Summary
Damaged DNA has a profound impact on mammalian health and overall survival. In addition to being the source of mutations that initiate cancer, the accumulation of toxic amounts of DNA damage can cause severe developmental diseases and accelerate ageing. Therefore, understanding how cells respond to DNA damage has become one of the most intense areas of biomedical research in the recent years. However, whereas most mechanistic studies derive from in vitro or in cellulo work, the impact of a given mutation on a living organism is largely unpredictable. For instance, why BRCA1 mutations preferentially lead to breast cancer whereas mutations compromising mismatch repair drive colon cancer is still not understood. In this context, evaluating the specific physiological impact of mutations that compromise genome integrity has become crucial for a better dimensioning of our knowledge. We here describe the various technologies that can be used for modeling mutations in mice, and provide a review of the genes and pathways that have been modeled so far in the context of DNA damage responses.
Keywords: Mouse models, DNA damage response, Cancer, Ageing, Knockout, Knock-in, Transgenic, Humanized Allele, Rare diseases
1. Introduction
Our genetic material is constantly challenged by lesions that distort its integrity. DNA can suffer from a myriad of alterations, which include base modifications, intra- or internucleotide covalent links, and the breakage of one or both strands of the double helix. Accordingly, many specialized DNA repair processes have evolved to be able to maintain genome integrity (Lindahl and Barnes, 2000). Among the different types of DNA lesions, DNA double strand breaks (DSB) are considered the most deleterious, since they may lead to irreversible loss of genetic material or to chromosomal rearrangements, being a serious threat to human health. As such, DSB are highly cytotoxic. The repair of a DSB can occur by a direct ligation of the ends regardless of homology or by a process that involves the processing of the broken ends and the repair of the damage by using the information from a homologous sequence, typically – but not always- this being the sister chromatid. The two canonical pathways are Non-Homologous End Joining (NHEJ) and Homologous Recombination (HR); although multiple variants of each process have been identified in the recent years (i.e. Microhomology Mediated End Joining or Break Induced Replication, among others). NHEJ and HR are exclusive mechanisms, and the decision of whether a DSB is channeled into one or the other is now an area of intense research. NHEJ, a simpler yet error-prone form of DSB repair, operates throughout the cell cycle and involves the KU70/KU80 complex that senses the ends, the Ligase IV/XRCC4 complex for end ligation, and the signaling kinase DNA-PK, among other activities. On the other hand, HR-mediated repair is more precise, yet it is only active during S and G2 phases of the cell cycle, once a sister chromatid is available as a template. In contrast to the limited number of proteins that have been involved in NHEJ, a large number of factors participate in the different activities associated to HR such as end processing (ATM, BRCA1), strand invasion (RAD51 and its paralogues).
Given that DSB constitute one of the most cytotoxic and dangerous genomic lesions, the term DNA damage response (DDR) is commonly used for the signaling machinery that is activated by DSB. In the presence of broken chromosomes, the DDR stimulates the repair of the lesion while it also initiates cytostatic or cytotoxic responses that limit the expansion of the damaged cells (Harper and Elledge, 2007). At the tip of the DDR, DSB activate a phosphorylation-based cascade that starts with the activation of the Ataxia Telangiectasia Mutated (ATM) kinase. If the ends are processed and single stranded DNA (ssDNA) is generated, then the ATM and Rad3-related kinase (ATR) is also activated. Whereas DNA-PK is also activated by DSB, it seems to mostly phosphorylate factors at the lesion to promote its repair, but with a minimal impact on cellular growth. Interestingly, DNAPK can also promote checkpoint responses in the absence of ATM, which might be an interesting concept for the targeting of ATM deficient human tumors (Callen et al., 2009). ATM and ATR can limit the expansion of the damaged cells by the phosphorylation and activation of their target checkpoint kinases CHK2 and CHK1, respectively, which in turn limit CDK activity and therefore cell growth. In addition, the DDR can also phosphorylate and activate apoptotic factors, most prominently p53, to promote the elimination of the damaged cell.
DSB not only arise from exogenous DNA damaging sources, but also as products of physiological recombination reactions during the generation of immune or meiotic diversity. Specifically, the processes of V(D)J and Class-Switch recombination (CRS) generate the repertoire of T-cell receptors (TCR) and Immunoglobulins (Ig) in developing lymphocytes (Dudley et al., 2005); and DSB generated by the SPO11 endonuclease initiate meiotic recombination during spermatogenesis and oogenesis (Zickler and Kleckner, 1998). Consequently, immune deficiencies and infertility are often found in diseases associated to DSB repair deficiencies. These phenotypes, as well as many others, are often faithfully recapitulated in mouse models of the respective human diseases, which consolidated the mouse as a valuable tool for the study of the impact of DNA repair deficiencies in a physiological context.
2. Modeling genomic instability in mice
During the past 30 years, the generation of mouse models has allowed to study the physiological function of many of DNA repair proteins and their role in cancer and other diseases. Several human diseases caused by inherited mutations that compromise genomic integrity have been successfully modeled in mouse, providing the scientific community with unique platforms to investigate the role of a given mutation in a living mammal (see Table 1). Whereas mouse models recapitulate many of the symptoms found in human diseases associated with deficient DNA repair, this is not the case for all symptoms. A paradigmatic example is Ataxia Telangiectasia, driven by deleterious mutations in the ATM kinase (Savitsky et al., 1995). Whereas ATM deficient mice also suffer from cancer and immune deficiencies like the human patients, they do not present obvert features of Ataxia nor Telangiectasia (Barlow et al., 1996, Elson et al., 1996, Xu et al., 1996). In general terms, mice with compromised DNA repair activities often present one or several of the following alterations: abnormal development, altered pigmentation, sub-Mendelian birth ratios, infertility, immunedeficiencies, premature ageing and/or cancer predisposition.
Table 1. Examples of mouse models related to DNA damage responses.
Not all the strains available for each gene are referenced due to space constrains, but only some of them as examples of the available models.
| Mutated gen | Human disease | Mouse models | References |
|---|---|---|---|
| Tp53bp1 | Not known | KO | (Manis et al., 2004) |
| Artemis | RS-SCID | KO | (Rooney et al., 2002) |
| Atm | Ataxia Telangiectasia (AT) | KO | (Barlow et al., 1996, Elson et al., 1996, Xu et al., 1996) |
| Atr | Seckel Syndrome | Humanized allele | (Murga et al., 2009) |
| Bard1 | Breast cancer susceptibility | KO, cKO | (McCarthy et al., 2003, Shakya et al., 2008) |
| Blm | Bloom’s Syndrome | KO, cKO | (Chester et al., 1998, Luo et al., 2000) |
| Brca1 | Familial breast cancer | KO, cKO, KI, cKI | (Hakem et al., 1996, Evers and Jonkers, 2006, Liu et al., 2007, Drost et al., 2011) |
| Chk1 | Not known | KO, cKO, BAC-Tg | (Takai et al., 2000, Zaugg et al., 2007, Lopez-Contreras et al., 2012) |
| Chk2 | Li Fraumeni síndrome | KO | (Hirao et al., 2000) |
| Csa (Ercc8) | Cockayne syndrome | KO | (van der Horst et al., 2002) |
| Csb (Ercc6) | Cockayne syndrome | KO | (van der Horst et al., 1997) |
| CtIP | Jawad síndrome | KO | (Chen et al., 2005) |
| DNA-PKcs | Not known | KO | (Kirchgessner et al., 1995) |
| Ercc1 | Cerebrooculofacioskeletal syndrome 4 | KO | (McWhir et al., 1993) |
| Ercc3 | Xenoderma Pigmentosum | KI | (Andressoo et al., 2009) |
| FancA | Fanconi anemia | KO | (Wong et al., 2003) |
| FancC | Fanconi anemia | KO | (Chen et al., 1996, Whitney et al., 1996) |
| FancD1 (Brca2) | Fanconi anemia; Familial breast cancer | KO, cKO | (Ludwig et al., 1997, Suzuki et al., 1997, Jonkers et al., 2001, Evers and Jonkers, 2006) |
| FancD2 | Fanconi anemia | KO | (Wong et al., 2003) |
| FancG | Fanconi anemia | KO | (Pulliam-Leath et al., 2010) |
| H2ax | Not known | KO | (Celeste et al., 2002) |
| Ku70 (Xrcc6) | Not known | KO | (Li et al., 1998) |
| Ku80 (Xrcc5) | Not known | KO | (Difilippantonio et al., 2000) |
| Ligase IV | RS-SCID | KO, Tg | (Frank et al., 1998, Rucci et al., 2010) |
| Mdc1 | Not Known | KO | (Lou et al., 2006) |
| Mlh1 | Hereditary nonpolyposis colorectal cancer | KO, Tg | (Baker et al., 1996) |
| Mlh3 | Colorectal cancer, Endometrial cancer | KO | (Lipkin et al., 2002, Avdievich et al., 2008) |
| Mre11 | AT-like disorder (ATLD) | KI | (Theunissen et al., 2003) |
| Msh2 | Hereditary nonpolyposis colorectal cancer | KO | (de Wind et al., 1995) |
| Msh6 | Hereditary nonpolyposis colorectal cancer; Familial endometrial cancer | KO | (Edelmann et al., 1997) |
| Mus81 | Not known | KO | (McPherson et al., 2004) |
| c-Myc | Burkitt lymphoma (when overexpressed in B cells) | Tg | (Harris et al., 1988) |
| Nbs1 | Nijmegen breakage síndrome | KO, cKO | (Zhu et al., 2001, Demuth et al., 2004, Stracker et al., 2007) |
| p21 | Not Known | KO, KI | (Wang et al., 1997, Barboza et al., 2006) |
| p53 | Li-Fraumeni | KO, cKO, KI, Tg, BAC-Tg | (Donehower et al., 1992, Godley et al., 1996, Ludwig et al., 1997, Allemand et al., 1999, Garcia-Cao et al., 2002, MacPherson et al., 2004, Sluss et al., 2004, Johnson and Attardi, 2006, Dickins et al., 2007) |
| Parp1 | Not known | KO, KI | (Pieper et al., 1999) |
| Pms2 | Hereditary nonpolyposis colorectal cancer | KO | (Baker et al., 1995) |
| Ptip | Susceptibility to Alzheimer disease | KO, cKO | (Cho et al., 2003, Daniel et al., 2010) |
| Rad51c | Fanconi anemia | KO, KI | (Kuznetsov et al., 2007) |
| Rad54b | Not known | KO | (Couedel et al., 2004, Mills et al., 2004) |
| Recql2 (Wrn) | Werner síndrome | KO | (Lebel and Leder, 1998) |
| Rnf168 | Riddle síndrome | KO | (Bohgaki et al., 2011) |
| Rnf8 | Not known | KO | (Santos et al., 2010) |
| Rpa1 | Not known | KI | (Wang et al., 2005) |
| Rtel | Dyskeratosis congénita | KO, cKO | (Ding et al., 2004, Wu et al., 2007) |
| Slx4 | Fanconi anemia | KO | (Crossan et al., 2011) |
| Topb1 | Not known | KO, cKO | (Yamane et al., 2002, Jeon et al., 2011) |
| Xlf | SCID with microcephaly, growth retardation, and sensitivity to ionizing radiation | KO, cKO | (Li et al., 2008, Zha et al., 2011) |
| Xpd (Ercc2) | Xeroderma pigmentosum | KI | (de Boer et al., 1998) |
| Xpf (Ercc4) | Xeroderma pigmentosum; Cockayne syndrome | KO | (Tian et al., 2004) |
| Xpg (Ercc5) | Xeroderma pigmentosum | KI | (Shiomi et al., 2004) |
| Xrcc1 | Not known | KO | (Tebbs et al., 1999) |
| Xrcc4 | Not known | KO, cKO | (Gao et al., 1998) |
| Zmpste24 (Face-1) | Hutchinson–Gilford progeria like syndrome | KO | (Pendas et al., 2002) |
Focusing on cancer, the connection with genomic instability was first noted by Theodor Boveri in the early XXth century (Boveri, 1914). In addition to the altered chromosome numbers detected by Boveri, tumor cells present important chromosomal rearrangements, which implies a defective DNA repair. Moreover, cancer cells acquire mutations at a faster pace than normal cells, (the so-called “mutator phenotype” (Jackson and Loeb, 1998)), which further supported that –at least some- cancerous mutations might relate to DNA Repair pathways. This concept was finally confirmed by the discovery of mutations affecting DNA Repair genes as drivers of hereditary cancers. Brca1 and Brca2 are the most frequently mutated genes in familiar breast cancer and their gene products are essential for the completion of HR (Apostolou and Fostira, 2013). In fact, given their relevance for human disease, dozens of BRCA1 and BRCA2 mouse models have been developed (Fig. 1) (Evers and Jonkers, 2006). Importantly, some of these models quite remarkably resemble the histopathological features of human BRCA1/2-deficient tumors. Yet, a remarkable difference between organisms exists as well; whereas full Brca1 or Brca2 heterozygosity are sufficient to drive tumorigenesis in humans, both alleles have to be mutated in mouse breast tissue for tumors to arise. In addition to HR, a fraction of hereditary colon cancers (Lynch Syndrome) is linked to deficiencies on Mismatch Repair (MMR), through mutations on genes such as Mlh1, Msh6 and Msh2. Whereas MMR-deficient mice are also tumor prone, tumors principally emerge on the lymphatic compartment and not on the intestinal tract (de Wind et al., 1998). Besides these examples that derived from previous knowledge in human tumors, it is fair to say that a common feature of mice deficient in DNA repair pathways is a higher propensity to develop malignancies (particularly lymphoid). Hence, cancer incidence is a common phenotype that is evaluated on mouse models of DNA Repair factors.
Figure 1. Examples of different gene targeting strategies that have been used to modify Brca1 in mice.
A) Structure of wild type Brca1. B) Example of a constitutive KO allele generated by introducing a neomycin-resistance cassette that replaces exon 5 and interrupts the reading frame. C) Conditional KO allele with loxP sites flanking exons 5 to 11. D) Conditional deletion of exon 11 results in a Δ11 version of BRCA1 protein with hypomorphic activity. This is a frequently used strain for BRCA1 studies given that animals lacking E11 almost complete embryogenesis, and embryonic fibroblasts can be obtained. E) A point mutation introduced in exon 5 (E5) replaces Cys 61 with Gly (C61G) in this constitutive KI allele, this being the mutation with the highest frequency detected in human Brca1 mutation carriers. F) An example of a conditional knockin allele of Brca1 generated by FLEX technology. Upon Cre recombination, the DNA fragment between loxP sites (black arrows) flips around and the E5 containing the C61G mutation is expressed. A second round of recombination between loxP155 sites (white arrows) excises the endogenous E5. The inversion-deletion process can also start with the inversion of the loxP155 sites, followed by the excision of the loxP-flanked sites. In both cases, the end product of the process is an allele that has lost the endogenous E5, and which has gained a C61G mutant version. The introduction of FRT sites flanking the neomycin-resistance cassette (grey arrows) enables the removal of this sequence upon crossing with a FLP expressing mouse strain. Note: Exon 4 of Brca1 gene was originally annotated improperly, which is why E4 is missing on the annotations.
In addition to hereditary cancers, mutations on DNA repair genes have also been associated to rare developmental disorders with a broader spectrum of clinical manifestations. As mentioned previously, Ataxia Telangiectasia (AT) is one of such syndromes and is caused by deleterious homozygous mutations in Atm. Again, whereas some of the symptoms of AT patients (i.e. radiosensitivity, immunedeficiency) were recapitulated on Atm deficient mice, the neurological symptoms which are arguably the most dramatic problem for these patients, are not present on the mutant mice. Another complication from modeling DNA repair-related diseases in mice arises from the fact that some of the mutations are essential on the rodents. This is the case for instance of the Bloom Syndrome (BS), which arises from the deficiency of the BLM helicase. Yet, BLM nullyzygosity is not compatible with mouse development (Chester et al., 1998). Perhaps one of the biggest challenges in the field has been the modeling of Fanconi Anemia (FA), a multigenic disease caused by mutations in more than a dozen genes involved on the repair of DNA interstrand crosslinks (ICL) (Moldovan and D'Andrea, 2009). FA is characterized by aplastic anemia, developmental problems and high predisposition to leukemia and other types of cancer. Whereas many of the FA genes have been mutated in mice, the FA phenotypes are very poorly recapitulated on the mouse mutants, if not inexistent (Parmar et al., 2009). On a positive note, however, it was only because of the use of mouse models that recent efforts revealed that endogenous aldehydes might be the natural genotoxic source that generates ICLs on FA patients, and thus responsible for their disease (Garaycoechea et al., 2012).
Intriguingly, rare diseases linked to accelerated ageing (progeria) also seem to have their basis on mutations that drive genomic instability. Two of the most widely known progeroid diseases, Hutchinson-Gilford Progeria (HGP) and the Werner Syndrome (WS) are linked to an accumulation of DNA damage. In the case of the WS, the mutation lies on WRN, a BLM-related helicase. Even if WRN deficient mice are not particularly progeroid, it has been proposed that this might be due to the longer telomere lengths of mice. Accordingly, work with mouse models revealed that WRN deficiency does in fact accelerate ageing in the context of short telomeres (Chang et al., 2004). As for the case of HGP, whereas the syndrome is caused by mutations on LmnA, related to the nuclear lamina and not to DNA repair (Eriksson et al., 2003), work with mouse models revealed that, for reasons yet to be understood, HGP cells accumulate substantial amounts of DNA damage (Liu et al., 2005). Other rare diseases in which patients with features of accelerated ageing are linked to mutations in DNA repair proteins are emerging rapidly with the help of the massive sequencing technologies. In our laboratory, we modeled one of those diseases known as the Seckel Syndrome (SS), which was shown to be driven by hypomorphic mutations on the ATR kinase (O'Driscoll et al., 2003). Based on the human mutation, we generated a humanized mouse model of the SS that faithfully recapitulated many of the diagnostic features of the human patients (Murga et al., 2009). Interestingly, ATR-Seckel mice accumulate high amounts of genomic instability during embryonic development, which then accelerates ageing on born animals. This finding led us to propose the concept of “Intrauterine Programming of Ageing”; namely that ageing rates can be determined by stresses to which we are exposed during embryonic development (Fernandez-Capetillo, 2010). Syndromes aside, the frequent finding of ageing features on DNA Repair mutant mice has been one of the essential facts to support the concept that ageing is driven by a progressive accumulation of DNA damage throughout our lives, particularly on stem cell compartments (Rossi et al., 2008).
We could not end this list of mouse models linked to ageing without mentioning the work done, significantly by the group of Jan Hoeijmakers, on mouse models of nucleotide excision repair (NER). Xeroderma Pigmentosum (XP) and Cockayne Syndrome (CS), among others, are hereditary diseases linked to NER deficiencies. In fact, it was XP and not breast or colon cancers the first link between DNA Repair mutations and cancer (Cleaver and Bootsma, 1975). XP is caused by mutations in NER genes (XPA, XPB, XPC and others), and patients present a high sensitivity to UV-light as well as a high cancer incidence. On the other hand, whereas CS is also caused by mutations in some NER genes, classically in ERCC8 and ERCC6, CS patients are more prone to accelerate ageing than cancer (Natale, 2011). A number of mouse models with mutations in NER genes have been developed presenting a broad spectrum of symptoms resembling XP and CS (Jaarsma et al., 2013). Noteworthy, different mutations on the very same NER gene can be pro-cancer or pro-ageing, further illustrating the intimate connection between these two outcomes. In addition to deepen our understanding of DNA Repair pathways, these mouse models also allowed Hoeijmakers to make the seminal discovery of a transcriptional signature of “aged” tissues (Niedernhofer et al., 2006).
In conclusion, the use of mouse models has allowed to solidify the concept that ageing and cancer can be initiated by DNA damage, and has allowed the generation of research models for human syndromes which can potentially be used for the discovery of new therapeutic strategies. Whereas we have also learned that many of the phenotypes linked to DNA Repair mutations are not shared between mice and humans, it is also true that many of the underlying mutations on the human patients are not simply deleterious, and some extra effort might be needed in order to generate models that faithfully recapitulate the human conditions. We here provide a broad overview, necessarily incomplete, of the most common technologies that have been used and can be used for mutagenizing the mouse genome for these purposes.
3. Technologies for the generation of mouse models
Techniques for the engineering of genetically modified (GM) mice were developed in the 1970ˋs-1980ˋs, fueled by mutual advances in developmental biology and molecular engineering (Fig. 2). In 1974, Rudolf Jaenisch generated the first transgenic mouse by retrovirus mediated infection of an early stage embryo, showing that this leads to integration of the foreign DNA into the host genome (Jaenisch and Mintz, 1974). Yet, it took until 1981 to show that injection of DNA into a pronucleus not only allows to engineer chimeric transgenic mice but also that this genetic material is passed on to subsequent generations (Brinster et al., 1981, Costantini and Lacy, 1981, Gordon and Ruddle, 1981). A key achievement for the field was the observation that murine embryonic stem (ES) cells can be isolated from blastocysts, cultured and re-introduced into a surrogate mother to generate chimeras (Evans and Kaufman, 1981, Martin, 1981, Bradley et al., 1984). Introduction of a selection marker gene (e.g. Neomycin) into ES-cells allowed to screen for successfully transfected clones prior to reintroducing them into the surrogate mother (Gossler et al., 1986) greatly enhancing efficiency of the method.
Figure 2. Timeline of the breakthroughs related to the use of genetically modified mouse models in research.
General breakthroughs are depicted in red, and some of the relevant models mentioned in this work in green.
With those technologies, GM mice were exclusively generated from constructs randomly integrating into the genome, giving rise to transgenic animals. The next breakthrough for gene targeting came, interestingly, with a better understanding of how a foreign DNA element can be introduced into a specific locus by exploiting recombination pathways (HR). The introduction of sequences into the foreign DNA that were identical to an endogenous locus allowed to introduce site-directed modifications and was first reported for the beta-globin locus by Smithies and colleagues (Smithies et al., 1985). Shortly thereafter, two groups reported the first successful generation of chimeric mice from gene-targeted ES-cells (Doetschman et al., 1987, Thomas and Capecchi, 1987). For these and previous findings, the 2007 Nobel prize in medicine was awarded to Capecchi, Sir Martin J. Evans and Oliver Smithies “for their discoveries of principles for introducing specific gene modifications in mice by the use of embryonic stem cells”. These were the seminal discoveries that enabled the explosion of mouse model research. Since then, the field has gradually improved by the creative design of targeting constructs and the introduction of in vivo recombination technologies. We here summarize some of the most frequent types of GM mice, exemplified through mouse models related to genome instability.
3.1. Transgenesis
Transgenic mice contain additional genetic material, ranging from additional copies of an endogenous gene to mutant versions of a gene of interest.
3.1.1. cDNA-based transgenesis
Most frequently, a transgenic mouse line consists of a cDNA of the gene of interest under the control of a short regulatory sequence. Genomic insertion of a transgene, generally delivered into a fertilized egg by pronuclear injection, occurs at random. As a consequence, identification of the insertion site and number of integrations is an important issue. In addition, comparison of the phenotype of various founder mouse lines should be conducted in order to exclude insertion site-specific phenotypes (e.g. embryonic lethality due to interruption of an essential gene). Working in heterozygosity is in these cases preferable to avoid phenotypes caused by transgene-integration.
The use of transgenic mice has dramatically increased our knowledge of (increased) gene function in the context of whole organisms. This was particularly relevant in cancer, due to the increased expression of oncogenes in human tumors. Hence, in the 1980’s, introducing oncogenes into the mouse genome generated the first so-called “oncomice”. Those early experiments led to the verification that ectopic expression of an oncogene in a healthy organism can lead to a heritable predisposition for cancer development (reviewed in (Hanahan et al., 2007). A prominent example of a cDNA-based transgenic mouse model is the Eμ-Myc mouse, a model for lymphomagenesis developed in 1985 by Brinster and colleagues (Adams et al., 1985). Ectopic overexpression of Myc is a frequent oncogenic mutation leading to aberrant regulation of proliferation and resulting in genomic instability (reviewed in (Campaner and Amati). The Eμ-Myc mouse carries a c-myc fusion gene under the control of the Ig heavy chain enhancer, which was based on the translocation found in Burkitt’s B-cell lymphoma (Adams, 1985). Mice carrying the Eμ-Myc transgene develop aggressive lymphoma or leukaemia and die within the first months of life (Adams, 1985, Langdon et al., 1986, Harris et al., 1988), and have been widely used as platforms for the discovery of genetic or chemical conditions that limit Myc-induced carcinogenesis.
Besides constitutive whole-body expression, transgenes can be placed under spatial (with the use of tissue specific promoters) or temporal (promoters that can be activated by orally available inert chemicals) control. Yet, one frequent limitation of cDNA based transgenesis is the lack of proper regulation of the studied gene.
3.1.2. BAC-transgenesis
cDNA-based transgenes lack the natural genomic regulation (introns, promoters, enhancers…), which may limit the interpretations of the results obtained. Moreover, unregulated expression might even limit the generation of the mouse lines. For instance, efforts to generate p53 transgenic mice repeatedly failed due to severe defects in embryogenesis, differentiation and apoptosis (Nakamura et al., 1995, Godley et al., 1996, Allemand et al., 1999). In contrast to classical transgenic vectors, bacterial artificial chromosomes (BAC) can harbor up to 300kb of DNA and can accommodate full genes along with their regulatory elements (Shizuya et al., 1992). The use of a BAC vector containing p53 in its natural environment omits the previously observed side effects of p53 overexpression, and allowed the generation of mice with enhanced p53 function (superP53) (Garcia-Cao et al., 2002). In the context of the DDR, our group showed that mice with just one extra allele of Chk1 showed a more efficient response against replication stress (RS) (Lopez-Contreras et al., 2012). The use of this strain allowed us to demonstrate that, unexpectedly, a modest increase in CHK1 levels (known as a tumor suppressor) could facilitate oncogenic transformation by limiting the genomic instability generated by the oncogenes. It should be noted that the use of these “super” mice, carrying a full extra copy of a gene might actually be physiologically relevant, given that duplications (some of which contain genes) are one of the most frequent source of diversity between human genomes. In fact, increased gene dosage of certain genes involved on the maintenance of genome stability, such as Replication Protein A (RPA), can counterintuitively lead to genomic instability in humans (Outwin et al., 2011).
3.1.3. shRNA-transgenesis
Inferring the function of a gene can also be done by turning it off, and looking at the consequences. This can also be done by transgenesis, taking advantage of the phenomenon of RNA interference (Hannon, 2002). Upon expression a transgenic shRNA triggers RNA interference of a specific messenger RNA in vivo, thus inducing a gene knockdown, which mimics genetic deletion to variable extents. shRNA transgenes can be made conditional by using the Tet system. The integrity of the endogenous target gene combined with the use of the Tet system allows for reversibility of the gene silencing. A well-known example of RNA interference-based mouse modeling is a strain carrying a transgene coding for a p53-targeting shRNA regulated by a tet-responsive promoter (rtTA) (Dickins et al., 2007). This strain develops lymphomas upon p53 downregulation, and lymphoma regression upon inactivation of the system. To follow-up with previous examples, this technology has also been used for proteins such as RPA, allowing the study of the consequences of sudden and broad genomic instability in adult mice (McJunkin et al., 2011).
3.2. Gene Targeting
3.2.1. Constitutive knockout
Knockout (KO) models are characterized by the inactivation of a specific gene, resulting in truncated, non-functional or absent protein expression. This method allows the study of gene function in vivo and is of notable interest since many human diseases, including some types of cancer, are caused by gene loss of function (for review see (Mak, 1998)). The vast majority of KO mice are generated from embryonic stem (ES) cell clones in which a particular gene was mutagenized through HR by a replacement vector lacking one or more essential coding exons. The replacement vector usually contains a drug resistances markers (e.g. Neomycin) to allow both negative and positive selection of the targeted cells (Hall et al., 2009). This being the most widely spread technology used, to date hundreds of different knockout mice have been generated. In fact, almost every discovery of a new DNA repair protein is followed up within the subsequent years with the generation of a knockout strain. There are thus too many knockout mice related to genomic instability to summarize. As examples of some of the most widely used we could single out ATM, Chk2, 53BP1 or H2AX deficient mice, perhaps ATM-/- being the most widely used as a DDR model by many investigators around the world.
3.2.2. Conditional knockout
Although traditional knockout mouse technology is a convenient way of studying protein function, the method has a number of important restrictions. Roughly 15 percent of constitutive KO models are unable to complete embryogenesis, this percentage being higher for proteins involved on essential repair pathways such as HR. This is the case for the ATM- and Rad3-related kinase ATR, which is the essential kinase that coordinates the response against replication stress (RS) (Brown and Baltimore, 2000, de Klein et al., 2000). Hence, in order to study ATR function at the organismal level, a conditional KO strain was generated (Brown and Baltimore, 2003). The use of these mice allowed to reveal that ATR deletion in adult mice accelerates ageing (Ruzankina et al., 2007). Of note, the ageing phenotype was not directly related to ATR in this case, but rather to the loss of adult stem cells that lost ATR, followed by a compensatory proliferation of the remaining ATR wild type cells. This phenotype of accelerated ageing is frequently observed upon the deletion of an essential gene in adult mice (which is usually not absolute, so that the wild type cells need to compensate for the loss of cells).
The conditional knockout technology relies on DNA recombinases such as Cre or FLP. These enzymes catalyze a recombination reaction between two target sequences, loxP in the case of Cre and FRT sites in the case of FLP (Sternberg et al., 1981, Sauer and Henderson, 1988). Depending on the relative orientation of the target sites, the result is the excision of the DNA contained in between the target sites (when the sequences are on the same orientation) or its inversion (for inverted sequences). For the generation of conditional KO mice, loxP sites are usually placed at both sides of an essential exon of the target gene, preferably by targeting an exon (or exons) that would lead to an altered reading frame upon deletion. The “floxed” mice are then bred with Cre expressing transgenic animals that can direct the expression to the recombinase, and thus the deletion, to a particular tissue. The use of conditional knockout models demands a similar development of Cre recombinases. We now have hundredths of tissue specific Cre lines (an updated database is mantained by the Nagy laboratory at http://nagy.mshri.on.ca/cre_new/) which can be used in combination of floxed alleles. The number of Flp carrying lines is also rapidly evolving. Importantly, a tamoxifen-inducible version of Cre (Cre-ERT2 in its most evolved version), allows for not only a tissue-specific expression of the Cre (given by the promoter), but also for a temporal control of its activation (in response to an inert derivative of tamoxifen, 4-hydroxy-tamoxifen).
The use of conditional knockout alleles is of particular relevance for the study of essential genes or for cases in which the phenotype of a KO strain is so aggressive that precludes the study of additional phenotypes. For instance, while p53 KO mice provided an important model for studying p53ˋs role in tumor development (Donehower et al., 1992), these mice almost invariably succumb to aggressive lymphomas. The incidence of lymphoma (and in some cases sarcoma) does not allow for other tumor types to develop before animals succumb to disease. However, the most frequently observed human cancer type carrying p53-mutations are carcinomas (for review see (Johnson and Attardi, 2006). Hence, in order to induce different types of tumors and study the consequences of p53 loss in other tissues, a number of tissue-specific p53 KO models have been employed. In regards to genomic instability, the use of conditional knockouts has been particularly useful for the study of BRCA1-associated breast tumors (reviewed in (Evers and Jonkers, 2006)). In contrast to the high incidence of tumors on human BRCA1 mutation carriers, mammary-specific deletion of BRCA1, together with p53, is needed in mice for tumors to develop.
3.2.3. Knockin
Besides deleting a protein, specific mutations can also be targeted or knocked in at a particular genomic location (knockin; KI). KI alleles can introduce point mutations, incorporate exogenous DNA fragments, or replace large endogenous DNA fragments with exogenous ones. These strategies provide a variety of possibilities for studying specific functions of proteins in vivo, modeling disease and generating tools for mouse engineering. Knockin strategies have been widely used to interrogate the role of specific protein residues, such as phosphorylation sites or key residues at catalytic sites. Relevant examples are the p53 phosphorylation defective mutant mice generated at the laboratories of Stephen N. Jones and Tyler Jacks (MacPherson et al., 2004, Sluss et al., 2004). Absence of p53 protein is known to be a predisposing condition to tumor development as shown in various KO models (Donehower et al., 1992). Serine residues 18 and 23 were proposed to be critical phosphorylation sites for p53-mediated tumor suppression (Oliner et al., 1993). To test this in a physiological context, several studies used KI mice carrying point mutations at the endogenous p53 locus that replaced serines 18 and 23 with alanines (MacPherson et al., 2004, Sluss et al., 2004, Chao et al., 2006, Armata et al., 2007). Unexpectedly, these studies showed that such phosphorylations only had a minor contribution to p53 tumor suppressive function. In contrast, in some other cases the use of knockin alleles of phosphorylation sites previously shown to be relevant in vitro revealed a lack of a phenotype in vivo. This was for instance the case of the ATM kinase, where previous studies had shown that autophosphorylation at S1981 was essential for its activation (Bakkenist and Kastan, 2003). However, mice with the corresponding residue mutated to Alanine do not show any obvious phenotype (Pellegrini et al., 2006).
3.2.3. a Humanized alleles
In some cases, the study of a mouse model might have the limitations that the orthologue sequence is not identical to the human sequence. Yet, strategies have also evolved that enable the study of the human sequence within the mouse genome. One of those strategies implies the modification of a mouse locus with the corresponding human sequences, so that the allele becomes “humanized”. This strategy was used to develop a mouse model for ATR-Seckel syndrome (Murga et al., 2009). The Seckel syndrome is a genetic disease characterized by a severe growth retardation, microcephaly. A subset of Seckel patients carry a synonymous point mutation in the ATR gene that leads to reduced protein levels due to abnormal splicing (O’Driscoll et al., 2003). On ATR-Seckel mice, the human region containing the mutated exon, together with the surrounding 2 exons and introns was swapped with the corresponding mouse region. This humanized allele rendered mice expressing very low levels of ATR, which faithfully recapitulated the symptoms observed in the patients (Murga et al., 2009). In addition, ATR-Seckel mice were used to reveal that low ATR levels are particularly toxic for cells expressing oncogenes (Murga et al., 2011, Toledo et al., 2011). Of note, other strategies have also been used within the genomic instability for the generation of humanized mice. For instance, transgene-introduction of mutated BACs containing the entire human NBS1 locus, in the context of the deletion of the murine NBS1 gene, generated humanized alleles of disease-relevant NBS1 mutations (Difilippantonio et al., 2005).
3.2.4. Conditional knockin
Some targeted mutations are early embryonic lethal in mice when expressed constitutively. One possibility to overcome this limitation is having one KI allele harboring the mutation and the other a cKO. Upon Cre-mediated recombination the conditional allele is excised and only the KI mutant allele is expressed. One example of this strategy was done with BRCA1 C61G mutant mice (Drost et al., 2011). The C61G mutation at the RING domain of BRCA is the most frequent missense mutation in familial breast cancer. Yet, homozygosity for this mutation is embryonic lethal in mice. Using this strategy the authors were able to express the C61G mutant protein specifically in mammary gland, which led to mammary tumors (Drost et al., 2011). However, it must be noted that using this approach only half of the mutant protein is expressed. In the recent years, genome-engineering technologies have led to the development of technologies that enable a controlled expression of the mutant allele. One of such approaches is the FLEX switch (Schnutgen et al., 2003) (Fig. 1e). This strategy relies on the use of heterotypic target sites for the Cre recombinase placed on opposite orientations, which promote the inversion of the flanked sequence. FLEX conditional mutant alleles are built around an exon that harbours the mutation of interest. The allele is composed of a forward-oriented endogenous version of the exon, followed by its mutated version in the reverse orientation. Both are flanked by two pairs of different heterotypic recombinase target sites with a specific orientation. While the endogenous version is expressed constitutively, an allele subjected to Cre recombination will “flip around” and stabilize the reverse orientation. In this manner the mutant version is switched on, while the endogenous one is switched off. Whereas no examples of the FLEX technology have been published on the field of genomic instability yet, we have developed a conditional mouse model of the BRCA1 C61G mutation described above which works as expected (Lopez-Contreras et al., unpublished).
3.2.5. Gene trapping strategies
While directed gene targeting provides an invaluable tool to study specific gene function, it remains a relatively labor intensive and expensive method. One alternative to directed mutagenesis is to randomly mutagenize murine ES cells with gene trapping constructs, which can then be used for the generation of mutant mice. A typical gene trap cassette contains a strong 5´splice acceptor sequence followed by a promotorless antibiotic resistance gene and/or a reporter gene (e.g. beta-gal). When inserted within an intron of an active gene, upstream exons are spliced to the resistance-reporter cassette, leading to a truncated protein. In this manner, it is preferable to obtain alleles that have been trapped at the beginning of the gene, to maximize the probability of getting severe protein dysfunction. Since the endogenous promoter of the trapped gene drives the expression of the reporter, these alleles can also be used to study the dynamics of gene expression in vivo. Mutagenesis of the ES cells is achieved by the integration of the genetrap construct, which has been mostly done by retroviral transduction, but also by other methods (i.e. use of transposons). Once mutant cells are selected with the use of the selectable markers, the identification of the trapped gene can be done either by identifying the exon previous to the genetrap (5’RACE), or by directly mapping the precise position of the insertion on the genome through inverse PCR and related protocols. Genetrap-mediated gene silencing in ES cells was already described in the early 90’s (von Melchner et al., 1992), and subsequently used for screenings in vitro (Forrester et al., 1996). Genetrap based screenings have the limitation that only one allele is trapped, so that the phenotypes have to be evident in heterozygosis. This limitation was recently overcome by the development of haploid murine ES cells (see below). Yet, the easiness of this process launched the creation of large libraries of mutant ES cells, which investigators could then obtain to generate their mouse strains. For instance, one of the earliest collections was that of the German Genetrap Consortium (http://www.genetrap.de), which already in 2003 reported the generation of more than five thousand characterized ES clones. Other consortia are available such as the International Gene Trap Consortium (http://www.genetrap.org), BayGenomics (https://www.mmrrc.org/catalog/overview_BG.php), the Center for Modelling Human Disease (http://www.cmhd.ca/genetrap/index.html) or the collections at the Japanese RIKEN Institute (http://www2.brc.riken.jp/lab/mouse_es/index.html.en). These collections have been of great use to investigators on many fields, including the DDR. For instance, 53BP1 (Morales et al., 2003), and its upstream regulators RNF8 (Minter-Dykhouse et al., 2008) and RNF168 (Bohgaki et al., 2011) were all modeled by genetrapped strains. Besides genetrap collections, it should be mentioned that the International Knockout Mouse Consortium has more recently took the challenge of developing constitutive and conditional gene targeting constructs for all mouse genes (Skarnes et al., 2011). The Consortium distributes modified ES cells, targeting vectors, or even living mice in the case that they have already been generated. In fact, the ultimate aim of the project is to phenotype mouse mutants for every murine gene, and the effort is now run under a collaborative project named the International Mouse Phenotyping Consortium (http://www.mousephenotype.org/martsearch_ikmc_project/). Many investigators, including us, have benefited from this effort that greatly speeds up the creation of KO or cKO strains.
4. Conclusions and future
The accumulation of DNA damage has its roots on pathologies such as cancer, immunedeficiencies or accelerated ageing, which can only be investigated in the context of a living animal. We have here summarized the technologies that can be used to mutagenize the mouse genome in order to generate mutant mice, with a focus on the efforts that have been placed on mice related to DNA damage repair. Given the large number of mice and technologies available, this work is not fully comprehensive, and we apologize for the many mice and technologies that have not been mentioned for space reasons. As for the future, among the many technologies that are emerging, two deserve particular attention. On one hand, the rapid development of sequence specific nucleases of the CRISPR-Cas9 system allows for a very effective site-specific manipulation of the mammalian genome, even in cell lines that are not particularly recombinogenic such as most of the human cell lines used in the laboratory (Ran et al., 2013). Not surprisingly, this method has been rapidly adapted to mutagenize murine ES cells. Its high targeting efficiency and ease of use allows for the simultaneous targeting of several genes, which will greatly reduce the time to generate compound mutants (Wang et al., 2013, Yang et al., 2013). In addition to CRISPR-Cas9, the generation of mammalian haploid cells, including murine ES cells, constitutes another recent breakthrough that will further facilitate the fast generation of mice homozygous for one or more mutations (Leeb and Wutz, 2011). With the incoming technologies, there has never been a time in which the generation of mutant mice has been easier than it is now. All fields of biomedicine should benefit from these developments that will help us dimension the relevance of our in vitro findings in the context of a living mammal.
Acknowledgements
Work in OF laboratory is supported by the Spanish Ministry of Science (SAF2011-23753), Association for International Cancer Research (12-0229), Fundació La Marató de TV3 (33/C(2013), Howard Hughes Medical Institute and the European Research Council (ERC-210520). JS is a recipient of a predoctoral fellowship from the Spanish Government (FPI-2012). MN is funded by a predoctoral fellowship from the La Caixa Foundation. AJL is a recipient of a postdoctoral fellowship from the Spanish Association Against Cancer (AECC).
References
- Adams JM. Oncogene activation by fusion of chromosomes in leukaemia. Nature. 1985;315:541–2. doi: 10.1038/315542a0. [DOI] [PubMed] [Google Scholar]
- Adams JM, Harris AW, Pinkert CA, Corcoran LM, Alexander WS, Cory S, et al. The c-myc oncogene driven by immunoglobulin enhancers induces lymphoid malignancy in transgenic mice. Nature. 1985;318:533–8. doi: 10.1038/318533a0. [DOI] [PubMed] [Google Scholar]
- Allemand I, Anglo A, Jeantet AY, Cerutti I, May E. Testicular wild-type p53 expression in transgenic mice induces spermiogenesis alterations ranging from differentiation defects to apoptosis. Oncogene. 1999;18:6521–30. doi: 10.1038/sj.onc.1203052. [DOI] [PubMed] [Google Scholar]
- Andressoo JO, Weeda G, de Wit J, Mitchell JR, Beems RB, van Steeg H, et al. An Xpb mouse model for combined xeroderma pigmentosum and cockayne syndrome reveals progeroid features upon further attenuation of DNA repair. Mol Cell Biol. 2009;29:1276–90. doi: 10.1128/MCB.01229-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Apostolou P, Fostira F. Hereditary breast cancer: the era of new susceptibility genes. Biomed Res Int. 2013;2013:747318. doi: 10.1155/2013/747318. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Armata HL, Garlick DS, Sluss HK. The ataxia telangiectasia-mutated target site Ser18 is required for p53-mediated tumor suppression. Cancer Res. 2007;67:11696–703. doi: 10.1158/0008-5472.CAN-07-1610. [DOI] [PubMed] [Google Scholar]
- Avdievich E, Reiss C, Scherer SJ, Zhang Y, Maier SM, Jin B, et al. Distinct effects of the recurrent Mlh1G67R mutation on MMR functions, cancer, and meiosis. Proc Natl Acad Sci U S A. 2008;105:4247–52. doi: 10.1073/pnas.0800276105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Baker SM, Bronner CE, Zhang L, Plug AW, Robatzek M, Warren G, et al. Male mice defective in the DNA mismatch repair gene PMS2 exhibit abnormal chromosome synapsis in meiosis. Cell. 1995;82:309–19. doi: 10.1016/0092-8674(95)90318-6. [DOI] [PubMed] [Google Scholar]
- Baker SM, Plug AW, Prolla TA, Bronner CE, Harris AC, Yao X, et al. Involvement of mouse Mlh1 in DNA mismatch repair and meiotic crossing over. Nat Genet. 1996;13:336–42. doi: 10.1038/ng0796-336. [DOI] [PubMed] [Google Scholar]
- Bakkenist CJ, Kastan MB. DNA damage activates ATM through intermolecular autophosphorylation and dimer dissociation. Nature. 2003;421:499–506. doi: 10.1038/nature01368. [DOI] [PubMed] [Google Scholar]
- Barboza JA, Liu G, Ju Z, El-Naggar AK, Lozano G. p21 delays tumor onset by preservation of chromosomal stability. Proc Natl Acad Sci U S A. 2006;103:19842–7. doi: 10.1073/pnas.0606343104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barlow C, Hirotsune S, Paylor R, Liyanage M, Eckhaus M, Collins F, et al. Atm-deficient mice: a paradigm of ataxia telangiectasia. Cell. 1996;86:159–71. doi: 10.1016/s0092-8674(00)80086-0. [DOI] [PubMed] [Google Scholar]
- Bohgaki T, Bohgaki M, Cardoso R, Panier S, Zeegers D, Li L, et al. Genomic instability, defective spermatogenesis, immunodeficiency, and cancer in a mouse model of the RIDDLE syndrome. PLoS Genet. 2011;7:e1001381. doi: 10.1371/journal.pgen.1001381. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boveri T. Zur Frage der Entstehung Maligner Tumoren. Gustav Fischer; Jena: 1914. [Google Scholar]
- Bradley A, Evans M, Kaufman MH, Robertson E. Formation of germ-line chimaeras from embryo-derived teratocarcinoma cell lines. Nature. 1984;309:255–6. doi: 10.1038/309255a0. [DOI] [PubMed] [Google Scholar]
- Brinster RL, Chen HY, Trumbauer M, Senear AW, Warren R, Palmiter RD. Somatic expression of herpes thymidine kinase in mice following injection of a fusion gene into eggs. Cell. 1981;27:223–31. doi: 10.1016/0092-8674(81)90376-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brown EJ, Baltimore D. ATR disruption leads to chromosomal fragmentation and early embryonic lethality. Genes Dev. 2000;14:397–402. [PMC free article] [PubMed] [Google Scholar]
- Brown EJ, Baltimore D. Essential and dispensable roles of ATR in cell cycle arrest and genome maintenance. Genes Dev. 2003;17:615–28. doi: 10.1101/gad.1067403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Callen E, Jankovic M, Wong N, Zha S, Chen HT, Difilippantonio S, et al. Essential role for DNA-PKcs in DNA double-strand break repair and apoptosis in ATM-deficient lymphocytes. Mol Cell. 2009;34:285–97. doi: 10.1016/j.molcel.2009.04.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Campaner S, Amati B. Two sides of the Myc-induced DNA damage response: from tumor suppression to tumor maintenance. Cell Div. 7:6. doi: 10.1186/1747-1028-7-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Celeste A, Petersen S, Romanienko PJ, Fernandez-Capetillo O, Chen HT, Sedelnikova OA, et al. Genomic instability in mice lacking histone H2AX. Science. 2002;296:922–7. doi: 10.1126/science.1069398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cleaver JE, Bootsma D. Xeroderma pigmentosum: biochemical and genetic characteristics. Annu Rev Genet. 1975;9:19–38. doi: 10.1146/annurev.ge.09.120175.000315. [DOI] [PubMed] [Google Scholar]
- Costantini F, Lacy E. Introduction of a rabbit beta-globin gene into the mouse germ line. Nature. 1981;294:92–4. doi: 10.1038/294092a0. [DOI] [PubMed] [Google Scholar]
- Couedel C, Mills KD, Barchi M, Shen L, Olshen A, Johnson RD, et al. Collaboration of homologous recombination and nonhomologous end-joining factors for the survival and integrity of mice and cells. Genes Dev. 2004;18:1293–304. doi: 10.1101/gad.1209204. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Crossan GP, van der Weyden L, Rosado IV, Langevin F, Gaillard PH, McIntyre RE, et al. Disruption of mouse Slx4, a regulator of structure-specific nucleases, phenocopies Fanconi anemia. Nat Genet. 2011;43:147–52. doi: 10.1038/ng.752. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chang S, Multani AS, Cabrera NG, Naylor ML, Laud P, Lombard D, et al. Essential role of limiting telomeres in the pathogenesis of Werner syndrome. Nat Genet. 2004;36:877–82. doi: 10.1038/ng1389. [DOI] [PubMed] [Google Scholar]
- Chao C, Herr D, Chun J, Xu Y. Ser18 and 23 phosphorylation is required for p53-dependent apoptosis and tumor suppression. EMBO J. 2006;25:2615–22. doi: 10.1038/sj.emboj.7601167. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen M, Tomkins DJ, Auerbach W, McKerlie C, Youssoufian H, Liu L, et al. Inactivation of Fac in mice produces inducible chromosomal instability and reduced fertility reminiscent of Fanconi anaemia. Nat Genet. 1996;12:448–51. doi: 10.1038/ng0496-448. [DOI] [PubMed] [Google Scholar]
- Chen PL, Liu F, Cai S, Lin X, Li A, Chen Y, et al. Inactivation of CtIP leads to early embryonic lethality mediated by G1 restraint and to tumorigenesis by haploid insufficiency. Mol Cell Biol. 2005;25:3535–42. doi: 10.1128/MCB.25.9.3535-3542.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chester N, Kuo F, Kozak C, O'Hara CD, Leder P. Stage-specific apoptosis, developmental delay, and embryonic lethality in mice homozygous for a targeted disruption in the murine Bloom’s syndrome gene. Genes Dev. 1998;12:3382–93. doi: 10.1101/gad.12.21.3382. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cho EA, Prindle MJ, Dressler GR. BRCT domain-containing protein PTIP is essential for progression through mitosis. Mol Cell Biol. 2003;23:1666–73. doi: 10.1128/MCB.23.5.1666-1673.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Daniel JA, Santos MA, Wang Z, Zang C, Schwab KR, Jankovic M, et al. PTIP promotes chromatin changes critical for immunoglobulin class switch recombination. Science. 2010;329:917–23. doi: 10.1126/science.1187942. [DOI] [PMC free article] [PubMed] [Google Scholar]
- de Boer J, de Wit J, van Steeg H, Berg RJ, Morreau H, Visser P, et al. A mouse model for the basal transcription/DNA repair syndrome trichothiodystrophy. Mol Cell. 1998;1:981–90. doi: 10.1016/s1097-2765(00)80098-2. [DOI] [PubMed] [Google Scholar]
- de Klein A, Muijtjens M, van Os R, Verhoeven Y, Smit B, Carr AM, et al. Targeted disruption of the cell-cycle checkpoint gene ATR leads to early embryonic lethality in mice. Curr Biol. 2000;10:479–82. doi: 10.1016/s0960-9822(00)00447-4. [DOI] [PubMed] [Google Scholar]
- de Wind N, Dekker M, Berns A, Radman M, te Riele H. Inactivation of the mouse Msh2 gene results in mismatch repair deficiency, methylation tolerance, hyperrecombination, and predisposition to cancer. Cell. 1995;82:321–30. doi: 10.1016/0092-8674(95)90319-4. [DOI] [PubMed] [Google Scholar]
- de Wind N, Dekker M, van Rossum A, van der Valk M, te Riele H. Mouse models for hereditary nonpolyposis colorectal cancer. Cancer Res. 1998;58:248–55. [PubMed] [Google Scholar]
- Demuth I, Frappart PO, Hildebrand G, Melchers A, Lobitz S, Stockl L, et al. An inducible null mutant murine model of Nijmegen breakage syndrome proves the essential function of NBS1 in chromosomal stability and cell viability. Hum Mol Genet. 2004;13:2385–97. doi: 10.1093/hmg/ddh278. [DOI] [PubMed] [Google Scholar]
- Dickins RA, McJunkin K, Hernando E, Premsrirut PK, Krizhanovsky V, Burgess DJ, et al. Tissue-specific and reversible RNA interference in transgenic mice. Nat Genet. 2007;39:914–21. doi: 10.1038/ng2045. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Difilippantonio MJ, Zhu J, Chen HT, Meffre E, Nussenzweig MC, Max EE, et al. DNA repair protein Ku80 suppresses chromosomal aberrations and malignant transformation. Nature. 2000;404:510–4. doi: 10.1038/35006670. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Difilippantonio S, Celeste A, Fernandez-Capetillo O, Chen HT, Reina San Martin B, Van Laethem F, et al. Role of Nbs1 in the activation of the Atm kinase revealed in humanized mouse models. Nat Cell Biol. 2005;7:675–85. doi: 10.1038/ncb1270. [DOI] [PubMed] [Google Scholar]
- Ding H, Schertzer M, Wu X, Gertsenstein M, Selig S, Kammori M, et al. Regulation of murine telomere length by Rtel: an essential gene encoding a helicase-like protein. Cell. 2004;117:873–86. doi: 10.1016/j.cell.2004.05.026. [DOI] [PubMed] [Google Scholar]
- Doetschman T, Gregg RG, Maeda N, Hooper ML, Melton DW, Thompson S, et al. Targetted correction of a mutant HPRT gene in mouse embryonic stem cells. Nature. 1987;330:576–8. doi: 10.1038/330576a0. [DOI] [PubMed] [Google Scholar]
- Donehower LA, Harvey M, Slagle BL, McArthur MJ, Montgomery CA, Jr, Butel JS, et al. Mice deficient for p53 are developmentally normal but susceptible to spontaneous tumours. Nature. 1992;356:215–21. doi: 10.1038/356215a0. [DOI] [PubMed] [Google Scholar]
- Drost R, Bouwman P, Rottenberg S, Boon U, Schut E, Klarenbeek S, et al. BRCA1 RING function is essential for tumor suppression but dispensable for therapy resistance. Cancer Cell. 2011;20:797–809. doi: 10.1016/j.ccr.2011.11.014. [DOI] [PubMed] [Google Scholar]
- Dudley DD, Chaudhuri J, Bassing CH, Alt FW. Mechanism and control of V(D)J recombination versus class switch recombination: similarities and differences. Adv Immunol. 2005;86:43–112. doi: 10.1016/S0065-2776(04)86002-4. [DOI] [PubMed] [Google Scholar]
- Edelmann W, Yang K, Umar A, Heyer J, Lau K, Fan K, et al. Mutation in the mismatch repair gene Msh6 causes cancer susceptibility. Cell. 1997;91:467–77. doi: 10.1016/s0092-8674(00)80433-x. [DOI] [PubMed] [Google Scholar]
- Elson A, Wang Y, Daugherty CJ, Morton CC, Zhou F, Campos-Torres J, et al. Pleiotropic defects in ataxia-telangiectasia protein-deficient mice. Proc Natl Acad Sci U S A. 1996;93:13084–9. doi: 10.1073/pnas.93.23.13084. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eriksson M, Brown WT, Gordon LB, Glynn MW, Singer J, Scott L, et al. Recurrent de novo point mutations in lamin A cause Hutchinson-Gilford progeria syndrome. Nature. 2003;423:293–8. doi: 10.1038/nature01629. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Evans MJ, Kaufman MH. Establishment in culture of pluripotential cells from mouse embryos. Nature. 1981;292:154–6. doi: 10.1038/292154a0. [DOI] [PubMed] [Google Scholar]
- Evers B, Jonkers J. Mouse models of BRCA1 and BRCA2 deficiency: past lessons, current understanding and future prospects. Oncogene. 2006;25:5885–97. doi: 10.1038/sj.onc.1209871. [DOI] [PubMed] [Google Scholar]
- Fernandez-Capetillo O. Intrauterine programming of ageing. EMBO Rep. 2010;11:32–6. doi: 10.1038/embor.2009.262. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Forrester LM, Nagy A, Sam M, Watt A, Stevenson L, Bernstein A, et al. An induction gene trap screen in embryonic stem cells: Identification of genes that respond to retinoic acid in vitro. Proc Natl Acad Sci U S A. 1996;93:1677–82. doi: 10.1073/pnas.93.4.1677. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Frank KM, Sekiguchi JM, Seidl KJ, Swat W, Rathbun GA, Cheng HL, et al. Late embryonic lethality and impaired V(D)J recombination in mice lacking DNA ligase IV. Nature. 1998;396:173–7. doi: 10.1038/24172. [DOI] [PubMed] [Google Scholar]
- Gao Y, Sun Y, Frank KM, Dikkes P, Fujiwara Y, Seidl KJ, et al. A critical role for DNA end-joining proteins in both lymphogenesis and neurogenesis. Cell. 1998;95:891–902. doi: 10.1016/s0092-8674(00)81714-6. [DOI] [PubMed] [Google Scholar]
- Garaycoechea JI, Crossan GP, Langevin F, Daly M, Arends MJ, Patel KJ. Genotoxic consequences of endogenous aldehydes on mouse haematopoietic stem cell function. Nature. 2012;489:571–5. doi: 10.1038/nature11368. [DOI] [PubMed] [Google Scholar]
- Garcia-Cao I, Garcia-Cao M, Martin-Caballero J, Criado LM, Klatt P, Flores JM, et al. “Super p53” mice exhibit enhanced DNA damage response, are tumor resistant and age normally. EMBO J. 2002;21:6225–35. doi: 10.1093/emboj/cdf595. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Godley LA, Kopp JB, Eckhaus M, Paglino JJ, Owens J, Varmus HE. Wild-type p53 transgenic mice exhibit altered differentiation of the ureteric bud and possess small kidneys. Genes Dev. 1996;10:836–50. doi: 10.1101/gad.10.7.836. [DOI] [PubMed] [Google Scholar]
- Gordon JW, Ruddle FH. Integration and stable germ line transmission of genes injected into mouse pronuclei. Science. 1981;214:1244–6. doi: 10.1126/science.6272397. [DOI] [PubMed] [Google Scholar]
- Gossler A, Doetschman T, Korn R, Serfling E, Kemler R. Transgenesis by means of blastocyst-derived embryonic stem cell lines. Proc Natl Acad Sci U S A. 1986;83:9065–9. doi: 10.1073/pnas.83.23.9065. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hakem R, de la Pompa JL, Sirard C, Mo R, Woo M, Hakem A, et al. The tumor suppressor gene Brca1 is required for embryonic cellular proliferation in the mouse. Cell. 1996;85:1009–23. doi: 10.1016/s0092-8674(00)81302-1. [DOI] [PubMed] [Google Scholar]
- Hall B, Limaye A, Kulkarni AB. Overview: generation of gene knockout mice. Curr Protoc Cell Biol. 2009:1–7. doi: 10.1002/0471143030.cb1912s44. Chapter 19 Unit 19 2 2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hanahan D, Wagner EF, Palmiter RD. The origins of oncomice: a history of the first transgenic mice genetically engineered to develop cancer. Genes Dev. 2007;21:2258–70. doi: 10.1101/gad.1583307. [DOI] [PubMed] [Google Scholar]
- Hannon GJ. RNA interference. Nature. 2002;418:244–51. doi: 10.1038/418244a. [DOI] [PubMed] [Google Scholar]
- Hansen J, Floss T, Van Sloun P, Fuchtbauer EM, Vauti F, Arnold HH, et al. A large-scale, gene-driven mutagenesis approach for the functional analysis of the mouse genome. Proc Natl Acad Sci U S A. 2003;100:9918–22. doi: 10.1073/pnas.1633296100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Harper JW, Elledge SJ. The DNA damage response: ten years after. Mol Cell. 2007;28:739–45. doi: 10.1016/j.molcel.2007.11.015. [DOI] [PubMed] [Google Scholar]
- Harris AW, Pinkert CA, Crawford M, Langdon WY, Brinster RL, Adams JM. The E mu-myc transgenic mouse. A model for high-incidence spontaneous lymphoma and leukemia of early B cells. J Exp Med. 1988;167:353–71. doi: 10.1084/jem.167.2.353. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hirao A, Kong YY, Matsuoka S, Wakeham A, Ruland J, Yoshida H, et al. DNA damage-induced activation of p53 by the checkpoint kinase Chk2. Science. 2000;287:1824–7. doi: 10.1126/science.287.5459.1824. [DOI] [PubMed] [Google Scholar]
- Jaarsma D, van der Pluijm I, van der Horst GT, Hoeijmakers JH. Cockayne syndrome pathogenesis: lessons from mouse models. Mech Ageing Dev. 2013;134:180–95. doi: 10.1016/j.mad.2013.04.003. [DOI] [PubMed] [Google Scholar]
- Jackson AL, Loeb LA. On the origin of multiple mutations in human cancers. Semin Cancer Biol. 1998;8:421–9. doi: 10.1006/scbi.1998.0113. [DOI] [PubMed] [Google Scholar]
- Jaenisch R, Mintz B. Simian virus 40 DNA sequences in DNA of healthy adult mice derived from preimplantation blastocysts injected with viral DNA. Proc Natl Acad Sci U S A. 1974;71:1250–4. doi: 10.1073/pnas.71.4.1250. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jeon Y, Ko E, Lee KY, Ko MJ, Park SY, Kang J, et al. TopBP1 deficiency causes an early embryonic lethality and induces cellular senescence in primary cells. J Biol Chem. 2011;286:5414–22. doi: 10.1074/jbc.M110.189704. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Johnson TM, Attardi LD. Dissecting p53 tumor suppressor function in vivo through the analysis of genetically modified mice. Cell Death Differ. 2006;13:902–8. doi: 10.1038/sj.cdd.4401902. [DOI] [PubMed] [Google Scholar]
- Jonkers J, Meuwissen R, van der Gulden H, Peterse H, van der Valk M, Berns A. Synergistic tumor suppressor activity of BRCA2 and p53 in a conditional mouse model for breast cancer. Nat Genet. 2001;29:418–25. doi: 10.1038/ng747. [DOI] [PubMed] [Google Scholar]
- Kirchgessner CU, Patil CK, Evans JW, Cuomo CA, Fried LM, Carter T, et al. DNA-dependent kinase (p350) as a candidate gene for the murine SCID defect. Science. 1995;267:1178–83. doi: 10.1126/science.7855601. [DOI] [PubMed] [Google Scholar]
- Kuznetsov S, Pellegrini M, Shuda K, Fernandez-Capetillo O, Liu Y, Martin BK, et al. RAD51C deficiency in mice results in early prophase I arrest in males and sister chromatid separation at metaphase II in females. J Cell Biol. 2007;176:581–92. doi: 10.1083/jcb.200608130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Langdon WY, Harris AW, Cory S, Adams JM. The c-myc oncogene perturbs B lymphocyte development in E-mu-myc transgenic mice. Cell. 1986;47:11–8. doi: 10.1016/0092-8674(86)90361-2. [DOI] [PubMed] [Google Scholar]
- Lebel M, Leder P. A deletion within the murine Werner syndrome helicase induces sensitivity to inhibitors of topoisomerase and loss of cellular proliferative capacity. Proc Natl Acad Sci U S A. 1998;95:13097–102. doi: 10.1073/pnas.95.22.13097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leeb M, Wutz A. Derivation of haploid embryonic stem cells from mouse embryos. Nature. 2011;479:131–4. doi: 10.1038/nature10448. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li G, Alt FW, Cheng HL, Brush JW, Goff PH, Murphy MM, et al. Lymphocyte-specific compensation for XLF/cernunnos end-joining functions in V(D)J recombination. Mol Cell. 2008;31:631–40. doi: 10.1016/j.molcel.2008.07.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li GC, Ouyang H, Li X, Nagasawa H, Little JB, Chen DJ, et al. Ku70: a candidate tumor suppressor gene for murine T cell lymphoma. Mol Cell. 1998;2:1–8. doi: 10.1016/s1097-2765(00)80108-2. [DOI] [PubMed] [Google Scholar]
- Lindahl T, Barnes DE. Repair of endogenous DNA damage. Cold Spring Harb Symp Quant Biol. 2000;65:127–33. doi: 10.1101/sqb.2000.65.127. [DOI] [PubMed] [Google Scholar]
- Lipkin SM, Moens PB, Wang V, Lenzi M, Shanmugarajah D, Gilgeous A, et al. Meiotic arrest and aneuploidy in MLH3-deficient mice. Nat Genet. 2002;31:385–90. doi: 10.1038/ng931. [DOI] [PubMed] [Google Scholar]
- Liu B, Wang J, Chan KM, Tjia WM, Deng W, Guan X, et al. Genomic instability in laminopathy-based premature aging. Nat Med. 2005;11:780–5. doi: 10.1038/nm1266. [DOI] [PubMed] [Google Scholar]
- Liu X, Holstege H, van der Gulden H, Treur-Mulder M, Zevenhoven J, Velds A, et al. Somatic loss of BRCA1 and p53 in mice induces mammary tumors with features of human BRCA1-mutated basal-like breast cancer. Proc Natl Acad Sci U S A. 2007;104:12111–6. doi: 10.1073/pnas.0702969104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lopez-Contreras AJ, Gutierrez-Martinez P, Specks J, Rodrigo-Perez S, Fernandez-Capetillo O. An extra allele of Chk1 limits oncogene-induced replicative stress and promotes transformation. J Exp Med. 2012;209:455–61. doi: 10.1084/jem.20112147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lou Z, Minter-Dykhouse K, Franco S, Gostissa M, Rivera MA, Celeste A, et al. MDC1 maintains genomic stability by participating in the amplification of ATM-dependent DNA damage signals. Mol Cell. 2006;21:187–200. doi: 10.1016/j.molcel.2005.11.025. [DOI] [PubMed] [Google Scholar]
- Ludwig T, Chapman DL, Papaioannou VE, Efstratiadis A. Targeted mutations of breast cancer susceptibility gene homologs in mice: lethal phenotypes of Brca1, Brca2, Brca1/Brca2, Brca1/p53, and Brca2/p53 nullizygous embryos. Genes Dev. 1997;11:1226–41. doi: 10.1101/gad.11.10.1226. [DOI] [PubMed] [Google Scholar]
- Luo G, Santoro IM, McDaniel LD, Nishijima I, Mills M, Youssoufian H, et al. Cancer predisposition caused by elevated mitotic recombination in Bloom mice. Nat Genet. 2000;26:424–9. doi: 10.1038/82548. [DOI] [PubMed] [Google Scholar]
- MacPherson D, Kim J, Kim T, Rhee BK, Van Oostrom CT, DiTullio RA, et al. Defective apoptosis and B-cell lymphomas in mice with p53 point mutation at Ser 23. EMBO J. 2004;23:3689–99. doi: 10.1038/sj.emboj.7600363. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mak TW. The Gene Knockout Facts Book. Academic; San Diego, CA: 1998. [Google Scholar]
- Manis JP, Morales JC, Xia Z, Kutok JL, Alt FW, Carpenter PB. 53BP1 links DNA damage-response pathways to immunoglobulin heavy chain class-switch recombination. Nat Immunol. 2004;5:481–7. doi: 10.1038/ni1067. [DOI] [PubMed] [Google Scholar]
- Martin GR. Isolation of a pluripotent cell line from early mouse embryos cultured in medium conditioned by teratocarcinoma stem cells. Proc Natl Acad Sci U S A. 1981;78:7634–8. doi: 10.1073/pnas.78.12.7634. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McCarthy EE, Celebi JT, Baer R, Ludwig T. Loss of Bard1, the heterodimeric partner of the Brca1 tumor suppressor, results in early embryonic lethality and chromosomal instability. Mol Cell Biol. 2003;23:5056–63. doi: 10.1128/MCB.23.14.5056-5063.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McJunkin K, Mazurek A, Premsrirut PK, Zuber J, Dow LE, Simon J, et al. Reversible suppression of an essential gene in adult mice using transgenic RNA interference. Proc Natl Acad Sci U S A. 2011;108:7113–8. doi: 10.1073/pnas.1104097108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McPherson JP, Lemmers B, Chahwan R, Pamidi A, Migon E, Matysiak-Zablocki E, et al. Involvement of mammalian Mus81 in genome integrity and tumor suppression. Science. 2004;304:1822–6. doi: 10.1126/science.1094557. [DOI] [PubMed] [Google Scholar]
- McWhir J, Selfridge J, Harrison DJ, Squires S, Melton DW. Mice with DNA repair gene (ERCC-1) deficiency have elevated levels of p53, liver nuclear abnormalities and die before weaning. Nat Genet. 1993;5:217–24. doi: 10.1038/ng1193-217. [DOI] [PubMed] [Google Scholar]
- Mills KD, Ferguson DO, Essers J, Eckersdorff M, Kanaar R, Alt FW. Rad54 and DNA Ligase IV cooperate to maintain mammalian chromatid stability. Genes Dev. 2004;18:1283–92. doi: 10.1101/gad.1204304. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Minter-Dykhouse K, Ward I, Huen MS, Chen J, Lou Z. Distinct versus overlapping functions of MDC1 and 53BP1 in DNA damage response and tumorigenesis. J Cell Biol. 2008;181:727–35. doi: 10.1083/jcb.200801083. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moldovan GL, D’Andrea AD. How the fanconi anemia pathway guards the genome. Annu Rev Genet. 2009;43:223–49. doi: 10.1146/annurev-genet-102108-134222. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Morales JC, Xia Z, Lu T, Aldrich MB, Wang B, Rosales C, et al. Role for the BRCA1 C-terminal repeats (BRCT) protein 53BP1 in maintaining genomic stability. J Biol Chem. 2003;278:14971–7. doi: 10.1074/jbc.M212484200. [DOI] [PubMed] [Google Scholar]
- Murga M, Bunting S, Montana MF, Soria R, Mulero F, Canamero M, et al. A mouse model of ATR-Seckel shows embryonic replicative stress and accelerated aging. Nat Genet. 2009;41:891–8. doi: 10.1038/ng.420. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Murga M, Campaner S, Lopez-Contreras AJ, Toledo LI, Soria R, Montana MF, et al. Exploiting oncogene-induced replicative stress for the selective killing of Myc-driven tumors. Nat Struct Mol Biol. 2011;18:1331–5. doi: 10.1038/nsmb.2189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nakamura T, Pichel JG, Williams-Simons L, Westphal H. An apoptotic defect in lens differentiation caused by human p53 is rescued by a mutant allele. Proc Natl Acad Sci U S A. 1995;92:6142–6. doi: 10.1073/pnas.92.13.6142. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Natale V. A comprehensive description of the severity groups in Cockayne syndrome. Am J Med Genet A. 2011;155A:1081–95. doi: 10.1002/ajmg.a.33933. [DOI] [PubMed] [Google Scholar]
- Niedernhofer LJ, Garinis GA, Raams A, Lalai AS, Robinson AR, Appeldoorn E, et al. A new progeroid syndrome reveals that genotoxic stress suppresses the somatotroph axis. Nature. 2006;444:1038–43. doi: 10.1038/nature05456. [DOI] [PubMed] [Google Scholar]
- O’Driscoll M, Ruiz-Perez VL, Woods CG, Jeggo PA, Goodship JA. A splicing mutation affecting expression of ataxia-telangiectasia and Rad3-related protein (ATR) results in Seckel syndrome. Nat Genet. 2003;33:497–501. doi: 10.1038/ng1129. [DOI] [PubMed] [Google Scholar]
- Oliner JD, Pietenpol JA, Thiagalingam S, Gyuris J, Kinzler KW, Vogelstein B. Oncoprotein MDM2 conceals the activation domain of tumour suppressor p53. Nature. 1993;362:857–60. doi: 10.1038/362857a0. [DOI] [PubMed] [Google Scholar]
- Outwin E, Carpenter G, Bi W, Withers MA, Lupski JR, O’Driscoll M. Increased RPA1 Gene Dosage Affects Genomic Stability Potentially Contributing to 17p13.3 Duplication Syndrome. PLoS Genet. 2011;7:e1002247. doi: 10.1371/journal.pgen.1002247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parmar K, D’Andrea A, Niedernhofer LJ. Mouse models of Fanconi anemia. Mutat Res. 2009;668:133–40. doi: 10.1016/j.mrfmmm.2009.03.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pellegrini M, Celeste A, Difilippantonio S, Guo R, Wang W, Feigenbaum L, et al. Autophosphorylation at serine 1987 is dispensable for murine Atm activation in vivo. Nature. 2006;443:222–5. doi: 10.1038/nature05112. [DOI] [PubMed] [Google Scholar]
- Pendas AM, Zhou Z, Cadinanos J, Freije JM, Wang J, Hultenby K, et al. Defective prelamin A processing and muscular and adipocyte alterations in Zmpste24 metalloproteinase-deficient mice. Nat Genet. 2002;31:94–9. doi: 10.1038/ng871. [DOI] [PubMed] [Google Scholar]
- Pieper AA, Brat DJ, Krug DK, Watkins CC, Gupta A, Blackshaw S, et al. Poly(ADP-ribose) polymerase-deficient mice are protected from streptozotocin-induced diabetes. Proc Natl Acad Sci U S A. 1999;96:3059–64. doi: 10.1073/pnas.96.6.3059. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pulliam-Leath AC, Ciccone SL, Nalepa G, Li X, Si Y, Miravalle L, et al. Genetic disruption of both Fancc and Fancg in mice recapitulates the hematopoietic manifestations of Fanconi anemia. Blood. 2010;116:2915–20. doi: 10.1182/blood-2009-08-240747. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ran FA, Hsu PD, Wright J, Agarwala V, Scott DA, Zhang F. Genome engineering using the CRISPR-Cas9 system. Nat Protoc. 2013;8:2281–308. doi: 10.1038/nprot.2013.143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rooney S, Sekiguchi J, Zhu C, Cheng HL, Manis J, Whitlow S, et al. Leaky Scid phenotype associated with defective V(D)J coding end processing in Artemis-deficient mice. Mol Cell. 2002;10:1379–90. doi: 10.1016/s1097-2765(02)00755-4. [DOI] [PubMed] [Google Scholar]
- Rossi DJ, Jamieson CH, Weissman IL. Stems cells and the pathways to aging and cancer. Cell. 2008;132:681–96. doi: 10.1016/j.cell.2008.01.036. [DOI] [PubMed] [Google Scholar]
- Rucci F, Notarangelo LD, Fazeli A, Patrizi L, Hickernell T, Paganini T, et al. Homozygous DNA ligase IV R278H mutation in mice leads to leaky SCID and represents a model for human LIG4 syndrome. Proc Natl Acad Sci U S A. 2010;107:3024–9. doi: 10.1073/pnas.0914865107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ruzankina Y, Pinzon-Guzman C, Asare A, Ong T, Pontano L, Cotsarelis G, et al. Deletion of the developmentally essential gene ATR in adult mice leads to age-related phenotypes and stem cell loss. Cell Stem Cell. 2007;1:113–26. doi: 10.1016/j.stem.2007.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Santos MA, Huen MS, Jankovic M, Chen HT, Lopez-Contreras AJ, Klein IA, et al. Class switching and meiotic defects in mice lacking the E3 ubiquitin ligase RNF8. J Exp Med. 2010;207:973–81. doi: 10.1084/jem.20092308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sauer B, Henderson N. Site-specific DNA recombination in mammalian cells by the Cre recombinase of bacteriophage P1. Proc Natl Acad Sci U S A. 1988;85:5166–70. doi: 10.1073/pnas.85.14.5166. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Savitsky K, Bar-Shira A, Gilad S, Rotman G, Ziv Y, Vanagaite L, et al. A single ataxia telangiectasia gene with a product similar to PI-3 kinase. Science. 1995;268:1749–53. doi: 10.1126/science.7792600. [DOI] [PubMed] [Google Scholar]
- Schnutgen F, Doerflinger N, Calleja C, Wendling O, Chambon P, Ghyselinck NB. A directional strategy for monitoring Cre-mediated recombination at the cellular level in the mouse. Nat Biotechnol. 2003;21:562–5. doi: 10.1038/nbt811. [DOI] [PubMed] [Google Scholar]
- Shakya R, Szabolcs M, McCarthy E, Ospina E, Basso K, Nandula S, et al. The basal-like mammary carcinomas induced by Brca1 or Bard1 inactivation implicate the BRCA1/BARD1 heterodimer in tumor suppression. Proc Natl Acad Sci U S A. 2008;105:7040–5. doi: 10.1073/pnas.0711032105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shiomi N, Kito S, Oyama M, Matsunaga T, Harada YN, Ikawa M, et al. Identification of the XPG region that causes the onset of Cockayne syndrome by using Xpg mutant mice generated by the cDNA-mediated knock-in method. Mol Cell Biol. 2004;24:3712–9. doi: 10.1128/MCB.24.9.3712-3719.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shizuya H, Birren B, Kim UJ, Mancino V, Slepak T, Tachiiri Y, et al. Cloning and stable maintenance of 300-kilobase-pair fragments of human DNA in Escherichia coli using an F-factor-based vector. Proc Natl Acad Sci U S A. 1992;89:8794–7. doi: 10.1073/pnas.89.18.8794. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Skarnes WC, Rosen B, West AP, Koutsourakis M, Bushell W, Iyer V, et al. A conditional knockout resource for the genome-wide study of mouse gene function. Nature. 2011;474:337–42. doi: 10.1038/nature10163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sluss HK, Armata H, Gallant J, Jones SN. Phosphorylation of serine 18 regulates distinct p53 functions in mice. Mol Cell Biol. 2004;24:976–84. doi: 10.1128/MCB.24.3.976-984.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smithies O, Gregg RG, Boggs SS, Koralewski MA, Kucherlapati RS. Insertion of DNA sequences into the human chromosomal beta-globin locus by homologous recombination. Nature. 1985;317:230–4. doi: 10.1038/317230a0. [DOI] [PubMed] [Google Scholar]
- Sternberg N, Hamilton D, Austin S, Yarmolinsky M, Hoess R. Site-specific recombination and its role in the life cycle of bacteriophage P1. Cold Spring Harb Symp Quant Biol. 1981;45(Pt 1):297–309. doi: 10.1101/sqb.1981.045.01.042. [DOI] [PubMed] [Google Scholar]
- Stracker TH, Morales M, Couto SS, Hussein H, Petrini JH. The carboxy terminus of NBS1 is required for induction of apoptosis by the MRE11 complex. Nature. 2007;447:218–21. doi: 10.1038/nature05740. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Suzuki A, de la Pompa JL, Hakem R, Elia A, Yoshida R, Mo R, et al. Brca2 is required for embryonic cellular proliferation in the mouse. Genes Dev. 1997;11:1242–52. doi: 10.1101/gad.11.10.1242. [DOI] [PubMed] [Google Scholar]
- Takai H, Tominaga K, Motoyama N, Minamishima YA, Nagahama H, Tsukiyama T, et al. Aberrant cell cycle checkpoint function and early embryonic death in Chk1(-/-) mice. Genes Dev. 2000;14:1439–47. [PMC free article] [PubMed] [Google Scholar]
- Tebbs RS, Flannery ML, Meneses JJ, Hartmann A, Tucker JD, Thompson LH, et al. Requirement for the Xrcc1 DNA base excision repair gene during early mouse development. Dev Biol. 1999;208:513–29. doi: 10.1006/dbio.1999.9232. [DOI] [PubMed] [Google Scholar]
- Theunissen JW, Kaplan MI, Hunt PA, Williams BR, Ferguson DO, Alt FW, et al. Checkpoint failure and chromosomal instability without lymphomagenesis in Mre11(ATLD1/ATLD1) mice. Mol Cell. 2003;12:1511–23. doi: 10.1016/s1097-2765(03)00455-6. [DOI] [PubMed] [Google Scholar]
- Thomas KR, Capecchi MR. Site-directed mutagenesis by gene targeting in mouse embryo-derived stem cells. Cell. 1987;51:503–12. doi: 10.1016/0092-8674(87)90646-5. [DOI] [PubMed] [Google Scholar]
- Tian M, Shinkura R, Shinkura N, Alt FW. Growth retardation, early death, and DNA repair defects in mice deficient for the nucleotide excision repair enzyme XPF. Mol Cell Biol. 2004;24:1200–5. doi: 10.1128/MCB.24.3.1200-1205.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Toledo LI, Murga M, Fernandez-Capetillo O. Targeting ATR and Chk1 kinases for cancer treatment: a new model for new (and old) drugs. Mol Oncol. 2011;5:368–73. doi: 10.1016/j.molonc.2011.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- van der Horst GT, Meira L, Gorgels TG, de Wit J, Velasco-Miguel S, Richardson JA, et al. UVB radiation-induced cancer predisposition in Cockayne syndrome group A (Csa) mutant mice. DNA Repair (Amst) 2002;1:143–57. doi: 10.1016/s1568-7864(01)00010-6. [DOI] [PubMed] [Google Scholar]
- van der Horst GT, van Steeg H, Berg RJ, van Gool AJ, de Wit J, Weeda G, et al. Defective transcription-coupled repair in Cockayne syndrome B mice is associated with skin cancer predisposition. Cell. 1997;89:425–35. doi: 10.1016/s0092-8674(00)80223-8. [DOI] [PubMed] [Google Scholar]
- von Melchner H, DeGregori JV, Rayburn H, Reddy S, Friedel C, Ruley HE. Selective disruption of genes expressed in totipotent embryonal stem cells. Genes Dev. 1992;6:919–27. doi: 10.1101/gad.6.6.919. [DOI] [PubMed] [Google Scholar]
- Wang H, Yang H, Shivalila CS, Dawlaty MM, Cheng AW, Zhang F, et al. One-step generation of mice carrying mutations in multiple genes by CRISPR/Cas-mediated genome engineering. Cell. 2013;153:910–8. doi: 10.1016/j.cell.2013.04.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y, Putnam CD, Kane MF, Zhang W, Edelmann L, Russell R, et al. Mutation in Rpa1 results in defective DNA double-strand break repair, chromosomal instability and cancer in mice. Nat Genet. 2005;37:750–5. doi: 10.1038/ng1587. [DOI] [PubMed] [Google Scholar]
- Wang YA, Elson A, Leder P. Loss of p21 increases sensitivity to ionizing radiation and delays the onset of lymphoma in atm-deficient mice. Proc Natl Acad Sci U S A. 1997;94:14590–5. doi: 10.1073/pnas.94.26.14590. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Whitney MA, Royle G, Low MJ, Kelly MA, Axthelm MK, Reifsteck C, et al. Germ cell defects and hematopoietic hypersensitivity to gamma-interferon in mice with a targeted disruption of the Fanconi anemia C gene. Blood. 1996;88:49–58. [PubMed] [Google Scholar]
- Wong JC, Alon N, McKerlie C, Huang JR, Meyn MS, Buchwald M. Targeted disruption of exons 1 to 6 of the Fanconi Anemia group A gene leads to growth retardation, strain-specific microphthalmia, meiotic defects and primordial germ cell hypoplasia. Hum Mol Genet. 2003;12:2063–76. doi: 10.1093/hmg/ddg219. [DOI] [PubMed] [Google Scholar]
- Wu X, Sandhu S, Ding H. Establishment of conditional knockout alleles for the gene encoding the regulator of telomere length (RTEL) Genesis. 2007;45:788–92. doi: 10.1002/dvg.20359. [DOI] [PubMed] [Google Scholar]
- Xu X, Wagner KU, Larson D, Weaver Z, Li C, Ried T, et al. Conditional mutation of Brca1 in mammary epithelial cells results in blunted ductal morphogenesis and tumour formation. Nat Genet. 1999;22:37–43. doi: 10.1038/8743. [DOI] [PubMed] [Google Scholar]
- Xu Y, Ashley T, Brainerd EE, Bronson RT, Meyn MS, Baltimore D. Targeted disruption of ATM leads to growth retardation, chromosomal fragmentation during meiosis, immune defects, and thymic lymphoma. Genes Dev. 1996;10:2411–22. doi: 10.1101/gad.10.19.2411. [DOI] [PubMed] [Google Scholar]
- Yamane K, Wu X, Chen J. A DNA damage-regulated BRCT-containing protein, TopBP1, is required for cell survival. Mol Cell Biol. 2002;22:555–66. doi: 10.1128/MCB.22.2.555-566.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang H, Wang H, Shivalila CS, Cheng AW, Shi L, Jaenisch R. One-step generation of mice carrying reporter and conditional alleles by CRISPR/Cas-mediated genome engineering. Cell. 2013;154:1370–9. doi: 10.1016/j.cell.2013.08.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zaugg K, Su YW, Reilly PT, Moolani Y, Cheung CC, Hakem R, et al. Cross-talk between Chk1 and Chk2 in double-mutant thymocytes. Proc Natl Acad Sci U S A. 2007;104:3805–10. doi: 10.1073/pnas.0611584104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zha S, Guo C, Boboila C, Oksenych V, Cheng HL, Zhang Y, et al. ATM damage response and XLF repair factor are functionally redundant in joining DNA breaks. Nature. 2011;469:250–4. doi: 10.1038/nature09604. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhu J, Petersen S, Tessarollo L, Nussenzweig A. Targeted disruption of the Nijmegen breakage syndrome gene NBS1 leads to early embryonic lethality in mice. Curr Biol. 2001;11:105–9. doi: 10.1016/s0960-9822(01)00019-7. [DOI] [PubMed] [Google Scholar]
- Zickler D, Kleckner N. The leptotene-zygotene transition of meiosis. Annu Rev Genet. 1998;32:619–97. doi: 10.1146/annurev.genet.32.1.619. [DOI] [PubMed] [Google Scholar]


