avian Models: Advancement in Avian Germline Transmission, Genome Editing, and Applications. Genes 2023, 14, 899. https://doi.org/10.3390/genes14040899

AMA Style

Kim Y-M, Woo S-J, Han J-Y. Strategies for the Generation of Gene Modified Avian Models: Advancement in Avian Germline Transmission, Genome Editing, and Applications. Genes. 2023; 14(4):899. https://doi.org/10.3390/genes14040899

Chicago/Turabian Style

Kim, Young-Min, Seung-Je Woo, and Jae-Yong Han. 2023. "Strategies for the Generation of Gene Modified Avian Models: Advancement in Avian Germline Transmission, Genome Editing, and Applications" Genes 14, no. 4: 899. https://doi.org/10.3390/genes14040899

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