丁香实验_LOGO
登录
提问
我要登录
|免费注册
点赞
收藏
wx-share
分享

Mouse Strains and Genetic Nomenclature

互联网

1466
  • Abstract
  • Table of Contents
  • Figures
  • Literature Cited

Abstract

 

In this article we describe the main characteristics and peculiarities of the different strains and stocks of laboratory animals from the genetic point of view. We explain how they are produced and maintained as well as their advantages and disadvantages in the context of animal experiments. We also provide some guidance to make the best possible choice when establishing an experimental protocol. Curr. Protoc. Mouse Biol. 1:213?238. © 2011 by John Wiley & Sons, Inc.

Keywords: inbred strains; congenic strains; recombinant inbred strains; recombinant congenic strains; outbred stocks

     
 
GO TO THE FULL PROTOCOL:
PDF or HTML at Wiley Online Library

Table of Contents

  • The Different Categories of Laboratory Strains and Stocks
  • Inbred Strains and Their Derivatives
  • Outbred and Randombred Stocks
  • Nomenclature Rules for Mouse and Rat Strains
  • Conclusions
  • Warning!
  • Literature Cited
  • Figures
  • Tables
     
 
GO TO THE FULL PROTOCOL:
PDF or HTML at Wiley Online Library

Materials

 
GO TO THE FULL PROTOCOL:
PDF or HTML at Wiley Online Library

Figures

  •   Figure 1. This drawing schematically represents the breeding system that is commonly used to produce an inbred strain: mating a male and a female from the same litter (brother × sister) in successive generations. Theoretical computation would indicate that exceptional matings between parent and offspring would not affect the progress towards homozygosity provided that the parent that is selected for mating is always the youngest. A male, for example, can be used for mating with one of its daughters but not with a female offspring born from this cross. Each generation of inbreeding is symbolized by the uppercase letter F, followed by the number of generations. When this number is not known, a question mark is often used; for example, F ? + 27 would indicate that the number of brother × sister matings was not known when the strain was acquired but 27 generations of unrelaxed inbreeding have been added since this time. B, brother; S, sister.
    View Image
  •   Figure 2. The curve was drawn based on the Fibonacci's series and represents faithfully the cumulated percentage of genes that have become fixed in the homozygous state as inbreeding proceeds. From generation F5 onwards, this percentage is incremented by 19.6% at each generation.
    View Image
  •   Figure 3. This table was captured from a window from the Mouse Phenome Database. It represents the complete set of data for blood cholesterol performed on both male and female mice of 43 different inbred strains in Dr. B. Paigen's laboratory at the Jackson Laboratory. These data (in mg/deciliter) correspond to baseline data from mice aged 7 to 9 weeks. Checking these data before embarking on a research project related to cholesterol metabolism is definitely of great help (http://phenome.jax.org/db/qp?rtn=views/measplot&brieflook=9904). Arrowheads indicate exceptionally high or low values.
    View Image
  •   Figure 4. Historical data, confirmed by sequence data, indicate that modern laboratory inbred strains derive from a small number of ancestors belonging to several different subspecies of the genus Mus . Today's classical laboratory inbred strains should be regarded as recombinant strains derived (in unequal percentages) from three parental components: Mus musculus domesticus , Mus musculus musculus , and Mus musculus castaneus . For this reason it would probably be more appropriate to designate them as Muslaboratorius ”! This heterogeneous and unnatural genetic constitution is detectable at the genomic/sequence level by variations in the density of polymorphisms (in particular SNPs), with sharp edges, as represented on the picture by color differences for some chromosomal regions. The pattern of DNA polymorphisms distribution along the different chromosomal regions varies according to the strain and can be used for the purpose of mapping complex (QTL) traits.
    View Image
  •   Figure 5. (A ) This picture represents a mouse family tree. The 102 inbred strains represented on this picture have been genotyped for a set of 1,638 informative SNP markers, evenly distributed over the whole genome (spaced on average <1.5 Mb). Applying a neighbor‐joining method to the data, the authors constructed a mouse strain family tree that could be organized into seven groups. The length and angle of the branches have been optimized for printing and do not reflect the actual evolutionary distances between strains. This family tree is in good agreement with most other existing genealogies. Adapted from Petkov et al. (). (B ) This picture represents a phylogenetic family tree of 93 rat inbred strains. It was developed with the same technique as above (heuristic search for maximum parsimony). Adapted from Mashimo et al. ().
    View Image
  •   Figure 6. This picture schematically represents the different breeding schemes that can be used for keeping an inbred strain with three cages. (A ) The first schemes consist of mating one male and one female in each of the three cages, at each generation, with the breeders being brother (B) and sister (S) born in the same cage. This system cannot be recommended because in case of infertility or death one line will be lost, and in case of full success (unlikely!), one would end up with three loosely related substrains instead of one single strain. (B ) This breeding scheme is the most reliable. It consists of selecting the progeny at generation F, which would allow making the largest number of pairs of breeders at generation F+1. It is very safe, but, unfortunately, it is not always applicable. (C ) This system is an intermediate between A and B, and is the most popular in practice.
    View Image
  •   Figure 7. This picture represents the successive steps in the establishment of a congenic strain. At each generation, a breeder carrying the targeted character ( marked with a dot ) is back‐crossed to a partner of the recipient (B) strain.
    View Image
  •   Figure 8. After each back‐cross generation, an average of 50% of the genomic DNA of the donor strain is replaced by the equivalent proportion of genomic DNA of the background strain. However, selecting at each back‐cross generation, the breeder with the lowest percentage of introgressed (donor) DNA greatly accelerates the establishment of a congenic strain. This picture represents the breeder (boxed) that was recognized as the most interesting (“best breeder”) after genotyping because of the lowest percentage of donor genome (shown in red). The mutation or region of interest is indicated by an arrowhead.
    View Image
  •   Figure 9. Selecting the best breeders at each back‐cross generation can save a lot of time when establishing a congenic strain. It is important to note that genotyping requires many polymorphic DNA markers only for the first back‐cross progenies. Once a marker is characterized homozygous, it is no longer necessary to type it in the forthcoming generations. The y axis represents the number of mice with a percentage of homozygous loci lower or equal to the value indicated on the x axis. Adapted from Wakeland et al. ().
    View Image
  •   Figure 10. The genomic DNA flanking the target (introgressed) gene or transgene can be estimated very precisely with a battery of molecular markers, generation after generation. When, by chance, a cross‐over reduces the flanking DNA on one side or the other, the mouse then becomes a privileged breeder. The genotyping of the flanking regions with the aim of selecting the “best breeder” can be integrated in the overall genotyping protocol. Briefly, it is recommended to select and type many more markers in the flanking regions than in other regions of the genome. The probability of finding a mouse with one crossing‐over on each side of the flanking region is extremely low. For this reason it is recommended to proceed in two discrete steps, sequentially breeding mice with cross‐overs in a flanking region of the targeted gene. If we consider a locus A, which is c Morgans distant from the selected (or targeted) locus T, on the same chromosome, the probability that there has been no recombination between the two loci A and T is e c per generation and therefore e nc after n generations (Johnson, ). After 10 generations of back‐crossing, there is a 37% chance that no recombination occurred between loci A and T if they are distant 10 centimorgans (cM). This will rise to 61% if A and T are 5 cM apart and up to 90% if the two loci are 1 cM apart. If we consider that there are approximately 27,000 genes in the mouse genome and that the genetic map in this species spans ∼1,520 cM on average, this would mean that in 90% of the cases two congenic strains differ by ∼17 genes on both sides of the introgressed locus T. Of course, this is far from being negligible even if it should be weighted by the fact that no more than 20% of the genes are polymorphic between any two inbred strains.
    View Image
  •   Figure 11. Reciprocal congenics allow making comparisons with a high degree of standardization because epistatic interactions may be detected by this procedure. For example, if a given allele in the background of a congenic strain interferes with the expression of the target gene, this might be detected when comparing the two reciprocal congenic strains but would presumably be corrected when comparing the two congenic F1s.
    View Image
  •   Figure 12. This picture represents a set of four consomic strains flanked by the parental strains A and B. In each strain, a complete chromosome pair has been replaced by the homologous chromosome of the other inbred strain via a series of marker‐assisted back‐crosses. A complete panel of consomic strains consists of 21 strains, each derived from the same donor and host but having a different chromosome (Chr 1‐19, X or Y) of the host replaced by its counterpart from the donor. A reciprocal panel can be produced by inverting the donor and host strains, respectively.
    View Image
  •   Figure 13. This diagram represents a set of four recombinant inbred (R.I.) strains flanked by the parental strains A and B. Individual RI strains have a unique combination of loci derived by recombination of the alleles present in the original parental strains. Since RI strains are inbred and each strain has a unique genotype, RI strains have a number of advantages over F2 or back‐cross mouse populations as tools for mapping genes or quantitative trait loci (QTL).
    View Image
  •   Figure 14. (A ) The Collaborative Cross is a randomized cross of eight unrelated mouse inbred strains designed by members of the Complex Trait Consortium. The lines are first crossed pair‐wise to make all 56 possible G1 parents. A set of possible 4‐way crosses is performed, keeping Y‐chromosome and mitochondrial balance. Finally, all 8 genomes are brought together in G2:F1 and the offspring of this cross are inbred. The Collaborative Cross is a community resource that was initially designed for the purpose of mapping complex traits. (B ) The initial previsions were to breed around 1000 inbred strains where all the alleles of the initial inbred strains would be associated in a wide and unique variety of combinations. Only one strain is represented in this illustration; other strains would be similar but with a different pattern of parental strain distribution.
    View Image

Videos

Literature Cited

   Bailey, D.W. 1971. Recombinant inbred strains, an aid to finding identity, linkage, and function of histocompatibility and other genes. Transplantation 11:325‐327.
   Beck, J.A., Lloyd, S., Hafezparast, M., Lennon‐Pierce, M., Eppig, J.T., Festing, M.F., and Fisher, E.M. 2000. Genealogies of mouse inbred strains. Nat. Genet. 24:23‐25.
   Bishop, C.E., Boursot, P., Baron, B., Bonhomme, F., and Hatat, D. 1985. Most classical Mus musculus domesticus laboratory mouse strains carry a Mus musculus musculus Y chromosome. Nature 315:70‐72.
   Bonhomme, F. and Guénet, J.‐L. 1996. The laboratory mouse and its wild relatives. In Genetic Variants and Strains of the Laboratory Mouse ( M.F. Lyon, S. Rastan. and S.D.M. Brown, eds.) pp. 1577‐1596. Oxford University Press, Oxford.
   Bonhomme, F., Guénet, J.‐L., Dod, B., Moriwaki, K., and Bulfield, G. 1987. The polyphyletic origin of laboratory inbred mice and their rate of evolution. Biol. J. Linnean Soc. 30:51‐58.
   Bulfield, G., Siller, W.G., Wight, P.A., and Moore, K.J. 1984. X chromosome‐linked muscular dystrophy (mdx) in the mouse. Proc. Natl. Acad. Sci. U.S.A. 81:1189‐1192.
   Burgio, G., Szatanik, M., Guénet, J.‐L., Arnau, M.R., Panthier, J.J., and Montagutelli, X. 2007. Interspecific recombinant congenic strains between C57BL/6 and mice of the Mus spretus species: A powerful tool to dissect genetic control of complex traits. Genetics 177:2321‐2333.
   Burgio, G., Baylac, M., Heyer, E., and Montagutelli, X. 2009. Genetic analysis of skull shape variation and morphological integration in the mouse using interspecific recombinant congenic strains between C57BL/6 and mice of the Mus spretus species. Evolution 63:2668‐2686.
   Charlesworth, D. and Willis, J.H. 2009. The genetics of inbreeding depression. Nat. Rev. Genet. 10:783‐796.
   Chen, S., Kadomatsu, K., Kondo, M., Toyama, Y., Toshimori, K., Ueno, S., Miyake, Y., and Muramatsu, T. 2004. Effects of flanking genes on the phenotypes of mice deficient in basigin/CD147. Biochem. Biophys. Res. Commun. 324:147‐153.
   Chesler, E.J., Miller, D.R., Branstetter, L.R., Galloway, L.D., Jackson, B.L., Philip, V.M., Voy, B.H., Culiat, C.T., Threadgill, D.W., Williams, R.W., Churchill, G.A., Johnson, D.K., and Manly, K.F. 2008. The Collaborative Cross at Oak Ridge National Laboratory: Developing a powerful resource for system genetics. Mamm. Genome. 19:382‐389.
   Chia, R., Achilli, F., Festing, M.F., and Fisher E.M.C. 2005. The origins and uses of mouse outbred stocks. Nat. Genet. 37:1181‐1186.
   Churchill, G.A., Airey, D.C., Allayee, H., Angel, J.M., Attie, A.D., Beatty, J., Beavis, W.D., Belknap, J.K., Bennett, B., Berrettini, W., Bleich, A., Bogue, M., Broman, K.W., Buck, K.J., Buckler, E., Burmeister, M., Chesler, E.J., Cheverud, J.M., Clapcote, S., Cook, M.N., Cox, R.D., Crabbe, J.C., Crusio, W.E., Darvasi, A., Deschepper, C.F., Doerge, R.W., Farber, C.R., Forejt, J., Gaile, D., Garlow, S.J., Geiger, H., Gershenfeld, H., Gordon, T., Gu, J., Gu, W., de Haan, G., Hayes, N.L., Heller, C., Him‐melbauer, H., Hitzemann, R., Hunter, K., Hsu, H.C., Iraqi, F.A., Ivandic, B., Jacob, H.J., Jansen, R.C., Jepsen, K.J., Johnson, D.K., Johnson, T.E., Kempermann, G., Kendziorski, C., Kotb, M., Kooy, R.F., Llamas, B., Lammert, F., Lassalle, J.M., Lowenstein, P.R., Lu, L., Lusis, A., Manly, K.F., Marcucio, R., Matthews, D., Medrano, J.F., Miller, D.R., Mittleman, G., Mock, B.A., Mogil, J.S., Montagutelli, X., Morahan, G., Morris, D.G., Mott, R., Nadeau, J.H., Nagase, H., Nowakowski, R.S., O'Hara, B.F., Osadchuk, A.V., Page, G.P., Paigen, B., Paigen, K., Palmer, A.A., Pan, H.J., Peltonen‐Palotie, L., Peirce, J., Pomp, D., Pravenec, M., Prows, D.R., Qi, Z., Reeves, R.H., Roder, J., Rosen, G.D., Schadt, E.E., Schalkwyk, L.C., Seltzer, Z., Shimomura, K., Shou, S., Sillanpaa, M.J., Siracusa, L.D., Snoeck, H.W., Spearow, J.L., Svenson, K., Tarantino, L.M., Threadgill, D., Toth, L.A., Valdar, W., de Villena, F.P., Warden, C., Whatley, S., Williams, R.W., Wiltshire, T., Yi, N., Zhang, D., Zhang, M., Zou, F., and the Complex Trait Consortium. 2004. The Collaborative Cross: A community resource for the genetic analysis of complex traits. Nat. Genet. 36:1133‐1137.
   Davisson, M.T. 1996. Rules for nomenclature of inbred strains In Genetic Variants and Strains of the Laboratory Mouse (M.F. Lyon, S. Rastan, and S.D.M. Brown, eds.) pp. 1532‐1536. Oxford University Press, Oxford.
   Dejager, L., Libert, C., and Montagutelli, X. 2009. Thirty years of Mus spretus: A promising future. Trends Genet. 25:234‐241.
   Demant, P. 2003. Cancer susceptibility in the mouse: Genetics, biology and implications for human cancer. Nat. Rev. Genet. 4:721‐734.
   Demant, P. and Hart, A.A.M. 1986. Recombinant congenic strains: A new tool for analyzing genetic traits determined by more than one gene. Immunogenetics 24:416‐422.
   Ferris, S.D., Sage, R.D., and Wilson, A.C. 1982. Evidence from mtDNA sequences that common laboratory strains of inbred mice are descended from a single female. Nature 14:163‐165.
   Festing, M.F. 2010. Inbred strains should replace outbred stocks in toxicology, safety testing, and drug development. Toxicol. Pathol. 38:681‐690.
   Flint, J., Valdar, W., Shifman, S., and Mott, R. 2005. Strategies for mapping and cloning quantitative trait genes in rodents. Nat. Rev. Genet. 4:271‐286.
   Frazer, K.A., Eskin, E., Kang, H.M., Bogue, M.A., Hinds, D.A., Beilharz, E.J., Gupta, R.V., Montgomery, J., Morenzoni, M.M., Nilsen, G.B., Pethiyagoda, C.L., Stuve, L.L., Johnson, F.M., Daly, M.J., Wade, C.M., and Cox, D.R. 2007. A sequence‐based variation map of 8.27 million SNPs in inbred mouse strains. Nature 448:1050‐1053.
   Gregorová, S., Divina, P., Storchova, R., Trachtulec, Z., Fotopulosova, V., Svenson, K.L., Donahue, L.R., Paigen, B., and Forejt, J. 2008. Mouse consomic strains: Exploiting genetic divergence between Mus M. musculus and Mus m. domesticus subspecies. Genome Res. 18:509‐515.
   Guénet, J.‐L. and Bonhomme, F. 2003. Wild mice: An ever‐increasing contribution to a popular mammalian model. Trends Genet. 19:24‐31.
   Hartl, D.L. 2001. Genetic management of outbred laboratory rodent populations. Charles River Genetic Literature. http://www.criver.com/sitecollectiondocuments/rm_gt_r_genetic_management_outbred_rodent.pdf
   Hummel, K.P., Coleman, D.L., and Lane, P.W. 1972. The influence of genetic background on expression of mutations at the diabetes locus in the mouse. Biochem. Genet. 7:1‐13.
   Jackson, A.U., Fornes, A., Galecki, A., Miller, R.A., and Burke, D.T. 1999. Multiple‐trait quantitative trait loci analysis using a large mouse sibship. Genetics 151:785‐795.
   Johnson, L.L. 1981. At how many histocompatibility loci do congenic mouse strains differ? Probability estimates and some implications. J. Hered. 72:27‐31.
   Li, R., Lyons, M.A., Wittenburg, H., Paigen, B., and Churchill, G.A. 2005. Combining data from multiple inbred line crosses improves the power and resolution of quantitative trait loci mapping. Genetics 169:1699‐1709.
   Linder, C.C. 2001. The influence of genetic background on spontaneous and genetically engineered mouse models of complex diseases. Lab Anim. 30:34‐39.
   Markel, P., Shu, P., Ebeling, C., Carlson, G.A., Nagle, D.L., Smutko, J.S., and Moore, K.J. 1997. Theoretical and empirical issues for marker‐ assisted breeding of congenic mouse strains. Nat. Genet. 17:280‐284.
   Mashimo, T., Lucas, M., Simon‐Chazottes, D., Frenkiel, M.P., Montagutelli, X., Ceccaldi, P.E., Deubel, V., Guénet, J.L., and Despres, P. 2002. A nonsense mutation in the gene encoding 2′‐5′‐oligoadenylate synthetase/L1 isoform is associated with West Nile virus susceptibility in laboratory mice. Proc. Natl. Acad. Sci. U.S.A. 99:11311‐11316.
   Mashimo, T., Voigt, B., Tsurumi, T., Naoi, K., Nakanishi, S., Yamasaki, K., Kuramoto, T., and Serikawa, T. 2006. A set of highly informative rat simple sequence length polymorphism (SSLP) markers and genetically defined rat strains. BMC Genet. 7:19.
   Mattson, D.L., Dwinell, M.R., Greene, A.S., Kwitek, A.E., Roman, R.J., Jacob, H.J., and Cowley, A.W. Jr. 2008. Chromosome substitution reveals the genetic basis of Dahl salt‐sensitive hypertension and renal disease. Am. J. Physiol. Renal Physiol. 295:837‐842.
   Morse, H. C. III. 1978. Origins of Inbred Mice, Academic Press, San Diego.
   Mott, R., Talbot, C.J, Turri, M.G., Collins, A.C., and Flint, J. 2000. A method for fine mapping quantitative trait loci in outbred animal stocks. Proc. Natl. Acad. Sci. U.S.A. 97:12649‐12654.
   Nadeau, J., Singer, J., Matin, A., and Lander, E. 2000. Analyzing complex genetic traits with chromosome substitution strains. Nat. Genet. 24:221‐225.
   Ogonuki, N., Inoue, K., Hirose, M., Miura, I., Mochida, K., Sato, T., Mise, N., Mekada, K., Yoshiki, A., Abe, K., Kurihara, H., Wakana, S., and Ogura, A. 2009. A high‐speed congenic strategy using first‐wave male germ cells. PLoS ONE. 4:e4943.
   Paigen, K. and Eppig, J.T. 2000. A mouse phenome project. Mamm. Genome 11:715‐717.
   Petkov, P.M., Ding, Y., Cassell, M.A., Zhang, W., Wagner, G., Sargent, E.E., Asquith, S., Crew, V., Johnson, K.A., Robinson, P., Scott, V.E., and Wiles, M.V. 2004. An efficient SNP system for mouse genome scanning and elucidating strain relationships. Genome Res. 14:1806‐1811.
   Poiley, S.M. 1960. A systematic method of breeder rotation for non‐inbred laboratory animals colonies. Proc. Anim. Care Panel 10:159.
   Poltorak, A., He, X., Smimova, I., Liu, M.Y., Van Huffel, C., Du, X., Birdwell, D., Alejos, E., Silva, M., Galanos, C., Freudenberg, M., Ricciardi‐Castagnoli, P., Layton, B., and Beutler, B. 1998. Defective LPS signaling in C3H/HeJ and C57BL/10ScCr mice: Mutations in Tlr4 gene. Science 282:2085‐2088.
   Rosenstreich, D.L. and Glode, L.M. 1975. Difference in B cell mitogen responsiveness between closely related strains of mice. J. Immunol. 115:777‐780.
   Schuster‐Gossler, K., Lee, A.W., Lerner, C.P., Parker, H.J., Dyer, V.W., Scott, V.E., Gossler, A., and Conover, J.C. 2001. Use of coisogenic host blastocysts for efficient establishment of germline chimeras with C57BL/6J ES cell lines. Biotechniques 31:1022‐1026.
   Silver, L.M. 1995. Mouse Genetics. Concepts and applications. Oxford University Press, Oxford. Also see Internet Resources.
   Snell, G.D. 1948. Methods for the study of histocompatibility genes. J. Genet. 49:86‐108.
   Specht, C.G. and Schoepfer, R. 2001. Deletion of the alphasynuclein locus in a subpopulation of C57BL/6J inbred mice. BMC Neurosci. 2:11.
   Stevens, J.C., Gareth T., Banks, G.T., Festing, M.F.W. and Fisher, E.M.C. 2007. Quiet mutations in inbred strains of mice. Trends Mol. Med. 13:512‐519.
   Sundberg, J.P. and Schofield, P.N. 2010. Commentary: Mouse genetic nomenclature: Standardization of strain, gene, and protein symbols. Vet. Pathol. 47:1100‐1104.
   Threadgill, D.W., Dlugosz, A.A., Hansen, L.A., Tennenbaum, T., Lichti, U., Yee, D., La Mantia, C., Mourton, T., Herrup, K., and Harris, R.C. 1995. Targeted disruption of mouse EGF receptor: Effect of genetic background on mutant phenotype. Science 269:230‐234.
   Threadgill, D.W., Hunter, K.W., and Williams, R.W. 2002. Genetic dissection of complex and quantitative traits: from fantasy to reality via a community effort. Mamm. Genome 13:175‐178.
   Tucker, P.K., Lee, B.K., Lundrigan, B.L., and Eicher, E.M. 1992. Geographic origin of the Y chromosomes in “old” inbred strains of mice. Mamm. Genome. 3:254‐261.
   Wade, C.M., Kulbokas, E.J. 3rd, Kirby, A.W., Zody, M.C., Mullikin, J.C., Lander, E.S., Lindblad‐Toh, K., and Daly, M.J. 2002. The mosaic structure of variation in the laboratory mouse genome. Nature 420:574‐578.
   Wakeland, E., Morel, L., Achey, K., Yui, M., and Longmate, J. 1997. Speed congenics: A classic technique in the fast lane (relatively speaking). Immunol. Today. 18:472‐477.
   Wolfer, D.P., Crusio, W.E. and Lipp, H.P. 2002. Knockout mice: Simple solutions to the problems of genetic background and flanking genes. Trends Neurosci. 25:336‐340.
   Yalcin, B., Nicod, J., Bhomra, A., Davidson, S., Cleak, J., Farinelli, L., Osterås, M., Whitley, A., Yuan, W., Gan, X., Goodson, M., Klenerman, P., Satpathy, A., Mathis, D., Benoist, C., Adams, D.J., Mott, R., and Flint, J. 2010. Commercially available outbred mice for genome‐wide association studies. PLoS Genet. 6:e1001085.
   Yonekawa, H., Moriwaki, K., Gotoh, O., Miyashita, N., Migita, S., Bonhomme, F., Hjorth, J.P., Petras, M.L., and Tagashira, Y. 1982. Origins of laboratory mice deduced from restriction patterns of mitochondrial DNA. Differentiation 22:222‐226.
   Zou, F., Gelfond, J.A., Airey, D.C., Lu, L., Manly, K.F., Williams, R.W., and Threadgill, D.W. 2005. Quantitative trait locus analysis using recombinant inbred intercrosses: Theoretical and empirical considerations. Genetics 170:1299‐1311.
INTERNET RESOURCES
   http://www.informatics.jax.org/mgihome/nomen/strains.shtml
   In 2001, the International Committee on Standardized Nomenclature for Mice (Chairperson: Dr. Janan T. Eppig) and the Rat Genome and Nomenclature Committee (Chairperson: Dr. Göran Levan) agreed to establish a joint set of rules for strain nomenclature, applicable to strains of both species. These guidelines are updated annually by the international nomenclature committees. The official Web site for these guidelines may be found at the above URL.
   http://phenome.jax.org/
   Official Web site of The Mouse Phenome Project, an international collaboration representing five countries in both the academic and corporate sectors. Its aim is to establish a collection of baseline phenotypic data on commonly used and genetically diverse inbred mouse strains through a coordinated effort.
   http://www.informatics.jax.org/external/festing/search_form.cgi
   An annotated list of mouse and rat inbred strain.
   http://www.isogenic.info/html/animal_models_in_research.html
   Dr. M. F. W. Festing's Web site about the best use of animal models.
   http://www.informatics.jax.org/morsebook/
   Electronic version of the book by Herbert C. Morse III, Origins of Inbred Mice, Academic Press. 1978.
   http://www.informatics.jax.org/silverbook/
   Electronic version of the book by Lee M. Silver. 1995. Mouse Genetics: Concepts and Applications, Oxford University Press.
GO TO THE FULL PROTOCOL:
PDF or HTML at Wiley Online Library
 
提问
扫一扫
丁香实验小程序二维码
实验小助手
丁香实验公众号二维码
关注公众号
反馈
TOP
打开小程序