Open access peer-reviewed chapter

Transgenesis and Genome Editing in Poultry

Written By

Young Sun Hwang and Jae Yong Han

Submitted: 20 November 2017 Reviewed: 03 April 2018 Published: 05 September 2018

DOI: 10.5772/intechopen.76961

From the Edited Volume

Application of Genetics and Genomics in Poultry Science

Edited by Xiaojun Liu

Chapter metrics overview

1,837 Chapter Downloads

View Full Metrics

Abstract

The transgenic approach and precise editing of specific loci in the genome have diverse practical uses in animal biotechnology. Recent advances in genome-editing technology, including clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated protein 9 (Cas9) have helped to generate highly valuable and quality-improved poultry. The production of transgenic and genome-edited birds mainly depends on primordial germ cells (PGCs), which are the progenitor cells of gametes, due to the unique system that is quite different from the mammalian system. This chapter introduces the basic physiology of avian PGCs and the latest PGC-mediated methodologies in transgenesis and genome editing of birds. Based on these techniques, future applications of precisely genome-modulated poultry are discussed to provide opportunities and benefits for humans.

Keywords

  • avian primordial germ cells
  • CRISPR/Cas9
  • genome editing
  • poultry
  • transgenesis

1. Introduction

The ability to genetically modify and precisely edit the genomes of animals has revolutionized various fields in which the genotypes, phenotypes, and traits of animals can be easily modified. Traditional animal breeding has been dependent on selective or artificial breeding for improvements in productivity, food quality, and other economical traits of the offspring [1]. However, transgenic and precise genome-editing tools facilitate improvements in genetic traits of animals when combined with conventional breeding systems. Recent technological progress of programmable nucleases, particularly the clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated protein 9 (Cas9) system, has enabled much higher frequencies of homologous recombination events and targeted mutagenesis through a highly efficient generation of double-stranded DNA breaks (DSBs) in specific regions and genetic modifications at targeted loci in the genome [2, 3]. This system has been adopted for programmable genome editing in various organisms, including humans [4]. More recently, programmable one-base pair conversion into another without DNA cleavage has been reported [5, 6]. This novel base editing system overcomes the low efficiency of correcting a point mutation using Cas9-mediated genome editing by delivering a homology-directed repair (HDR) donor template.

Germ-line modification is essential for the application of transgenic and genome-editing technologies in animals and to transmit modified and improved genetic traits from generation to generation. Germ-line modification methods differ between mammals and birds. The first transgenic mouse was generated by microinjecting the target DNA into the pro-nucleus of a fertilized embryo [7]. Livestock, including rabbits, sheep, and pigs, have been genetically modified using this technique [8]. This classical strategy is still widely used in animal transgenesis despite several disadvantages, such as the low efficiency of producing founder animals and the random integration of foreign DNA. Another popular method is the use of embryonic stem cells (ESCs) to modify the germ line, especially in mice (Figure 1a). Following genetic modification of ESCs in vitro, the cells are injected into the recipient blastocyst. Then, germ-line chimeras composed of germ cells originating from both endogenous and exogenous sources are produced to generate transgenic offspring derived from genetically modified ESCs [9, 10].

Figure 1.

The different transgenic and genome-editing system between mammalian and avian species. (a) In mammals, transgenic and genome-edited systems are based on direct introduction of genome-editing tool into the zygote or microinjection of genome-edited ESCs into the recipient blastocyst. (b) In aves, those systems can be applied via injection of genome-edited primordial germ cells (PGCs) into the blood vessel of recipient. This figure is adopted from [135].

Unlike mammals, a unique system is used for transgenesis and genetic modification in avian species (Figure 1b) due to their oviparity and the physiological properties of the ovum [11]. As the avian zygote is surrounded by a large amount of yolk and a small germinal disc, introducing foreign DNA or microinjecting avian ESCs into the zygote is quite difficult [12, 13, 14]. Alternatively, the first transgenic chicken was produced via retroviral injection into the subgerminal cavity of Eyal-Giladi and Kochav (EGK) [15] stage X embryos [16]. Then, various strategies, including viral infection of stage X embryos [17, 18, 19], microinjection of transgenes into fertilized eggs [14, 19], and ESCs [20], have been applied to produce genetically modified transgenic birds. However, because of the low efficiency of germ-line transmission, these strategies have not been appropriate for the production of genome-modified birds until recently. As an alternative cell source comparable to ESCs in mammals, primordial germ cells (PGCs) in avian species have overcome this limitation [21]. Here we present an overview of PGC physiology, recent advances in transgenesis and genome editing, and potential strategies for programmable genetic modulation in poultry.

Advertisement

2. Physiological overview of avian PGCs

2.1. Early development of avian PGCs

Since the first examination of the origin of PGCs in chicken germinal epithelium [22], chickens have been used as a valuable germ cell model (Figure 2). In initial studies on the origin of avian PGCs, only the central region of the blastoderm was considered to give rise to PGCs [23, 24], until the discovery and tracing of the chicken VASA homolog (CVH) in 2000 [25]. CVH is used as a PGC marker during the early developmental stages in chickens. CVH mRNA and protein expression can be consistently detected during early embryogenesis, from functional oocyte to fertilized embryo. The CVH protein was observed in granulofibrillar structures surrounding the mitochondrial cloud and the spectrin protein-enriched structure of oocytes, suggesting a CVH-containing structure in the germplasm of chickens. During early cleavage, CVH was found in cleavage furrows and restricted to about 6–8 cells at the 300-cell stage. According to these observations, the specification of germ cells in chickens seems to follow a pre-determined model from maternally inherited material. More recently, another germ cell marker detected in various species, deleted in azoospermia-like (DAZL), was identified in pre-PGCs of chicken embryos during intrauterine development prior to oviposition [26]. Using DAZL as an early germplasm marker, the germ granule was determined to be asymmetrically localized in oocytes, with a shift to a diffused form during early cleavage when the zygotic genome is activated. Moreover, knockdown of DAZL expression in chicken PGCs affects germ-cell integrity, such as proliferation, gene expression, and apoptosis. These findings further demonstrate that the origin of PGCs in birds is mediated by maternally inherited determinants, which is required to examine specific functions of germplasm components and to clarify the mechanisms.

Figure 2.

Schematic representation of the development and migration of PGCs in chicken. Chicken PGCs (cPGCs) are dispersed at stage X and move to the germinal crescent at HH stage 4. They then undergo circulation via extra-embryonic blood vessels until settlement in embryonic gonads at HH stage 17. This figure is adopted from [135].

In vertebrates, germ cells arise in a specific region of the embryo and then migrate to the genital ridges during early development [27]. Avian PGCs are clustered and derived from the epiblast layer [28, 29]. Then, the PGCs migrate toward the germinal crescent region at Hamburger and Hamilton (HH) stage 4 [30, 31, 32]. The PGCs are located in this extraembryonic region from HH stages 4–10 during formation of the primitive streak [15, 31, 33]. Using chicken fibroblast cell-line DF-1 and PGC transplantation into the embryos, PGCs are shown to passively reach the anterior region but, later, are actively incorporated into the germinal crescent compared to DF-1 [32]. However, the detailed mechanism of active migration, which may be guided by attractive and repulsive cues, remains poorly understood. Several studies have shown that migrating and in vitro-cultured PGCs generate pseudopodia, suggesting germ-cell migration in birds occurs via amoeboid movement [34, 35, 36]. Subsequently, PGCs enter the embryonic blood vessels through the anterior vitelline vein during HH stages 10–12 [37, 38], contrary to mammalian PGC migration from the hindgut endoderm to the mesentery [27, 39]. The PGCs enter blood vessels and are most abundant at HH stage 12 [35, 40]. PGCs circulating in the embryonic bloodstream start to settle in the genital ridge and invade the thickened coelomic epithelium during HH stages 15–18 [41, 42]. Research suggests that coelomic epithelium releases a chemical cue to attract PGCs to the gonads [43]. Later, the main factors guiding chicken PGCs to the genital ridges are chemokine stromal cell-derived factor 1 (SDF1) and its receptor, C-X-C motif chemokine receptor 4 (CXCR4) expressed by migrating PGCs [44]. One study revealed that the anterior vitelline vein plays a key role directly accumulating migrating PGCs, which reach the genital ridges during circulation [37]. Thus, compared with mammalian PGCs, the unique migratory pathway of avian PGCs through the bloodstream enables the generation of germ-line chimeras or genetically modified birds via an exogenous injection of PGCs into the blood vessel of a recipient embryo.

2.2. Germ-line chimera production via PGCs

Due to difficulties in the application of the mammalian system for highly efficient production of transgenic birds, many researchers have focused on improving the efficiency of germ-line transmission. In 1976, the colonization of germinal crescent-derived donor turkey PGCs was examined in recipient chicken gonads following intravascular injection, and a germ-line chimera chicken was produced from functional gametes derived from turkey PGCs [45]. In addition, PGCs from the germinal crescent have been successfully transplanted into recipient embryos to produce germ-line chimeras in quail [46]. Germ-line chimeras and donor-derived progeny in birds have been generated by transplanted blood PGCs from the HH stage 14−16 embryos [47, 48] and gonadal PGCs (gPGCs) of HH stage 26−28 embryos [49, 50] in chickens and quail. Germ-line chimeras using cryopreserved PGCs or interspecific germ-line chimera enables the preservation of avian genetic resources and restoration of endangered bird species [51, 52]. Furthermore, endogenous PGCs are depleted in recipient embryos to improve the efficiency of germ-line chimera production. Various approaches, such as exposure to gamma rays [53], administration of busulfan to embryos [54], and removal of blood from recipient embryos at HH stages 14–15 [55], have been used to eliminate endogenous germ cells in birds. One report showed that the germ-line chimera efficiency of a busulfan-treated founder was approximately 99% [56]. These efforts have promoted the development of transgenic and genome-edited birds.

Many attempts have been made to develop an alternative system for producing germ-line chimera using other germ-line competent cells, such as blastodermal cells [57], embryonic germ cells [58], germ-line stem cells, and spermatogonial stem cells [59]; however, the efficiency of using these cells is lower than that of the PGC-mediated method. In addition to efficient germ-line chimera production, a PGC culture system, which has been optimized and proven to maintain germ-line competency after expansion in vitro [36, 60, 61, 62] despite differences in efficiency, provides many advantages of the use of PGCs in terms of transgenesis and precise genome modulation in birds. Although there are challenges to overcome, including the relationship between the in vitro culture of PGCs and germ-line competency and the absence of germ-line competency-associated markers, the PGC-mediated germ-line transmission system is the most efficient method to establish transgenic and genome-edited birds, until now.

Advertisement

3. Transgenesis and programmable genome editing in poultry

3.1. Primordial germ-cell isolation and in vitro culture in birds

Avian PGCs are generally obtained from three different stages, such as the germinal crescent in HH stage 4–8 embryos, embryonic blood in HH stage 14–16 embryos, and gonads in HH 26–28 embryos. Before the PGC cell-surface antigens were identified, PGCs were isolated using a density gradient-dependent centrifugation method [63, 64]. However, this method was limited due to low yield rates, purity, and viability of isolated PGCs. After the discovery of PGC-specific surface markers such as chicken stage-specific embryonic antigen-1 (SSEA-1) and the quail germ-cell-specific marker, QCR1, magnetic-activated cell sorting or fluorescence-activated cell sorting systems were used to isolate highly pure avian PGCs [65, 66, 67]. Nevertheless, isolating PGCs from wild or endangered birds, in which PGC-specific markers have not yet been fully determined, is difficult using such cell-sorting systems. To overcome this problem, a Transwell-mediated size-dependent isolation method was recently developed in various avian PGCs from HH stage 14–16 embryonic blood based on the larger size of PGCs compared to whole blood cells at that stage [68].

Since in vitro culture of PGCs without loss of germ-line competency was successfully established in 2006 [60], many studies have focused on optimizing PGC culture systems and revealing the detailed signaling mechanisms related to the proliferation and maintenance of germ-line competency in vitro. For example, basic fibroblast growth factor is essential for in vitro proliferation and survival through the MEK/ERK signaling pathway in chicken PGCs [36, 61]. Furthermore, in vitro self-renewal of chicken PGCs requires complex pathways composed of MEK1, AKT, and SMAD3 signaling to maintain germ-line competency [69], and Wnt/β-catenin signaling is also required for the proliferation of PGCs in vitro [70]. This system could be developed for various avian species and would be useful to apply to PGC-mediated avian transgenesis and genome editing.

3.2. PGC-mediated transgenesis in birds

Before the establishment of in vitro PGC culture systems, the major method for transgenesis in birds relied on injecting viruses into EGK stage X embryos. The first transgenic chicken was generated by microinjecting recombinant avian leukosis virus into the subgerminal cavity of EGK stage X embryos [16]. In addition, transgenic quail were produced using direct injection of a replication-defective retrovirus into the embryonic blastoderm [71]. Due to frequent silencing of the transgene, which is randomly integrated in the genome of the transgenic animal [17, 71, 72, 73], the lentivirus-mediated method has been recognized as the most efficient viral transduction system for avian transgenesis. This system successfully produces diverse transgenic chickens without silencing gene expression [74, 75, 76, 77]. In the case of zebra finch, microinjecting lentivirus into blastodermal stage embryos generated the first transgenic finch expressing green fluorescent protein (GFP) [78]. The transgenic birds, using PGCs from the germinal crescent of HH stage 5 chicken embryos, was firstly produced [79]. Furthermore, gPGC-mediated transgenesis in quail was successfully established via the lentiviral system [80]. In that study, although the efficiency of the gPGC-mediated method was comparable to the blastoderm-mediated method in quail, the production of transgenic birds through viral infection into non-cultivated PGCs after purification has been demonstrated.

Much effort has also been made to develop non-viral transgenic systems without PGCs, such as sperm-mediated gene transfection [81, 82] and direct microinjection of transgenes into fertilized eggs [14]. However, these approaches appear to have low germ-line chimerism and transmission efficiency compared with that of the PGC-mediated method. Furthermore, because of long-term in vitro PGC culture systems, it is possible to develop more optimal approaches to produce genetically modified birds compared with other germ-line cells. Combined with a culture system, the highly efficient non-viral transposable systems, such as piggyBac and Tol2, have been developed for stable transgene integration into the genome of chicken PGCs [83, 84]. The transgenic efficiency in cultured PGCs using lipofection or electroporation is remarkably higher than that of the virus-mediated methods to produce transgenic chickens. Moreover, using site-specific gene cassette exchange in transgenic chicken genomes via PGCs with the flippase recombinase system was introduced [85]. Alternatively, transgenic birds have been produced by direct transfection into circulating PGCs at HH stages 14–16 [86, 87, 88], although transgenic efficiency is usually lower than that of the cultured PGC-mediated method. This approach could be applied to establish a transgenesis system in avian species, as PGCs are difficult to manipulate in vitro in birds.

3.3. Precise genome-editing tools

Programmable DNA nucleases, such as ZFN and TALEN, have made genome editing in the target region possible over the last decade. Briefly, ZFN is the first programmable genome-editing tool. ZFN is a fusion protein with zinc finger proteins that bind to specific DNA and nuclease domains, such as the FokI endonuclease [89]. The second generation of programmable genome-editing tools is TALEN, in which the TAL effector of a DNA-binding domain derived from Xanthomonas is fused with a DNA cleavage nuclease domain [90]. Generally, ZFN and TALEN specifically recognize target sequences, resulting in the generation of DSBs to enable efficient gene targeting in specific genomic loci compared with natural homologous recombination, although the DNA binding mechanism is different between the systems. Compared to ZFN, TALEN is a more flexible tool for editing genome sequences in the target site because the TAL effector contains one repeat domain that binds to one nucleotide each [91]. Furthermore, customized TALEN can be easily synthesized using an assembly kit for precise genetic modifications [92, 93, 94].

The CRISPR/Cas9 system is considered the most revolutionary tool and has been developed to carry out highly efficient and specific genome editing as a simple RNA-guided platform. This system is derived from prokaryotic DNA, which is involved in acquired resistance against exogenous plasmid DNA and phages. These bacteria possess clustered repeats called CRISPRs that bind to the viral RNA to disrupt it with the Cas9 protein to defend [95]. CRISPR/Cas9 also leads to DNA cleavage at a specifically recognized target site, resulting in the generation of DSBs, similar to ZFN and TALEN. However, unlike ZFN and TALEN, which require paired units to induce DSBs at the target region, the CRISPR/Cas9 system, a type-II CRISPR system, includes the Cas protein, CRISPR RNA (crRNA), trans-activating CRISPR RNA (tracrRNA), and a protospacer adjacent motif (PAM) sequence for targeted genome editing [2] (Figure 3a). The PAM sequence is in the upstream region of the crRNA-binding site and guides the Cas9 protein to the target site. The target DNA sequence is specifically recognized by the CRISPR/Cas9 complex through base pairing with a guide RNA and subsequently induces DSBs at the targeted genomic loci. Then, these DSBs activate the cell’s DNA repair system, which includes random indels at the site of DNA cleavage via non-homologous end joining or replacement of a homologous DNA template in the DNA surrounding the cleavage site via HDR. Compared to ZFN and TALEN, which rely on DNA-binding specificity and were developed through expensive and time-consuming processes, the CRISPR/Cas9 system is convenient for targeted genome editing because it is extremely easy to synthesize crRNA and tracrRNA and it is easy to construct thousands of customized CRISPR/Cas9 systems depending on the targeted genes. Furthermore, the CRISPR/Cas9 system is simple and practically easy to use with robust cutting activity, leading to a fast and cost-effective system for modifying the genomes of various organisms [96, 97]. Among the ZFN, TALEN, and CRISPR/Cas9 systems, CRISPR/Cas9 is now the most powerful method to precisely edit in a targeted manner and has been applied in diverse organisms, including animals, plants, and humans [4, 98, 99].

Figure 3.

Principles of clustered regularly interspaced short palindromic repeats (CRISPR)-CRISPR-associated protein 9 (Cas9) and development of the base editing system. (a) CRISPR/Cas9 system enhances the DNA mutation through the creation of a double-stranded DNA breaks (DSBs) at a specific locus in the genome and generates highly efficient genetic modification in a targeted manner. (b) A single base editing system is generated by fusion of cytidine deaminase to the catalytically inactivated Cas9 (dead Cas9) and provides a valuable tool for precise genome editing with regard to highly targeted single-base changes. This figure is modified from [136].

More recently, a programmable base editing system leading to precise and efficient nucleotide conversion was developed and applied to various species to minimize DNA damage and acquire point-mutation corrections without HDR donors during genome editing. The strategies for modifying a single base pair by CRISPR/Cas9 are difficult to apply for various purposes, although efficiency is 60% in cultured cells [100]. However, the base editing system is composed of a group of cytidine deaminases, including the apolipoprotein B mRNA editing enzyme (APOBEC) 1–4 or activation-induced deaminase, resulting in deamination of cytidine to uridine [101] (Figure 3b). The deaminases fuse with CRISPR/Cas9 substitute C in a target site with T (or G to A) without breaking the DNA [5, 102, 103]. A base editing system is advantageous because indel formation rates are <0.1% [5]. Until now, the improved base editing systems have been advanced up to four generations. The first-generation base editor (BE1) involves the rat APOBEC1 fused with the N-terminus of catalytically dead Cas9 by a 16-residue peptide of the XTEN linker [5]. BE1 converts C to U with an activity window of approximately five nucleotides. However, base excision repair removing U from DNA decreases intracellular efficiency. To increase the low-editing efficiency of BE1, BE2, which fuses uracil glycosylase inhibitor (UGI) to the C-terminus of BE1, was created. In human cells, BE2 increases the editing efficiency threefold compared with BE1. Moreover, BE3, which involves Cas9 nickase (A840H)-UGI-generated nicks in an unmodified DNA strand, results in 37% efficiency, which is from three to sixfold to that of BE2. Subsequently, BE4 was developed to increase efficiency to 50% compared with BE3 and decrease undesired products [104]. Moreover, the Mu protein Gam, which binds DSBs and protects their ends from degradation during base editing, was fused to the N-terminus of BE4 resulting in a reduction of indel frequency [104]. In addition, A∙T to G∙C conversion was recently developed to broaden the application of the base editing system [105]. This simple system for base conversion has been applied to precisely modify the human and mice genomes [106]. In addition, base editing has been successfully applied in various plants [5, 6, 99, 107, 108, 109, 110, 111, 112, 113]. In the near future, the ability to modify single-base changes in the base editing system will be widely used for precise genome editing and specialized purposes by substituting amino acids.

3.4. The recent progress of genome-edited poultry

In birds, combining an in vitro culture system for PGCs and an efficient genome-editing system can produce programmable genome-edited poultry, especially chickens. Although total germ-line transmission efficiency from targeted PGCs is approximately 0.1% due to natural homologous recombination that occurs with very low frequency, the immunoglobulin gene knockout chicken was first produced via the PGC-mediated method in 2013 [114]. However, applying TALEN technology to in vitro-cultured PGCs improved germ-line transmission efficiency of mutant progeny to 8% of the donor-derived knockout chicks in the ovalbumin locus in 2014 [115]. This case is the first programmed DNA nuclease-mediated knockout chicken, and the TALEN-mediated gene knockout appeared to be much more efficient than the conventional homologous recombination-mediated system. Later, the CRISPR/Cas9 system was used to efficiently generate ovomucoid gene-targeted chickens by transferring transiently drug-selected PGCs into recipient embryos using gamma-ray irradiation to deplete endogenous PGCs [116]. Here, the G0 founders had 93% mutant sperm and produced 53% ovomucoid gene mutant offspring, indicating a highly efficient CRISPR/Cas9 system in birds. Furthermore, through HDR insertion of an additional loxP site into the loxP variable region segment previously inserted into the joining gene segment of the chicken immunoglobulin heavy chain (IgH) locus and Cre recombination, a 28-kb genomic DNA sequence at the IgH locus was deleted in CRISPR/Cas9-mediated genome-modified chickens [114, 117]. More recently, CVH gene-targeted chickens via the TALEN-mediated HDR system were produced using 2-week-recovered PGCs with GFP transgene knockin at the CVH locus with 8.1% efficiency [118]. They generated 6% CVH-targeted progeny from one G0 male founder showing 10% of genomic equivalents in its semen. Germ-line transmission efficiency varies among genome-edited PGC lines compared with TALEN- and CRISPR-mediated genome modification. Because of possible loss of germ-line competency during long-term in vitro culture and genetic modification, it is crucial to optimize the conditions for establishing stable PGC lines during genome editing. On the other hand, a recent method, called sperm transfection-assisted gene-editing based on direct delivery of the CRISPR/Cas9 complex, is a potential alternative for avian transgenesis and genome editing without culturing PGCs, despite the low efficiency of genome editing and germ-line transmission [119].

3.5. Further applications of genome editing in poultry

After completion of the chicken genome sequencing project in 2004 and the subsequently available genomic sequences of the zebra finch and turkey, infinite possibilities and multiple opportunities are available to access invaluable genetic information from birds [120]. The bird 10 K genome sequencing project was initiated in 2015 based on recent next-generation sequencing technology. The progress of efficient genome-editing technologies in birds synergizes the value of avian genetic information by avian genome manipulation for the development of beneficial poultry breeds (Figure 4).

Figure 4.

Strategies for the production of genome-edited poultries. PGCs in poultry can be obtained from embryonic blood and embryonic gonads. After the delivery of genome editing tools, genome-edited poultry can be established by microinjection of directly isolated or in vitro cultured PGCs into the blood vessels of recipient embryos. Avian genome editing systems can be applied to produce various avian models and poultry. This figure is adopted from [135].

We expect to establish an efficient bioreactor system to produce valuable proteins through genome-editing technology in chickens, which has the well-known advantage that egg white protein is easy to purify and eggs are produced daily by chickens [11, 121]. Bioreactors producing target proteins under the strong ovalbumin promoter have interested researchers for a long time [77, 122]. Thus, HDR-mediated target gene insertion into the ovalbumin locus could be an ideal bioreactor system to cost-effectively produce more than 1 g of target protein from an egg. In addition, genome editing has been used to remove or enhance targeted nutrients in meat and eggs of chickens. Knocking out allergen-related genes, such as ovalbumin and ovomucoid, has been achieved and could be used to generate allergen-free chicken meat and eggs [115, 116]. Additionally, muscle-related genes, such as myostatin, could be used to generate double-muscled and muscle hypertrophied chickens via genome editing, as in other livestock [123, 124, 125]. The conventional genetically modified organism (GMO) containing a foreign gene has been a concern due to a safety issue from unknown allergic reactions or the use of antibiotic resistance genes. Genome-edited poultry can be produced with a controllable genome-editing system, which is similar to natural mutations rather than foreign gene insertion as in conventional GMO. Moreover, an advanced base editing system may be more suitable for slight modifications of nucleotides without HDR in some cases. After scientists convince the public that genome-edited animals are similar to naturally selected animals, genome-edited poultry will be profitable for consumers.

In addition, it will be possible to control avian-specific diseases, such as avian influenza and Marek’s disease, which cause serious problems in the poultry industry. Although understanding the mechanism of avian virus pathogenesis is essential for the application [126, 127] and limited in vitro results have been achieved [98, 128], avian genome-editing technology is expected to be used to develop avian disease-resistant birds by eliminating host factors or receptors of avian viruses. Lastly, because birds lay a large number of eggs and have a short ovulation cycle, they are considered the best model organism for studying human ovarian cancer [129]. With precise genome modulation of ovarian cancer-related genes in an avian model, especially chickens, it is possible to reveal the genetic mechanisms of ovarian cancer. In addition, avian genome-editing tools will gradually be applied to other birds, such as zebra finch, which is an exclusive non-human model organism for investigating the biological basis of speech learning and neurobehavioral research and disease [130, 131, 132, 133, 134]. Until now, direct injection of virus-mediated transgenesis into embryos has been used in zebra finch [132]. Genome-editing technology delivered by both in vivo and in vitro strategies will be widely applied to reveal the function and mechanism of neuron-related genes in zebra finch.

Advertisement

4. Conclusions

Poultry is important not only as a food resource but also as a valuable model animal for diverse disciplines, such as human disease, neurological research, and developmental biology. Until a few years ago, the difficulties in transgenesis and genome editing of birds limited their use as model animals. State-of-the-art technologies, such as CRISPR/Cas9 and the base editing system, have provided new insights into avian models when combined with PGC culture and other reliable germ-line systems. The novel genome-edited birds, including specific-gene knockout, human disease models, allergen-free, and disease-resistant poultry and egg-based bioreactors, are expected to be developed. Although the challenges in improving germ-line transmission strategies remain for many poultry species, programmable genome-editing tools will be useful in the development of genetically modulated poultry, together with efficient delivery and germ-line modification. Therefore, applying genome editing technology to birds will contribute to the poultry industry and ultimately provide benefits to humans.

Advertisement

Acknowledgments

This work was supported by the National Research Foundation of Korea (NRF), grant funded by the Korean government (MSIP) (No. 2015R1A3A2033826).

Advertisement

Conflict of interest

The authors declare that they have no competing interests.

References

  1. 1. Andersson L, Georges M. Domestic-animal genomics: Deciphering the genetics of complex traits. Nature Reviews Genetics. 2004;5(3):202-212
  2. 2. Jinek M et al. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science. 2012;337(6096):816-821
  3. 3. Kim JS. Genome editing comes of age. Nature Protocols. 2016;11(9):1573-1578
  4. 4. Ma H et al. Correction of a pathogenic gene mutation in human embryos. Nature. 2017;548(7668):413-419
  5. 5. Komor AC et al. Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage. Nature. 2016;533(7603):420-424
  6. 6. Nishida K et al. Targeted nucleotide editing using hybrid prokaryotic and vertebrate adaptive immune systems. Science. 2016;353(6305)
  7. 7. Gordon JW et al. Genetic transformation of mouse embryos by microinjection of purified DNA. Proceedings of the National Academy of Sciences of the United States of America. 1980;77(12):7380-7384
  8. 8. Hammer RE et al. Production of transgenic rabbits, sheep and pigs by microinjection. Nature. 1985;315(6021):680-683
  9. 9. Thomas KR, Capecchi MR. Site-directed mutagenesis by gene targeting in mouse embryo-derived stem-cells. Cell. 1987;51(3):503-512
  10. 10. Smithies O et al. Insertion of DNA sequences into the human chromosomal beta-globin locus by homologous recombination. Nature. 1985;317(6034):230-234
  11. 11. Han JY. Germ cells and transgenesis in chickens. Comparative Immunology, Microbiology and Infectious Diseases. 2009;32(2):61-80
  12. 12. Lee HC et al. Cleavage events and sperm dynamics in chick intrauterine embryos. PLoS One. 2013;8(11):e80631
  13. 13. Bellairs R, Lorenz FW, Dunlap T. Cleavage in the chick embryo. Journal of Embryology and Experimental Morphology. 1978;43:55-69
  14. 14. Love J et al. Transgenic birds by DNA microinjection. Bio-Technology (NY). 1994;12(1):60-63
  15. 15. Eyalgiladi H, Kochav S. From cleavage to primitive streak formation—Complementary normal table and a new look at 1st stages of development of chick. 1. General morphology. Developmental Biology. 1976;49(2):321-337
  16. 16. Salter DW et al. Gene insertion into the chicken germ line by retroviruses. Poultry Science. 1986;65(8):1445-1458
  17. 17. Bosselman RA et al. Germline transmission of exogenous genes in the chicken. Science. 1989;243(4890):533-535
  18. 18. Thoraval P et al. Germline transmission of exogenous genes in chickens using helper-free ecotropic avian-leukosis virus-based vectors. Transgenic Research. 1995;4(6):369-377
  19. 19. Sherman A. Transposition of the Drosophila element mariner into the chicken germ line (vol 16, pg 1050, 1998). Nature Biotechnology. 1999;17(1):81-81
  20. 20. Zhu L et al. Production of human monoclonal antibody in eggs of chimeric chickens. Nature Biotechnology. 2005;23(9):1159-1169
  21. 21. Capecchi MR. Gene targeting in mice: Functional analysis of the mammalian genome for the twenty-first century. Nature Reviews Genetics. 2005;6(6):507-512
  22. 22. Waldeyer W. Eirstock und Ei. Eine Beitrag zur Anatomie und Entwicklungsgeschichte der Sexualorgane. Leipzig; 1870
  23. 23. 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(2):209-219
  24. 24. Ginsburg M, Hochman J, Eyal-Giladi H. Immunohistochemical analysis of the segregation process of the quail germ cell lineage. The International Journal of Developmental Biology. 1989;33(3):389-395
  25. 25. Tsunekawa N et al. Isolation of chicken vasa homolog gene and tracing the origin of primordial germ cells. Development. 2000;127(12):2741-2750
  26. 26. Lee HC et al. DAZL expression explains origin and central formation of primordial germ cells in chickens. Stem Cells and Development. 2016;25(1):68-79
  27. 27. Richardson BE, Lehmann R. Mechanisms guiding primordial germ cell migration: Strategies from different organisms. Nature Reviews Molecular Cell Biology. 2010;11(1):37-49
  28. 28. 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. Developmental Biology. 1976;49(2):321-337
  29. 29. Eyal-Giladi H, Ginsburg M, Farbarov A. Avian primordial germ cells are of epiblastic origin. Journal of Embryology and Experimental Morphology. 1981;65:139-147
  30. 30. Tagami T, Kagami H. Developmental origin of avian primordial germ cells and its unique differentiation in the gonads of mixed-sex chimeras. Molecular Reproduction and Development. 1998;50(3):370-376
  31. 31. Ginsburg M, Eyalgiladi H. Temporal and spatial-aspects of the gradual migration of primordial germ-cells from the epiblast into the germinal crescent in the avian embryo. Journal of Embryology and Experimental Morphology. 1986;95:53-71
  32. 32. Kang KS et al. Spatial and temporal action of chicken primordial germ cells during initial migration. Reproduction. 2015;149(2):179-187
  33. 33. Hamburger V, Hamilton HL. A series of normal stages in the development of the chick embryo. Journal of Morphology. 1951;88(1):49-92
  34. 34. Kuwana T et al. Behavior of chick primordial germ cells moving toward gonadal primordium in vitro: Scanning electron microscopic study. The Anatomical Record. 1987;219(2):164-170
  35. 35. Fujimoto T, Ukeshima A, Kiyofuji R. The origin, migration and morphology of the primordial germ cells in the chick embryo. The Anatomical Record. 1976;185(2):139-145
  36. 36. Choi JW et al. Basic fibroblast growth factor activates MEK/ERK cell signaling pathway and stimulates the proliferation of chicken primordial germ cells. PLoS One. 2010;5(9):e12968
  37. 37. De Bernardo A et al. Chicken primordial germ cells use the anterior vitelline veins to enter the embryonic circulation. Biology Open. 2012;1(11):1146-1152
  38. 38. Niewkoop P, Sutasurya L. Primordial Germ Cells in the Chordates. Cambridge: Cambridge University; 1979. pp. 118-123
  39. 39. Saitou M, Yamaji M. Primordial germ cells in mice. Cold Spring Harbor Perspectives in Biology. 2012;4(11):a008375
  40. 40. Lee H, Karasanyi N, Nagele RG Jr. The role of the cell surface in the migration of primordial germ cells in early chick embryos: Effects of concanavalin A. Journal of Embryology and Experimental Morphology. 1978;46:5-20
  41. 41. Nakamura Y et al. Migration and proliferation of primordial germ cells in the early chicken embryo. Poultry Science. 2007;86(10):2182-2193
  42. 42. Ukeshima A, Kudo M, Fujimoto T. Relationship between genital ridge formation and settlement site of primordial germ cells in chick embryos. The Anatomical Record. 1987;219(3):311-314
  43. 43. Kuwana T, Maeda-Suga H, Fujimoto T. Attraction of chick primordial germ cells by gonadal anlage in vitro. The Anatomical Record. 1986;215(4):403-406
  44. 44. Stebler J et al. Primordial germ cell migration in the chick and mouse embryo: The role of the chemokine SDF-1/CXCL12. Developmental Biology. 2004;272(2):351-361
  45. 45. Reynaud G. Reproductive capacity and offspring of chickens submitted to a transfer of primordial germ cells during embryonic life. Wilehm Roux’s Archives of Developmental Biology. 1976;179(2):85-110
  46. 46. Wentworth BC et al. Manipulation of avian primordial germ cells and gonadal differentiation. Poultry Science. 1989;68(7):999-1010
  47. 47. Tajima A et al. Production of germ-line chimera by transfer of primordial germ-cells in the domestic chicken (Gallus-Domesticus). Theriogenology. 1993;40(3):509-519
  48. 48. Ono T, Matsumoto T, Arisawa Y. Production of donor-derived offspring by transfer of primordial germ cells in Japanese quail. Experimental Animals. 1998;47(4):215-219
  49. 49. Chang IK et al. Production of germline chimeric chickens by transfer of cultured primordial germ cells. Cell Biology International. 1997;21(8):495-499
  50. 50. Kim MA et al. Production of quail (Coturnix japonica) germline chimeras by transfer of gonadal primordial germ cells into recipient embryos. Theriogenology. 2005;63(3):774-782
  51. 51. Kang SJ et al. Reproduction of wild birds via interspecies germ cell transplantation. Biology of Reproduction. 2008;79(5):931-937
  52. 52. Wernery U et al. Primordial germ cell-mediated chimera technology produces viable pure-line Houbara bustard offspring: Potential for repopulating an endangered species. PLoS One. 2010;5(12):e15824
  53. 53. Carsience RS et al. Germline chimeric chickens from dispersed donor blastodermal cells and compromised recipient embryos. Development. 1993;117(2):669-675
  54. 54. Aige-Gil V, Simkiss K. Sterilisation of avian embryos with busulphan. Research in Veterinary Science. 1991;50(2):139-144
  55. 55. Naito M et al. Preservation of chick primordial germ cells in liquid nitrogen and subsequent production of viable offspring. Journal of Reproduction and Fertility. 1994;102(2):321-325
  56. 56. Nakamura Y et al. Germline replacement by transfer of primordial germ cells into partially sterilized embryos in the chicken. Biology of Reproduction. 2010;83(1):130-137
  57. 57. Petitte JN et al. Production of somatic and germline chimeras in the chicken by transfer of early blastodermal cells. Development. 1990;108(1):185-189
  58. 58. Park TS et al. Birth of germline chimeras by transfer of chicken embryonic germ (EG) cells into recipient embryos. Molecular Reproduction and Development. 2003;65(4):389-395
  59. 59. Jung JG et al. The reversible developmental unipotency of germ cells in chicken. Reproduction. 2010;139(1):113-119
  60. 60. van de Lavoir MC et al. Germline transmission of genetically modified primordial germ cells. Nature, 2006. 441(7094):766-769
  61. 61. Macdonald J et al. Characterisation and germline transmission of cultured avian primordial germ cells. PLoS One. 2010;5(11):e15518
  62. 62. Song Y et al. Characteristics of long-term cultures of avian primordial germ cells and gonocytes. Biology of Reproduction. 2014;90(1):15
  63. 63. Yasuda Y et al. A method to obtain avian germ-line chimeras using isolated primordial germ-cells. Journal of Reproduction and Fertility. 1992;96(2):521-528
  64. 64. Zhao DF, Kuwana T. Purification of avian circulating primordial germ cells by Nycodenz density gradient centrifugation. British Poultry Science. 2003;44(1):30-35
  65. 65. Chang IK et al. Simple method for isolation of primordial germ-cell from Chick-embryos. Cell Biology International Reports. 1992;16(9):853-857
  66. 66. Ono T, Machida Y. Immunomagnetic purification of viable primordial germ cells of Japanese quail (Coturnix japonica). Comparative Biochemistry and Physiology a-Molecular and Integrative Physiology. 1999;122(2):255-259
  67. 67. Mozdziak PE et al. Isolation of chicken primordial germ cells using fluorescence-activated cell sorting. Poultry Science. 2005;84(4):594-600
  68. 68. Jung K et al. Size-dependent isolation of primordial germ cells from avian species. Molecular Reproduction and Development. 2017;9999:1-9
  69. 69. Whyte J et al. FGF, insulin, and SMAD signaling cooperate for avian primordial germ cell self-renewal. Stem Cell Reports. 2015;5(6):1171-1182
  70. 70. Lee HC, Lim S, Han JY. Wnt/beta-catenin signaling pathway activation is required for proliferation of chicken primordial germ cells in vitro. Scientific Reports. 2016;6:34510
  71. 71. Mizuarai S et al. Production of transgenic quails with high frequency of germ-line transmission using VSV-G pseudotyped retroviral vector. Biochemical and Biophysical Research Communications. 2001;286(3):456-463
  72. 72. Jahner D et al. De novo methylation and expression of retroviral genomes during mouse embryogenesis. Nature. 1982;298(5875):623-628
  73. 73. Challita PM, Kohn DB. Lack of expression from a retroviral vector after transduction of murine hematopoietic stem cells is associated with methylation in vivo. Proceedings of the National Academy of Sciences of the United States of America. 1994;91(7):2567-2571
  74. 74. McGrew MJ et al. Efficient production of germline transgenic chickens using lentiviral vectors. EMBO Reports. 2004;5(7):728-733
  75. 75. Chapman SC et al. Ubiquitous GFP expression in transgenic chickens using a lentiviral vector. Development. 2005;132(5):935-940
  76. 76. Scott BB, Lois C. Generation of tissue-specific transgenic birds with lentiviral vectors. Proceedings of the National Academy of Sciences of the United States of America. 2005;102(45):16443-16447
  77. 77. Lillico SG et al. Oviduct-specific expression of two therapeutic proteins in transgenic hens. Proceedings of the National Academy of Sciences of the United States of America. 2007;104(6):1771-1776
  78. 78. Agate R et al. Transgenic songbirds offer an opportunity to develop a genetic model for vocal learning. Proceedings of the National Academy of Sciences of the United States of America. 2009;106(42):17963-17967
  79. 79. Vick L, Li Y, Simkiss K. Transgenic birds from transformed primordial germ cells. Proceedings of the Biological Sciences. 1993;251(1332):179-182
  80. 80. Shin SS et al. Generation of transgenic quail through germ cell-mediated germline transmission. The FASEB Journal. 2008;22(7):2435-2444
  81. 81. Collares T et al. Transgene transmission in chickens by sperm-mediated gene transfer after seminal plasma removal and exogenous DNA treated with dimethylsulfoxide or N,N-dimethylacetamide. Journal of Biosciences. 2011;36(4):613-620
  82. 82. Nakanishi A, Iritani A. Gene transfer in the chicken by sperm-mediated methods. Molecular Reproduction and Development. 1993;36(2):258-261
  83. 83. Macdonald J et al. Efficient genetic modification and germ-line transmission of primordial germ cells using piggyBac and Tol2 transposons. Proceedings of the National Academy of Sciences of the United States of America. 2012;109(23):E1466-E1472
  84. 84. Park TS, Han JY. PiggyBac transposition into primordial germ cells is an efficient tool for transgenesis in chickens. Proceedings of the National Academy of Sciences of the United States of America. 2012;109(24):9337-9341
  85. 85. Lee HJ et al. Site-specific recombination in the chicken genome using Flipase recombinase-mediated cassette exchange. The FASEB Journal. 2016;30(2):555-563
  86. 86. Tyack SG et al. A new method for producing transgenic birds via direct in vivo transfection of primordial germ cells. Transgenic Research. 2013;22(6):1257-1264
  87. 87. Lambeth LS et al. Transgenic chickens overexpressing aromatase have high estrogen levels but maintain a predominantly male phenotype. Endocrinology. 2016;157(1):83-90
  88. 88. Zhang Z et al. Transgenic quail production by microinjection of lentiviral vector into the early embryo blood vessels. PLoS One. 2012;7(12):e50817
  89. 89. Kim YG, Cha J, Chandrasegaran S. Hybrid restriction enzymes: Zinc finger fusions to Fok I cleavage domain. Proceedings of the National Academy of Sciences of the United States of America. 1996;93(3):1156-1160
  90. 90. Boch J et al. Breaking the code of DNA binding specificity of TAL-type III effectors. Science. 2009;326(5959):1509-1512
  91. 91. Miller JC et al. A TALE nuclease architecture for efficient genome editing. Nature Biotechnology. 2011;29(2):143-148
  92. 92. Engler C et al. Golden gate shuffling: A one-pot DNA shuffling method based on type IIs restriction enzymes. PLoS One. 2009;4(5):e5553
  93. 93. Cermak T et al. Efficient design and assembly of custom TALEN and other TAL effector-based constructs for DNA targeting. Nucleic Acids Research. 2011;39(12):e82
  94. 94. Sanjana NE et al. A transcription activator-like effector toolbox for genome engineering. Nature Protocols. 2012;7(1):171-192
  95. 95. Barrangou R et al. CRISPR provides acquired resistance against viruses in prokaryotes. Science. 2007;315(5819):1709-1712
  96. 96. Hwang WY et al. Efficient genome editing in zebrafish using a CRISPR-Cas system. Nature Biotechnology. 2013;31(3):227-229
  97. 97. Wang H et al. One-step generation of mice carrying mutations in multiple genes by CRISPR/Cas-mediated genome engineering. Cell. 2013;153(4):910-918
  98. 98. Lee HJ et al. Precise gene editing of chicken Na+/H+ exchange type 1 (chNHE1) confers resistance to avian leukosis virus subgroup J (ALV-J). Developmental and Comparative Immunology. 2017;77:340-349
  99. 99. Shimatani Z et al. Targeted base editing in rice and tomato using a CRISPR-Cas9 cytidine deaminase fusion. Nature Biotechnology. 2017;35(5):441-443
  100. 100. Richardson CD et al. Non-homologous DNA increases gene disruption efficiency by altering DNA repair outcomes. Nature Communications. 2016;7:12463
  101. 101. Conticello SG. The AID/APOBEC family of nucleic acid mutators. Genome Biology. 2008;9(6):229
  102. 102. Mali P et al. CAS9 transcriptional activators for target specificity screening and paired nickases for cooperative genome engineering. Nature Biotechnology. 2013;31(9):833-838
  103. 103. Ran FA et al. Double nicking by RNA-guided CRISPR Cas9 for enhanced genome editing specificity. Cell. 2013;154(6):1380-1389
  104. 104. Komor AC et al. Improved base excision repair inhibition and bacteriophage mu gam protein yields C:G-to-T: A base editors with higher efficiency and product purity. Science Advances. 2017;3(8):eaao4774
  105. 105. Gaudelli NM et al. Programmable base editing of A*T to G*C in genomic DNA without DNA cleavage. Nature. 2017;551(7681):464-471
  106. 106. Hess GT et al. Methods and applications of CRISPR-mediated base editing in eukaryotic genomes. Molecular Cell. 2017;68(1):26-43
  107. 107. Rees HA et al. Improving the DNA specificity and applicability of base editing through protein engineering and protein delivery. Nature Communications. 2017;8:15790
  108. 108. Kim K et al. Highly efficient RNA-guided base editing in mouse embryos. Nature Biotechnology. 2017;35(5):435-437
  109. 109. Liang P et al. Effective gene editing by high-fidelity base editor 2 in mouse zygotes. Protein & Cell. 2017;8(8):601-611
  110. 110. Chadwick AC, Wang X, Musunuru K. In vivo base editing of PCSK9 (proprotein convertase subtilisin/kexin type 9) as a therapeutic alternative to genome editing. Arteriosclerosis, Thrombosis, and Vascular Biology. 2017;37(9):1741-1747
  111. 111. Ding Q et al. Permanent alteration of PCSK9 with in vivo CRISPR-Cas9 genome editing. Circulation Research. 2014;115(5):488-492
  112. 112. Lu Y, Zhu J. Precise editing of a target base in the rice genome using a modified CRISPR/Cas9 system. Molecular Plant. 2017;10(3):523-525
  113. 113. Zong Y et al. Precise base editing in rice, wheat and maize with a Cas9-cytidine deaminase fusion. Nature Biotechnology. 2017;35(5):438-440
  114. 114. Schusser B et al. Immunoglobulin knockout chickens via efficient homologous recombination in primordial germ cells. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(50):20170-20175
  115. 115. Park TS et al. Targeted gene knockout in chickens mediated by TALENs. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(35):12716-12721
  116. 116. Oishi I et al. Targeted mutagenesis in chicken using CRISPR/Cas9 system. Scientific Reports. 2016;6:23980
  117. 117. Dimitrov L et al. Germline gene editing in chickens by efficient CRISPR-mediated homologous recombination in primordial germ cells. PLoS One. 2016;11(4):e0154303
  118. 118. Taylor L et al. Efficient TALEN-mediated gene targeting of chicken primordial germ cells. Development. 2017;144(5):928-934
  119. 119. Cooper CA et al. Generation of gene edited birds in one generation using sperm transfection assisted gene editing (STAGE). Transgenic Research. 2017;26(3):331-347
  120. 120. Zhang G et al. Genomics: Bird sequencing project takes off. Nature. 2015;522(7554):34
  121. 121. Lillico SG et al. Transgenic chickens as bioreactors for protein-based drugs. Drug Discovery Today. 2005;10(3):191-196
  122. 122. Park TS et al. Deposition of bioactive human epidermal growth factor in the egg white of transgenic hens using an oviduct-specific minisynthetic promoter. FASEB Journal. 2015;29(6):2386-2396
  123. 123. Lv Q et al. Efficient generation of myostatin gene mutated rabbit by CRISPR/Cas9. Scientific Reports. 2016;6:25029
  124. 124. Crispo M et al. Efficient generation of myostatin knock-out sheep using CRISPR/Cas9 technology and microinjection into zygotes. PLoS One. 2015;10(8):e0136690
  125. 125. Wang K et al. Efficient generation of myostatin mutations in pigs using the CRISPR/Cas9 system. Scientific Reports. 2015;5:16623
  126. 126. Biggs PM, Nair V. The long view: 40 years of Marek’s disease research and avian pathology. Avian Pathology. 2012;41(1):3-9
  127. 127. Long JS et al. Species difference in ANP32A underlies influenza A virus polymerase host restriction. Nature. 2016;529(7584):101-104
  128. 128. Lee HJ et al. Acquisition of resistance to avian leukosis virus subgroup B through mutations on tvb cysteine-rich domains in DF-1 chicken fibroblasts. Veterinary Research. 2017;48(1):48
  129. 129. Johnson PA, Giles JR. The hen as a model of ovarian cancer. Nature Reviews Cancer. 2013;13(6):432-436
  130. 130. Petkov CI, Jarvis ED. Birds, primates, and spoken language origins: Behavioral phenotypes and neurobiological substrates. Frontiers in Evolutionary Neuroscience. 2012;4:12
  131. 131. Spierings MJ, Ten Cate C. Zebra finches as a model species to understand the roots of rhythm. Frontiers in Neuroscience. 2016;10:345
  132. 132. Velho TA, Lois C. Generation of transgenic zebra finches with replication-deficient lentiviruses. Cold Spring Harbor Protocols. 2014;2014(12):1284-1289
  133. 133. Liu WC et al. Human mutant huntingtin disrupts vocal learning in transgenic songbirds. Nature Neuroscience. 2015;18(11):1617-1622
  134. 134. Abe K, Matsui S, Watanabe D. Transgenic songbirds with suppressed or enhanced activity of CREB transcription factor. Proceedings of the National Academy of Sciences of the United States of America. 2015;112(24):7599-7604
  135. 135. Han JY, Park YH. Primordial germ cell-mediated transgenesis and genome editing in birds. Journal of Animal Science and Biotechnology. 2018;9(1):19
  136. 136. Lee BR, et al. Recent progress toward precise genome editing in animals. Journal of Animal Breeding and Genomics. 2017;1(2):85-101

Written By

Young Sun Hwang and Jae Yong Han

Submitted: 20 November 2017 Reviewed: 03 April 2018 Published: 05 September 2018