Next Article in Journal
Peptidomic Approaches and Observations in Neurodegenerative Diseases
Next Article in Special Issue
A Pan-RNase Inhibitor Enabling CRISPR-mRNA Platforms for Engineering of Primary Human Monocytes
Previous Article in Journal
Controlling Gas Generation of Li-Ion Battery through Divinyl Sulfone Electrolyte Additive
Previous Article in Special Issue
Correction of Beta-Thalassemia IVS-II-654 Mutation in a Mouse Model Using Prime Editing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Enhancing Animal Disease Resistance, Production Efficiency, and Welfare through Precise Genome Editing

1
State Key Laboratory of Animal Nutrition, Key Laboratory of Animal Genetics Breeding and Reproduction of Ministry of Agriculture and Rural Affairs of China, Institute of Animal Sciences, Chinese Academy of Agricultural Sciences, Beijing 100193, China
2
Guangdong Provincial Key Laboratory of Animal Molecular Design and Precise Breeding, Key Laboratory of Animal Molecular Design and Precise Breeding of Guangdong Higher Education Institutes, School of Life Science and Engineering, Foshan University, Foshan 528231, China
3
Agricultural Genome Institute at Shenzhen, Chinese Academy of Agricultural Sciences, Shenzhen 518124, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(13), 7331; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23137331
Submission received: 10 June 2022 / Revised: 25 June 2022 / Accepted: 28 June 2022 / Published: 30 June 2022
(This article belongs to the Special Issue Gene Editing and Delivery in Animal Genetic Engineering)

Abstract

:
The major goal of animal breeding is the genetic enhancement of economic traits. The CRISPR/Cas system, which includes nuclease-mediated and base editor mediated genome editing tools, provides an unprecedented approach to modify the mammalian genome. Thus, farm animal genetic engineering and genetic manipulation have been fundamentally revolutionized. Agricultural animals with traits of interest can be obtained in just one generation (and without long time selection). Here, we reviewed the advancements of the CRISPR (Clustered regularly interspaced short palindromic repeats)/Cas (CRISPR associated proteins) genome editing tools and their applications in animal breeding, especially in improving disease resistance, production performance, and animal welfare. Additionally, we covered the regulations on genome-edited animals (GEAs) and ways to accelerate their use. Recommendations for how to produce GEAs were also discussed. Despite the current challenges, we believe that genome editing breeding and GEAs will be available in the near future.

1. Introduction

Genetic improvement of economic traits, such as disease resistance, meat production, and meat quality, is the main focus of animal breeding. Through selective breeding and crossbreeding, some economic traits, such as growth and reproduction, have been greatly enhanced over the past several decades. However, these conventional breeding methods have been costly and painstakingly slow, and some polygenetic traits, such as disease resistance, have not been dramatically improved. The development of large animal genetic manipulation technology, particularly CRISPR (Clustered regularly interspaced short palindromic repeats)/Cas (CRISPR associated proteins) system mediated genome editing, has provided efficient ways to affect traits of interest to produce agricultural animals in just one generation.
The CRISPR/Cas system has led to a revolution in the field of genetic manipulation and tremendously expanded its range, providing great tools for animal biotechnology research and livestock breeding. Rapid advancements have been made not only in gene editing, base editing [1], and prime editing [2] but also in transcriptional regulation and post-transcriptional engineering using CRISPR/Cas system-based tools. In recent years, CRISPR/Cas tools have revolutionized the field of large animal breeding. They have shown great promise not only for reducing selection time and production costs, but also for improving characteristics difficult to achieve or not amenable by traditional breeding methods. Valuable traits such as disease resistance, meat production, meat quality, and traits that can improve animal welfare can now be efficiently achieved by CRISPR/Cas tools. In this review, we systematically presented the innovations in CRISPR/Cas genome editing tools and their application in agricultural animal breeding. We also discussed some of the challenges encountered with genome-edited animals (GEAs) and its future applications.

2. CRISPR/Cas System Mediated Genetic Manipulation

Precision genomic DNA modification can now be achieved with much greater simplicity and efficiency using the CRISPR/Cas system (Figure 1A). RNA-guided Cas9, Cas9 nickase (Cas9n), and Cas12 can cleave the genome in a specific position resulting in double-stranded breaks (DSBs), which trigger the cell to repair the DNA damage by either non-homologous end-joining (NHEJ) or homology-directed repair (HDR). NHEJ can introduce small insertions or deletions, HDR can introduce a targeted mutation with a donor sequence. A base editor can perform precise point-mutations without DSBs; it consists of a Cas9n and nucleobase deaminase enzyme that catalyzes targeted deamination reactions. Several different base editors have been developed so far [2]: cytosine base editors (CBEs) catalyze C > T transitions, adenosine base editors (ABEs) facilitate transitions of A > G, guanosine base editors can attain C > G transversions [3], and dual adenine and cytosine base editors [4,5] can achieve C > T and A > G conversions at the same time.
Another alternative gene-editing technology named ‘prime editing’ has further expanded the range of precision genomic DNA modification [6]. Prime editors not only can catalyze all twelve possible transition and transversion mutations but can also mediate small insertion or deletion mutations. It consists of a Cas9n fused with a reverse transcriptase (RT) domain and a pegRNA (prime editing guide RNA). The pegRNA not only directs primer editor binding to the target DNA sequence and then facilitates nicking of the target strand, but also works as a template for the RT domain to synthesize a new DNA strand. This new strand, which contains the modification site, is then inserted into the target site following DNA repair.
Beyond genomic DNA modification, CRISPR/Cas system can be used to activate or repress RNA transcription (Figure 1B). Catalytically deficient Cas9 (dead Cas9, dCas9) can be used as a modular platform, coupling with transcriptional regulators or domains, such as VP64, KRAB [7,8], or DNMT3A [9], can attain specific, rapid, and multiplexed transcriptional activation, repression, or epigenetic modification in a range of cell types [10,11].
At the post-transcriptional level, gene expression regulation can also be manipulated by CRISPR/Cas system (Figure 1C). Class 2 type VI Cas proteins Cas13 family have programmable RNase activity [12,13,14], which can cleavage the RNA molecular by the binding of a guide RNA. Therefore, Cas13 proteins have been employed in RNA interference [12], targeted RNA splicing [12,15], and RNA base editing [16,17] in mammalian cells. It has been proved that Cas13 proteins can efficiently and specifically mediate mRNAs and long non-coding RNAs (lncRNAs) knock-down in mammalian cells [12], embryos [18], and mouse models [19,20]. Furthermore, catalytically inactivated Cas13 (dead Cas13, dCas13) protein fused with splicing regulatory domains can be directed by a crRNA to activate or perturb specific exons [12,15]. Dead Cas13 fuses with nucleobase deaminase enzymes instead of Cas9n, and can be used to edit a single RNA base in cells [16,17], providing a new method for correcting point mutations at the post-transcriptional level.

3. Disease Resistance Breeding

One of the most important applications of genome editing on farm animals is to improve the resistance or tolerance to pathogens. Farm animal infectious diseases not only cause huge economic losses to the animal husbandry industry but also threaten human health. It has always been an insurmountable industrial problem that bothers animal breeders and veterinary experts. However, disease resistance is a complex and polygenic trait; using traditional genetic selection for disease resistance breeding is much more costly, time-consuming, and inefficient. Moreover, the broad application of vaccines and antibiotics weakened the urgency of disease resistance selection programming to a certain extent. But transgenic and gene targeting have been successfully used in breeding antiviral animals such as cattle free of prion protein [21] and pigs that express anti-foot-and-mouth disease virus (FMDV) shRNA [22]. Recently, genome editing technology, especially CRISPR/Cas mediated gene knock-out/knock-in and precise modification, highly improved the efficiency of disease resistance animal breeding.
Leukotoxin secreted by Mannheimia (Pasteurella) haemolytica binds to the uncleaved signal peptide of CD18 protein and causes cytolysis of ruminant leukocytes [23], resulting in acute inflammation and lung tissue damage, inflicting a huge economic loss to the world-wide cattle industry. The first-generation gene editing tool zinc finger nuclease (ZFNs) have been applied to introduce a single amino acid mutation in the bovine CD18 protein [24]. Leukocytes from the CD18-gene-edited cattle expressed CD18 protein without the signal peptide, thus these leukocytes were resistant to leukotoxin-induced cytolysis. Another example is the CD163 gene. Porcine reproductive and respiratory syndrome (PRRS) is a worldwide infectious disease that costs million dollars a year to the swine industry. Due to the genetic diversity of PRRS virus (PRRSV), vaccines have not been able to control the disease. In 2010, Van Breedam et al. identified that SIGLEC1 (also known as CD169) and CD163 were necessary surface receptors for PRRSV entrance and uncoat in porcine cells [25,26]. Based on this finding, SIGLEC1 knock-out pigs [27] and CD163 knock-out pigs [28,29] were generated by CRISPR/Cas technology. A follow-up study on CD163 biallelic knock-out pigs showed that they process full resistance to PRRSV [30], no matter when they were infected directly or exposed to infected pen mates. Further studies show that CD163 knock-out pigs are also completely resistant to the infection of highly pathogenic PRRSV, which is more virulent than classical type 2 PRRSV [31].
However, CD163 gene has a variety of biological functions, knock-out may have a negative physiological impact on the economic traits of pigs. Therefore, a further precision modification was performed by either replacing the domain with orthologous CD163 protein domain [32,33] or by deleting SRCR domain 5 [34,35,36,37]. These studies providing a basis for further investigation of the essential region or even several amino acids associated with PRRSV infection. Once the essential amino acids for PRRSV infection were identified, by changing only one or several amino acids, PRRSV-resistant pig lines that retaining the biological functions of CD163 protein can be obtained. This might be achieved by the application of base editing technologies and tools in the near future.
Similar to CD163, porcine aminopeptidase N (pAPN, also known as ANPEP, CD13) gene was identified as a candidate receptor of transmissible gastroenteritis virus (TGEV) and porcine epidemic diarrhea virus (PEDV). Inhibition or direct knock-out of pAPN in cells can moderate TGEV infection [38,39]. Pigs with null pAPN are resistant to TGEV, but retained susceptibility to infection with PEDV [40,41]. Xu et al. (2020) successfully generated CD163 and pAPN gene knock-out pigs using CRISPR/Cas9 and somatic cell nuclear transfer (SCNT) [42]. These double-gene-knock-out (DKO) pigs not only exhibit complete resistance to both PRRSV and TGEV but also exhibit decreased susceptibility to porcine deltacoronavirus (PDCoV) infection. Furthermore, there are no differences in the production performance, reproductive performance, or pork nutrient content between DKO pigs and wild-type control pigs. This study shows that multiple sites editing in a pig genome is feasible by CRISPR/Cas9 and cellular screening, and also enlightened that breeding animals with multiple desirable traits, like disease-resistant, can be achieved by cellular surface receptors editing. In another research, Tu et al. (2019) obtained CMP-N glycolylneuraminic acid hydroxylase (CMAH) gene knock-out pigs by microinjection of two single-guide RNA and Cas9 mRNA. Although CMAH knock-out piglets exhibited delayed PEDV onset and diminished disease severity, they are not immune to PEDV [43]. Therefore, there is still a lot of work to be carried out in order to find the PEDV receptor.
Besides gene knock-out, site-specific expressing of resistance-associated genes could also be a strategy to breed disease resistance livestock. Bovine tuberculosis, which is caused by Mycobacterium bovis, is a serious threat to the agricultural economy and human health. Currently, no effective programs exist to control bovine tuberculosis in many less-developed areas of the world. The mouse SP110 (also known as Ipr1) gene can limit the growth of Mycobacterium in macrophages and inducing apoptosis in infected cells [44]. Research proved that integrating the mouse SP110 gene into the bovine genome by TALEN can control the growth of Mycobacterium and limit the transmission of tuberculosis in pen mates [45]. The natural resistance-associated macrophage protein-1 gene (NRAMP1, also known as SLC11A1) of bovine is associated with innate resistance to intracellular pathogens such as Mycobacterium, Leishmania, Salmonella, and Brucella. Adding a copy of bovine NRAMP1 gene to the specific locus of bovine genome by single Cas9 nickase can provide cattle with increased resistance to tuberculosis [46]. Histone deacetylase 6 (HDAC6) has anti-viral activities during the viral life cycle, overexpressing HDAC6 enhances the resistance to PRRSV infection both in vitro and in vivo [47].
As CRISPR/Cas system is the adaptive and heritable immune system of bacteria and archaea. It is reasonable that CRISPR/Cas system could also be used as a designed immune system to eliminate or inhibit the replication of animal viruses. This strategy was already tested in animal cells. Tang et al. (2017) designed 75 single guide RNAs (sgRNA) targeting both essential and nonessential genes across the genome of pseudorabies virus (PRV) [48], which is a swine herpesvirus that causes significant economic losses in the worldwide swine industry, in vitro experiments found that most of the sgRNAs significantly inhibited PRV replication. More importantly, they also demonstrated that targeting PRV with sgRNA pools that contain multiple sgRNAs can completely abolish the production of infectious viruses in cells. African swine fever (ASF) is another economically important infectious disease of swine with high mortality rates, which threatens pig production across the globe. It is caused by the African swine fever virus (ASFV), which is a double-stranded DNA virus, and no essential cell surface receptors have been identified. Hubner et al. (2018) observed complete abrogation of ASFV yields by targeting the viral phosphoprotein p30. However, they also found ASFV mutants with one or two nucleotides can escape CRISPR/Cas9 inhibition [49]. These data proved the possibility that Cas9 and multi-targeting sgRNA could be developed as an efficient antiviral strategy. With multiple sgRNAs that targeting multiple viral sequences, Cas9 or Cas13d can be directed to recognize and degrade viral DNA or RNA, providing broad-spectrum antiviral capabilities for animals (Figure 2). Cas9 or Cas13d and sgRNAs could be integrated into the genome of transgenic animals to breed antiviral animals, or delivered to animal cells as novel antivirus agents.
However, there seem to be many hurdles to overcome before its realization. First, it is necessary to evaluate the therapeutic potentials in vivo, and a detailed analysis of the host immune response to the Cas9 and Cas13 proteins and multiple sgRNAs needs to be conducted for the prediction of potential side effects associated with this antiviral therapy. Second, the potential off-target activity of this CRISPR/Cas9-based antiviral therapy needs to be evaluated, especially when multiple sgRNAs were employed to degrade the viral genome. Various measures including high-infidelity Cas9 variants should be taken to minimize the potential off-target activity before in vivo application. Third, the cost of this CRISPR/Cas9-based antiviral therapy needs to be fairly inexpensive to applicate on farms, at least it should be cost-effective than biosafety measures that are already widely used in animal farms to control ASFV, PRV, PRRSV, etc. Despite these challenges, the full healing potential of CRISPR/Cas system based antiviral therapy should be able to motivate its development.

4. Improving Production Performance

Increasing animal lean meat rates is the mainstream breeding goal of food animals for a very long time. Myostatin (MSTN, also known as GDF8) is a negative regulator of skeletal muscle mass [50]. Natural mutations of MSTN have been reported in cattle [51,52], sheep [53], dogs [54], pigs [55,56,57], and human [58]. These animals show a double-muscled phenotype of dramatically increased muscle mass. In adult tissues, MSTN is expressed almost exclusively in skeletal muscle, but detectable levels of MSTN RNA are also present in adipose tissue [59]. MSTN knock-out confers a remarkable increase in lean meat yield [60,61,62] in many mammals and increased levels of polyunsaturated fatty acids in pigs [62,63]. Although MSTN knock-out significantly increased lean meat production of pigs, severe hindlimb weakness was observed among MSTN−/− newborns in Western commercial pig breeds [55,64]. This congenital hindlimb weakness defect is a prohibitive bottleneck to the safe and ethical application of MSTN-editing in pigs. Fan et al. (2021) performed a long-term and multidomain evaluation for multiple MSTN-edited pig breeds. They demonstrated a practical alternative edit-site-based solution to overcome the hindlimb weakness and illustrated that MSTN-editing can sustainably increase the yields of breed-specific lean meat and the levels of desirable lipids without deleteriously affecting feed-conversion rates or litter size.
Another strategy to overcome the hindlimb weakness is to mimic the naturally existing MSTN mutations, such as the Piedmontese c.G938A mutation and the Belgian Blue mutation (821del11). The Piedmontese c.G938A mutation at the MSTN results in the substitution of a highly conserved cysteine to tyrosine (p.C313Y) in the mature region of the protein. Wang et al. (2016) introduce a missense point mutation, which mimicking the orthologous p.C313Y mutation, and generated one cloned piglet harboring the p.C313Y mutation via SCNT [65]. Zou et al. (2019) introduced an 11-bp deletion, which is orthologous to the natural Belgian Blue MSTN mutation, at the exon 3 of pig MSTN gene and obtained two cloned Duroc piglets [66]. In addition, site-specific insertion of MSTN inhibitor was also performed to enhance the growth performance [67]. These works expand the range of modifying MSTN gene, holding great promise for animal breeding and disease modeling.
Another potential candidate gene for improving meat production in pigs is the insulin-like growth factor 2 (IGF2) gene. IGF2 regulates cellular proliferation, differentiation, and apoptosis in both fetal and post-natal growth. The transcription and expression of IGF2 are downregulated by the zinc finger BED domain-containing protein 6 (ZBED6), mutations in the IGF2 intron 3-3072, which is a ZBED6 binding site, can upregulate the expression of IGF2 and improve muscle development [68]. Modifying this ZBED6 binding site by CRISPR/Cas9 genome editing tools will greatly enhance the muscle development in indigenous Chinese pig breeds such as the Chinese Bama pig and the Liang Guang Small Spotted pig [69,70]. In sheep, the suppressor cytokine signaling 2 (SOCS2) gene plays a vital role in the control of bone mass and body weight. A point mutation g.C1901T in SOCS2 is highly associated with increased body weight and size in sheep [71]. Zhou et al. (2019) obtained gene-edited lambs with a C to T point mutation in SOCS2 gene by micro-injection of programmable deaminases BE3 into sheep zygotes and without inducing unintended off-target mutations at the genome-wide scale [72].
The fat-1 gene is a fatty acid desaturase gene that originates in Caenorhabditis elegans. In the last two decades, a large number of fat-1 transgenic animals have been developed to convert n-6 polyunsaturated fatty acids (n-6PUFAs) to n-3 polyunsaturated fatty acids (n-3PUFAs) and increase meat quality. However, the random integration transgene strategy frequently results in fluctuating transgene expression and the insertion of selected marker genes. Li et al. (2018) used the CRISPR/Cas9 technology to introduce a single copy of the fat-1 gene into the porcine Rosa26 locus, resulting in site-specific fat-1 knock-in pigs with a considerable rise in n-3PUFAs levels [73]. Zhang et al. (2018) used CRISPR/Cas9 to insert the fat-1 gene into the goat MSTN locus, resulting in simultaneous deletion of endogenous genes and site-specific fat-1 gene knock-in [74]. You et al. (2021) created double-gene knock-in pigs by inserting single copies of the fat-1 and IGF-1 genes into the porcine Rosa26 locus at the same time [75]. These pigs have a great potential for boosting pork’s nutritional value.
Uncoupling protein 1 (UCP1) is a key element of nonshivering thermogenesis and is important for preventing cold and regulating body adiposity. However, domestic pigs lack a functional UCP1 gene, making them susceptible to cold and prone to fat deposition. Zheng et al. (2017) used a CRISPR/Cas9-mediated approach to efficiently insert mouse UCP1 cDNA into the porcine endogenous UCP1 locus [76] and obtained UCP1-KI pigs, which had less fat deposition, higher carcass lean percentage, and, most importantly, improved ability to maintain body temperature in cold environments. These pigs provide a potentially valuable resource for farm animal production.

5. Improving Animal Welfare

Genome editing in livestock will not only improve farm animals’ resistance or tolerance to pathogens, increase production performance, but also will prevent unnecessary animal suffering, which may encourage public support of GEAs for food chain production.
In modern livestock, daily management of horned cattle poses a high risk of injury for each other as well as for the farmers. Physical dehorning of cattle is used to protect animals and farmers from accidental injury but is associated with stress and pain for the calves. Naturally occurring structural variants causing hornlessness are known for most beef cattle. The polled Celtic variant from the genome of an Angus cow was integrated into dairy cattle using genome editing tools and somatic cell cloning [77]. The presence of some substances in male pork such as androstenedione and methylindole will make the ‘boar taint’ and affect the taste of pork. To improve pork quality and to facilitate production management, boars are usually castrated after birth. In 2016, the Australian Society of Animal Production (ASAP) conference presented a new method of using CRISPR/Cas9 technology to knock-out the KISSR gene (responsible for testicular development in pigs) to block testicular development and achieve the effect of depopulation [78]. Taken together, these research works show that genome editing is precise, sustainable, and directly applicable to improved animal well-being.
In summary, with the rapid development of genome editing tools, GEAs with desirable features can now be efficiently obtained (Table 1). Before these GEAs are commercialized, their breeding potential and/or safety should be assessed.

6. Regulations on GEAs

The growth of green and sustainable agriculture to feed the world’s growing population will be substantially aided by GEAs, but how to control GEAs and genome-edited animal products remains an unsolved dilemma. As a result of the technological breakthroughs, regulations created for the control of transgenic animals appear to no longer be adequate for GEAs.
On 14 December 2020, the U.S. Food and Drug Administration (FDA) approved GalSafeTM pigs for medical and/or food use, it is the first-of-its-kind intentional genomic alteration (IGA) in livestock [87]. On 7 March 2022, FDA announced it has made a low-risk determination for the marketing of products, including food, from two “PRLR-SLICK” genome-edited beef cattle and their offspring after determining that the intentional genomic alteration does not raise any safety concerns [88]. The IGA results in the equivalent genotype (genetic make-up) and short-hair coat trait seen in some conventionally bred cattle, known as a “slick” coat. This is the FDA’s first low-risk determination for enforcement discretion for an IGA in an animal for food use. On 24 January 2022, Ministry of Agriculture and Rural Affairs of China officially issued the “Guidelines for Safety Evaluation of Gene Edited Plants for Agricultural Use (for Trial Implementation)”, marking a significant simplification of China’s regulatory policy on the safety evaluation of gene edited plants, which is expected to further accelerate the industrialization of biological breeding in China [89]. In countries such as Argentina, Australia, and Brazil, regulation is not required if genome-edited animals do not contain any foreign DNA [90,91].
Realizing the enormous potential of GEAs for long-term solutions to serious environmental and food security challenges will be the first step toward allowing their broad usage in most countries in the world. Governments, developers, manufacturers, and consumers must all work together and communicate effectively. For governments, a case-by-case GEA review and regulatory method will result in a system that is more efficient, objective, complete, and operational [92,93]. For developers, using DNA-free editing techniques to create GEAs, such as gRNA/Cas9 ribonucleoproteins (RNP) [94] and base editor RNP, should help gain public trust.

7. Prospects and Challenges

Genome editing technology offers ground-breaking tools and methodologies for manipulating the genomes of large animals, opening up new possibilities for livestock breeding and animal husbandry. It is expected that other applications will be developed, and that genome-edited livestock-derived meat will be available for consumption in the near future. However, when producing animals for agriculture, the off-target effect remains a key concern. Off-target mutations may result in knock-out events or silent mutations in protein-coding genes, or interference with transcriptional regulation. Mutations in protein-coding regions may cause aberrant form of proteins, which may induce food allergenicity. Changes in translational regulation may have an impact on animal health, reproduction, and growth performance. Therefore, low-risk genome editing methods, such as DNA-free genome editing, have gained much attention recently. DNA-free genome editing strategies could profoundly reduce the risk of off-target mutations [94]. Somatic cell nuclear transfer (SCNT) could also be used to eliminate off-target mutations before GEAs breeding. By the SCNT approach, it is possible to verify the genotype and off-target mutations of donor cells before live animal production takes place. It can also avoid the occurrence of genetic mosaicism and reducing the overall cost of genome edited livestock production. These aspects are especially critical for application in large domestic animals that have particularly long generation intervals.
Another trend is to use base editor tools to create point mutation in large animal instead of Cas9 tools. Base editing tools have several advantages over Cas9 tools. The first one is that base editors do not cause DSBs. DSBs caused by Cas9 has been shown to result in excessive DNA damage and cell death, using base editor tools instead of Cas9 in animal breeding can avoid these unpredictable risks, especially when editing multiple genes simultaneously. Additionally, Base editors have higher efficiency in making point mutations. For example, the efficiency of making point mutation at porcine MSTN gene by Cas9 and single-stranded oligonucleotides mediated HDR was 10.3% [64], while 46.3% target point mutation were achieved at porcine MSTN locus by using an optimized cytosine base editor (hA3A-BE3-NG) [95]. Furthermore, by only one transfection, a three genes base-editing rate of 71.4% can be achieved at the porcine GGTA1, B4galNT2, and CMAH loci [96]. This research indicates that base editors are more efficient in achieving point mutations, especially for multiple gene editing.
In summary, the rapid progress of gene editing technologies will greatly accelerate their application in domestic animal husbandry. GEAs with valuable traits will contribute to the goal of developing green and sustainable agriculture for the world.

Author Contributions

Conceptualization, Z.L., Y.M. and K.L.; writing—original draft preparation, Z.L. and T.W.; writing—review and editing, Z.L., T.W., G.X., H.W., B.W., Z.F., Y.M. and K.L.; supervision, Z.F., Y.M. and K.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (32130102), the GuangDong Basic and Applied Basic Research Foundation (2020B1515120016), the Beijing Municipal Natural Science Foundation (6173034), the Major Scientific Research Tasks for Scientific and Technological Innovation Projects of the Chinese Academy of Agricultural Sciences (CAAS-ZDRW202006), the Central Public Interest Scientific Institution Basal Research Fund (2021-YWF-ZYSQ-08) and the Agricultural Science and Technology Innovation Program of Chinese Academy of Agricultural Sciences (ASTIP-IAS05).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Kantor, A.; McClements, M.E.; MacLaren, R.E. CRISPR-Cas9 DNA Base-Editing and Prime-Editing. Int. J. Mol. Sci. 2020, 21, 6240. [Google Scholar] [CrossRef] [PubMed]
  2. Zeballos, C.M.; Gaj, T. Next-Generation CRISPR Technologies and Their Applications in Gene and Cell Therapy. Trends Biotechnol. 2021, 39, 692–705. [Google Scholar] [CrossRef] [PubMed]
  3. Kurt, I.C.; Zhou, R.; Iyer, S.; Garcia, S.P.; Miller, B.R.; Langner, L.M.; Grünewald, J.; Joung, J.K. CRISPR C-to-G base editors for inducing targeted DNA transversions in human cells. Nat. Biotechnol. 2021, 39, 41–46. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, X.; Zhu, B.; Chen, L.; Xie, L.; Yu, W.; Wang, Y.; Li, L.; Yin, S.; Yang, L.; Hu, H.; et al. Dual base editor catalyzes both cytosine and adenine base conversions in human cells. Nat. Biotechnol. 2020, 38, 856–860. [Google Scholar] [CrossRef] [PubMed]
  5. Grunewald, J.; Zhou, R.; Lareau, C.A.; Garcia, S.P.; Iyer, S.; Miller, B.R.; Langner, L.M.; Hsu, J.Y.; Aryee, M.J.; Joung, J.K. A dual-deaminase CRISPR base editor enables concurrent adenine and cytosine editing. Nat. Biotechnol. 2020, 38, 861–864. [Google Scholar] [CrossRef]
  6. 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]
  7. Gilbert, L.A.; Larson, M.H.; Morsut, L.; Liu, Z.; Brar, G.A.; Torres, S.E.; Stern-Ginossar, N.; Brandman, O.; Whitehead, E.H.; Doudna, J.A.; et al. CRISPR-mediated modular RNA-guided regulation of transcription in eukaryotes. Cell 2013, 154, 442–451. [Google Scholar] [CrossRef] [Green Version]
  8. Gilbert, L.A.; Horlbeck, M.A.; Adamson, B.; Villalta, J.E.; Chen, Y.; Whitehead, E.H.; Guimaraes, C.; Panning, B.; Ploegh, H.L.; Bassik, M.C.; et al. Genome-Scale CRISPR-Mediated Control of Gene Repression and Activation. Cell 2014, 159, 647–661. [Google Scholar] [CrossRef] [Green Version]
  9. Vojta, A.; Dobrinić, P.; Tadić, V.; Bočkor, L.; Korać, P.; Julg, B.; Klasić, M.; Zoldoš, V. Repurposing the CRISPR-Cas9 system for targeted DNA methylation. Nucleic Acids Res. 2016, 44, 5615–5628. [Google Scholar] [CrossRef] [Green Version]
  10. Nunez, J.K.; Chen, J.; Pommier, G.C.; Cogan, J.Z.; Replogle, J.M.; Adriaens, C.; Ramadoss, G.N.; Shi, Q.; Hung, K.L.; Samelson, A.J.; et al. Genome-wide programmable transcriptional memory by CRISPR-based epigenome editing. Cell 2021, 184, 2503–2519. [Google Scholar] [CrossRef]
  11. Engreitz, J.; Abudayyeh, O.; Gootenberg, J.; Zhang, F. CRISPR Tools for Systematic Studies of RNA Regulation. Cold Spring Harb. Perspect. Biol. 2019, 11, a035386. [Google Scholar] [CrossRef] [Green Version]
  12. Konermann, S.; Lotfy, P.; Brideau, N.J.; Oki, J.; Shokhirev, M.N.; Hsu, P.D. Transcriptome Engineering with RNA-Targeting Type VI-D CRISPR Effectors. Cell 2018, 173, 665–676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Cox, D.B.T.; Gootenberg, J.S.; Abudayyeh, O.O.; Franklin, B.; Kellner, M.J.; Joung, J.; Zhang, F. RNA editing with CRISPR-Cas13. Science 2017, 358, 1019–1027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Abudayyeh, O.O.; Gootenberg, J.S.; Essletzbichler, P.; Han, S.; Joung, J.; Belanto, J.J.; Verdine, V.; Cox, D.B.T.; Kellner, M.J.; Regev, A.; et al. RNA targeting with CRISPR-Cas13. Nature 2017, 550, 280–284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Du, M.; Jillette, N.; Zhu, J.J.; Li, S.; Cheng, A.W. CRISPR artificial splicing factors. Nat. Commun. 2020, 11, 2973. [Google Scholar] [CrossRef] [PubMed]
  16. Abudayyeh, O.O.; Gootenberg, J.S.; Franklin, B.; Koob, J.; Kellner, M.J.; Ladha, A.; Joung, J.; Kirchgatterer, P.; Cox, D.B.T.; Zhang, F. A cytosine deaminase for programmable single-base RNA editing. Science 2019, 365, 382–386. [Google Scholar] [CrossRef]
  17. Sinnamon, J.R.; Kim, S.Y.; Corson, G.M.; Song, Z.; Nakai, H.; Adelman, J.P.; Mandel, G. Site-directed RNA repair of endogenous Mecp2 RNA in neurons. Proc. Natl. Acad. Sci. USA 2017, 114, E9395–E9402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Kushawah, G.; Hernandez-Huertas, L.; Abugattas-Nunez Del Prado, J.; Martinez-Morales, J.R.; DeVore, M.L.; Hassan, H.; Moreno-Sanchez, I.; Tomas-Gallardo, L.; Diaz-Moscoso, A.; Monges, D.E.; et al. CRISPR-Cas13d Induces Efficient mRNA Knockdown in Animal Embryos. Dev. Cell 2020, 54, 805–817. [Google Scholar] [CrossRef]
  19. He, B.; Peng, W.; Huang, J.; Zhang, H.; Zhou, Y.; Yang, X.; Liu, J.; Li, Z.; Xu, C.; Xue, M.; et al. Modulation of metabolic functions through Cas13d-mediated gene knockdown in liver. Protein Cell 2020, 11, 518–524. [Google Scholar] [CrossRef] [Green Version]
  20. Zhao, X.; Liu, L.; Lang, J.; Cheng, K.; Wang, Y.; Li, X.; Shi, J.; Wang, Y.; Nie, G. A CRISPR-Cas13a system for efficient and specific therapeutic targeting of mutant KRAS for pancreatic cancer treatment. Cancer Lett. 2018, 431, 171–181. [Google Scholar] [CrossRef]
  21. Richt, J.A.; Kasinathan, P.; Hamir, A.N.; Castilla, J.; Sathiyaseelan, T.; Vargas, F.; Sathiyaseelan, J.; Wu, H.; Matsushita, H.; Koster, J.; et al. Production of cattle lacking prion protein. Nat. Biotechnol. 2007, 25, 132–138. [Google Scholar] [CrossRef] [PubMed]
  22. Hu, S.; Qiao, J.; Fu, Q.; Chen, C.; Ni, W.; Wujiafu, S.; Ma, S.; Zhang, H.; Sheng, J.; Wang, P.; et al. Transgenic shRNA pigs reduce susceptibility to foot and mouth disease virus infection. Elife 2015, 4, e06951. [Google Scholar] [CrossRef] [PubMed]
  23. Shanthalingam, S.; Srikumaran, S. Intact signal peptide of CD18, the beta-subunit of beta2-integrins, renders ruminants susceptible to Mannheimia haemolytica leukotoxin. Proc. Natl. Acad. Sci. USA 2009, 106, 15448–15453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Shanthalingam, S.; Tibary, A.; Beever, J.E.; Kasinathan, P.; Brown, W.C.; Srikumaran, S. Precise gene editing paves the way for derivation of Mannheimia haemolytica leukotoxin-resistant cattle. Proc. Natl. Acad. Sci. USA 2016, 113, 13186–13190. [Google Scholar] [CrossRef] [Green Version]
  25. Van Breedam, W.; Delputte, P.L.; Van Gorp, H.; Misinzo, G.; Vanderheijden, N.; Duan, X.; Nauwynck, H.J. Porcine reproductive and respiratory syndrome virus entry into the porcine macrophage. J. Gen. Virol. 2010, 91, 1659–1667. [Google Scholar] [CrossRef]
  26. Van Gorp, H.; Van Breedam, W.; Van Doorsselaere, J.; Delputte, P.L.; Nauwynck, H.J. Identification of the CD163 protein domains involved in infection of the porcine reproductive and respiratory syndrome virus. J. Virol. 2010, 84, 3101–3105. [Google Scholar] [CrossRef] [Green Version]
  27. Prather, R.S.; Rowland, R.R.; Ewen, C.; Trible, B.; Kerrigan, M.; Bawa, B.; Teson, J.M.; Mao, J.; Lee, K.; Samuel, M.S.; et al. An intact sialoadhesin (Sn/SIGLEC1/CD169) is not required for attachment/internalization of the porcine reproductive and respiratory syndrome virus. J. Virol. 2013, 87, 9538–9546. [Google Scholar] [CrossRef] [Green Version]
  28. Whitworth, K.M.; Lee, K.; Benne, J.A.; Beaton, B.P.; Spate, L.D.; Murphy, S.L.; Samuel, M.S.; Mao, J.; O’Gorman, C.; Walters, E.M.; et al. Use of the CRISPR/Cas9 system to produce genetically engineered pigs from in vitro-derived oocytes and embryos. Biol. Reprod. 2014, 91, 78. [Google Scholar] [CrossRef] [Green Version]
  29. Wei, Y.; Liu, Z.; Xu, K.; Evanna, H.; Dyce, P.; Li, J.; Zhou, W.; Dong, S.; Feng, B.; Mu, Y.; et al. Generation and Propagation of Cluster of Differentiation 163 Biallelic Gene Editing Pigs. Sci. Agric. Sin. 2018, 51, 770–777. [Google Scholar] [CrossRef]
  30. Whitworth, K.M.; Rowland, R.R.; Ewen, C.L.; Trible, B.R.; Kerrigan, M.A.; Cino-Ozuna, A.G.; Samuel, M.S.; Lightner, J.E.; McLaren, D.G.; Mileham, A.J.; et al. Gene-edited pigs are protected from porcine reproductive and respiratory syndrome virus. Nat. Biotechnol. 2016, 34, 20–22. [Google Scholar] [CrossRef]
  31. Yang, H.; Zhang, J.; Zhang, X.; Shi, J.; Pan, Y.; Zhou, R.; Li, G.; Li, Z.; Cai, G.; Wu, Z. CD163 knockout pigs are fully resistant to highly pathogenic porcine reproductive and respiratory syndrome virus. Antiviral Res. 2018, 151, 63–70. [Google Scholar] [CrossRef] [PubMed]
  32. Wells, K.D.; Bardot, R.; Whitworth, K.M.; Trible, B.R.; Fang, Y.; Mileham, A.; Kerrigan, M.A.; Samuel, M.S.; Prather, R.S.; Rowland, R.R.R. Replacement of Porcine CD163 Scavenger Receptor Cysteine-Rich Domain 5 with a CD163-Like Homolog Confers Resistance of Pigs to Genotype 1 but Not Genotype 2 Porcine Reproductive and Respiratory Syndrome Virus. J. Virol. 2017, 91, e01521-16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Chen, J.; Wang, H.; Bai, J.; Liu, W.; Liu, X.; Yu, D.; Feng, T.; Sun, Z.; Zhang, L.; Ma, L.; et al. Generation of Pigs Resistant to Highly Pathogenic-Porcine Reproductive and Respiratory Syndrome Virus through Gene Editing of CD163. Int. J. Biol. Sci. 2019, 15, 481–492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Burkard, C.; Lillico, S.G.; Reid, E.; Jackson, B.; Mileham, A.J.; Ait-Ali, T.; Whitelaw, C.B.; Archibald, A.L. Precision engineering for PRRSV resistance in pigs: Macrophages from genome edited pigs lacking CD163 SRCR5 domain are fully resistant to both PRRSV genotypes while maintaining biological function. PLoS Pathog. 2017, 13, e1006206. [Google Scholar] [CrossRef] [PubMed]
  35. Burkard, C.; Opriessnig, T.; Mileham, A.J.; Stadejek, T.; Ait-Ali, T.; Lillico, S.G.; Whitelaw, C.B.A.; Archibald, A.L. Pigs Lacking the Scavenger Receptor Cysteine-Rich Domain 5 of CD163 Are Resistant to Porcine Reproductive and Respiratory Syndrome Virus 1 Infection. J. Virol. 2018, 92, e00415–e00418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Wang, H.; Shen, L.; Chen, J.; Liu, X.; Tan, T.; Hu, Y.; Bai, X.; Li, Y.; Tian, K.; Li, N.; et al. Deletion of CD163 Exon 7 Confers Resistance to Highly Pathogenic Porcine Reproductive and Respiratory Viruses on Pigs. Int. J. Biol. Sci. 2019, 15, 1993–2005. [Google Scholar] [CrossRef] [Green Version]
  37. Guo, C.; Wang, M.; Zhu, Z.; He, S.; Liu, H.; Liu, X.; Shi, X.; Tang, T.; Yu, P.; Zeng, J.; et al. Highly Efficient Generation of Pigs Harboring a Partial Deletion of the CD163 SRCR5 Domain, Which Are Fully Resistant to Porcine Reproductive and Respiratory Syndrome Virus 2 Infection. Front. Immunol. 2019, 10, 1846. [Google Scholar] [CrossRef] [Green Version]
  38. Zhu, X.; Liu, S.; Wang, X.; Luo, Z.; Shi, Y.; Wang, D.; Peng, G.; Chen, H.; Fang, L.; Xiao, S. Contribution of porcine aminopeptidase N to porcine deltacoronavirus infection. Emerg Microbes Infect. 2018, 7, 65. [Google Scholar] [CrossRef]
  39. Ji, C.M.; Wang, B.; Zhou, J.; Huang, Y.W. Aminopeptidase-N-independent entry of porcine epidemic diarrhea virus into Vero or porcine small intestine epithelial cells. Virology 2018, 517, 16–23. [Google Scholar] [CrossRef]
  40. Whitworth, K.M.; Rowland, R.R.R.; Petrovan, V.; Sheahan, M.; Cino-Ozuna, A.G.; Fang, Y.; Hesse, R.; Mileham, A.; Samuel, M.S.; Wells, K.D.; et al. Resistance to coronavirus infection in amino peptidase N-deficient pigs. Transgenic Res. 2019, 28, 21–32. [Google Scholar] [CrossRef] [Green Version]
  41. Luo, L.; Wang, S.; Zhu, L.; Fan, B.; Liu, T.; Wang, L.; Zhao, P.; Dang, Y.; Sun, P.; Chen, J.; et al. Aminopeptidase N-null neonatal piglets are protected from transmissible gastroenteritis virus but not porcine epidemic diarrhea virus. Sci. Rep. 2019, 9, 13186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Xu, K.; Zhou, Y.; Mu, Y.; Liu, Z.; Hou, S.; Xiong, Y.; Fang, L.; Ge, C.; Wei, Y.; Zhang, X.; et al. CD163 and pAPN double-knockout pigs are resistant to PRRSV and TGEV and exhibit decreased susceptibility to PDCoV while maintaining normal production performance. Elife 2020, 9, e57132. [Google Scholar] [CrossRef] [PubMed]
  43. Tu, C.F.; Chuang, C.K.; Hsiao, K.H.; Chen, C.H.; Chen, C.M.; Peng, S.H.; Su, Y.H.; Chiou, M.T.; Yen, C.H.; Hung, S.W.; et al. Lessening of porcine epidemic diarrhoea virus susceptibility in piglets after editing of the CMP-N-glycolylneuraminic acid hydroxylase gene with CRISPR/Cas9 to nullify N-glycolylneuraminic acid expression. PLoS ONE 2019, 14, e0217236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Pan, H.; Yan, B.-S.; Rojas, M.; Shebzukhov, Y.V.; Zhou, H.; Kobzik, L.; Higgins, D.E.; Daly, M.J.; Bloom, B.R.; Kramnik, I. Ipr1 gene mediates innate immunity to tuberculosis. Nature 2005, 434, 767–772. [Google Scholar] [CrossRef] [PubMed]
  45. Wu, H.; Wang, Y.; Zhang, Y.; Yang, M.; Lv, J.; Liu, J.; Zhang, Y. TALE nickase-mediated SP110 knockin endows cattle with increased resistance to tuberculosis. Proc. Natl. Acad. Sci. USA 2015, 112, E1530–E1539. [Google Scholar] [CrossRef] [Green Version]
  46. Gao, Y.; Wu, H.; Wang, Y.; Liu, X.; Chen, L.; Li, Q.; Cui, C.; Liu, X.; Zhang, J.; Zhang, Y. Single Cas9 nickase induced generation of NRAMP1 knockin cattle with reduced off-target effects. Genome Biol. 2017, 18, 13. [Google Scholar] [CrossRef] [Green Version]
  47. Lu, T.; Song, Z.; Li, Q.; Li, Z.; Wang, M.; Liu, L.; Tian, K.; Li, N. Overexpression of Histone Deacetylase 6 Enhances Resistance to Porcine Reproductive and Respiratory Syndrome Virus in Pigs. PLoS ONE 2017, 12, e0169317. [Google Scholar] [CrossRef]
  48. Tang, Y.-D.; Liu, J.-T.; Wang, T.-Y.; Sun, M.-X.; Tian, Z.-J.; Cai, X.-H. CRISPR/Cas9-mediated multiple single guide RNAs potently abrogate pseudorabies virus replication. Arch. Virol. 2017, 162, 3881–3886. [Google Scholar] [CrossRef]
  49. Hubner, A.; Petersen, B.; Keil, G.M.; Niemann, H.; Mettenleiter, T.C.; Fuchs, W. Efficient inhibition of African swine fever virus replication by CRISPR/Cas9 targeting of the viral p30 gene (CP204L). Sci. Rep. 2018, 8, 1449. [Google Scholar] [CrossRef] [Green Version]
  50. McPherron, A.C.; Lawler, A.M.; Lee, S.-J. Regulation of skeletal muscle mass in mice by a new TGF-beta superfamily member. Nature 1997, 387, 83–90. [Google Scholar] [CrossRef]
  51. Grobet, L.; Royo Martin, L.J.; Poncelet, D.; Pirottin, D.; Brouwers, B.; Riquet, J.; Schoeberlein, A.; Dunner, S.; Ménissier, F.; Massabanda, J.; et al. A deletion in the bovine myostatin gene causes the double-muscled phenotype in cattle. Nat. Genet. 1997, 17, 71–74. [Google Scholar] [CrossRef]
  52. Grobet, L.; Poncelet, D.; Royo, L.J.; Brouwers, B.; Pirottin, D.; Michaux, C.; Ménissier, F.; Zanotti, M.; Dunner, S.; Georges, M. Molecular definition of an allelic series of mutations disrupting the myostatin function and causing double-muscling in cattle. Mamm. Genome 1998, 9, 210–213. [Google Scholar] [CrossRef] [PubMed]
  53. Boman, I.A.; Klemetsdal, G.; Blichfeldt, T.; Nafstad, O.; Våge, D.I. A frameshift mutation in the coding region of the myostatin gene (MSTN) affects carcass conformation and fatness in Norwegian White Sheep (Ovis aries). Anim. Genet. 2009, 40, 418–422. [Google Scholar] [CrossRef] [PubMed]
  54. Mosher, D.S.; Quignon, P.; Bustamante, C.D.; Sutter, N.B.; Mellersh, C.S.; Parker, H.G.; Ostrander, E.A. A mutation in the myostatin gene increases muscle mass and enhances racing performance in heterozygote dogs. PLoS Genet. 2007, 3, e79. [Google Scholar] [CrossRef]
  55. Matika, O.; Robledo, D.; Pong-Wong, R.; Bishop, S.C.; Riggio, V.; Finlayson, H.; Lowe, N.R.; Hoste, A.E.; Walling, G.A.; del-Pozo, J.; et al. Balancing selection at a premature stop mutation in the myostatin gene underlies a recessive leg weakness syndrome in pigs. PLoS Genet. 2019, 15, e1007759. [Google Scholar] [CrossRef] [Green Version]
  56. Stinckens, A.; Luyten, T.; Bijttebier, J.; Van den Maagdenberg, K.; Dieltiens, D.; Janssens, S.; De Smet, S.; Georges, M.; Buys, N. Characterization of the complete porcine MSTN gene and expression levels in pig breeds differing in muscularity. Anim. Genet. 2008, 39, 586–596. [Google Scholar] [CrossRef] [PubMed]
  57. Yu, L.; Tang, H.; Wang, J.; Wu, Y.; Zou, L.; Jiang, Y.; Wu, C.; Li, N. Polymorphisms in the 5′ regulatory region of myostatin gene are associated with early growth traits in Yorkshire pigs. Sci. China Ser. C 2007, 50, 642–647. [Google Scholar] [CrossRef]
  58. Schuelke, M.; Wagner, K.R.; Stolz, L.E.; Hübner, C.; Riebel, T.; Kömen, W.; Braun, T.; Tobin, J.F.; Lee, S.-J. Myostatin Mutation Associated with Gross Muscle Hypertrophy in a Child. N. Engl. J. Med. 2004, 350, 2682–2688. [Google Scholar] [CrossRef] [Green Version]
  59. Roberts, S.B.; Goetz, F.W. Myostatin protein and RNA transcript levels in adult and developing brook trout. Mol. Cell. Endocrinol. 2003, 210, 9–20. [Google Scholar] [CrossRef]
  60. Ding, Y.; Zhou, S.-W.; Ding, Q.; Cai, B.; Zhao, X.-E.; Zhong, S.; Jin, M.-H.; Wang, X.-L.; Ma, B.-H.; Chen, Y.-L. The CRISPR/Cas9 induces large genomic fragment deletions of MSTN and phenotypic changes in sheep. J. Integr. Agric. 2020, 19, 1065–1073. [Google Scholar] [CrossRef]
  61. He, Z.; Zhang, T.; Jiang, L.; Zhou, M.; Wu, D.; Mei, J.; Cheng, Y. Use of CRISPR/Cas9 technology efficiently targetted goat myostatin through zygotes microinjection resulting in double-muscled phenotype in goats. Biosci. Rep. 2018, 38, BSR20180742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Fan, Z.; Liu, Z.; Xu, K.; Wu, T.; Ruan, J.; Zheng, X.; Bao, S.; Mu, Y.; Sonstegard, T.; Li, K. Long-term, multidomain analyses to identify the breed and allelic effects in MSTN-edited pigs to overcome lameness and sustainably improve nutritional meat production. Sci. China Life Sci. 2022, 65, 362–375. [Google Scholar] [CrossRef]
  63. Ren, H.; Xiao, W.; Qin, X.; Cai, G.; Chen, H.; Hua, Z.; Cheng, C.; Li, X.; Hua, W.; Xiao, H.; et al. Myostatin regulates fatty acid desaturation and fat deposition through MEF2C/miR222/SCD5 cascade in pigs. Commun. Biol. 2020, 3, 612. [Google Scholar] [CrossRef]
  64. Wang, K.; Ouyang, H.; Xie, Z.; Yao, C.; Guo, N.; Li, M.; Jiao, H.; Pang, D. Efficient Generation of Myostatin Mutations in Pigs Using the CRISPR/Cas9 System. Sci. Rep. 2015, 5, 16623. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, K.; Tang, X.; Liu, Y.; Xie, Z.; Zou, X.; Li, M.; Yuan, H.; Ouyang, H.; Jiao, H.; Pang, D. Efficient Generation of Orthologous Point Mutations in Pigs via CRISPR-assisted ssODN-mediated Homology-directed Repair. Mol. Ther. Nucleic Acids 2016, 5, e396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Zou, Y.; Li, Z.; Zou, Y.; Hao, H.; Hu, J.; Li, N.; Li, Q. Generation of pigs with a Belgian Blue mutation in MSTN using CRISPR/Cpf1-assisted ssODN-mediated homologous recombination. J. Integr. Agric. 2019, 18, 1329–1336. [Google Scholar] [CrossRef]
  67. Li, M.; Tang, X.; You, W.; Wang, Y.; Chen, Y.; Liu, Y.; Yuan, H.; Gao, C.; Chen, X.; Xiao, Z.; et al. HMEJ-mediated site-specific integration of a myostatin inhibitor increases skeletal muscle mass in porcine. Mol. Ther. Nucleic Acids 2021, 26, 49–62. [Google Scholar] [CrossRef]
  68. Younis, S.; Schonke, M.; Massart, J.; Hjortebjerg, R.; Sundstrom, E.; Gustafson, U.; Bjornholm, M.; Krook, A.; Frystyk, J.; Zierath, J.R.; et al. The ZBED6-IGF2 axis has a major effect on growth of skeletal muscle and internal organs in placental mammals. Proc. Natl. Acad. Sci. USA 2018, 115, E2048–E2057. [Google Scholar] [CrossRef] [Green Version]
  69. Xiang, G.; Ren, J.; Hai, T.; Fu, R.; Yu, D.; Wang, J.; Li, W.; Wang, H.; Zhou, Q. Editing porcine IGF2 regulatory element improved meat production in Chinese Bama pigs. Cell. Mol. Life Sci. 2018, 75, 4619–4628. [Google Scholar] [CrossRef]
  70. Liu, X.; Liu, H.; Wang, M.; Li, R.; Zeng, J.; Mo, D.; Cong, P.; Liu, X.; Chen, Y.; He, Z. Disruption of the ZBED6 binding site in intron 3 of IGF2 by CRISPR/Cas9 leads to enhanced muscle development in Liang Guang Small Spotted pigs. Transgenic Res. 2019, 28, 141–150. [Google Scholar] [CrossRef]
  71. Rupp, R.; Senin, P.; Sarry, J.; Allain, C.; Tasca, C.; Ligat, L.; Portes, D.; Woloszyn, F.; Bouchez, O.; Tabouret, G.; et al. A Point Mutation in Suppressor of Cytokine Signalling 2 (Socs2) Increases the Susceptibility to Inflammation of the Mammary Gland while Associated with Higher Body Weight and Size and Higher Milk Production in a Sheep Model. PLoS Genet. 2015, 11, e1005629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Zhou, S.; Cai, B.; He, C.; Wang, Y.; Ding, Q.; Liu, J.; Liu, Y.; Ding, Y.; Zhao, X.; Li, G.; et al. Programmable Base Editing of the Sheep Genome Revealed No Genome-Wide Off-Target Mutations. Front. Genet. 2019, 10, 215. [Google Scholar] [CrossRef] [PubMed]
  73. Li, M.; Ouyang, H.; Yuan, H.; Li, J.; Xie, Z.; Wang, K.; Yu, T.; Liu, M.; Chen, X.; Tang, X.; et al. Site-Specific Fat-1 Knock-In Enables Significant Decrease of n-6PUFAs/n-3PUFAs Ratio in Pigs. G3-Genes Genom Genet. 2018, 8, 1747–1754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Zhang, J.; Cui, M.L.; Nie, Y.W.; Dai, B.; Li, F.R.; Liu, D.J.; Liang, H.; Cang, M. CRISPR/Cas9-mediated specific integration of fat-1 at the goat MSTN locus. FEBS J. 2018, 285, 2828–2839. [Google Scholar] [CrossRef] [Green Version]
  75. You, W.; Li, M.; Qi, Y.; Wang, Y.; Chen, Y.; Liu, Y.; Li, L.; Ouyang, H.; Pang, D. CRISPR/Cas9-Mediated Specific Integration of Fat-1 and IGF-1 at the pRosa26 Locus. Genes 2021, 12, 1027. [Google Scholar] [CrossRef]
  76. Zheng, Q.; Lin, J.; Huang, J.; Zhang, H.; Zhang, R.; Zhang, X.; Cao, C.; Hambly, C.; Qin, G.; Yao, J.; et al. Reconstitution of UCP1 using CRISPR/Cas9 in the white adipose tissue of pigs decreases fat deposition and improves thermogenic capacity. Proc. Natl. Acad. Sci. USA 2017, 114, E9474–E9482. [Google Scholar] [CrossRef] [Green Version]
  77. Carlson, D.F.; Lancto, C.A.; Zang, B.; Kim, E.S.; Walton, M.; Oldeschulte, D.; Seabury, C.; Sonstegard, T.S.; Fahrenkrug, S.C. Production of hornless dairy cattle from genome-edited cell lines. Nat. Biotechnol. 2016, 34, 479–481. [Google Scholar] [CrossRef]
  78. Sonstegard, T.S.; Carlson, D.; Lancto, C.A.; Fahrenkrug, S.C. Precision animal breeding as a sustainable, non-GMO solution for improving animal production and welfare. ASAP Anim. Prod. 2016, 31, 316–317. [Google Scholar]
  79. McCleary, S.; Strong, R.; McCarthy, R.R.; Edwards, J.C.; Howes, E.L.; Stevens, L.M.; Sanchez-Cordon, P.J.; Nunez, A.; Watson, S.; Mileham, A.J.; et al. Substitution of warthog NF-κB motifs into RELA of domestic pigs is not sufficient to confer resilience to African swine fever virus. Sci. Rep. 2020, 10, 8951. [Google Scholar] [CrossRef] [PubMed]
  80. Zhu, X.; Wei, Y.; Zhan, Q.; Yan, A.; Feng, J.; Liu, L.; Tang, D. CRISPR/Cas9-Mediated Biallelic Knockout of IRX3 Reduces the Production and Survival of Somatic Cell-Cloned Bama Minipigs. Animals 2020, 10, 501. [Google Scholar] [CrossRef] [Green Version]
  81. Zou, Y.; Li, Z.; Zou, Y.; Hao, H.; Li, N.; Li, Q. An FBXO40 knockout generated by CRISPR/Cas9 causes muscle hypertrophy in pigs without detectable pathological effects. Biochem. Biophys. Res. Commun. 2018, 498, 940–945. [Google Scholar] [CrossRef] [PubMed]
  82. Huang, J.; Wang, A.; Huang, C.; Sun, Y.; Song, B.; Zhou, R.; Li, L. Generation of Marker-Free pbd-2 Knock-in Pigs Using the CRISPR/Cas9 and Cre/loxP Systems. Genes 2020, 11, 951. [Google Scholar] [CrossRef] [PubMed]
  83. Xie, Z.; Jiao, H.; Xiao, H.; Jiang, Y.; Liu, Z.; Qi, C.; Zhao, D.; Jiao, S.; Yu, T.; Tang, X.; et al. Generation of pRSAD2 gene knock-in pig via CRISPR/Cas9 technology. Antivir. Res. 2020, 174, 104696. [Google Scholar] [CrossRef] [PubMed]
  84. Ikeda, M.; Matsuyama, S.; Akagi, S.; Ohkoshi, K.; Nakamura, S.; Minabe, S.; Kimura, K.; Hosoe, M. Correction of a Disease Mutation using CRISPR/Cas9-assisted Genome Editing in Japanese Black Cattle. Sci. Rep. 2017, 7, 17827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Zhou, S.; Yu, H.; Zhao, X.; Cai, B.; Ding, Q.; Huang, Y.; Li, Y.; Li, Y.; Niu, Y.; Lei, A.; et al. Generation of gene-edited sheep with a defined Booroola fecundity gene (FecBB) mutation in bone morphogenetic protein receptor type 1B (BMPR1B) via clustered regularly interspaced short palindromic repeat (CRISPR)/CRISPR-associated (Cas) 9. Reprod. Fertil. Dev. 2018, 30, 1616–1621. [Google Scholar] [CrossRef]
  86. Niu, Y.; Zhao, X.; Zhou, J.; Li, Y.; Huang, Y.; Cai, B.; Liu, Y.; Ding, Q.; Zhou, S.; Zhao, J.; et al. Efficient generation of goats with defined point mutation (I397V) in GDF9 through CRISPR/Cas9. Reprod. Fertil. Dev. 2018, 30, 307–312. [Google Scholar] [CrossRef]
  87. U.S. Food and Drug Administration. FDA Approves First-of-its-Kind Intentional Genomic Alteration in Line of Domestic Pigs for Both Human Food, Potential Therapeutic Uses. Available online: https://www.fda.gov/news-events/press-announcements/fda-approves-first-its-kind-intentional-genomic-alteration-line-domestic-pigs-both-human-food (accessed on 14 December 2020).
  88. U.S. Food and Drug Administration. FDA Makes Low-Risk Determination for Marketing of Products from Genome-Edited Beef Cattle After Safety Review. Available online: https://www.fda.gov/news-events/press-announcements/fda-makes-low-risk-determination-marketing-products-genome-edited-beef-cattle-after-safety-review (accessed on 7 March 2022).
  89. Ministry of Agriculture and Rural Affairs of China. Guidelines for Safety Evaluation of Gene Edited Plants for Agricultural Use (for Trial Implementation). Available online: http://www.moa.gov.cn/ztzl/zjyqwgz/sbzn/202201/t20220124_6387561.htm (accessed on 24 January 2022).
  90. Lema, M.A. Regulatory aspects of gene editing in Argentina. Transgenic Res. 2019, 28, 147–150. [Google Scholar] [CrossRef]
  91. Thygesen, P. Clarifying the regulation of genome editing in Australia: Situation for genetically modified organisms. Transgenic Res. 2019, 28, 151–159. [Google Scholar] [CrossRef]
  92. Fan, Z.; Mu, Y.; Sonstegard, T.; Zhai, X.; Li, K.; Hackett, P.B.; Zhu, Z. Social Acceptance for Commercialization of Genetically Modified Food Animals. Natl. Sci. Rev. 2021, 8, nwab067. [Google Scholar] [CrossRef]
  93. Fan, Z.; Wu, T.; Wu, K.; Mu, Y.; Li, K. Reflections on the system of evaluation of gene-edited livestock. Front. Agr. Sci. Eng. 2020, 7, 211–217. [Google Scholar] [CrossRef]
  94. Xu, K.; Zhang, X.; Liu, Z.; Ruan, J.; Xu, C.; Che, J.; Fan, Z.; Mu, Y.; Li, K. A transgene-free method for rapid and efficient generation of precisely edited pigs without monoclonal selection. Sci. China Life Sci. 2022, 1–12. [Google Scholar] [CrossRef] [PubMed]
  95. Wang, Y.; Bi, D.; Qin, G.; Song, R.; Yao, J.; Cao, C.; Zheng, Q.; Hou, N.; Wang, Y.; Zhao, J. Cytosine Base Editor (hA3A-BE3-NG)-Mediated Multiple Gene Editing for Pyramid Breeding in Pigs. Front. Genet. 2020, 11, 592623. [Google Scholar] [CrossRef] [PubMed]
  96. Yuan, H.; Yu, T.; Wang, L.; Yang, L.; Zhang, Y.; Liu, H.; Li, M.; Tang, X.; Liu, Z.; Li, Z.; et al. Efficient base editing by RNA-guided cytidine base editors (CBEs) in pigs. Cell. Mol. Life Sci. 2020, 77, 719–733. [Google Scholar] [CrossRef] [PubMed]
Figure 1. CRISPR/Cas systems mediated genetic manipulation. CRISPR-Cas systems allow multiple levels of genetic manipulation. (A) At the DNA level, Cas9, Cas9n and Cas12a are used for inducing dsDNA breaks for knock-out/deletion or knock-in/insertion. Cas9n can also be fused to base editors or primer editor to modify nucleotides in dsDNA for base substitution or base rewriting without introducing a dsDNA break. (B) During transcription, dCas9 can be fused to transcriptional activators, repressors, or epigenetic modifiers to activate or repress the transcription of single or multiple genes. (C) At the level of RNA, Cas13 can be used for targeted RNA manipulation. Cas13 can knock down specific RNA molecules by catalyzing RNA cleavage. Cas13 fused to base editors can be used to modify nucleotides in RNA molecules to achieve RNA base substitution or base rewriting. A single or multiple crRNA that bind to splice site motifs such as SA and SD combined with dCas13 protein can block specific exon from recognition by the splicing machinery, resulting in targeted RNA splicing. E1, exon 1; E2, exon 2; E3, exon 3; SA, splice acceptor site; SD, splice donor site.
Figure 1. CRISPR/Cas systems mediated genetic manipulation. CRISPR-Cas systems allow multiple levels of genetic manipulation. (A) At the DNA level, Cas9, Cas9n and Cas12a are used for inducing dsDNA breaks for knock-out/deletion or knock-in/insertion. Cas9n can also be fused to base editors or primer editor to modify nucleotides in dsDNA for base substitution or base rewriting without introducing a dsDNA break. (B) During transcription, dCas9 can be fused to transcriptional activators, repressors, or epigenetic modifiers to activate or repress the transcription of single or multiple genes. (C) At the level of RNA, Cas13 can be used for targeted RNA manipulation. Cas13 can knock down specific RNA molecules by catalyzing RNA cleavage. Cas13 fused to base editors can be used to modify nucleotides in RNA molecules to achieve RNA base substitution or base rewriting. A single or multiple crRNA that bind to splice site motifs such as SA and SD combined with dCas13 protein can block specific exon from recognition by the splicing machinery, resulting in targeted RNA splicing. E1, exon 1; E2, exon 2; E3, exon 3; SA, splice acceptor site; SD, splice donor site.
Ijms 23 07331 g001
Figure 2. Various antiviral strategies mediated by CRISPR/Cas systems. (1) Receptor deletion or modification: To enter the host cell, certain viruses use cell membrane receptors. CRISPR/Cas technologies can eliminate or modify these receptor genes, preventing virus-receptor binding and providing full resistance to animals. (2) Integration of antiviral genes or sgRNA pools and Cas9 or Cas13 proteins into animal genomes: Using the CRISPR/Cas technique, single or multiple antiviral genes or sgRNA pools and Cas9 or Cas13 genes can be integrated into the safe harbors of animal genome, the constant expression or spatio-temporal specific expression of these genes will interfere with or degrade viral RNA or DNA. (3) Using sgRNA/Cas as novel antivirus agents: sgRNAs polls targeting multiple viral sequences together with Cas9 or Cas13d proteins can be delivered to animal cells via viral delivery systems (Lentivirus or adeno-associated virus system) or non-viral delivery systems (Nanoparticles) to interfere with or degrade viral RNA or DNA.
Figure 2. Various antiviral strategies mediated by CRISPR/Cas systems. (1) Receptor deletion or modification: To enter the host cell, certain viruses use cell membrane receptors. CRISPR/Cas technologies can eliminate or modify these receptor genes, preventing virus-receptor binding and providing full resistance to animals. (2) Integration of antiviral genes or sgRNA pools and Cas9 or Cas13 proteins into animal genomes: Using the CRISPR/Cas technique, single or multiple antiviral genes or sgRNA pools and Cas9 or Cas13 genes can be integrated into the safe harbors of animal genome, the constant expression or spatio-temporal specific expression of these genes will interfere with or degrade viral RNA or DNA. (3) Using sgRNA/Cas as novel antivirus agents: sgRNAs polls targeting multiple viral sequences together with Cas9 or Cas13d proteins can be delivered to animal cells via viral delivery systems (Lentivirus or adeno-associated virus system) or non-viral delivery systems (Nanoparticles) to interfere with or degrade viral RNA or DNA.
Ijms 23 07331 g002
Table 1. List of genome-edited livestock animals.
Table 1. List of genome-edited livestock animals.
SpeciesGeneModification *Method *ApplicationsReferences
PigRELAKOZFNDisease resistance[79]
CD169KOHRDisease resistance[27]
CD163KOCRISPR/Cas9Disease resistance[29,30,31]
CD163 (SRCR 5 domain)hCD163L1 SRCR domain 8 replacementCRISPR/Cas9Disease resistance[32,33]
CD163 (SRCR 5 domain)Domain deletionCRISPR/Cas9Disease resistance[34,35,36]
HDAC6TGSCNTDisease resistance[47]
CD163Partial domain deletionCRISPR/Cas9Disease resistance[37]
pAPNKOCRISPR/Cas9Disease resistance[40]
CD163, pAPNDouble KOCRISPR/Cas9Disease resistance[42]
shRNATGInjectionDisease resistance[22]
CMAHKOCRISPR/Cas9Disease resistance[43]
IGF2Regulatory element mutationCRISPR/Cas9Meat production[69,70]
IRX3KOCRISPR/Cas9Fat content[80]
FBXO40KOCRISPR/Cas9Meat production[81]
PBD-2KICRISPR/Cas9Disease resistance[82]
MSTNPoint mutationCRISPR/Cas9Meat production[65]
MSTNPartial deletionCRISPR/Cpf1Meat production[66]
FSTKICRISPR/Cas9Meat production[67]
MSTNKOCRISPR/Cas9Meat production[62]
RSAD2KICRISPR/Cas9Disease resistance[83]
fat-1KICRISPR/Cas9Meat quality[73]
fat-1, IGF1Double KICRISPR/Cas9Meat quality and meat production[75]
UCP1KICRISPR/Cas9Fat content and animal welfare[76]
KISSRKOCRISPR/Cas9Animal welfare[78]
CattleSP110KITALENsDisease resistance[45]
NRAMP1/SLC11A1KICRISPR/Cas9Disease resistance[46]
CD18Point mutationZFNDisease resistance[24]
IARSKICRISPR/Cas9Animal welfare[84]
POLLED alleleKITALENsAnimal welfare[77]
SheepBMPR1BPoint mutationCRISPR/Cas9Reproductive traits[85]
MSTNKOCRISPR/Cas9Meat production[60]
SOCS2Point mutationBase EditorGrowth rate[72]
GoatGDF9KICRISPR/Cas9Reproductive traits[86]
fat-1, MSTNKI and KOCRISPR/Cas9Meat quality and meat production[74]
MSTNKOCRISPR/Cas9Meat production[61]
* KO, knock out. KI, knock in. TG, transgene. HR, homologous recombination.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Z.; Wu, T.; Xiang, G.; Wang, H.; Wang, B.; Feng, Z.; Mu, Y.; Li, K. Enhancing Animal Disease Resistance, Production Efficiency, and Welfare through Precise Genome Editing. Int. J. Mol. Sci. 2022, 23, 7331. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23137331

AMA Style

Liu Z, Wu T, Xiang G, Wang H, Wang B, Feng Z, Mu Y, Li K. Enhancing Animal Disease Resistance, Production Efficiency, and Welfare through Precise Genome Editing. International Journal of Molecular Sciences. 2022; 23(13):7331. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23137331

Chicago/Turabian Style

Liu, Zhiguo, Tianwen Wu, Guangming Xiang, Hui Wang, Bingyuan Wang, Zheng Feng, Yulian Mu, and Kui Li. 2022. "Enhancing Animal Disease Resistance, Production Efficiency, and Welfare through Precise Genome Editing" International Journal of Molecular Sciences 23, no. 13: 7331. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23137331

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