Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 57
Filter
Add more filters










Publication year range
1.
J Biol Chem ; 276(20): 16786-96, 2001 May 18.
Article in English | MEDLINE | ID: mdl-11278801

ABSTRACT

Both prokaryotic and eukaryotic cells have the capacity to repair DNA damage preferentially in the transcribed strand of actively expressed genes. However, we have found that several types of DNA damage, including cyclobutane pyrimidine dimers (CPDs) are repaired with equal efficiency in both the transcribed and nontranscribed strands of the adenine phosphoribosyltransferase (APRT) gene in Chinese hamster ovary cells. We further found that, in two mutant cell lines in which the entire APRT promoter region has been deleted, CPDs are still efficiently repaired in both strands of the promoterless APRT gene, even though neither strand appears to be transcribed. These results suggest that efficient repair of both strands at this locus does not require transcription of the APRT gene. We have also mapped CPD repair in exon 3 of the APRT gene in each cell line at single nucleotide resolution. Again, we found similar rates of CPD repair in both strands of the APRT gene domain in both APRT promoter-deletion mutants and their parental cell line. Our findings suggest that current models of transcription-coupled repair and global genomic repair may underestimate the importance of factors other than transcription in governing the efficiency of nucleotide excision repair.


Subject(s)
7,8-Dihydro-7,8-dihydroxybenzo(a)pyrene 9,10-oxide/metabolism , Adenine Phosphoribosyltransferase/genetics , Adenine Phosphoribosyltransferase/metabolism , DNA Adducts/metabolism , DNA Repair , Escherichia coli Proteins , Promoter Regions, Genetic , Pyrimidine Dimers/metabolism , Transcription, Genetic , Animals , CHO Cells , Cricetinae , DNA/genetics , DNA/radiation effects , DNA Damage , Endodeoxyribonucleases/metabolism , Exons , Kinetics , RNA, Messenger/genetics , Recombinant Proteins/metabolism , Restriction Mapping , Sequence Deletion , Transfection , Ultraviolet Rays
2.
EMBO J ; 19(20): 5552-61, 2000 Oct 16.
Article in English | MEDLINE | ID: mdl-11032822

ABSTRACT

The XpF/Ercc1 structure-specific endonuclease performs the 5' incision in nucleotide excision repair and is the apparent mammalian counterpart of the Rad1/Rad10 endonuclease from Saccharomyces cerevisiae. In yeast, Rad1/Rad10 endonuclease also functions in mitotic recombination. To determine whether XpF/Ercc1 endonuclease has a similar role in mitotic recombination, we targeted the APRT locus in Chinese hamster ovary ERCC1(+) and ERCC1(-) cell lines with insertion vectors having long or short terminal non-homologies flanking each side of a double-strand break. No substantial differences were evident in overall recombination frequencies, in contrast to results from targeting experiments in yeast. However, profound differences were observed in types of APRT(+) recombinants recovered from ERCC1(-) cells using targeting vectors with long terminal non-homologies-almost complete ablation of gap repair and single-reciprocal exchange events, and generation of a new class of aberrant insertion/deletion recombinants absent in ERCC1(+) cells. These results represent the first demonstration of a requirement for ERCC1 in targeted homologous recombination in mammalian cells, specifically in removal of long non-homologous tails from invading homologous strands.


Subject(s)
DNA Repair/genetics , DNA-Binding Proteins , Endonucleases , Proteins/metabolism , Recombination, Genetic/genetics , Sequence Homology, Nucleic Acid , Adenine Phosphoribosyltransferase/genetics , Adenine Phosphoribosyltransferase/metabolism , Animals , Blotting, Southern , CHO Cells , Cell Line , Cricetinae , DNA/genetics , DNA/metabolism , Electroporation , Gene Deletion , Gene Targeting , Genetic Vectors/genetics , Mutagenesis, Insertional/genetics , Proteins/genetics
3.
Nucleic Acids Res ; 28(19): 3771-8, 2000 Oct 01.
Article in English | MEDLINE | ID: mdl-11000269

ABSTRACT

Spontaneous recombination between direct repeats at the adenine phosphoribosyltransferase (APRT) locus in ERCC1-deficient cells generates a high frequency of rearrangements that are dependent on the process of homologous recombination, suggesting that rearrangements are formed by misprocessing of recombination intermediates. Given the specificity of the structure-specific Ercc1/Xpf endonuclease, two potential recombination intermediates are substrates for misprocessing in ERCC1(-) cells: heteroduplex loops and heteroduplex intermediates with non-homologous 3' tails. To investigate the roles of each, we constructed repeats that would yield no heteroduplex loops during spontaneous recombination or that would yield two non-homologous 3' tails after treatment with the rare-cutting endonuclease I-SCE:I. Our results indicate that misprocessing of heteroduplex loops is not the major source of recombination-dependent rearrangements in ERCC1-deficient cells. Our results also suggest that the Ercc1/Xpf endonuclease is required for efficient removal of non-homologous 3' tails, like its Rad1/Rad10 counterpart in yeast. Thus, it is likely that misprocessing of non-homologous 3' tails is the primary source of recombination-dependent rearrangements in mammalian cells. We also find an unexpected effect of ERCC1 deficiency on I-SCE:I-stimulated rearrangements, which are not dependent on homologous recombination, suggesting that the ERCC1 gene product may play a role in generating the rearrangements that arise after I-SCE:I-induced double-strand breaks.


Subject(s)
DNA Repair/genetics , DNA-Binding Proteins , DNA/chemistry , DNA/metabolism , Endonucleases , Proteins/metabolism , Recombination, Genetic/genetics , Adenine Phosphoribosyltransferase/genetics , Animals , Blotting, Southern , Cell Line , Crossing Over, Genetic/genetics , DNA/genetics , DNA Damage/genetics , Gene Deletion , Nucleic Acid Conformation , Nucleic Acid Heteroduplexes/chemistry , Nucleic Acid Heteroduplexes/genetics , Nucleic Acid Heteroduplexes/metabolism , Proteins/genetics , Repetitive Sequences, Nucleic Acid/genetics , Substrate Specificity , Transfection
5.
Mutagenesis ; 13(4): 357-65, 1998 Jul.
Article in English | MEDLINE | ID: mdl-9717172

ABSTRACT

We isolated and characterized the ERCC1 coding sequence from three Chinese hamster ovary (CHO) parental (CHO-AA8, CHO-AT3-2 and CHO-9) and 10 ERCC1 mutant cell lines. Two general classes of mutations were observed: two mutant cell lines exhibited nucleotide additions or deletions to produce frameshift mutations and seven mutant cell lines exhibited point mutations that resulted in transitions or transversions, including nonsense mutations and mutations that generated intron/exon splicing errors. One mutant (UV201) which had been provisionally assigned to ERCC1 complementation group 1 (CG1) had no detectable mutation in its coding sequence. Of the nine ERCC1 mutant alleles characterized two mutations were identified in the XpA binding region of the Ercc1 protein; no mutations were found in the N-terminal portion of the Ercc1 protein. Results of Northern hybridization analysis showed that the relative levels of ERCC1 mRNA differed significantly both among the parental cell lines and among the mutant cell lines derived from each parental cell line. Western analysis with a CHO Ercc1-specific antibody detected Ercc1 protein in each of the parental cell lines and also in UV201. The marked reduction in Ercc1 protein levels observed in all the other mutants examined supports the hypothesis that ERCC1 mutations may destabilize this polypeptide.


Subject(s)
CHO Cells/physiology , DNA-Binding Proteins , Endonucleases , Mutation , Proteins/genetics , Amino Acid Sequence , Animals , Base Sequence , Blotting, Western , Cricetinae , Gene Expression Regulation , Genetic Variation , Molecular Sequence Data , Proteins/metabolism , Reverse Transcriptase Polymerase Chain Reaction , Sequence Homology, Amino Acid
6.
Somat Cell Mol Genet ; 24(2): 91-105, 1998 Mar.
Article in English | MEDLINE | ID: mdl-9919309

ABSTRACT

In this study, we have examined the effects of targeting vector configuration and site of vector linearization on the frequency of targeted recombination at the endogenous CHO APRT locus, and have analyzed the types and class distributions of APRT+ recombinants obtained in APRT targeting experiments employing uncut circular, insertion-type (ends-in), and replacement-type (ends-out) configurations of the same pAG7 targeting vector, including configurations produced by introduction of a double-strand break (DSB) at sites either within, or at the 5' or 3' boundaries of APRT targeting homology. Our results suggest that: 1) plasmid-chromosome targeted recombination in mammalian cells may not be stimulated to the same degree by a DSB in the targeting vector as by a DSB in the chromosomal target; 2) recombinant class distributions are highly dependent upon targeting vector configuration; and 3) one-sided invasion mechanisms may play a significant role in homologous recombination in mammalian cells.


Subject(s)
Adenine Phosphoribosyltransferase/genetics , Genetic Vectors/genetics , Recombination, Genetic , Animals , Blotting, Southern , CHO Cells/physiology , Cricetinae , DNA Damage/genetics , Genetic Techniques
7.
Proc Natl Acad Sci U S A ; 94(24): 13122-7, 1997 Nov 25.
Article in English | MEDLINE | ID: mdl-9371810

ABSTRACT

Nucleotide excision repair proteins have been implicated in genetic recombination by experiments in Saccharomyces cerevisiae and Drosophila melanogaster, but their role, if any, in mammalian cells is undefined. To investigate the role of the nucleotide excision repair gene ERCC1, the hamster homologue to the S. cerevisiae RADIO gene, we disabled the gene by targeted knockout. Partial tandem duplications of the adenine phosphoribosyltransferase (APRT) gene then were constructed at the endogenous APRT locus in ERCC1- and ERCC1+ cells. To detect the full spectrum of gene-altering events, we used a loss-of-function assay in which the parental APRT+ tandem duplication could give rise to APRT- cells by homologous recombination, gene rearrangement, or point mutation. Measurement of rates and analysis of individual APRT- products indicated that gene rearrangements (principally deletions) were increased at least 50-fold, whereas homologous recombination was affected little. The formation of deletions is not caused by a general effect of the ERCC1 deficiency on gene stability, because ERCC1- cell lines with a single wild-type copy of the APRT gene yielded no increase in deletions. Thus, deletion formation is dependent on the tandem duplication, and presumably the process of homologous recombination. Recombination-dependent deletion formation in ERCC1- cells is supported by a significant decrease in a particular class of crossover products that are thought to arise by repair of a heteroduplex intermediate in recombination. We suggest that the ERCC1 gene product in mammalian cells is involved in the processing of heteroduplex intermediates in recombination and that the misprocessed intermediates in ERCC1- cells are repaired by illegitimate recombination.


Subject(s)
DNA Repair , DNA-Binding Proteins , Endonucleases , Proteins/genetics , Recombination, Genetic , Sequence Deletion , Adenine Phosphoribosyltransferase/genetics , Animals , Cell Line , Cricetinae , Thymidine Kinase/genetics
8.
Cancer Res ; 57(18): 4048-56, 1997 Sep 15.
Article in English | MEDLINE | ID: mdl-9307292

ABSTRACT

To investigate the role of DNA strand breakage as the molecular lesion responsible for initiating genomic instability, five different strand-breaking agents, bleomycin, neocarzinostatin, hydrogen peroxide, restriction endonucleases, and ionizing radiation, were examined for their capacity to induce delayed chromosomal instability. These studies used GM10115 human-hamster hybrid cells, which contain one copy of human chromosome 4 in a background of 20-24 hamster chromosomes. Chromosomal instability was investigated using fluorescence in situ hybridization to visualize chromosomal rearrangements involving the human chromosome. Rearrangements are detected multiple generations after treatment, in clonal populations derived from single progenitor cells surviving treatment of the specified DNA-damaging agents. Clastogenic and cytotoxic activities of all agents were tested by examining chromosome aberration yields in first-division metaphases and by clonogenic survival assays. Analysis of over 250 individual clones representing over 50,000 metaphases demonstrates that when compared at comparable levels of cell kill, ionizing radiation, bleomycin, and neocarzinostatin are equally effective at eliciting delayed genomic instability. These observations document, for the first time, the persistent destabilization of chromosomes following chemical treatment. In contrast, the analysis of nearly 300 clones and 60,000 metaphases, involving treatment with four different restriction endonucleases and/or hydrogen peroxide, did not show any delayed chromosomal instability. These data indicate that DNA strand breakage per se does not necessarily lead to chromosomal instability but that the complexity or quality of DNA strand breaks are important in initiating this phenotype.


Subject(s)
DNA Damage/drug effects , Adenine Phosphoribosyltransferase/genetics , Animals , Bleomycin/toxicity , Cell Division/drug effects , Cell Survival/drug effects , Chromosomes, Human, Pair 4 , Cricetinae , DNA Restriction Enzymes/pharmacology , Humans , Hybrid Cells/radiation effects , Hydrogen Peroxide/toxicity , In Situ Hybridization, Fluorescence , Radiation, Ionizing , Time Factors , Zinostatin/pharmacology
9.
Mutagenesis ; 12(4): 277-83, 1997 Jul.
Article in English | MEDLINE | ID: mdl-9237774

ABSTRACT

Positive selection-negative selection gene targeting was used to disrupt the nucleotide excision repair gene ERCC1 in a Chinese hamster ovary cell line, CHO-K1. Southern and Northern analysis showed that a cell clone isolated by this targeting approach, CHO-7-27, had an ERCC1 gene structure consistent with targeted disruption of ERCC1 exon V, and did not express ERCC1 mRNA. CHO-7-27 was further characterized with respect to UV and mitomycin C sensitivities, and was shown to exhibit severe mutagen sensitivity phenotypes consistent with those of other CHO cell ERCC1 mutants. Mutation frequency experiments showed that CHO-7-27 was UV-hypermutable at the hypoxanthine-guanine phosphoribosyltransferase locus. Experiments assessing host cell reactivation of viral DNA synthesis for UV-irradiated adenovirus showed that CHO7-27 exhibited a severely deficient HCR phenotype similar to that of UV20 cells. Our results demonstrate that CHOK1 cells are hemizygous for the ERCC1 gene, and show that the comparatively mild mutagen sensitivities and lack of severely deficient HCR phenotypes of conventionally derived CHO-K1 ERCC1 mutants, in contrast to the severe phenotypes of CHO-AA8-derived mutants, are not due to any intrinsic genetic differences between CHO-K1 and CHO-AA8 parental cell lines.


Subject(s)
Adenoviridae/genetics , CHO Cells/drug effects , CHO Cells/radiation effects , DNA-Binding Proteins , Endonucleases , Mutagenesis , Proteins/genetics , Adenoviridae/pathogenicity , Adenoviridae/radiation effects , Animals , CHO Cells/virology , Cell Survival/drug effects , Cell Survival/radiation effects , Cricetinae , DNA, Viral/biosynthesis , DNA, Viral/radiation effects , Homozygote , Hypoxanthine Phosphoribosyltransferase/drug effects , Hypoxanthine Phosphoribosyltransferase/genetics , Hypoxanthine Phosphoribosyltransferase/radiation effects , Mitomycin/pharmacology , Mutation , Phenotype , Proteins/drug effects , Proteins/radiation effects , Ultraviolet Rays
10.
J Immunol ; 157(10): 4464-73, 1996 Nov 15.
Article in English | MEDLINE | ID: mdl-8906823

ABSTRACT

The identity and abundance of self-peptide/MHC class I complexes that serve as ligands for alloreactive T cells remain largely unknown. Using the Kb-restricted, alloreactive T cells as a probe, the Ag precursor gene, adenosine phosphoribosyl transferase (APRT), was isolated by expression cloning. Its naturally processed product was identified as the SLVELTSL (SEL8) octapeptide. The SEL8 peptide shared five residues with the previously identified SVVEFSSL (JAL8) peptide that stimulated the same T cell, but lacked the critical phenylalanine/tyrosine residue at the primary p5 anchor position. Despite the absence of this key conserved anchor residue, SEL8 was bound tightly by Kb MHC and yet was expressed at less than 10 copies/cell. Mutations in the donor APRT gene in the APC caused a concomitant loss in the ability of APCs to stimulate T cells. The results confirm that the display of peptide/MHC complexes in cells exceeds the predictions based upon consensus motifs, and that CD8+ alloreactive and conventional Ag-specific T cells are indistinguishable in their ability to recognize unique and rare peptide/MHC class I complexes.


Subject(s)
CD8-Positive T-Lymphocytes/immunology , Epitopes/immunology , Histocompatibility Antigens Class I/immunology , Isoantigens/immunology , Peptides/immunology , Animals , Cell Line , Mice , Peptides/metabolism , Poly(ADP-ribose) Polymerases/analysis , Protein Binding/immunology
11.
Mutat Res ; 352(1-2): 87-96, 1996 Jun 10.
Article in English | MEDLINE | ID: mdl-8676921

ABSTRACT

We describe an apparent hotspot for spontaneous deletions and base substitution mutations at a TTC trinucleotide direct repeat/MboII restriction site in exon 5 of the Chinese hamster APRT gene, in a region with the potential to form a relatively stable, quasipalindromic, stem-loop structure. The recurrent 3 bp TTC deletions observed at this site, which account for approx. 20% of the characterized spontaneous APRT deletions in hemizygous CHO cell lines, represent the only spontaneous deletion events that have been recovered more than once at this locus. A total of 11 independently derived, spontaneous CHO cell APRT mutants with identical 3 bp TTC deletions at this exon 5 MboII site, plus another five mutants that have single base substitutions at this site have been identified among spontaneous mutant collections in several different laboratories. Intriguingly, each of the frequently deleted or mutated bases at this exon 5 deletion hotspot site would correspond to one of the unpaired bases within a single-stranded 'loop' region of a stable, quasipalindromic, stem-loop structure that can be formed by intrastrand pairing of inverted repeats in this portion of the APRT gene sequence. An identical TTC trinucleotide direct repeat sequence at the same site in exon 5 of the human APRT gene also appears to be a hotspot for spontaneous deletion.


Subject(s)
Adenine Phosphoribosyltransferase/genetics , Exons/genetics , Mutation/genetics , Amino Acid Sequence , Animals , Base Sequence , CHO Cells , Cricetinae , DNA/chemistry , DNA Mutational Analysis , Genes/genetics , Humans , Molecular Sequence Data , Mutagenesis/physiology , Nucleic Acid Conformation , Point Mutation/genetics , Sequence Deletion/genetics
12.
Nucleic Acids Res ; 24(4): 746-53, 1996 Feb 15.
Article in English | MEDLINE | ID: mdl-8604319

ABSTRACT

Several DNA sequence elements are thought to stimulate homologous recombination, illegitimate recombination, or both in mammalian cells. Some are implicated by their recurrence around rearrangement breakpoints, others by their effects on recombination of extrachromosomal plasmids. None of these sequences, however, has been tested on the chromosome in a defined context. In this paper we show how the adenine phosphoribosyltransferase locus in CHO cells can be used to study the recombinogenic potential of defined DNA sequences. As an example we have measured the effect on homologous recombination of a dinucleotide repeat, (GT)29, which has been shown to stimulate homologous recombination in extrachromosomal vectors 3-20 fold. On the chromosome at the adenine phosphoribosyltransferase locus, however, this sequence shows no capacity to stimulate recombination or to influence the distribution of recombination events.


Subject(s)
Adenine Phosphoribosyltransferase/genetics , DNA, Satellite/genetics , DNA/genetics , Microsatellite Repeats/genetics , Recombination, Genetic , Animals , Base Sequence , CHO Cells , Cricetinae , Molecular Sequence Data , Sequence Analysis
13.
Mol Cell Biol ; 14(10): 6663-73, 1994 Oct.
Article in English | MEDLINE | ID: mdl-7935385

ABSTRACT

Using simple linear fragments of the Chinese hamster adenine phosphoribosyltransferase (APRT) gene as targeting vectors, we have investigated the homology dependence of targeted recombination at the endogenous APRT locus in Chinese hamster ovary (CHO) cells. We have examined the effects of varying either the overall length of targeting sequence homology or the length of 5' or 3' flanking homology on both the frequency of targeted homologous recombination and the types of recombination events that are obtained. We find an exponential (logarithmic) relationship between length of APRT targeting homology and the frequency of targeted recombination at the CHO APRT locus, with the frequency of targeted recombination dependent upon both the overall length of targeting homology and the length of homology flanking each side of the target gene deletion. Although most of the APRT+ recombinants analyzed reflect simple targeted replacement or conversion of the target gene deletion, a significant fraction appear to have arisen by target gene-templated extension and correction of the targeting fragment sequences. APRT fragments with limited targeting homology flanking one side of the target gene deletion yield proportionately fewer target gene conversion events and proportionately more templated extension and vector correction events than do fragments with more substantial flanking homology.


Subject(s)
Adenine Phosphoribosyltransferase/genetics , Cricetulus/genetics , Recombination, Genetic , Sequence Homology, Nucleic Acid , Animals , CHO Cells , Cricetinae , Gene Conversion , Gene Deletion , Genetic Vectors , Mutagenesis, Insertional , Nucleic Acid Hybridization , Restriction Mapping , Selection, Genetic
14.
Somat Cell Mol Genet ; 19(4): 363-75, 1993 Jul.
Article in English | MEDLINE | ID: mdl-8105543

ABSTRACT

A 21-bp deletion in the third exon of the APRT gene in Chinese hamster ovary (CHO) cells was corrected by transfection with a plasmid containing hamster APRT sequences. Targeted correction frequencies in the range of 0.3-3.0 x 10(-6) were obtained with a vector containing 3.2 kb of APRT sequence homology. To examine the influence of vector configuration on targeted gene correction, a double-strand break was introduced at one of two positions in the vector prior to transfection by calcium phosphate-DNA coprecipitation or electroporation. A double-strand break in the region of APRT homology contained in the vector produced an insertion-type vector, while placement of the break just outside the region of homology produced a replacement-type vector. Gene targeting with both linear vector configurations yielded equivalent ratios of targeted recombinants to nontargeted vector integrants; however, targeting with the two different vector configurations resulted in different distributions of targeted recombination products. Analysis of 66 independent APRT+ recombinant clones by Southern hybridization showed that targeting with the vector in a replacement-type configuration yielded fewer targeted integrants and more target gene convertants than did the integration vector configuration. Targeted recombination was about fivefold more efficient with electroporation than with calcium phosphate-DNA coprecipitation; however, both gene transfer methods produced similar distributions of targeted recombinants, which depended only on targeting vector configuration. Our results demonstrate that insertion-type and replacement-type gene targeting vectors produce similar overall targeting frequencies in gene correction experiments, but that vector configuration can significantly influence the yield of particular recombinant types.


Subject(s)
Adenine Phosphoribosyltransferase/genetics , Gene Transfer Techniques , Genetic Vectors , Animals , Base Sequence , Blotting, Southern , CHO Cells , Cell Line , Cricetinae , DNA , Molecular Sequence Data , Plasmids , Polymorphism, Restriction Fragment Length , Recombination, Genetic , Sequence Homology, Nucleic Acid
15.
Mutat Res ; 261(4): 267-79, 1991 Dec.
Article in English | MEDLINE | ID: mdl-1722282

ABSTRACT

The initiation of carcinogenesis by carcinogens such as 7r,8t-dihydroxy-9,10t-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE-I) is thought to involve the formation of DNA adducts. However, the diastereomeric diol epoxide, 7r,8t-dihydroxy-9,10c-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE-II), also forms DNA adducts but is inactive in standard carcinogenesis models. We have measured the formation and loss of DNA adducts derived from BPDE-II in a DNA-repair-proficient line of Chinese hamster ovary (CHO) cells, AT3-2, and in two derived mutant cell lines, UVL-1 and UVL-10, which are unable to repair bulky DNA adducts. BPDE-II adducts were lost from cellular DNA in AT3-2 cells with a half-life of 13.8 h; this was about twice the rate found for BPDE-I adducts. BPDE-II adducts were also lost from DNA in UVL-1 and UVL-10 cells, but at a much slower rate. When purified DNA was modified in vitro with BPDE-II and then held at 37 degrees C, DNA adducts were removed at a rate identical to that seen in UVL-1 and UVL-10 cells, suggesting that the loss in these cells was not due to enzymatic DNA-repair processes but to chemical lability of the adducts. Mutant frequencies at the APRT and HPRT loci were measured at BPDE-II doses that resulted in greater than 20% survival, and were found to increase linearly with dose. In the DNA-repair-deficient cells, the HPRT locus was moderately hypermutable compared with AT3-2 cells (about 5-fold); the APRT locus was extremely hypermutable, giving about 25-fold higher mutant fractions in UVL-1 and UVL-10 than in AT3-2 cells at equal initial levels of binding. When we compared the mutational efficiency of BPDE-II at both loci in AT3-2 cells (the mutant frequency in mutants/10(6) survivors at a dose that resulted in one adduct per 10(6) base pairs) with our previous studies of BPDE-1, we found that BPDE-II was 4-5 times less efficient as a mutagen than BPDE-I. This difference in mutational efficiency could be explained in part by the increased rate of loss of BPDE-II adducts from the cellular DNA, part of which was due to an increased rate of enzymatic removal of these lesions compared with the removal of BPDE-I adducts.


Subject(s)
7,8-Dihydro-7,8-dihydroxybenzo(a)pyrene 9,10-oxide/metabolism , Benzopyrenes/toxicity , DNA Adducts , DNA Repair/drug effects , DNA/metabolism , Epoxy Compounds/toxicity , Mutagenesis/genetics , Adenine Phosphoribosyltransferase/genetics , Animals , CHO Cells , Carcinogenicity Tests , Cricetinae , Dose-Response Relationship, Drug , Epoxy Compounds/metabolism , Hypoxanthine Phosphoribosyltransferase/genetics , Kinetics , Mutagenicity Tests , Mutagens/toxicity , Mutation/genetics , Stereoisomerism
16.
Mutat Res ; 261(4): 281-93, 1991 Dec.
Article in English | MEDLINE | ID: mdl-1722283

ABSTRACT

Insights into the mechanisms of chemical carcinogenesis can sometimes be gained by comparing the effects of closely related chemicals which differ in carcinogenic potency. We have treated Chinese hamster ovary (CHO) cells with a non-carcinogenic metabolite of benzo[a]pyrene, 9r,10t-dihydroxy-7c,8c-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE-III), and measured the formation and persistence of DNA adducts. We have correlated this binding data with cytotoxicity and mutagenicity in a DNA-repair-proficient CHO cell line (AT3-2) and in two derived lines, UVL-1 and UVL-10, which are unable to repair bulky DNA adducts. These data are compared with similar studies of the effects of the carcinogenic metabolite, 7r,8t-dihydroxy-9t,10t-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE-I). Synchronous fluorescence spectroscopy was used to measure the levels of BPDE-III-DNA adducts in treated cells. Adduct levels increased linearly with dose, but the absolute binding levels were about 30-fold lower than in comparable incubations with BPDE-I. Measurements of the removal of adducts derived from these two diol epoxides indicated no significant difference in the rate of repair measured 24 h post-treatment. When cells were treated with increasing doses of BPDE-III, survival curves were obtained which exhibited a shoulder region at low doses and an exponential decrease in plating efficiency at higher doses. By comparison of the D0's, the DNA-repair-deficient cell lines were found to be 4-5-fold more sensitive to the killing effects of BPDE-III than were the repair-proficient AT3-2 cells.


Subject(s)
7,8-Dihydro-7,8-dihydroxybenzo(a)pyrene 9,10-oxide/metabolism , Benzopyrenes/toxicity , DNA Adducts , DNA Repair/drug effects , DNA/metabolism , Epoxy Compounds/toxicity , Mutagenesis/genetics , Adenine Phosphoribosyltransferase/genetics , Animals , CHO Cells , Carcinogenicity Tests , Cell Survival/drug effects , Cricetinae , Fluorescence , Hypoxanthine Phosphoribosyltransferase/genetics , Kinetics , Mutagenicity Tests , Mutation/genetics
17.
Plasmid ; 25(3): 208-16, 1991 May.
Article in English | MEDLINE | ID: mdl-1924558

ABSTRACT

A plasmid was constructed by fusion of a selectable mammalian gene, hamster adenine phosphoribosyltransferase (APRT), to the Zn(2+)-inducible sheep metallothionein I (MT I) promoter. This plasmid was used to produce stable Chinese hamster ovary (CHO) cell transformants by electroporation to study the effects of induced gene expression on DNA-mediated transformation. The sheep MT Ia promoter was chosen for these experiments because it regulates gene expression differently than murine MT promoters, exhibiting low basal levels of gene expression in uninduced conditions. We have shown that in the absence of Zn2+, there is very low expression of a sheep MT I-APRT fusion gene in stable CHO cells transformants; induction of APRT mRNA and enzyme activity by Zn2+ produced a "threshold" response, from low basal levels to high induced levels, in Zn2+ responsive stable transformant clones. In electroporation experiments, transformation frequencies were unaffected by Zn2+ treatments during a preselection period, but the presence of Zn2+ during selection increased the recovery of stable transformant clones 8- to 10-fold. All stable transformants analyzed displayed Zn(2+)-inducible APRT enzyme activity. Our results indicate that stable mammalian cell transformants with inducible genes under regulation of the sheep MT I promoter should be useful, because of low basal and high induced expression, for studies in which modulation of transcriptional activity is required.


Subject(s)
Adenine Phosphoribosyltransferase/genetics , Plasmids , Adenine Phosphoribosyltransferase/metabolism , Animals , CHO Cells , Cell Division , Cell Line, Transformed , Chromosome Deletion , Clone Cells , Cricetinae , Metallothionein/genetics , Promoter Regions, Genetic , Restriction Mapping , Transfection
18.
Mol Carcinog ; 4(6): 519-26, 1991.
Article in English | MEDLINE | ID: mdl-1793489

ABSTRACT

We previously showed that ultraviolet (UV) irradiation of cotransfected plasmid DNA molecules stimulated genetic transformation that depended on intermolecular homologous recombination in Chinese hamster ovary (CHO) cells. Repair-proficient cells and an excision repair complementation class 1 (ERCC1) UV-sensitive DNA repair-deficient mutant responded similarly to UV stimulation in cotransfections with plasmids containing linker insertion-disrupted copies of the herpes simplex virus thymidine kinase (HSV-TK) gene. In this study, we cotransfected homologous DNA molecules containing nonoverlapping deletions of the hamster adenine phosphoribosyltransferase (APRT) gene into APRT-deficient CHO ERCC1 (UVL-10) and ERCC2 (UVL-1) excision-repair mutants and parental repair-proficient CHO cells. UV damage in cotransfected circular plasmid molecules stimulated transformation in repair-proficient cells and an ERCC1 mutant, but not in an ERCC2 mutant. Linearization of plasmids prior to cotransfection greatly enhanced transformation frequencies in all three cell lines, but UV stimulation using linear recombination substrates was no longer evident. Our results suggest (i) that the ERCC1 gene defect in CHO UVL-10 cells does not affect UV stimulation of homology-dependent extra-chromosomal recombination, and (ii) that a CHO cell ERCC2 excision-repair mutant, although recombination proficient, may exhibit altered recombination in response to UV damage.


Subject(s)
DNA Repair , Recombination, Genetic/radiation effects , Adenine Phosphoribosyltransferase/genetics , Animals , CHO Cells , Cricetinae , DNA Damage , Plasmids , Transfection , Ultraviolet Rays
19.
Somat Cell Mol Genet ; 16(5): 437-41, 1990 Sep.
Article in English | MEDLINE | ID: mdl-2237639

ABSTRACT

We demonstrate the feasibility of targeted gene replacement at an endogenous, chromosomal gene locus in cultured mammalian cells, employing a two-step strategy similar to an approach routinely used for genetic manipulation in yeast. Utilizing an APRT+ recombinant generated by targeted integration of plasmid sequences (including a functional copy of the gpt gene) at the CHO APRT locus, we have been able to select gpt- "pop-out" recombinants that have arisen by intrachromosomal recombination between APRT direct repeats at the targeted integration site. Reciprocal exchanges leading to "pop-out" of integrated plasmid/gpt gene sequences occur at a rate of approximately 6.3 x 10(-6) per cell generation. Depending on the site of crossover, such "pop-out" events result in either replacement or restoration of the original APRT target gene sequence.


Subject(s)
Adenine Phosphoribosyltransferase/genetics , Genetic Engineering/methods , Adenine/analogs & derivatives , Animals , Blotting, Southern , Cell Line , Chromosome Deletion , Cricetinae , Cricetulus , Drug Resistance/genetics , Female , Mutagenesis, Site-Directed , Ovary/cytology , Recombination, Genetic , Thioguanine
20.
Nucleic Acids Res ; 18(17): 5173-80, 1990 Sep 11.
Article in English | MEDLINE | ID: mdl-2169607

ABSTRACT

In order to identify a poison sequence that might be useful in studying illegitimate recombination of mammalian cell chromosomes, several DNA segments were tested for their ability to interfere with gene expression when placed in an intron. A tRNA gene and its flanking sequences (267 bp) were shown to inhibit SV40 plaque formation 100-fold, when inserted into the intron in the T-antigen gene. Similarly, when the same DNA segment was placed in the second intron of the adenosine phosphoribosyl transferase (APRT) gene from CHO cells, it inhibited transformation of APRT-CHO cells 500-fold. These two tests indicated that the 267-bp DNA segment contained a poison sequence. The poison sequence did not affect replication since the replication of poisoned SV40 genomes was complemented by viable SV40 genomes and poisoned APRT genes were stably integrated into cell chromosomes. Cleavage of the poison sequence in the SV40 T-antigen intron by restriction enzymes indicated that the tRNA structural sequences and the 5' flanking sequences were not required for inhibition of SV40 plaque formation. Sequence analysis of viable mutant SV40, which arose after transfection of poisoned genomes, localized the poison sequence to a 35 bp segment immediately 3' of the tRNA structural sequences.


Subject(s)
Gene Expression Regulation , Introns , Mutation , Recombination, Genetic , Adenine Phosphoribosyltransferase/genetics , Animals , Antigens, Viral, Tumor/genetics , Base Sequence , Cell Line , Gene Expression Regulation, Enzymologic , Genes , Genes, Viral , Haplorhini , Molecular Sequence Data , RNA, Transfer/genetics , Simian virus 40/genetics , Simian virus 40/physiology , Transfection , Virus Replication/genetics
SELECTION OF CITATIONS
SEARCH DETAIL
...