Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-29T00:28:49.335Z Has data issue: false hasContentIssue false

Functional analysis of bovine Nramp1 and production of transgenic cloned embryos in vitro

Published online by Cambridge University Press:  17 May 2013

Xiang Cheng
Affiliation:
Department of Animal Biotechnology, College of Veterinary Medicine, Northwest A&F University, Yangling, Shaanxi 712100, China.
Xiaoli Yu
Affiliation:
Department of Animal Biotechnology, College of Veterinary Medicine, Northwest A&F University, Yangling, Shaanxi 712100, China.
Yajun Liu
Affiliation:
Department of Animal Biotechnology, College of Veterinary Medicine, Northwest A&F University, Yangling, Shaanxi 712100, China.
Jie Deng
Affiliation:
Department of Animal Biotechnology, College of Veterinary Medicine, Northwest A&F University, Yangling, Shaanxi 712100, China.
Xiaoling Ma
Affiliation:
Department of Animal Biotechnology, College of Veterinary Medicine, Northwest A&F University, Yangling, Shaanxi 712100, China.
Huayan Wang*
Affiliation:
Department of Animal Biotechnology, College of Veterinary Medicine, Northwest A&F University, Yangling, Shaanxi 712100, China.
*
All correspondence to: Huayan Wang. Department of Animal Biotechnology, College of Veterinary Medicine, Northwest A&F University, Yangling, Shaanxi 712100, China. Tel: +86 029 87080069. Fax: +86 029 87080068. E-mail: hhwang101@163.com

Summary

Natural resistance-associated macrophage protein 1 (Nramp1) plays an important role in restraining the growth of intracellular pathogens within macrophages. In this study, Nramp1 cDNA was cloned from Qinchuan cattle and its anti-bacterial activity was demonstrated as being able to significantly inhibit the growth of Salmonella abortusovis and Brucella abortus in macrophages. Calf fibroblasts stably transfected with pSP–NRAMP1–HA vector were used to reconstruct bovine embryos by somatic cell nuclear transfer (SCNT). Reconstructed embryos were maturated in vitro and the blastocyst formation rate (14.0%) was similar to that of control embryos (14.5%). Transgenic blastocysts were transplanted into 43 recipient cattle, of which 14 recipients became pregnant as evidenced by non-return estrus and by rectal palpation. One fetus was aborted after 6½ months of pregnancy and transgene integration was confirmed by semi-quantitative polymerase chain reaction. Together, this study showed that bovine Nramp1 retains biological function against the growth of intracellular bacteria and can be used to reconstruct embryos and produce Nramp1 transgenic cattle, which may benefit the animal and enhance their ability to prevent attack by intracellular pathogens.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2013 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Barthel, R., Feng, J., Piedrahita, J.A., McMurray, D.N., Templeton, J.W. & Adams, L.G. (2001). Stable transfection of the bovine NRAMP1 gene into murine RAW264.7 cells: effect on Brucella abortus survival. Infect. Immun. 69, 3110–9.CrossRefGoogle ScholarPubMed
Barton, C.H., White, J.K., Roach, T.I. & Blackwell, J.M. (1994). NH2-terminal sequence of macrophage-expressed natural resistance-associated macrophage protein (Nramp) encodes a proline/serine-rich putative Src homology 3-binding domain. J. Exp. Med. 179, 1683–7.CrossRefGoogle ScholarPubMed
Blackwell, J.M., Barton, C.H., White, J.K., Roach, T.I., Shaw, M.A., Whitehead, S.H., Mock, B.A., Searle, S., Williams, H. & Baker, A.M. (1994). Genetic regulation of leishmanial and mycobacterial infections: the Lsh /Ity/Bcg gene story continues. Immunol. Lett. 43, 99107.CrossRefGoogle ScholarPubMed
Campbell, K.H., Fisher, P., Chen, W.C., Choi, I., Kelly, R.D., Lee, J.H. & Xhu, J. (2007). Somatic cell nuclear transfer: past, present and future perspectives. Theriogenology 68 (Suppl 1), S21431.CrossRefGoogle ScholarPubMed
Canonne-Hergaux, F., Calafat, J., Richer, E., Cellier, M., Grinstein, S., Borregaard, N. & Gros, P. (2002). Expression and subcellular localization of NRAMP1 in human neutrophil granules. Blood 100, 268–75.CrossRefGoogle ScholarPubMed
Cellier, M., Belouchi, A. & Gros, P. (1996). Resistance to intracellular infections: comparative genomic analysis of Nramp. Trends Genet. 12, 201–4.CrossRefGoogle ScholarPubMed
Cellier, M., Shustik, C., Dalton, W., Rich, E., Hu, J., Malo, D., Schurr, E. & Gros, P. (1997). Expression of the human NRAMP1 gene in professional primary phagocytes: studies in blood cells and in HL-60 promyelocytic leukemia. J. Leukoc. Biol. 61, 96105.CrossRefGoogle ScholarPubMed
Cheng, X., Deng, J., Meng, S.Y., Lai, W.F. & Wang, H.Y. (2011). The cloning and its expression in the cells and the tissues of the bovine natural resistance-associated macrophage protein 1 gene isoform (NRAMP1-ISO). Chin. J. Agric. Biotech. 19, 685–91.Google Scholar
Cheng, X. & Wang, H.Y. (2012). Multiple targeting motifs direct NRAMP1 into lysosomes. Biochem. Biophys. Res. Commun. 19, 578–83.CrossRefGoogle Scholar
Farin, P.W., Piedrahita, J.A. & Farin, C.E. (2006). Errors in development of fetuses and placentas from in vitro-produced bovine embryos. Theriogenology 65, 178–91.CrossRefGoogle ScholarPubMed
Feng, J., Li, Y., Hashad, M., Schurr, E., Gros, P., Adams, L.G. & Templeton, J.W. (1996). Bovine natural resistance associated macrophage protein 1 (Nramp1) gene. Genome Res. 6, 956–64.CrossRefGoogle ScholarPubMed
Forbes, J.R. & Gros, P. (2001). Divalent-metal transport by NRAMP proteins at the interface of host-pathogen interactions. Trends Microbiol. 9, 397403.CrossRefGoogle ScholarPubMed
Fritsche, G., Nairz, M., Theurl, I., Mair, S., Bellmann-Weiler, R., Barton, H.C. & Weiss, G. (2007). Modulation of macrophage iron transport by Nramp1 (Slc11a1). Immunobiology 212, 751–7.CrossRefGoogle ScholarPubMed
Govoni, G., Gauthier, S., Billia, F., Iscove, N.N. & Gros, P. (1997). Cell-specific and inducible Nramp1 gene expression in mouse macrophages in vitro and in vivo. J. Leukoc. Biol. 62, 277–86.CrossRefGoogle ScholarPubMed
Govoni, G., Canonne-Hergaux, F., Pfeifer, C.G., Marcus, S.L., Mills, S.D., Hackam, D.J., Grinstein, S., Malo, D., Finlay, B.B. & Gros, P. (1999). Functional expression of Nramp1 in vitro in the murine macrophage line RAW264.7. Infect. Immun. 67, 2225–32.CrossRefGoogle ScholarPubMed
Gruenheid, S., Pinner, E., Desjardins, M. & Gros, P. (1997). Natural resistance to infection with intracellular pathogens: the Nramp1 protein is recruited to the membrane of the phagosome. J. Exp. Med. 185, 717–30.CrossRefGoogle Scholar
He, W., Qiang, M., Ma, W., Valente, A.J., Quinones, M.P., Wang, W., Reddick, R.L., Xiao, Q., Ahuja, S.S., Clark, R.A., Freeman, G.L. & Li, S. (2006). Development of a synthetic promoter for macrophage gene therapy. Hum. Gene Ther. 17, 949–59.CrossRefGoogle ScholarPubMed
He, Y.L., Wu, Y.H., He, X.N., Liu, F.J., He, X.Y. & Zhang, Y. (2009). An immortalized goat mammary epithelial cell line induced with human telomerase reverse transcriptase (hTERT) gene transfer. Theriogenology 71, 1417–24.CrossRefGoogle ScholarPubMed
Jang, G., Bhuiyan, M.M., Jeon, H.Y., Ko, K.H., Park, H.J., Kim, M.K., Kim, J.J., Kang, S.K., Lee, B.C. & Hwang, W.S. (2006). An approach for producing transgenic cloned cows by nuclear transfer of cells transfected with human alpha 1-antitrypsin gene. Theriogenology 65, 1800–12.CrossRefGoogle ScholarPubMed
Kato, Y., Tani, T. & Tsunoda, Y. (2000). Cloning of calves from various somatic cell types of male and female adult, newborn and fetal cows. J. Reprod. Fertil. 120, 231–7.CrossRefGoogle ScholarPubMed
Lam-Yuk-Tseung, S., Picard, V. & Gros, P. (2006). Identification of a tyrosine-based motif (YGSI) in the amino terminus of Nramp1 (Slc11a1) that is important for lysosomal targeting. J. Biol. Chem. 281, 31677–88.CrossRefGoogle ScholarPubMed
Lee, S.L., Kumar, B.M., Kim, J.G., Ock, S.A., Jeon, B.G., Balasubramanian, S., Choe, S.Y. & Rho, G.J. (2007). Cellular composition and viability of cloned bovine embryos using exogene-transfected somatic cells. Reprod. Domest. Anim. 42, 4452.CrossRefGoogle ScholarPubMed
Lipiński, D., Duszewska, A.M., Zeyland, J., Mały, E., Gawron, W., Rynkowska, A., Reklewski, Z., & Słomski, R. (2007). Obtaining transgenic bovine skin fibroblasts containing human interferon alpha gene. Anim. Sci. Pap. Rep. 25, 211–20.Google Scholar
Malo, D., Vogan, K., Vidal, S., Hu, J., Cellier, M., Schurr, E., Fuks, A., Bumstead, N., Morgan, K. & Gros, P. (1994). Haplotype mapping and sequence analysis of the mouse Nramp gene predict susceptibility to infection with intracellular parasites. Genomics 23, 5161.CrossRefGoogle ScholarPubMed
Matsumoto, H.N., Tamura, M., Denhardt, D.T., Obinata, M. & Noda, M. (1995). Establishment and characterization of bone marrow stromal cell lines that support osteoclastogenesis. Endocrinology 136, 4084–91.CrossRefGoogle ScholarPubMed
Melo, E.O., Sousa, R.V., Iguma, L.T., Franco, M.M., Rech, E.L. & Rumpf, R. (2005). Isolation of transfected fibroblast clones for use in nuclear transfer and transgene detection in cattle embryos Genet. Mol. Res. 4, 812–21.Google ScholarPubMed
Peterson, A.J. & Lee, R.S. (2003). Improving successful pregnancies after embryo transfer. Theriogenology 59, 687–97.CrossRefGoogle ScholarPubMed
Powell, A.M., Talbot, N.C., Wells, K.D., Kerr, D.E., Pursel, V.G. & Wall, R.J. (2004). Cell donor influences success of producing cattle by somatic cell nuclear transfer. Biol. Reprod. 71, 210–6.CrossRefGoogle ScholarPubMed
Searle, S., Bright, N.A., Roach, T.I., Atkinson, P.G., Barton, C.H., Meloen, R.H. & Blackwell, J.M. (1998). Localisation of Nramp1 in macrophages: modulation with activation and infection. J. Cell Sci. 111 (Pt 19), 2855–66.CrossRefGoogle ScholarPubMed
Stabel, J.R. & Stabel, T.J. (1995). Immortalization and characterization of bovine peritoneal macrophages transfected with SV40 plasmid DNA. Vet. Immunol. Immunopathol. 45, 211–20.CrossRefGoogle ScholarPubMed
Stober, C.B., White, J.W., Popoff, J-F. & Blackwell, J.M. (2007). Slc11a1, formerly Nramp1, is expressed in dendritic cells and influences major histocompatibility complex class II expression and antigen-presenting cell function Infect. Immun. 75, 5059–67.CrossRefGoogle ScholarPubMed
Vidal, S.M., Malo, D., Vogan, K., Skamene, E. & Gros, P. (1993). Natural resistance to infection with intracellular parasites: isolation of a candidate for Bcg. Cell 73, 469–85.CrossRefGoogle ScholarPubMed
Vidal, S., Tremblay, M.L., Govoni, G., Gauthier, S., Sebastiani, G., Malo, D., Skamene, E., Olivier, M., Jothy, S. & Gros, P. (1995). The Ity/Lsh/Bcg locus: natural resistance to infection with intracellular parasites is abrogated by disruption of the Nramp1 gene. J. Exp. Med. 182, 655–66.CrossRefGoogle ScholarPubMed
Wen, Y.A., Guo, J.Y., Yu, X.L., Cen, D., Liu, J.B., Chen, Y., Zhang, J. & Tu, Z.G. (2008). Construction and targeting study of eukaryotic expression vector modulated by a macrophage-specific promoter. Xi Bao Yu Fen Zi Mian Yi Xue Za Zhi 24, 441–3.Google ScholarPubMed
Wen, Y.A., Yu, X.L., Xia, Q.F., Cen, D., Liu, J.B. & Tu, Z.G. (2009). Macrophage-specific overexpression of antimicrobial peptide PR-39 inhibits intracellular growth of Salmonella enterica serovar typhimurium in macrophage cells. Int. J. Antimicrob. Agents 34, 315–21.CrossRefGoogle ScholarPubMed
Zhang, Y.L., Wan, Y.J., Wang, Z.Y., Xu, D., Pang, X.S., Meng, L., Wang, L.H., Zhong, B.S. & Wang, F. (2010). Production of dairy goat embryos, by nuclear transfer, transgenic for human acid beta-glucosidase. Theriogenology 73, 681–90.CrossRefGoogle ScholarPubMed
Zhao, M.T., Lin, H., Liu, F.J., Quan, F.S., Wang, G.H., Liu, J., Hua, S. & Zhang, Y. (2009). Efficiency of human lactoferrin transgenic donor cell preparation for SCNT. Theriogenology 71, 376–84.CrossRefGoogle ScholarPubMed