Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Maximizing mouse cancer models

Key Points

  • The laboratory mouse (Mus musculus) is one of the best model systems for investigations of cancer biology in vivo, ranging from basic models such as xenograft tumours derived from tumour cell lines or explants, to highly complex genetically engineered mice (GEM).

  • We suggest that xenografts should no longer be referred to as mouse cancer models. Xenografts represent an intermediate step between cell culture and mouse cancer models, and could be more accurately termed 'animal culture'.

  • GEM can be classified as either transgenic or endogenous. This distinction is not merely semantic but is highly relevant, because the type of GEM can determine the experimental outcome in certain situations.

  • Transgenic GEM are mutant mice that express oncogenes or dominant-negative tumour-suppressor genes (TSGs) in a non-physiological manner owing to ectopic promoter and enhancer elements. Advantages of transgenic GEM include the ability to reversibly control target-gene expression with exogenous ligands. One disadvantage is that it might be difficult to elicit the exquisite control necessary to express oncogenes at physiological levels.

  • Endogenous GEM represent mutant mice that lose the expression of TSGs or express dominant-negative TSGs or oncogenes from their native promoters through the use of knockout and knockin technology. Conditional GEM models rely on the use of site-specific recombinases, such as Cre, to control the spatiotemporal mutation of the mouse genome. The use of these conditional models will prove to be key in addressing important molecular and therapeutic questions.

  • Modern GEM are poised to explore facets of cancer biology and medicine that are difficult or impossible to pursue clinically. However, all GEM described so far have certain shortcomings in mimicking human malignancy. Several issues (such as humanizing mice) and practical considerations concerning GEM will need to be addressed in order to meet our objectives.

Abstract

Animal models of cancer provide an alternative means to determine the causes of and treatments for malignancy, thus representing a resource of immense potential for cancer medicine. The sophistication of modelling cancer in mice has increased to the extent that investigators can both observe and manipulate a complex disease process in a manner impossible to perform in patients. However, owing to limitations in model design and technology development, and the surprising underuse of existing models, only now are we realising the full potential of mouse models of cancer and what new approaches are needed to derive the maximum value for cancer patients from this investment.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Transgenic GEM.
Figure 2: Endogenous GEM.
Figure 3: Sequential mutations in mouse cancer models.
Figure 4: Hallmarks of humanized mouse models.

Similar content being viewed by others

References

  1. Hanahan, D. & Weinberg, R. A. The hallmarks of cancer. Cell 100, 57–70 (2000).

    Article  CAS  PubMed  Google Scholar 

  2. Kendall, S. D., Adam, S. J. & Counter, C. M. Genetically engineered human cancer models utilizing mammalian transgene expression. Cell Cycle 5, 1074–1079 (2006).

    CAS  PubMed  Google Scholar 

  3. Lock, R. B. et al. The nonobese diabetic/severe combined immunodeficient (NOD/SCID) mouse model of childhood acute lymphoblastic leukemia reveals intrinsic differences in biologic characteristics at diagnosis and relapse. Blood 99, 4100–4108 (2002).

    CAS  PubMed  Google Scholar 

  4. Rubio-Viqueira, B. et al. An in vivo platform for translational drug development in pancreatic cancer. Clin. Cancer Res 12, 4652–4661 (2006).

    CAS  PubMed  Google Scholar 

  5. Sikder, H. et al. Disruption of Id1 reveals major differences in angiogenesis between transplanted and autochthonous tumors. Cancer Cell 4, 291–299 (2003).

    CAS  PubMed  Google Scholar 

  6. Becher, O. J. & Holland, E. C. Genetically engineered models have advantages over xenografts for preclinical studies. Cancer Res. 66, 3355–3358 (2006).

    CAS  PubMed  Google Scholar 

  7. De Both, N. J., Vermey, M., Groen, N., Dinjens, W. N. & Bosman, F. T. Clonal growth of colorectal-carcinoma cell lines transplanted to nude mice. Int. J. Cancer 72, 1137–1141 (1997).

    CAS  PubMed  Google Scholar 

  8. Staroselsky, A. N. et al. The use of molecular genetic markers to demonstrate the effect of organ environment on clonal dominance in a human renal-cell carcinoma grown in nude mice. Int. J. Cancer 51, 130–138 (1992).

    CAS  PubMed  Google Scholar 

  9. Kang, Y. et al. A multigenic program mediating breast cancer metastasis to bone. Cancer Cell 3, 537–549 (2003).

    CAS  PubMed  Google Scholar 

  10. Johnson, J. I. et al. Relationships between drug activity in NCI preclinical in vitro and in vivo models and early clinical trials. Br. J. Cancer 84, 1424–1431 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Fiebig, H. H., Berger, D. P., Winterhalter, B. R. & Plowman, J. In vitro and in vivo evaluation of US-NCI compounds in human tumor xenografts. Cancer Treat. Rev. 17, 109–117 (1990).

    CAS  PubMed  Google Scholar 

  12. Hardisty, J. F. Factors influencing laboratory animal spontaneous tumor profiles. Toxicol. Pathol. 13, 95–104 (1985).

    CAS  PubMed  Google Scholar 

  13. van Kranen, H. J. et al. Frequent p53 alterations but low incidence of ras mutations in UV-B-induced skin tumors of hairless mice. Carcinogenesis 16, 1141–1147 (1995).

    CAS  PubMed  Google Scholar 

  14. Balmain, A. & Pragnell, I. B. Mouse skin carcinomas induced in vivo by chemical carcinogens have a transforming Harvey-ras oncogene. Nature 303, 72–74 (1983).

    CAS  PubMed  Google Scholar 

  15. Cardiff, R. D. & Kenney, N. Mouse mammary tumor biology: a short history. Adv. Cancer Res. 98, 53–116 (2007).

    CAS  PubMed  Google Scholar 

  16. Enno, A. et al. MALToma-like lesions in the murine gastric mucosa after long-term infection with Helicobacter felis. A mouse model of Helicobacter pylori-induced gastric lymphoma. Am. J. Pathol. 147, 217–222 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Jonkers, J. & Berns, A. Conditional mouse models of sporadic cancer. Nature Rev. Cancer 2, 251–265 (2002).

    CAS  Google Scholar 

  18. Tuveson, D. A. & Jacks, T. Technologically advanced cancer modeling in mice. Curr. Opin. Genet. Dev. 12, 105–110 (2002).

    CAS  PubMed  Google Scholar 

  19. Gannon, M., Gamer, L. W. & Wright, C. V. Regulatory regions driving developmental and tissue-specific expression of the essential pancreatic gene pdx1. Dev. Biol. 238, 185–201 (2001).

    CAS  PubMed  Google Scholar 

  20. Mattick, J. S. & Makunin, I. V. Non-coding RNA. Hum. Mol. Genet. 15 Spec. No 1, R17–R29 (2006).

    CAS  PubMed  Google Scholar 

  21. Robertson, G. et al. Position-dependent variegation of globin transgene expression in mice. Proc. Natl Acad. Sci. USA 92, 5371–5375 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Palomo, C., Zou, X., Nicholson, I. C., Butzler, C. & Bruggemann, M. B-cell tumorigenesis in mice carrying a yeast artificial chromosome-based immunoglobulin heavy/c-myc translocus is independent of the heavy chain intron enhancer (Emu). Cancer Res. 59, 5625–5628 (1999).

    CAS  PubMed  Google Scholar 

  23. Kaufman, R. M., Pham, C. T. & Ley, T. J. Transgenic analysis of a 100-kb human b-globin cluster-containing DNA fragment propagated as a bacterial artificial chromosome. Blood 94, 3178–3184 (1999).

    CAS  PubMed  Google Scholar 

  24. Schonig, K., Schwenk, F., Rajewsky, K. & Bujard, H. Stringent doxycycline dependent control of CRE recombinase in vivo. Nucleic Acids Res. 30, e134 (2002).

    PubMed  PubMed Central  Google Scholar 

  25. Kuhn, R., Schwenk, F., Aguet, M. & Rajewsky, K. Inducible gene targeting in mice. Science 269, 1427–1429 (1995).

    CAS  PubMed  Google Scholar 

  26. Weinstein, I. B. Cancer. Addiction to oncogenes — the Achilles heal of cancer. Science 297, 63–64 (2002).

    CAS  PubMed  Google Scholar 

  27. Chin, L. et al. Essential role for oncogenic Ras in tumour maintenance. Nature 400, 468–472 (1999). The first mouse solid tumour models to experimentally demonstrate a requirement for oncogene expression in tumour maintenance.

    CAS  PubMed  Google Scholar 

  28. Fisher, G. H. et al. Induction and apoptotic regression of lung adenocarcinomas by regulation of a K-Ras transgene in the presence and absence of tumor suppressor genes. Genes Dev. 15, 3249–3262 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Felsher, D. W. & Bishop, J. M. Reversible tumorigenesis by MYC in hematopoietic lineages. Mol. Cell 4, 199–207 (1999).

    CAS  PubMed  Google Scholar 

  30. Huettner, C. S., Zhang, P., Van Etten, R. A. & Tenen, D. G. Reversibility of acute B-cell leukaemia induced by BCR-ABL1. Nature Genet. 24, 57–60 (2000).

    CAS  PubMed  Google Scholar 

  31. Moody, S. E. et al. Conditional activation of Neu in the mammary epithelium of transgenic mice results in reversible pulmonary metastasis. Cancer Cell 2, 451–461 (2002).

    CAS  PubMed  Google Scholar 

  32. Sarkisian, C. J. et al. Dose-dependent oncogene-induced senescence in vivo and its evasion during mammary tumorigenesis. Nature Cell Biol. 9, 493–505 (2007). This paper demonstrates that the tetracycline system can be used to modulate ectopic HrasG12V levels in vivo . The authors find that maximal Hras G12V transgene expression causes mammary gland proliferative arrest (premature senescence), whereas low level Hras G12V expression promotes proliferation. It also raises the concern that inducible ectopic oncogenes (and constitutive ectopic oncogenes) may inadvertently cause a non-physiological toxic response in cells in vivo.

    CAS  PubMed  Google Scholar 

  33. Blakely, C. M. et al. Developmental stage determines the effects of MYC in the mammary epithelium. Development 132, 1147–1160 (2005).

    CAS  PubMed  Google Scholar 

  34. Blyth, K. et al. Sensitivity to myc-induced apoptosis is retained in spontaneous and transplanted lymphomas of CD2-mycER mice. Oncogene 19, 773–782 (2000).

    CAS  PubMed  Google Scholar 

  35. Martins, C. P., Brown-Swigart, L. & Evan, G. I. Modeling the therapeutic efficacy of p53 restoration in tumors. Cell 127, 1323–1334 (2006).

    CAS  PubMed  Google Scholar 

  36. Ventura, A. et al. Restoration of p53 function leads to tumour regression in vivo. Nature 445, 661–665 (2007).

    CAS  PubMed  Google Scholar 

  37. Rusk, N. Making mice at high speed. Nature Methods 4, 196–197 (2007).

    Google Scholar 

  38. Holland, E. C., Hively, W. P., DePinho, R. A. & Varmus, H. E. A constitutively active epidermal growth factor receptor cooperates with disruption of G1 cell-cycle arrest pathways to induce glioma-like lesions in mice. Genes Dev. 12, 3675–3685 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Orsulic, S. et al. Induction of ovarian cancer by defined multiple genetic changes in a mouse model system. Cancer Cell 1, 53–62 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Lewis, B. C., Klimstra, D. S. & Varmus, H. E. The c-myc and PyMT oncogenes induce different tumor types in a somatic mouse model for pancreatic cancer. Genes Dev. 17, 3127–3138 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Lewis, B. C. et al. The absence of p53 promotes metastasis in a novel somatic mouse model for hepatocellular carcinoma. Mol. Cell Biol. 25, 1228–1237 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Lewis, B. C., Chinnasamy, N., Morgan, R. A. & Varmus, H. E. Development of an avian leukosis-sarcoma virus subgroup A pseudotyped lentiviral vector. J. Virol. 75, 9339–9344 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Theodorou, V. et al. MMTV insertional mutagenesis identifies genes, gene families and pathways involved in mammary cancer. Nature Genet. 39, 759–769 (2007).

    CAS  PubMed  Google Scholar 

  44. Thomas, K. R. & Capecchi, M. R. Site-directed mutagenesis by gene targeting in mouse embryo-derived stem cells. Cell 51, 503–512 (1987).

    CAS  PubMed  Google Scholar 

  45. Cichowski, K. et al. Mouse models of tumor development in neurofibromatosis type 1. Science 286, 2172–2176 (1999).

    CAS  PubMed  Google Scholar 

  46. Clarke, A. R. et al. Requirement for a functional Rb-1 gene in murine development. Nature 359, 328–330 (1992).

    CAS  PubMed  Google Scholar 

  47. Jacks, T. et al. Effects of an Rb mutation in the mouse. Nature 359, 295–300 (1992).

    CAS  PubMed  Google Scholar 

  48. Lee, E. Y. et al. Mice deficient for Rb are nonviable and show defects in neurogenesis and haematopoiesis. Nature 359, 288–294 (1992).

    CAS  PubMed  Google Scholar 

  49. Lakso, M. et al. Targeted oncogene activation by site-specific recombination in transgenic mice. Proc. Natl Acad. Sci. USA 89, 6232–6236 (1992).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Rubin, B. P. et al. A knock-in mouse model of gastrointestinal stromal tumor harboring kit K641E. Cancer Res. 65, 6631–6639 (2005).

    CAS  PubMed  Google Scholar 

  51. Sotillo, R. et al. Invasive melanoma in Cdk4-targeted mice. Proc. Natl Acad. Sci. USA 98, 13312–13317 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Tuveson, D. A. et al. Endogenous oncogenic K-ras(G12D) stimulates proliferation and widespread neoplastic and developmental defects. Cancer Cell 5, 375–387 (2004).

    CAS  PubMed  Google Scholar 

  53. Schubbert, S. et al. Germline KRAS mutations cause Noonan syndrome. Nature Genet. 38, 331–336 (2006).

    CAS  PubMed  Google Scholar 

  54. de Alboran, I. M. et al. Analysis of C-MYC function in normal cells via conditional gene-targeted mutation. Immunity 14, 45–55 (2001).

    CAS  PubMed  Google Scholar 

  55. Jackson, E. L. et al. Analysis of lung tumor initiation and progression using conditional expression of oncogenic K-ras. Genes Dev. 15, 3243–3248 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Lewandoski, M. & Martin, G. R. Cre-mediated chromosome loss in mice. Nature Genet. 17, 223–225 (1997).

    CAS  PubMed  Google Scholar 

  57. Dewald, G. W., Noel, P., Dahl, R. J. & Spurbeck, J. L. Chromosome abnormalities in malignant hematologic disorders. Mayo Clin. Proc. 60, 675–689 (1985).

    CAS  PubMed  Google Scholar 

  58. Mitelman, F., Johansson, B. & Mertens, F. The impact of translocations and gene fusions on cancer causation. Nature Rev. Cancer 7, 233–245 (2007).

    CAS  Google Scholar 

  59. Smith, A. J., Xian, J., Richardson, M., Johnstone, K. A. & Rabbitts, P. H. Cre-loxP chromosome engineering of a targeted deletion in the mouse corresponding to the 3p21. 3 region of homozygous loss in human tumours. Oncogene 21, 4521–4529 (2002).

    CAS  PubMed  Google Scholar 

  60. Kmita, M., Kondo, T. & Duboule, D. Targeted inversion of a polar silencer within the HoxD complex re-allocates domains of enhancer sharing. Nature Genet. 26, 451–454 (2000).

    CAS  PubMed  Google Scholar 

  61. Langer, S. J., Ghafoori, A. P., Byrd, M. & Leinwand, L. A genetic screen identifies novel non-compatible loxP sites. Nucleic Acids Res. 30, 3067–3077 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Forster, A. et al. Engineering de novo reciprocal chromosomal translocations associated with Mll to replicate primary events of human cancer. Cancer Cell 3, 449–458 (2003).

    CAS  PubMed  Google Scholar 

  63. Johnson, L. et al. Somatic activation of the K-ras oncogene causes early onset lung cancer in mice. Nature 410, 1111–1116 (2001). Describes the generation of a stochastic model for lung cancer that accurately mimics the mutation of a single cell in the context of normal surrounding tissue.

    CAS  PubMed  Google Scholar 

  64. Coste, I., Freund, J. N., Spaderna, S., Brabletz, T. & Renno, T. Precancerous lesions upon sporadic activation of b-catenin in mice. Gastroenterology 132, 1299–1308 (2007).

    CAS  PubMed  Google Scholar 

  65. Rudolph, K. L. et al. Longevity, stress response, and cancer in aging telomerase-deficient mice. Cell 96, 701–712 (1999).

    CAS  PubMed  Google Scholar 

  66. Chin, L. et al. p53 deficiency rescues the adverse effects of telomere loss and cooperates with telomere dysfunction to accelerate carcinogenesis. Cell 97, 527–538 (1999). References 65 and 66 demonstrate that mutation of telomerase predisposes mice to cancers that exhibit genomic instability, a hallmark of human cancers.

    CAS  PubMed  Google Scholar 

  67. Seibler, J. et al. Reversible gene knockdown in mice using a tight, inducible shRNA expression system. Nucleic Acids Res. 35, e54 (2007).

    PubMed  PubMed Central  Google Scholar 

  68. Xue, W. et al. Senescence and tumour clearance is triggered by p53 restoration in murine liver carcinomas. Nature 445, 656–660 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Kumar, M. S., Lu, J., Mercer, K. L., Golub, T. R. & Jacks, T. Impaired microRNA processing enhances cellular transformation and tumorigenesis. Nature Genet. 39, 673–677 (2007).

    CAS  PubMed  Google Scholar 

  70. Collier, L. S., Carlson, C. M., Ravimohan, S., Dupuy, A. J. & Largaespada, D. A. Cancer gene discovery in solid tumours using transposon-based somatic mutagenesis in the mouse. Nature 436, 272–276 (2005).

    CAS  PubMed  Google Scholar 

  71. Dupuy, A. J., Akagi, K., Largaespada, D. A., Copeland, N. G. & Jenkins, N. A. Mammalian mutagenesis using a highly mobile somatic Sleeping Beauty transposon system. Nature 436, 221–226 (2005). References 70 and 71 report the use of a genetically encoded transposon-based mutagenesis system to perform forward genetic screens in mice.

    CAS  PubMed  Google Scholar 

  72. Grippo, P. J., Nowlin, P. S., Demeure, M. J., Longnecker, D. S. & Sandgren, E. P. Preinvasive pancreatic neoplasia of ductal phenotype induced by acinar cell targeting of mutant Kras in transgenic mice. Cancer Res. 63, 2016–2019 (2003).

    CAS  PubMed  Google Scholar 

  73. Guerra, C. et al. Chronic pancreatitis is essential for induction of pancreatic ductal adenocarcinoma by K-Ras oncogenes in adult mice. Cancer Cell 11, 291–302 (2007).

    CAS  PubMed  Google Scholar 

  74. Cohen, S. J. et al. Phase II and pharmacodynamic study of the farnesyltransferase inhibitor R115777 as initial therapy in patients with metastatic pancreatic adenocarcinoma. J. Clin. Oncol. 21, 1301–1306 (2003).

    CAS  PubMed  Google Scholar 

  75. Kohl, N. E. et al. Inhibition of farnesyltransferase induces regression of mammary and salivary carcinomas in ras transgenic mice. Nature Med. 1, 792–797 (1995).

    CAS  PubMed  Google Scholar 

  76. Bos, J. L. ras oncogenes in human cancer: a review. Cancer Res. 49, 4682–4689 (1989).

    CAS  PubMed  Google Scholar 

  77. Lerner, E. C., Qian, Y., Hamilton, A. D. & Sebti, S. M. Disruption of oncogenic K-Ras4B processing and signaling by a potent geranylgeranyltransferase I inhibitor. J. Biol. Chem. 270, 26770–26773 (1995).

    CAS  PubMed  Google Scholar 

  78. Majumder, P. K. & Sellers, W. R. Akt-regulated pathways in prostate cancer. Oncogene 24, 7465–7474 (2005).

    CAS  PubMed  Google Scholar 

  79. Shen, W. H. et al. Essential role for nuclear PTEN in maintaining chromosomal integrity. Cell 128, 157–170 (2007).

    CAS  PubMed  Google Scholar 

  80. James, R. M. et al. K-ras proto-oncogene exhibits tumor suppressor activity as its absence promotes tumorigenesis in murine teratomas. Mol. Cancer Res. 1, 820–825 (2003).

    CAS  PubMed  Google Scholar 

  81. Zhang, Z. et al. Wildtype Kras2 can inhibit lung carcinogenesis in mice. Nature Genet. 29, 25–33 (2001).

    CAS  PubMed  Google Scholar 

  82. Bayascas, J. R., Sakamoto, K., Armit, L., Arthur, J. S. & Alessi, D. R. Evaluation of approaches to generation of tissue-specific knock-in mice. J. Biol. Chem. 281, 28772–28781 (2006).

    CAS  PubMed  Google Scholar 

  83. Forster, A. et al. The invertor knock-in conditional chromosomal translocation mimic. Nature Methods 2, 27–30 (2005).

    CAS  PubMed  Google Scholar 

  84. Dankort, D. et al. A new mouse model to explore the initiation, progression, and therapy of BRAFV600E-induced lung tumors. Genes Dev. 21, 379–384 (2007). An excellent example of a conditional oncogene mouse model in which the animal is functionally wild type before induction of the mutation.

    CAS  PubMed  PubMed Central  Google Scholar 

  85. Oberdoerffer, P., Otipoby, K. L., Maruyama, M. & Rajewsky, K. Unidirectional Cre-mediated genetic inversion in mice using the mutant loxP pair lox66/lox71. Nucleic Acids Res. 31, e140 (2003).

    PubMed  PubMed Central  Google Scholar 

  86. Zhang, Z. & Lutz, B. Cre recombinase-mediated inversion using lox66 and lox71: method to introduce conditional point mutations into the CREB-binding protein. Nucleic Acids Res. 30, e90 (2002).

    PubMed  PubMed Central  Google Scholar 

  87. Olive, K. P. et al. Mutant p53 gain of function in two mouse models of Li-Fraumeni syndrome. Cell 119, 847–860 (2004).

    CAS  PubMed  Google Scholar 

  88. Song, H., Hollstein, M. & Xu, Y. p53 gain-of-function cancer mutants induce genetic instability by inactivating ATM. Nature Cell Biol. 9, 573–580 (2007).

    CAS  PubMed  Google Scholar 

  89. Ahmed, B. Y. et al. Efficient delivery of Cre-recombinase to neurons in vivo and stable transduction of neurons using adeno-associated and lentiviral vectors. BMC Neurosci. 5, 4 (2004).

    PubMed  PubMed Central  Google Scholar 

  90. Vooijs, M., Jonkers, J. & Berns, A. A highly efficient ligand-regulated Cre recombinase mouse line shows that LoxP recombination is position dependent. EMBO Rep. 2, 292–297 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Kemp, R. et al. Elimination of background recombination: somatic induction of Cre by combined transcriptional regulation and hormone binding affinity. Nucleic Acids Res. 32, e92 (2004). A mouse model in which Cre activity is regulated at both the transcriptional and translational level, therefore preventing the background recombination associated with other inducible Cre models.

    PubMed  PubMed Central  Google Scholar 

  92. Hameyer, D. et al. Toxicity of ligand-dependant Cre-recombinases and generation of a conditional Cre-deleter mouse allowing mosaic recombination in peripheral tissues. Physiol. Genomics 12 June 2007 (doi:10. 1152/physiolgenomics. 00019. 2007)

    Google Scholar 

  93. Bjerkvig, R., Tysnes, B. B., Aboody, K. S., Najbauer, J. & Terzis, A. J. Opinion: the origin of the cancer stem cell: current controversies and new insights. Nature Rev. Cancer 5, 899–904 (2005).

    CAS  Google Scholar 

  94. Peto, R. in Origins of Human Cancer 1403–1428 (Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY, 1977) (eds Hiett, H. H., Watson, J. D & Winsten, J. A.).

    Google Scholar 

  95. Ijichi, H. et al. Aggressive pancreatic ductal adenocarcinoma in mice caused by pancreas-specific blockade of transforming growth factor-beta signaling in cooperation with active Kras expression. Genes Dev. 20, 3147–3160 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  96. Izeradjene, K. et al. Kras(G12D) and Smad4/Dpc4 haploinsufficiency cooperate to induce mucinous cystic neoplasms and invasive adenocarcinoma of the pancreas. Cancer Cell 11, 229–243 (2007).

    CAS  PubMed  Google Scholar 

  97. Hruban, R. H., Wilentz, R. E. & Kern, S. E. Genetic progression in the pancreatic ducts. Am. J. Pathol. 156, 1821–1825 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Kondo, S. et al. Efficient sequential gene regulation via FLP-and Cre-recombinase using adenovirus vector in mammalian cells including mouse ES cells. Microbiol. Immunol. 50, 831–843 (2006).

    CAS  PubMed  Google Scholar 

  99. Rangarajan, A. & Weinberg, R. A. Opinion: Comparative biology of mouse versus human cells: modelling human cancer in mice. Nature Rev. Cancer 3, 952–959 (2003).

    CAS  Google Scholar 

  100. McClatchey, A. I. et al. Mice heterozygous for a mutation at the Nf2 tumor suppressor locus develop a range of highly metastatic tumors. Genes Dev. 12, 1121–1133 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  101. Trofatter, J. A. et al. A novel moesin-, ezrin-, radixin-like gene is a candidate for the neurofibromatosis 2 tumor suppressor. Cell 72, 791–800 (1993).

    CAS  PubMed  Google Scholar 

  102. Giovannini, M. et al. Conditional biallelic Nf2 mutation in the mouse promotes manifestations of human neurofibromatosis type 2. Genes Dev. 14, 1617–1630 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  103. Liu, K. et al. Recombinase-mediated cassette exchange to rapidly and efficiently generate mice with human cardiac sodium channels. Genesis 44, 556–564 (2006).

    CAS  PubMed  Google Scholar 

  104. Zhuang, Y., Barndt, R. J., Pan, L., Kelley, R. & Dai, M. Functional replacement of the mouse E2A gene with a human HEB cDNA. Mol. Cell Biol. 18, 3340–3349 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  105. Wallace, H. A. et al. Manipulating the mouse genome to engineer precise functional syntenic replacements with human sequence. Cell 128, 197–209 (2007). An example of recombinase-mediated cassette exchange in which a large human genomic region was introduced into the syntenic mouse locus to elicit a disease phenotype that accurately mimics the human condition.

    CAS  PubMed  Google Scholar 

  106. Kipling, D. & Cooke, H. J. Hypervariable ultra-long telomeres in mice. Nature 347, 400–402 (1990).

    CAS  PubMed  Google Scholar 

  107. Harley, C. B., Futcher, A. B. & Greider, C. W. Telomeres shorten during ageing of human fibroblasts. Nature 345, 458–460 (1990).

    CAS  PubMed  Google Scholar 

  108. Prowse, K. R. & Greider, C. W. Developmental and tissue-specific regulation of mouse telomerase and telomere length. Proc. Natl Acad. Sci. USA 92, 4818–4822 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Lee, H. W. et al. Essential role of mouse telomerase in highly proliferative organs. Nature 392, 569–574 (1998).

    CAS  PubMed  Google Scholar 

  110. Ding, H. et al. Regulation of murine telomere length by Rtel: an essential gene encoding a helicase-like protein. Cell 117, 873–886 (2004).

    CAS  PubMed  Google Scholar 

  111. Martignoni, M., Groothuis, G. M. & de Kanter, R. Species differences between mouse, rat, dog, monkey and human CYP-mediated drug metabolism, inhibition and induction. Expert Opin. Drug Metab. Toxicol. 2, 875–894 (2006).

    CAS  PubMed  Google Scholar 

  112. Gonzalez, F. J. & Kimura, S. Study of P450 function using gene knockout and transgenic mice. Arch. Biochem. Biophys. 409, 153–158 (2003).

    CAS  PubMed  Google Scholar 

  113. Gonzalez, F. J. Cytochrome P450 humanised mice. Hum. Genomics 1, 300–306 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  114. Xie, W. & Evans, R. M. Pharmaceutical use of mouse models humanized for the xenobiotic receptor. Drug Discov. Today 7, 509–515 (2002).

    CAS  PubMed  Google Scholar 

  115. Corchero, J. et al. The CYP2D6 humanized mouse: effect of the human CYP2D6 transgene and HNF4α on the disposition of debrisoquine in the mouse. Mol. Pharmacol. 60, 1260–1267 (2001).

    CAS  PubMed  Google Scholar 

  116. Granvil, C. P. et al. Expression of the human CYP3A4 gene in the small intestine of transgenic mice: in vitro metabolism and pharmacokinetics of midazolam. Drug Metab. Dispos. 31, 548–558 (2003).

    CAS  PubMed  Google Scholar 

  117. Gonzalez, F. J. & Yu, A. M. Cytochrome P450 and xenobiotic receptor humanized mice. Annu. Rev. Pharmacol. Toxicol. 46, 41–64 (2006). An excellent review of currently available transgenic mice expressing human metabolic enzymes.

    CAS  PubMed  PubMed Central  Google Scholar 

  118. Bui, J. D. & Schreiber, R. D. Cancer immunosurveillance, immunoediting and inflammation: independent or interdependent processes? Curr. Opin. Immunol. 19, 203–208 (2007).

    CAS  PubMed  Google Scholar 

  119. de Visser, K. E., Eichten, A. & Coussens, L. M. Paradoxical roles of the immune system during cancer development. Nature Rev. Cancer 6, 24–37 (2006).

    CAS  Google Scholar 

  120. Shultz, L. D., Ishikawa, F. & Greiner, D. L. Humanized mice in translational biomedical research. Nature Rev. Immunol. 7, 118–130 (2007).

    CAS  Google Scholar 

  121. Chang, C. H., Fodor, W. L. & Flavell, R. A. Reactivation of a major histocompatibility complex class II gene in mouse plasmacytoma cells and mouse T cells. J. Exp. Med. 176, 1465–1469 (1992).

    CAS  PubMed  Google Scholar 

  122. Rehli, M. Of mice and men: species variations of Toll-like receptor expression. Trends Immunol. 23, 375–8 (2002).

    CAS  PubMed  Google Scholar 

  123. Hayakawa, Y., Huntington, N. D., Nutt, S. L. & Smyth, M. J. Functional subsets of mouse natural killer cells. Immunol. Rev. 214, 47–55 (2006).

    CAS  PubMed  Google Scholar 

  124. Chen, Z. et al. A 320-kilobase artificial chromosome encoding the human HLA DR3-DQ2 MHC haplotype confers HLA restriction in transgenic mice. J. Immunol. 168, 3050–3056 (2002).

    CAS  PubMed  Google Scholar 

  125. Chen, Z. et al. Humanized transgenic mice expressing HLA DR4-DQ3 haplotype: reconstitution of phenotype and HLA-restricted T-cell responses. Tissue Antigens 68, 210–219 (2006).

    CAS  PubMed  Google Scholar 

  126. Madsen, L. et al. Mice lacking all conventional MHC class II genes. Proc. Natl Acad. Sci. USA 96, 10338–10343 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  127. Kobata, A. The third chains of living organisms-a trail of glycobiology that started from the third floor of building 4 in NIH. Arch. Biochem. Biophys. 426, 107–121 (2004).

    CAS  PubMed  Google Scholar 

  128. Raju, T. S., Briggs, J. B., Borge, S. M. & Jones, A. J. Species-specific variation in glycosylation of IgG: evidence for the species-specific sialylation and branch-specific galactosylation and importance for engineering recombinant glycoprotein therapeutics. Glycobiology 10, 477–486 (2000).

    CAS  PubMed  Google Scholar 

  129. Carpelan-Holmstrom, M., Louhimo, J., Stenman, U. H., Alfthan, H. & Haglund, C. CEA, CA 19–9 and CA 72–4 improve the diagnostic accuracy in gastrointestinal cancers. Anticancer Res. 22, 2311–2316 (2002).

    CAS  PubMed  Google Scholar 

  130. Gion, M. et al. CA27. 29: a valuable marker for breast cancer management. A confirmatory multicentric study on 603 cases. Eur. J. Cancer 37, 355–363 (2001).

    CAS  PubMed  Google Scholar 

  131. Gadducci, A. et al. Serum half-life of CA 125 during early chemotherapy as an independent prognostic variable for patients with advanced epithelial ovarian cancer: results of a multicentric Italian study. Gynecol. Oncol. 58, 42–47 (1995).

    CAS  PubMed  Google Scholar 

  132. Falk, P. G., Bry, L., Holgersson, J. & Gordon, J. I. Expression of a human alpha-1, 3/4-fucosyltransferase in the pit cell lineage of FVB/N mouse stomach results in production of Leb-containing glycoconjugates: a potential transgenic mouse model for studying Helicobacter pylori infection. Proc. Natl Acad. Sci. USA 92, 1515–1519 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Houghton, J. & Wang, T. C. Helicobacter pylori and gastric cancer: a new paradigm for inflammation-associated epithelial cancers. Gastroenterology 128, 1567–1578 (2005).

    CAS  PubMed  Google Scholar 

  134. Itzkowitz, S. H. & Yio, X. Inflammation and cancer IV. Colorectal cancer in inflammatory bowel disease: the role of inflammation. Am. J. Physiol. Gastrointest. Liver Physiol. 287, G7–G17 (2004).

    CAS  PubMed  Google Scholar 

  135. Seitz, H. K. & Stickel, F. Risk factors and mechanisms of hepatocarcinogenesis with special emphasis on alcohol and oxidative stress. Biol. Chem. 387, 349–360 (2006).

    CAS  PubMed  Google Scholar 

  136. Jura, N., Archer, H. & Bar-Sagi, D. Chronic pancreatitis, pancreatic adenocarcinoma and the black box in-between. Cell Res. 15, 72–77 (2005).

    CAS  PubMed  Google Scholar 

  137. Engle, S. J. et al. Elimination of colon cancer in germ-free transforming growth factor b 1-deficient mice. Cancer Res. 62, 6362–6366 (2002).

    CAS  PubMed  Google Scholar 

  138. Coussens, L. M., Tinkle, C. L., Hanahan, D. & Werb, Z. MMP-9 supplied by bone marrow-derived cells contributes to skin carcinogenesis. Cell 103, 481–490 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  139. de Visser, K. E., Korets, L. V. & Coussens, L. M. De novo carcinogenesis promoted by chronic inflammation is B lymphocyte dependent. Cancer Cell 7, 411–423 (2005).

    CAS  PubMed  Google Scholar 

  140. Olumi, A. F. et al. Carcinoma-associated fibroblasts direct tumor progression of initiated human prostatic epithelium. Cancer Res. 59, 5002–5011 (1999).

    CAS  PubMed  Google Scholar 

  141. Orimo, A. et al. Stromal fibroblasts present in invasive human breast carcinomas promote tumor growth and angiogenesis through elevated SDF-1/CXCL12 secretion. Cell 121, 335–348 (2005).

    CAS  PubMed  Google Scholar 

  142. Kurose, K. et al. Frequent somatic mutations in PTEN and TP53 are mutually exclusive in the stroma of breast carcinomas. Nature Genet. 32, 355–357 (2002).

    CAS  PubMed  Google Scholar 

  143. Fukino, K., Shen, L., Patocs, A., Mutter, G. L. & Eng, C. Genomic instability within tumor stroma and clinicopathological characteristics of sporadic primary invasive breast carcinoma. JAMA 297, 2103–2111 (2007).

    CAS  PubMed  Google Scholar 

  144. Hu, M. et al. Distinct epigenetic changes in the stromal cells of breast cancers. Nature Genet. 37, 899–905 (2005).

    CAS  PubMed  Google Scholar 

  145. Hill, R., Song, Y., Cardiff, R. D. & Van Dyke, T. Selective evolution of stromal mesenchyme with p53 loss in response to epithelial tumorigenesis. Cell 123, 1001–1011 (2005).

    CAS  PubMed  Google Scholar 

  146. Lyons, S. K. Advances in imaging mouse tumour models in vivo. J. Pathol. 205, 194–205 (2005).

    CAS  PubMed  Google Scholar 

  147. Olive, K. P. & Tuveson, D. A. The use of targeted mouse models for preclinical testing of novel cancer therapeutics. Clin. Cancer Res. 12, 5277–5287 (2006).

    CAS  PubMed  Google Scholar 

  148. Sharpless, N. E. & Depinho, R. A. The mighty mouse: genetically engineered mouse models in cancer drug development. Nature Rev. Drug Discov. 5, 741–754 (2006).

    CAS  Google Scholar 

  149. Ji, H. et al. The impact of human EGFR kinase domain mutations on lung tumorigenesis and in vivo sensitivity to EGFR-targeted therapies. Cancer Cell 9, 485–495 (2006).

    CAS  PubMed  Google Scholar 

  150. Politi, K. et al. Lung adenocarcinomas induced in mice by mutant EGF receptors found in human lung cancers respond to a tyrosine kinase inhibitor or to down-regulation of the receptors. Genes Dev. 20, 1496–1510 (2006). References 149 and 150 demonstrate that mouse cancer models with relevant mutations respond accordingly to currently available therapies.

    CAS  PubMed  PubMed Central  Google Scholar 

  151. Lallemand-Breitenbach, V. et al. Retinoic acid and arsenic synergize to eradicate leukemic cells in a mouse model of acute promyelocytic leukemia. J. Exp. Med. 189, 1043–1052 (1999). A seminal publication in which therapeutic intervention in a mouse cancer model led to the successful development of an arsenic trioxide as a treatment for childhood acute promyelocytic leukaemia.

    CAS  PubMed  PubMed Central  Google Scholar 

  152. Soignet, S. & Maslak, P. Therapy of acute promyelocytic leukemia. Adv. Pharmacol. 51, 35–58 (2004).

    CAS  PubMed  Google Scholar 

  153. Gu, G., Dubauskaite, J. & Melton, D. A. Direct evidence for the pancreatic lineage: NGN3+ cells are islet progenitors and are distinct from duct progenitors. Development 129, 2447–2457 (2002).

    CAS  PubMed  Google Scholar 

  154. Kawaguchi, Y. et al. The role of the transcriptional regulator Ptf1a in converting intestinal to pancreatic progenitors. Nature Genet. 32, 128–134 (2002).

    CAS  PubMed  Google Scholar 

  155. Obata, J. et al. p48 subunit of mouse PTF1 binds to RBP-Jκ/CBF-1, the intracellular mediator of Notch signalling, and is expressed in the neural tube of early stage embryos. Genes Cells 6, 345–360 (2001).

    CAS  PubMed  Google Scholar 

  156. Bardeesy, N. et al. Smad4 is dispensable for normal pancreas development yet critical in progression and tumor biology of pancreas cancer. Genes Dev. 20, 3130–3146 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  157. Loonstra, A. et al. Growth inhibition and DNA damage induced by Cre recombinase in mammalian cells. Proc. Natl Acad. Sci. USA 98, 9209–9214 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  158. Pfeifer, A., Brandon, E. P., Kootstra, N., Gage, F. H. & Verma, I. M. Delivery of the Cre recombinase by a self-deleting lentiviral vector: efficient gene targeting in vivo. Proc. Natl Acad. Sci. USA 98, 11450–11455 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  159. Schmidt, E. E., Taylor, D. S., Prigge, J. R., Barnett, S. & Capecchi, M. R. Illegitimate Cre-dependent chromosome rearrangements in transgenic mouse spermatids. Proc. Natl Acad. Sci. USA 97, 13702–13707 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  160. Raymond, C. S. & Soriano, P. High-efficiency FLP and PhiC31 site-specific recombination in mammalian cells. PLoS ONE 2, e162 (2007).

    PubMed  PubMed Central  Google Scholar 

  161. Rodriguez, C. I. et al. High-efficiency deleter mice show that FLPe is an alternative to Cre-loxP. Nature Genet. 25, 139–140 (2000).

    CAS  PubMed  Google Scholar 

  162. Ehrhardt, A., Xu, H., Huang, Z., Engler, J. A. & Kay, M. A. A direct comparison of two nonviral gene therapy vectors for somatic integration: in vivo evaluation of the bacteriophage integrase phiC31 and the Sleeping Beauty transposase. Mol. Ther. 11, 695–706 (2005).

    CAS  PubMed  Google Scholar 

  163. Safran, M. et al. Mouse reporter strain for noninvasive bioluminescent imaging of cells that have undergone Cre-mediated recombination. Mol. Imaging 2, 297–302 (2003).

    CAS  PubMed  Google Scholar 

  164. Uhrbom, L., Nerio, E. & Holland, E. C. Dissecting tumor maintenance requirements using bioluminescence imaging of cell proliferation in a mouse glioma model. Nature Med. 10, 1257–1260 (2004).

    CAS  PubMed  Google Scholar 

  165. Safran, M. et al. Mouse model for noninvasive imaging of HIF prolyl hydroxylase activity: assessment of an oral agent that stimulates erythropoietin production. Proc. Natl Acad. Sci. USA 103, 105–110 (2006).

    CAS  PubMed  Google Scholar 

  166. Carlsen, H., Moskaug, J. O., Fromm, S. H. & Blomhoff, R. In vivo imaging of NF-κB activity. J. Immunol. 168, 1441–1446 (2002).

    CAS  PubMed  Google Scholar 

  167. Meuwissen, R. et al. Induction of small cell lung cancer by somatic inactivation of both Trp53 and Rb1 in a conditional mouse model. Cancer Cell 4, 181–189 (2003).

    CAS  PubMed  Google Scholar 

  168. Sansom, O. J. et al. Loss of Apc allows phenotypic manifestation of the transforming properties of an endogenous K-ras oncogene in vivo. Proc. Natl Acad. Sci. USA 103, 14122–14127 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  169. Edelmann, W. et al. Mutation in the mismatch repair gene Msh6 causes cancer susceptibility. Cell 91, 467–477 (1997).

    CAS  PubMed  Google Scholar 

  170. Jonkers, J. et al. Synergistic tumor suppressor activity of BRCA2 and p53 in a conditional mouse model for breast cancer. Nature Genet. 29, 418–425 (2001).

    CAS  PubMed  Google Scholar 

  171. Derksen, P. W. et al. Somatic inactivation of E-cadherin and p53 in mice leads to metastatic lobular mammary carcinoma through induction of anoikis resistance and angiogenesis. Cancer Cell 10, 437–449 (2006).

    CAS  PubMed  Google Scholar 

  172. Aguirre, A. J. et al. Activated Kras and Ink4a/Arf deficiency cooperate to produce metastatic pancreatic ductal adenocarcinoma. Genes Dev. 17, 3112–3126 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  173. Hingorani, S. R. et al. Trp53R172H and KrasG12D cooperate to promote chromosomal instability and widely metastatic pancreatic ductal adenocarcinoma in mice. Cancer Cell 7, 469–483 (2005).

    CAS  PubMed  Google Scholar 

  174. Wang, S. et al. Prostate-specific deletion of the murine Pten tumor suppressor gene leads to metastatic prostate cancer. Cancer Cell 4, 209–221 (2003).

    CAS  PubMed  Google Scholar 

  175. Kim, M. J. et al. Cooperativity of Nkx3. 1 and Pten loss of function in a mouse model of prostate carcinogenesis. Proc. Natl Acad. Sci. USA 99, 2884–2889 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  176. Zhou, Z. et al. Synergy of p53 and Rb deficiency in a conditional mouse model for metastatic prostate cancer. Cancer Res. 66, 7889–7898 (2006).

    CAS  PubMed  Google Scholar 

  177. Colnot, S. et al. Liver-targeted disruption of Apc in mice activates β-catenin signaling and leads to hepatocellular carcinomas. Proc. Natl Acad. Sci. USA 101, 17216–17221 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  178. Zender, L. et al. Identification and validation of oncogenes in liver cancer using an integrative oncogenomic approach. Cell 125, 1253–1267 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  179. Santoni-Rugiu, E., Nagy, P., Jensen, M. R., Factor, V. M. & Thorgeirsson, S. S. Evolution of neoplastic development in the liver of transgenic mice co-expressing c-myc and transforming growth factor-α. Am. J. Pathol. 149, 407–428 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  180. Dinulescu, D. M. et al. Role of K-ras and Pten in the development of mouse models of endometriosis and endometrioid ovarian cancer. Nature Med. 11, 63–70 (2005).

    CAS  PubMed  Google Scholar 

  181. Wu, R. et al. Mouse model of human ovarian endometrioid adenocarcinoma based on somatic defects in the Wnt/β-catenin and PI3K/Pten signaling pathways. Cancer Cell 11, 321–333 (2007).

    CAS  PubMed  Google Scholar 

  182. Teng, Y. et al. Synergistic function of Smad4 and PTEN in suppressing forestomach squamous cell carcinoma in the mouse. Cancer Res. 66, 6972–6981 (2006).

    CAS  PubMed  Google Scholar 

  183. Opitz, O. G. et al. A mouse model of human oral-esophageal cancer. J. Clin. Invest. 110, 761–769 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  184. Mo, L. et al. Hyperactivation of Ha-ras oncogene, but not Ink4a/Arf deficiency, triggers bladder tumorigenesis. J. Clin. Invest. 117, 314–325 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  185. Sansom, O. J., Griffiths, D. F., Reed, K. R., Winton, D. J. & Clarke, A. R. Apc deficiency predisposes to renal carcinoma in the mouse. Oncogene 24, 8205–8210 (2005).

    CAS  PubMed  Google Scholar 

  186. Xiao, A. et al. Somatic induction of Pten loss in a preclinical astrocytoma model reveals major roles in disease progression and avenues for target discovery and validation. Cancer Res. 65, 5172–5180 (2005).

    CAS  PubMed  Google Scholar 

  187. Shih, A. H. et al. Dose-dependent effects of platelet-derived growth factor-B on glial tumorigenesis. Cancer Res. 64, 4783–4789 (2004).

    CAS  PubMed  Google Scholar 

  188. Oshima, H. et al. Carcinogenesis in mouse stomach by simultaneous activation of the Wnt signaling and prostaglandin E2 pathway. Gastroenterology 131, 1086–1095 (2006).

    CAS  PubMed  Google Scholar 

  189. Andressoo, J. O. et al. An Xpd mouse model for the combined xeroderma pigmentosum/Cockayne syndrome exhibiting both cancer predisposition and segmental progeria. Cancer Cell 10, 121–132 (2006).

    CAS  PubMed  Google Scholar 

  190. Haramis, A. P. et al. De novo crypt formation and juvenile polyposis on BMP inhibition in mouse intestine. Science 303, 1684–1686 (2004).

    CAS  PubMed  Google Scholar 

  191. Su, L. K. et al. Multiple intestinal neoplasia caused by a mutation in the murine homolog of the APC gene. Science 256, 668–670 (1992).

    CAS  PubMed  Google Scholar 

  192. Schulze-Garg, C., Lohler, J., Gocht, A. & Deppert, W. A transgenic mouse model for the ductal carcinoma in situ (DCIS) of the mammary gland. Oncogene 19, 1028–1037 (2000).

    CAS  PubMed  Google Scholar 

  193. Hingorani, S. R. et al. Preinvasive and invasive ductal pancreatic cancer and its early detection in the mouse. Cancer Cell 4, 437–450 (2003).

    CAS  PubMed  Google Scholar 

  194. Kelavkar, U. P., Parwani, A. V., Shappell, S. B. & Martin, W. D. Conditional expression of human 15-lipoxygenase-1 in mouse prostate induces prostatic intraepithelial neoplasia: the FLiMP mouse model. Neoplasia 8, 510–522 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  195. Abdulkadir, S. A. et al. Conditional loss of Nkx3. 1 in adult mice induces prostatic intraepithelial neoplasia. Mol. Cell Biol. 22, 1495–1503 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  196. Conner, E. A. et al. Dual functions of E2F-1 in a transgenic mouse model of liver carcinogenesis. Oncogene 19, 5054–5062 (2000).

    CAS  PubMed  Google Scholar 

  197. Rankin, E. B., Tomaszewski, J. E. & Haase, V. H. Renal cyst development in mice with conditional inactivation of the von Hippel-Lindau tumor suppressor. Cancer Res. 66, 2576–2583 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  198. Reilly, K. M., Loisel, D. A., Bronson, R. T., McLaughlin, M. E. & Jacks, T. Nf1;Trp53 mutant mice develop glioblastoma with evidence of strain-specific effects. Nature Genet. 26, 109–113 (2000).

    CAS  PubMed  Google Scholar 

  199. Silberg, D. G. et al. Cdx2 ectopic expression induces gastric intestinal metaplasia in transgenic mice. Gastroenterology 122, 689–696 (2002).

    CAS  PubMed  Google Scholar 

  200. Pelengaris, S., Littlewood, T., Khan, M., Elia, G. & Evan, G. Reversible activation of c-Myc in skin: induction of a complex neoplastic phenotype by a single oncogenic lesion. Mol. Cell 3, 565–577 (1999).

    CAS  PubMed  Google Scholar 

  201. Bhowmick, N. A. et al. TGF-b signalling in fibroblasts modulates the oncogenic potential of adjacent epithelia. Science 303, 848–851 (2004).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank members of the Tuveson lab for helpful comments on the manuscript. Work is supported by US National Institutes of Health grants 1F32CA123887-01 (K.K.F.), CA101973 (D.A.T.), CA111292 (D.A.T.), CA084291 (D.A.T.), CA105490 (D.A.T.) and the Lustgarten Foundation for Pancreatic Cancer Research (D.A.T.). D.A.T. is a group leader of Cancer Research UK.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to David A. Tuveson.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

Related links

DATABASES

National Cancer Institute

FURTHER INFORMATION

Author's homepage

Glossary

Xenograft

Tumour tissue or cell lines from one species propagated in immunodeficient mice in ectopic or orthotopic sites.

Genetically engineered mice

(GEM). Mice harbouring genetic modifications designed to express either exogenous or endogenous mutated genes. In cancer modelling, these are frequently oncogenes or tumour-suppressor genes.

Cell autonomous

A trait engendered only in cells harbouring the mutation. In the case of a carcinoma, a cell autonomous function occurs only in the mutant tumour epithelial cells.

Non-cell autonomous

A trait engendered in cells that do not harbour the mutation. In the case of a carcinoma, a non-cell autonomous function occurs in stromal, immune and endothelial cells.

Autochthonous tumour

An endogenous or in situ tumour that evolves from normal cells of a tumour-bearing animal. This is in contrast to animal cultures in which exogenous tumour cells are implanted into a non-tumour-bearing animal.

Tumour microenvironment

The stroma and supporting milieu surrounding the tumour that consists of fibroblasts, immune cells and endothelial cells.

Oncogene addiction

The hypothesis that tumours arising as a result of a particular oncogenic lesion are exquisitely dependent on continued expression of that oncogene.

Loss of heterozygosity

Mutation or 'loss' of the remaining wild-type allele in a heterozygous GEM.

Haploinsufficiency

A phenotypic state that results from loss of one functional allele of any given gene in diploid cells. Sometimes also called allelic insufficiency.

Conditional models

GEM that rely on site-specific recombinase systems to engender gene expression in a spatially and/or temporally restricted manner.

Site-specific recombinase

(SSR). An enzyme, such as bacterial Cre, that catalyses recombination between two specific inverted repeat sequences (such as LoxP).

CreERT

A general term for a tamoxifen-inducible Cre recombinase in which the Cre cDNA is fused to the oestrogen receptor ligand-binding domain. This fusion protein is sequestered in the cytoplasm by heat shock proteins until exposure to Tamoxifen promotes nuclear translocation.

LoxStopLox

A LoxP-flanked sequence, often containing an antibiotic resistance marker, that prevents expression of a gene when placed between the promoter and coding exons.

Latent allele

An endogenous GEM strategy in which a stochastic recombination event is required to activate expression of the oncogenic allele. This is one of the few endogenous GEM that does not require a site-specific recombinase.

Hypomorphic allele

A mutation conveying decreased activity, either through reduced expression or partial loss of function.

Neomorphic allele

A mutation conveying a novel activity not present in the wild type protein.

Adeno-associated virus

(AAV) Small DNA viruses often used for gene therapy due to their broad host range and the ability to infect non-dividing cells.

Xenobiotic receptors

A family of enzymes involved in the metabolism of drugs by sensing the presence of a drug and initiating a response.

Cytochrome P450

A family of metabolic enzymes responsible for detoxifying and modifying drugs and other foreign compounds.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Frese, K., Tuveson, D. Maximizing mouse cancer models. Nat Rev Cancer 7, 654–658 (2007). https://doi.org/10.1038/nrc2192

Download citation

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrc2192

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing