BCAR1, a Human hom*ologue of the Adapter Protein p130Cas, and Antiestrogen Resistance in Breast Cancer Cells (2024)

Article Navigation

Volume 92 Issue 2 19 January 2000

Article Contents

  • Materials and Methods

  • Results

  • Discussion

  • < Previous
  • Next >

Journal Article

,

Arend Brinkman

1Affiliation of authors: Department of Pathology/Divisionof Molecular Biology, Josephine Nefkens Institute, University HospitalRotterdam, The Netherlands.

Correspondence to: ArendBrinkman, Ph.D., Department of Pathology/Division of Molecular Biology, Josephine NefkensInstitute, University Hospital Rotterdam, Rm. Be 435a, P.O. Box 1738, 3000 DR Rotterdam, TheNetherlands (e-mail: brinkman@bidh.azr.nl).

Search for other works by this author on:

Oxford Academic

,

Silvia van der Flier

1Affiliation of authors: Department of Pathology/Divisionof Molecular Biology, Josephine Nefkens Institute, University HospitalRotterdam, The Netherlands.

Search for other works by this author on:

Oxford Academic

,

Elisabeth M. Kok

1Affiliation of authors: Department of Pathology/Divisionof Molecular Biology, Josephine Nefkens Institute, University HospitalRotterdam, The Netherlands.

Search for other works by this author on:

Oxford Academic

Lambert C. J. Dorssers

1Affiliation of authors: Department of Pathology/Divisionof Molecular Biology, Josephine Nefkens Institute, University HospitalRotterdam, The Netherlands.

Search for other works by this author on:

Oxford Academic

JNCI: Journal of the National Cancer Institute, Volume 92, Issue 2, 19 January 2000, Pages 112–120, https://doi.org/10.1093/jnci/92.2.112

Published:

19 January 2000

Article history

Received:

09 June 1999

Revision received:

03 November 1999

Accepted:

09 November 1999

Published:

19 January 2000

  • PDF
  • Split View
  • Views
    • Article contents
    • Figures & tables
    • Video
    • Audio
    • Supplementary Data
  • Cite

    Cite

    Arend Brinkman, Silvia van der Flier, Elisabeth M. Kok, Lambert C. J. Dorssers, BCAR1, a Human hom*ologue of the Adapter Protein p130Cas, and AntiestrogenResistance in Breast Cancer Cells, JNCI: Journal of the National Cancer Institute, Volume 92, Issue 2, 19 January 2000, Pages 112–120, https://doi.org/10.1093/jnci/92.2.112

    Close

Search

Close

Search

Advanced Search

Search Menu

BACKGROUND: Treatment of breast cancer with the antiestrogen tamoxifen is effective inapproximately one half of the patients with estrogen receptor-positive disease, but tumors recurfrequently because of the development of metastases that are resistant to tamoxifen. We havepreviously shown that mutagenesis of human estrogen-dependent ZR-75-1 breast cancer cells byinsertion of a defective retrovirus genome caused the cells to become antiestrogen resistant. Inthis study, we isolated and characterized the crucial gene at the breast cancer antiestrogenresistance 1 (BCAR1) locus. METHODS/RESULTS: Transfer of the BCAR1 locus fromretrovirus-mutated, antiestrogen-resistant cells to estrogen-dependent ZR-75-1 cells by cell fusionconferred an antiestrogen-resistant phenotype on the recipient cells. The complete codingsequence of BCAR1 was isolated by use of exon-trapping and complementary DNA (cDNA)library screening. Sequence analysis of human BCAR1 cDNA predicted a protein of 870 aminoacids that was strongly hom*ologous to rat p130Cas-adapter protein. Genomic analysis revealedthat BCAR1 consists of seven exons and is located at chromosome 16q23.1. BCAR1 transcriptswere detected in multiple human tissues and were similar in size to transcripts produced byretrovirus-mutated ZR-75-1 cells. Transfection of BCAR1 cDNA into ZR-75-1 cells againresulted in sustained cell proliferation in the presence of antiestrogens, confirming that BCAR1was the responsible gene in the locus. CONCLUSIONS: Overexpression of the BCAR1 geneconfers antiestrogen resistance on human ZR-75-1 breast cancer cells. Overexpression of BCAR1in retrovirus-mutated cells appears to result from activation of the gene's promoter. Theisolation and characterization of this gene open new avenues to elucidating mechanisms by whichthe growth of human breast cancer becomes independent of estrogen.

The steroid hormone estradiol plays a pivotal role in breastcancer initiation and cell proliferation (1,2). Estradiol andits receptor participate in a multiprotein complex that acts as atranscription regulator for various genes (3-6). Inapproximately two thirds of all breast tumors, the estrogen receptor(ER) is present and functionally active (7-9). The presenceof the ER in breast carcinoma cells and their responsiveness toestrogens is clinically exploited by the administration of antagonistsof estrogens or antiestrogens (10-12). On binding,antiestrogens induce an aberrant conformation of the ER, resulting inan inactive transcription complex (13). As a consequence, theantiestrogen-receptor complex may bind DNA but can no longer modulatethe expression of its target genes (14). This inhibition ofthe estrogen response pathway (15) may block cellularproliferation (19,20).

During the last 2 decades, antiestrogens, particularly tamoxifen, have proved to be effective inthe treatment of hormone-responsive breast cancer (21-23). Adjuvanttreatment with tamoxifen reduces tumor recurrence and increases survival of patients withER-positive breast cancer (24,25). In metastatic breast cancer, tamoxifenleads to objective response in nearly one half of patients with ER-positive primary tumors (26). Resistance to antiestrogens, however, is a serious obstacle in themanagement of breast cancer. About 40% of ER-positive tumors fail to respond toantiestrogen therapy (intrinsic resistance) (27), whereas eventually most,if not all, breast tumors that initially respond to antiestrogens develop resistant metastases(acquired resistance).

The mechanisms underlying intrinsic or acquired antiestrogen resistance of breast tumors arestill poorly understood. Altered pharmacology of the antiestrogens (30-32), andchanges in the interactions between tumor cells and their environment (paracrine interactions) (33) have been proposed to contribute to the development of antiestrogenresistance. In addition, genetic or epigenetic changes in the tumor cells that promote thedevelopment of antiestrogen resistance have been postulated (34,35).Although each of these possibilities may account for or contribute to the resistant phenotype inindividual patients, none so far has been shown to explain antiestrogen resistance in a majority ofpatients. It is likely that progression to antiestrogen resistance is a multifactorial process.

We attempted to identify the genetic factors that may lead to antiestrogen resistance.Therefore, we applied retroviral-insertion mutagenesis to the human estrogen-dependent breastcancer cell line ZR-75-1. We demonstrated that random integration of a defective murineretrovirus in the genome of ZR-75-1 cells transformed the cells from an estrogen-dependent to atamoxifen-resistant phenotype (36). Mapping of common integration sitesand cell fusion-mediated gene transfer experiments so far have identified three independent locithat are involved in antiestrogen resistance (BCAR1, BCAR2, and BCAR3) (37). Here we report the isolation and characterization of the target gene of the BCAR1locus and its functional involvement in antiestrogen resistance in vitro.

Materials and Methods

Cell Lines

The human breast carcinoma cell line ZR-75-1 and derivatives thereof were cultured asdescribed elsewhere (36,38). Hybrid cell lines and BCAR1-transfected celllines were cultured in RPMI-1640 medium supplemented with 10% heat-inactivated bovinecalf serum and 1 nM 17β-estradiol. For BCAR1-transfected cell lines, 500μg/mL of the neomycin analogue G418 (Life Technologies, B.V., Breda, The Netherlands)was added to the culture medium.

The tamoxifen-resistant cell line XI-13, with two copies of the defective LN retroviralgenome (36,39), one located within the BCAR1 locus, was lethally γirradiated (40 Gy) and subsequently fused to hygromycin B-resistant ZR-75-1 (ZH3D7) cells withpolyethylene glycol as described previously (37). G418 plus hygromycinselection was performed in estradiol-containing culture medium (37).

Molecular Biology Techniques

A human genomic cosmid library in pJB8 (40) was obtained from W.J. M. van de Ven (University of Leuven, Leuven, Belgium) and screened by use of standardprotocols (41).

Exon trapping was performed according to the manufacturer's instructions (LifeTechnologies, Inc). In short, genomic DNA of the cosmid clones HC26 and HC34 was partiallydigested with the enzyme MboI. Fragments of approximately 5-10 kilobases (kb)(average, 7.5 kb) were isolated and cloned in the exon-trap vector pSPL3. Batches of pooledpurified plasmid DNA from five clones each were transfected into the simian kidney cell line(COS-1) (41). After transient expression, RNA was isolated andcomplementary DNA (cDNA) was synthesized. Following two rounds of polymerase chainreaction (PCR), the trapped sequences were analyzed, cloned, and sequenced.

RNA was isolated from the somatic hybrid D4E5 (containing the BCAR1 locus) and culturedfor 2 days with 4-hydroxy-tamoxifen by use of the CsCl-guanidine thiocyanate procedure (41). Poly(A) RNA was selected with the use of the PolyATtract mRNAIsolation System (Promega Corp., Madison, WI) and converted into cDNA (Zap Express cDNAand cloning kit; Stratagene Cloning Systems, La Jolla, CA) as recommended by the manufacturer.A cDNA library of approximately 2 × 106 phages was created and screened.Twenty-one phage clones hybridizing with a BCAR1 exon-trap probe were isolated andsequenced by use of standard procedures (41). DNA and protein sequenceanalyses and alignments were performed with the use of DNASIS (Hitachi Software EngineeringAmerica Ltd., Brisbane, CA), Blast (www.ncbi.nlm.gov/blast), and Clustalw(www2.ebi.ac.uk/clustalw) software. The BCAR1 sequence has been assigned GenBankAccession No. AJ 242987.

Northern blots were prepared by use of 1% agarose-formaldehyde gels as describedpreviously (41). Blots were hybridized with random-primer-labeled probesfor BCAR1, the estrogen-regulated PS2 gene, and the glyceraldehyde-3-phosphatedehydrogenase gene for loading control as described (42,43)

Mapping of the BCAR1 locus was performed on genomic DNA from a panel of 30 somaticcell hybrids obtained from D. F. Callen (Women's and Children's Hospital, Adelaide,Australia). PCR was performed with the locus-specific primers 11405′-CCCCACATACCCAGCACA-3′ and 11425′-CCCAGCTCCTCCTTATTCA-3′. The amplified product was verified byhybridization with the BCAR1 locus-specific probe 18/1 (36).

Expression Constructs and Transfection of BCAR1 in ZR-75-1 Cells

The full-length BCAR1 cDNA was inserted into the EcoRI-XhoI site of thelong terminal repeat-promoted pLXSN expression vector (39) and into amodified version of the episomal vector LZRS-IRES-Neo (LZRSpBMN-IRES) (44). In the latter vector, the BCAR1 cDNA and the neomycin resistance gene (to conferselectability) were separated by an IRES sequence, so that both genes were under control of asingle long terminal repeat (LTR) promoter and transcribed as a polycistronic messenger.

Stable transfectants were generated with the pLXSN/BCAR1 construct by the use of calciumphosphate precipitation (41). Transfection of ZR-75-1 cells with theepisomal LZRSpBMN-IRES/BCAR1 vector construct was performed by use of Lipofectinreagent (Life Technologies, Inc.), according to the manufacturer's instructions.Transfectants were selected for neomycin resistance (500 μg of G418/mL) inestradiol-containing medium.

Proliferation Assay

Somatic hybrid cells, BCAR1, and mock-transfected cells were grown in culture mediumsupplemented with 1 nM estradiol. Cells were harvested at 80% confluence bytrypsinization. The cells (0.7 × 106) were seeded in 25-cm2 flasks, in triplicate, and grown in the presence of either 1 μM OH-tamoxifen or 100nM of the pure antiestrogen ICI 182,780 (Zeneca Pharma, Ridderkerk, TheNetherlands). After 4-6 days, the cells were again trypsinized, counted using a Coulter Z1 cellcounter (Coulter Electronics Ltd, Luton, U.K.), and reseeded in triplicate as above. To follow theproliferation of the cells, the procedure was repeated at several time points and for each cell line.In case of limited recovery of cells at the end of the assay period, only two flasks were reseededor the culture was terminated. Proliferation was determined as the fold multiplication (mean of atleast two flasks) relative to day 0.

Drug-sensitivity tests on ZR-75-1-derived cell lines were performed in complete mediumcontaining estradiol in 96-well plates. Five thousand cells were plated with threefold dilutions(four wells each) of the drug. After 9 days of culture in the presence of the drug, viable cells weremeasured by use of the MTT assay as described (38)

Bcar1 Protein Detection

BCAR1 and mock-transfected cells were trypsinized, rinsed with phosphate-buffered saline,sonicated for 10 seconds, and lysed for 10 minutes at 100 °C in 1% sodium dodecylsulfate (SDS) and 10 mM Tris-HCl (pH 7.5). The protein concentration of the crudelysates was determined (BCA™, Protein assay reagent; Pierce Chemical Co., Rockford, IL).Equal amounts (3.3 μg) of protein were used for SDS-polyacrylamide gel electrophoresis(PAGE) (8% acrylamide) and subjected to western blot analysis as described (45). Bcar1 was visualized with the mouse monoclonal antibody to rat p130Cas(Transductional Laboratories, Lexington, KY). In addition, polyclonal antibodies raised againstthe N-terminus and C-terminus of rat p130Cas (Cas N-17 and Cas C-20; Santa CruzBiotechnology Inc., Santa Cruz, CA) were used to identify Bcar1 in an independent controlexperiment.

Results

Transfer of the BCAR1 Locus by Cell Fusion

We have used retrovirus insertion-mediated mutagenesis on the humanestrogen-dependent breast cancer cell line ZR-75-1 to generate 80 celllines showing resistance to 4-hydroxy-tamoxifen (36).Integration of the defective murine retrovirus (LN) into the host cellgenome could transform these cells to a tamoxifen-resistant phenotype.With the use of integration site-specific probes, we previously found acommon integration locus for the retrovirus in a subset of this panelof cell lines. Four independent cell clones were identified by Southernblot analysis, with a viral integration in the same orientation atdifferent positions within a 2.2-kb genomic region (36)(see also Fig. 2, A). We termed this locus BCAR1, for breastcancer antiestrogen resistance 1 locus—i.e., the first BCAR locusidentified.

To establish whether retroviral integration in the BCAR1 locus induces tamoxifen resistancein ZR-75-1 cells, we transferred the genomic region encompassing the BCAR1 locus fromtamoxifen-resistant cells into estrogen-dependent ZR-75-1 cells by cell fusion. XI-13 cells carrytwo copies of the defective retroviral genome, one copy of which is located within the BCAR1locus. On lethal γ irradiation, the cells were fused to hygromycin B-resistant ZR-75-1 cells(ZH3D7) with polyethylene glycol as described previously (37). Theneomycin resistance gene in the integrated retrovirus allowed the isolation of somatic cell hybridsby dual selection with G418 plus hygromycin in estradiol-containing culture medium. Twelve cellhybrids containing either one of the integration loci in the ZR-75-1 background were rescued andanalyzed. Southern blotting analysis by use of a specific probe for the neomycin resistance generevealed that five hybrids had retained the BCAR1 integration locus (a 15-kb BglIIfragment), whereas the remaining seven carried the other integration locus (a 10-kb BglII fragment). Both the virus-induced BCAR1 cell line XI-13 and the BCAR1 locus-containing cellhybrids derived therefrom (e.g., D4E5 and D4E6) showed enhanced proliferation when culturedin the presence of the antiestrogen ICI 182,780 (Fig. 1, A). Hybrid celllines carrying the other locus, as well as the ZH3D7 recipient cells, were growth inhibited. In theabsence of antiestrogens, BCAR1 locus-containing hybrids also demonstrated increasedproliferation compared with the control cell lines (Fig. 1, B). Note that theproliferation of most cell lines is more efficient in the absence of antiestrogen, which is likelybecause of some residual estrogens in the serum-containing medium. In the presence ofantiestrogens, the ER pathway is more effectively blocked. The lack of the ER in XI-13 cells (36) can explain the comparable growth levels of these cells in the presenceas well as in the absence of antiestrogens (Fig. 1). These fusionexperiments demonstrated unambiguously that the BCAR1 locus induced antiestrogen-resistantproliferation of the somatic cell hybrids and that the BCAR1 phenotype is dominant.

Cloning of the BCAR1 Locus

A human cosmid library was screened with genomic probes that havebeen employed to identify the BCAR1 locus (36). The probesincluded a 500 base-pair (bp) ApaLI-PstI fragment(14B2) and a 450-bp SphI fragment (18/1). Both probes resideat the center of the viral integration site in the BCAR1 locus (Fig. 2, A). Cosmids were isolated and analyzed by restrictionmapping, resulting in a continuous contig of approximately 80 kb (Fig.2, A). Unique probes derived from the 80-kb contig identified nofurther proviral integrations events in the 80 antiestrogen-resistantcell lines from our panel.

Identification of the BCAR1 Gene

Two of 10 overlapping cosmid clones covering the major part of theBCAR1 locus—HC26 and HC34—were selected for identification oftranscripts by exon trapping. A total of 31 trapped sequences wererecovered, of which 16 were related to known genes. Eight clonescorresponded to exons 2-7 of the chymotrypsin gene (accession No.NM_001906), and the remaining eight appeared to be hom*ologous tosequences of the rat p130Cas gene (accession No. D29766) (46).Of the remaining 15 clones, two seemed to be generated because ofcryptic splicing events in frequently occurring alu-type repetitivesequences in the BCAR1 locus, while the other clones did not showhom*ology to any known sequences in available databases.

The trapped exons were hybridized to polyA messenger RNAs (mRNAs) isolated from theantiestrogen-resistant BCAR1 cell lines (XII-13, XI-14, and XI-13) and from the parental cell lineZR-75-1. The chymotrypsin gene sequences did not hybridize in a northern blot analysis,indicating that this gene is not expressed in our ZR-75-1-derived cell lines. In contrast, exonsequences that were hom*ologous to the rat p130Cas gene hybridized to a single mRNA of 3.2 kb(Fig. 2, B). This transcript appeared to be increased in theantiestrogen-resistant cell lines but not in the parental cell line. As shown in Fig. 2, B, ZR-75-1 exhibits only low levels of an equally sized transcript. The differentialexpression of this transcript suggested that the rat p130Cas-hom*ologous gene was the candidatebreast cancer antiestrogen resistance gene, BCAR1. Conclusive evidence for its role inantiestrogen resistance required the isolation of the BCAR1 cDNA and its functional transfer toZR-75-1 cells by transfection (see below).

Primary Structure of the BCAR1 Gene

A cDNA library from the antiestrogen-resistant BCAR1 cell hybridD4E5 was screened with p130Cas hom*ologous sequences retrieved from exontrapping. Twenty-one positive cDNA clones were identified and assembledby sequence alignment into a continuous cDNA of 3208 bp. The BCAR1 cDNAencloses an open-reading frame of 2610 nucleotides. At position 122, aninitiation codon is flanked by sequences matching Kozak's criteria(47). The open-reading frame is flanked at the 3′ end by atranslation termination codon (TGA) at position 2732 and a 3′untranslated sequence of 468 nucleotides that contains multiple stopcodons. A canonic polyadenylation site is located 14 bp in front of thestart of the polyA tail in the cDNAs. The GenBank Accession No. for theBCAR1 sequence is AJ 242987.

The open-reading frame has a coding capacity for a protein of 870 amino acid residues, with acalculated molecular mass of approximately 93 kd. The predicted protein features a Srchom*ology 3 (SH3) domain in the N-terminal part and multiple potential tyrosine phosphorylationsites (Tyr-Ala/Pro-Xxx-Pro) in the central part of the protein (Fig. 3). TheC-terminal part encloses a proline-rich stretch (Arg-Pro-Leu-Pro-Ser-Pro-Pro) that may interactwith the SH3-binding site of the Src protein (48). Sequence alignment ofBcar1 revealed extensive hom*ology (91%) with rat and mouse p130Cas protein (Fig. 3).

When hybridized with BCAR1 cDNA probes, RNA preparations of BCAR1 cell linesdisplayed a single transcript of approximately 3.2 kb (Fig. 2, B). The sizeof the mRNA is in agreement with that of the cDNA, suggesting that the cDNA represents thefull-length transcript. Northern blots of mRNAs derived from a panel of multiple human tissueshybridized to the BCAR1 cDNA probe showed that the BCAR1 transcript is widely expressed(Fig. 4). The 3.2-kb mRNA was present in all tissues tested, being mostabundant in testis and with only a low representation in liver, thymus, and peripheral bloodleukocytes.

Additionally, in vitro transcription and translation of the BCAR1 cDNAdemonstrated that the cDNA encodes a protein with an apparent molecular mass of approximately116 kd on SDS-PAGE. The discrepancy in observed and calculated (93-kd) molecular mass isprobably because of folding of the molecule, causing aberrant gel mobility. Both rat p130Cas andthe Bcar1 protein exhibit the same size. As expected, a monoclonal antibody directed against ratp130Cas identified the BCAR1 gene product (Fig. 5, A) as well aspolyclonal antibodies that are directed against either the N-terminus or the C-terminus of the ratp130Cas (data not shown).

Genomic Organization of BCAR1

The genomic organization of the BCAR1 gene was delineated by use ofthe BCAR1 cDNA as a template. Restriction mapping and partialsequencing of cosmid clones HC26 and HC3 revealed that the transcribedregion of the BCAR1 gene consists of seven exons (I-VII) spanningapproximately 25 kb of genomic DNA (Fig 2, B). The size of the exonsvaries from 76 to 1101 bp, and the size of the introns varies from 182bp to approximately 8.5 kb. Sequencing through the exon-intronboundaries demonstrated that all splice junctions conform to the GT/AGrule for exon-intron junctions. Southern blot analysis showed that allrestriction fragments from the cosmid clones that hybridized to BCAR1cDNA were also present in genomic DNA. More important, the codingsequence of the BCAR1 gene was found to be identical to that of thecDNA, excluding cloning artifacts.

Exon 1 of BCAR1 contained the first 12 bp of the coding sequence. It is of interest that thisposition of the exon-intron boundary exactly matched the position where alternative splicing hasbeen observed in rat as well as mouse p130Cas (46,49). Exons 2-6 allmapped within the coding sequence, and exon 7 contained the last 509 coding nucleotides and thecomplete 3′ nontranslated region. The 5′ flanking region of the BCAR1 gene had ahigh G + C content (80%) and lacks suppression of CpG dinucleotides, which ischaracteristic for CpG islands. Such a configuration, which is often found for housekeeping genes,is consistent with the ubiquitous expression of the BCAR1 gene (Fig. 4).

In situ hybridization with the complete cosmid HC26 suggested that the BCAR1gene was located at human chromosome 16q22-23 (Van Agthoven T: personal communication).Primers specific for the integration region were used to analyze a panel of radiation hybridscontaining fragments of human chromosome 16. PCR analysis on genomic DNA derived fromthese hybrid cell lines determined the position of the BCAR1 locus to 16q23.1. Hybrids lackingthe region 16q23.2 to telomere (e.g., CY145) gave no BCAR1-specific product, whereas hybridsthat retained the 16q23.1 region (e.g., CY116 and CY117) gave specific products (50). Evidence for this position of the BCAR1 gene was further supported by theconcurrent trapping of exons of the chymotrypsin gene that had been assigned to chromosome16q23.1 (50). Hybridization with chymotrypsin gene probes on theBCAR1 cosmid clones confirmed the position of the chymotrypsin gene 4-5 kb downstream of theBCAR1 gene (Fig. 2, A).

Considering the vast hom*ology of the BCAR1 cDNA with rat and murine p130Cas sequencesand the strict conservation of the first exon-intron boundary, we postulate that the BCAR1 gene isthe human p130Cas hom*ologue, and we will further refer to this gene as BCAR1/p130Cas.

Transfection of BCAR1/p130Cas cDNA and Estrogen Independence in ZR-75-1Cells

To demonstrate that Bcar1/p130Cas is functionally involved inantiestrogen resistance, the full-length BCAR1/p130Cas cDNA wasintroduced in antiestrogen-sensitive ZR-75-1 cells by transfection.Five independent LXSN/BCAR1/p130Cas transfectants were generated,carrying an intact transgene integrated in the host genome asdetermined by Southern blot analysis. Similarly, eight independent celllines carrying the BCAR1/p130Cas cDNA in an episomal vector wereestablished. In addition, more than 22 vector-control cell lines havebeen raised.

All transfected cell lines were generated in estrogen-containing medium and subsequentlytested for proliferation in the presence of either 4-hydroxy-tamoxifen or the pure antiestrogen ICI182,780. A typical example of such an assay with the use of 4-hydroxy-tamoxifen is shown in Fig.5, A. Aliquots of the cultured cells have been used to measure the amountof Bcar1/p130Cas protein with the use of western blotting and immunodetection (Fig. 5, A). All cell lines exhibiting high expression of Bcar1/p130Cas (e.g.,4A12, B4, B6, and C4 and the somatic cell hybrid D4E6) showed net proliferation in the presenceof 4-hydroxy-tamoxifen. In contrast, ZR-75-1 and all vector control cell lines (e.g., C2, C6, D2,and D4) expressed very low levels of or no Bcar1/p130Cas and failed to proliferate in thepresence of the antiestrogen. Similar results were obtained in the assays by the use of the pureantiestrogen ICI 182,780 (data not shown). Of interest, Bcar1/p130Cas-transfected cellsunderwent a morphologic change when cultured in the presence of antiestrogens (data notshown). The cells displayed a typical flattened shape with multiple protrusions (similar to thevirus-induced cell lines), which may be suggestive of a reorganization of the cytoskeleton. Furtherexperiments with the use of immunofluorescence are needed to verify these observations. Toinvestigate the possibility that Bcar1/p130Cas overexpression also generates resistance to otherdrugs, we performed cell survival assays by use of doxorubicin, 5-fluorouracil, and methotrexate.These pilot experiments showed no relation between sensitivity to these drugs and overexpressionof Bcar1/p130Cas; i.e., comparable ID50 (dose showing 50% inhibition) valueswere obtained for overexpressing cell lines and cell lines without overexpression (data notshown).

Northern blotting analysis of Bcar1/p130Cas-transfected cells grown under antiestrogenselection revealed an impaired expression of the estrogen-regulated PS2 gene (43) similar to the parental cells (Fig. 5, B), suggesting that theproliferation of these cells was independent of ER function. Because integrin and growthfactor-receptor pathways may signal through p130Cas to the mitogen-activated protein kinase(MAPK) pathways (51-54), we have performed preliminary experimentsto monitor the activation state of various MAPKs in BCAR1/p130Cas-transfected cell lines.Specific antibodies directed against activated MAPKs (ERK, p38, and JNK) failed to documentconstitutive activation of these kinases in the presence of antiestrogens (not shown). Because ofthe slow growth and response kinetics of the transfected cells in antiestrogen-supplementedcultures, the levels of activated MAPK may have been below the threshold of detection in ourtests.

Discussion

The mechanisms underlying antiestrogen resistance in breast cancerare the subject of intensive study. Our working hypothesis was based onthe assumption that genetic or epigenetic alterations in the tumorcells contribute to the failure of the therapy. To identify targetgenes, we have employed a model system of estrogen dependenceresembling that seen in human breast cancer. In this article, wedescribe the cloning and characterization of the BCAR1 gene. On thebasis of our findings, BCAR1 appears to be the human hom*ologue of therat and mouse p130Cas genes; their function is described below.Overexpression of Bcar1/p130Cas in human estrogen-dependent ZR-75-1breast cancer cells is sufficient to drive cell proliferation in thepresence of antiestrogens. No difference in structure or size of theBCAR1/p130Cas transcript was seen in the virus-infected cell lines.This suggests that overexpression of the BCAR1/p130Cas gene is achievedby increased activity of the BCAR1/p130Cas gene promoter, likely due tothe presence of strong enhancer elements within the viral LTRs(55). This mechanism of transcription enhancement is differentfrom the observed viral promoter-insertion activation of the BCAR3 genethat resulted in an altered BCAR3 transcript and protein (43).It is possible that such events are selected against, in the case ofthe BCAR1/p130Cas gene, to ensure an intact and functional protein.

Our experiments indicate that BCAR1/p130Cas-mediated antiestrogen resistance is notassociated with resistance to other drugs relevant to the treatment of breast cancer. In addition,the ER does not play a crucial role in the resistant phenotype. Cell lines obtained after viralintegration in the BCAR1 locus (e.g., XI-13) completely lack ER expression and function (36) and show comparable growth rates in the absence or presence ofantiestrogens (Fig. 1). The somatic cell hybrids containing the BCAR1locus and the BCAR1/p130Cas transfectants are estrogen responsive with respect to growth andregulation of PS2 expression. These variants exhibit antiestrogen-resistant cell proliferation butare not dependent on antiestrogen for growth (Fig. 1, B). Theseobservations suggest a cell proliferation regulatory mechanism involving Bcar1/p130Cas, whichbypasses the ER-regulated pathway.

The observation that the BCAR1 gene is highly hom*ologous to the well-characterized rat andmouse p130Cas genes may help to understand the mechanism of antiestrogen-resistant breastcancer cell proliferation. Rat p130Cas has been identified as the major tyrosine-phosphorylatedprotein in v-Src- and v-Crk-transformed rat cells (46,56). Subsequentreports have implicated p130Cas in various processes in different cell types, including celltransformation (57,58), linking the extracellular matrix with the actincytoskeleton (59,60), integrin signaling (51,61-64), growth factor-receptor signaling (53,54,65-68), antigen-receptorsignaling (73,74), and essential cardiovascular development (58). This varied role of p130Cas may be explained by the presence ofseveral protein-protein interaction domains (46,75), which is the hallmarkof this novel class of adapter proteins. Family members are HEF1 (also designated CasL) (76,77) and Efs (also termed Sin) (78,79), whichmay have distinct functions (80). The SH3 domain, in particular, is wellconserved between these family members and has been shown to interact with proline-rich targetsequences from FAK (49), RAFTK (81),PTP-PEST (82), CAKbeta (83), and PTP1B (52). The central part of p130Cas contains multiple potentialtyrosine-phosphorylation sites capable of interacting with SH2 domain-containing proteins likeCrk (84,85), Src (48), and Nck (51). The C-terminal part of the p130Cas family members is well conserved but has notyet been implicated in a particular function in mammalian cells. Together, the numerousdocumented interactions and functions of p130Cas in various biologic processes and in differentcell types suggest a dynamic role for p130Cas. The outcome of these complex interactions willdepend on the availability of binding partners (including their activation status and affinity) andthe specific process in that cell type.

So far, little information is available regarding the role of Bcar1/p130Cas in regulation ofproliferation of breast epithelial cells. The observation that expression of Bcar1/p130Cas inprimary breast cancer predicted disease prognosis and response to tamoxifen therapy (45) suggests a genuine role for Bcar1/p130Cas in growth control of breast cancercells. Further elucidation of the pathway involved in Bcar1/p130Cas-mediated cell proliferation inbreast cancer cells may identify in vitro key regulators and possibly novel clinical targets.The recognition that both BCAR1/p130Cas and BCAR3 (42) can controlproliferation of breast cancer cells further supports the validity of our model system to identifybreast cancer antiestrogen resistant genes.

BCAR1, a Human hom*ologue of the Adapter Protein p130Cas, and AntiestrogenResistance in Breast Cancer Cells (3)

Fig 1.

Antiestrogen resistance assay of somaticcell hybrids. Antiestrogen-sensitive recipient cells (ZH3D7),antiestrogen-resistant donor cells (XI-13), and somatic cell hybridswith (D4E5 and D4E6) and without (D1F5 and D3D0) BCAR1 locus werecultured in triplicate in serum-containing medium with 100 nM of ICI 182,780 (panel A) or without (panel B). At thetimes indicated, cells were harvested, counted, and reseeded intriplicate. The numbers of cells recovered underantiestrogen at day 18 were insufficient to continue the culture ofD1F5 and only allowed for reseeding in duplicate for ZH3D7 and D3D0(panel A). Culture of the steadily growing XI-13 cells underantiestrogen was terminated at day 27 (panel A). Indicated isthe fold multiplication (mean of at least two flasks ± standarddeviation) relative to day 0. Error bars were smaller than symbols in cases where they do not appear.

Open in new tabDownload slide

BCAR1, a Human hom*ologue of the Adapter Protein p130Cas, and AntiestrogenResistance in Breast Cancer Cells (4)

Fig. 2.

A) Physical map of the human BCAR1 locus. At the top, the relative position of the BCAR1 gene and part of thechymotrypsin gene (CTRB) is shown (black bars) at a scale inkilobases (kb). The positions are shown of five representativeoverlapping cosmid clones, forming a continuous contig of 80 kb. At the bottom is an exploded view of the 2.2-kb integration region. Arrows indicate integration sites in varioustamoxifen-resistant BCAR1 cell lines. Note that each arrow represents an integration event in an individual BCAR1 cell line.The BCAR1 locus-specific probes 18/1 and 14B2 are shown as shadedboxes. (Sp = SphI, S = SstI, P = PstI, A= ApaLI, E = EcoRI, K = KpnI, and B= BamHI). B) Schematic representation of the BCAR1gene (not to scale). Shown are the relative positions of the 5′viral LTR, a putative CpG island, and the complete BCAR1 geneconsisting of seven exons. Underneath, the BCAR1 complementary DNA isdrawn encompassing an 870 amino acid open-reading frame with an SH3domain in the N-terminal part. The inserton the left presents a northern blot containing approximately 3 μg of polyAmessenger RNA of three representative BCAR1 antiestrogen-resistant cellclones raised by retroviral insertion mutagenesis (XII-13, XI-14, andXI-13) and ZR-75-1 cells. The blot was hybridized with trappedsequences hom*ologous to rat p130Cas (top panel) or withglyceraldehyde-3-phosphate dehydrogenase (GAPDH) probe to control forloading (bottom panel). Blots were washed until 0.3 ×standard saline citrate at 65 °C for 30 minutes. LTR = longterminal repeats; bp = base pair.

Open in new tabDownload slide

BCAR1, a Human hom*ologue of the Adapter Protein p130Cas, and AntiestrogenResistance in Breast Cancer Cells (5)

Fig. 3.

Bcar1 alignment with p130Cas. A Blast(www.ncbi.nlm.gov/blast) search with BCAR1 complementary DNAsequencesin available databases demonstrated hom*ology to rat and murine p130Cassequences. An alignment is shown of the predicted protein sequences ofhuman Bcar1 with rat and mouse p130Cas. A single gap of four amino acidresidues has been introduced by the Clustalw program(www2.ebi.ac.uk/clustalw) for optimal alignment. An SH3 hom*ologydomain(boxed), a proline-rich sequence region (bold and underlined), and multiple potential tyrosine phosphorylationsites (shaded) are indicated. The BCAR1 sequence has beenassigned GenBank accession No. AJ 242987.

Open in new tabDownload slide

BCAR1, a Human hom*ologue of the Adapter Protein p130Cas, and AntiestrogenResistance in Breast Cancer Cells (6)

Fig. 5.

Characterization of BCAR1/p130Cas transfectants. A) Left panel—Growth curves of a representativeselection of BCAR1-transfected cell lines grown in the presence ofantiestrogen. Three flasks of each LXSN/BCAR1/p130Cas-transfected cellline (4A12, B4, B6, and C4), the somatic cell hybrid (D4E6), andvector-only control cell lines (C2, C6, D2, and D4) were cultured inthe presence of OH-tamoxifen (OH-Tam) (1 μM). Atindicated time points, three flasks were harvested, processed for cellcounting, and reseeded. The number of cells recovered for C2 onlyallowed for reseeding in duplicate for the last time point. Indicatedis the fold multiplication (mean of at least two flasks ± standarddeviation) relative to day 0 (see Fig. 1). Rightpanel—A typical western blot of cell lysates (3.3 μg protein)separated on sodium dodecyl sulfate-polyacrylamide gel electrophoresis(SDS-PAGE) (8%), transferred to nitrocellulose filters, andvisualized with mouse monoclonal antibody to rat p130Cas. The blotcontained cell lysates isolated from these cell lines at day 12/13. An arrow indicates the direction of migration in SDS-PAGE gels,and an arrowhead marks the position of Bcar1/p130Cas. Similarexpression levels were observed on day 0 and on day 5/6. B) RNAwas prepared from the indicated cell lines after culture in bovine calfserum-containing medium (BCS) supplemented with 1 nM ofestradiol (E2) or 1 μM of OH-Tam for 5 days.Approximately 5 μg of RNA was blotted and hybridized with probesfor glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and PS2. Blotswere washed until 0.3 × standard saline citrate at 65 °Cfor 30 minutes.

Open in new tabDownload slide

BCAR1, a Human hom*ologue of the Adapter Protein p130Cas, and AntiestrogenResistance in Breast Cancer Cells (7)

Fig. 4.

Expression of BCAR1 in various human tissues.Commercially available northern blots containing polyA messenger RNA (2μg) of multiple human tissues (MTN blots; Clontech Laboratories,Inc, Palo Alto, CA) and checked for loading by the supplier werehybridized with a labeled BCAR1 total complementary DNA, washed until0.3 × standard saline citrate at 65 °C for 30 minutes,and exposed to an x-ray film for 4 days at−70 °C with an intensifying screen. Size markers areindicated.

Open in new tabDownload slide

Supported by grant DDHK 94-847 from the Dutch Cancer Foundation.

We thank Drs. M. Look, J. Veldscholte, and M. Schutte for critically reviewing thismanuscript. We also thank Drs. J. A. Foekens and T. van Agthoven for their stimulatingdiscussions, Dr. K. Nooter and K. van Wingerden for their contribution to the testing of drugs,Dr. J. G. Collard for the gift of the episomal expression vector, and Dr. A. M. Cleton-Jansen forher assistance with the analysis of the somatic cell hybrid panel.

(1)

Henderson BE, Ross R, Bernstein L. Estrogens as a cause ofhuman cancer: the Richard and Hinda Rosenthal Foundation Award Lecture.

Cancer Res

1988

;

48

:

246

-53.

(2)

King RJ. William L. McGuire Memorial Symposium. Estrogenand progestin effects in human breast carcinogenesis.

Breast Cancer Res Treat

1993

;

27

:

3

-15.

(3)

Evans RM. The steroid and thyroid hormone receptorsuperfamily.

Science

1988

;

240

:

889

-95.

(4)

Truss M, Beato M. Steroid hormone receptors: interaction withdeoxyribonucleic acid and transcription factors.

Endocr Rev

1993

;

14

:

459

-79.

(5)

Mangelsdorf DJ, Thummel C, Beato M, Herrlich P, Schutz G,Umesono K, et al. The nuclear receptor superfamily: the second decade.

Cell

1995

;

83

:

835

-9.

(6)

Beato M, Sanchez-Pacheco A. Interaction of steroid hormonereceptors with the transcription initiation complex.

Endocr Rev

1996

;

17

:

587

-609.

(7)

Horwitz KB, Wei LL, Sedlacek SM, D'Arville CN.Progestin action and progesterone receptor structure in human breast cancer: a review.

Recent Prog Horm Res

1985

;

41

:

249

-316.

(8)

Foekens JA, Rio MC, Seguin P, Van Putten WL, Fauque J, NapM, et al. Prediction of relapse and survival in breast cancer patients by pS2 protein status.

Cancer Res

1990

;

50

:

3832

-7.

(9)

Clarke RB, Howell A, Potten CS, Anderson E. Dissociationbetween steroid receptor expression and cell proliferation in the human breast.

Cancer Res

1997

;

57

:

4987

-91.

(10)

Jordan VC. Molecular mechanisms of antiestrogen action inbreast cancer.

Breast Cancer Res Treat

1994

;

31

:

41

-52.

(12)

Katzenellenbogen BS, Montano MM, Ekena K, Herman ME,McInerney EM. William L. McGuire Memorial Lecture. Antiestrogens: mechanisms of action andresistance in breast cancer.

Breast Cancer Res Treat

1997

;

44

:

23

-38.

(13)

Brzozowski AM, Pike AC, Dauter Z, Hubbard RE, Bonn T,Engstrom O, et al. Molecular basis of agonism and antagonism in the oestrogen receptor.

Nature

1997

;

389

:

753

-8.

(14)

Reese JC, Katzenellenbogen BS. Examination of theDNA-binding ability of ER in whole cells: implications for hormone-independent transactivationand the actions of antiestrogens.

Mol Cell Biol

1992

;

12

:

4531

-8.

(15)

Parker MG. Action of “pure” antiestrogens ininhibiting ER action.

Breast Cancer Res Treat

1993

;

26

:

131

-7.

(16)

Sutherland RL, Lee CS, Feldman RS, Musgrove EA. Regulationof breast cancer cell cycle progression by growth factors, steroids and steroid antagonists.

JSter Biochem Mol Biol

1992

;

41

:

315

-21.

(17)

Watts CK, Brady A, Sarcevic B, DeFazio A, Musgrove EA,Sutherland RL. Antiestrogen inhibition of cell cycle progression in breast cancer cells is associatedwith inhibition of cyclin-dependent kinase activity and decreased retinoblastoma proteinphosphorylation.

Mol Endocrinol

1995

;

9

:

1804

-13.

(18)

Guvakova MA, Surmacz E. Tamoxifen interferes with theinsulin-like growth factor I receptor (IGF-IR) signaling pathway in breast cancer cells.

Cancer Res

1997

;

57

:

2606

-10.

(19)

Jordan VC, Murphy CS. Endocrine pharmacology ofantiestrogens as antitumor agents.

Endocr Rev

1990

;

11

:

578

-610.

(20)

Musgrove EA, Hamilton JA, Lee CS, Sweeney KJ, Watts CK,Sutherland RL. Growth factor, steroid, and steroid antagonist regulation of cyclin gene expressionassociated with changes in T-47D human breast cancer cell cycle progression.

Mol CellBiol

1993

;

13

:

3577

-87.

(21)

Horwitz KB, McGuire WL, Pearson OH, Segaloff A. Predictingresponse to endocrine therapy in human breast cancer: a hypothesis.

Science

1975

;

189

:

726

-7.

(22)

Pritchard KI, Sutherland DJ. The use of endocrine therapy.

Hematol Oncol Clin North Am

1989

;

3

:

765

-805.

(23)

Santen RJ, Manni A, Harvey H, Redmond C. Endocrinetreatment of breast cancer in women.

Endocr Rev

1990

;

11

:

221

-65.

(24)

Early Breast Cancer Trialists' Collaborative Group.Systemic treatment of early breast cancer by hormonal, cytotoxic, or immune therapy. 133Randomised trials involving 31,000 recurrences and 24,000 deaths among 75,000 women.

Lancet

1992

;

339

:

1

-15;71-85.

(25)

Early Breast Cancer Trialists' Collaborative Group.Tamoxifen for early breast cancer: an overview of the randomised trials.

Lancet

1998

;

351

:

1451

-67.

(26)

Jaiyesimi IA, Buzdar AU, Decker DA, Hortobagyi GN. Use oftamoxifen for breast cancer: twenty-eight years later.

J Clin Oncol

1995

;

13

:

513

-29.

(27)

Foekens JA, Schmitt M, Van Putten WL, Peters HA, KramerMD, Janicke F, et al. Plasminogen activator inhibitor-1 and prognosis in primary breast cancer.

J Clin Oncol

1994

;

12

:

1648

-58.

(28)

Osborne CK, Coronado E, Allred DC, Wiebe V, DeGregorio M.Acquired tamoxifen resistance: correlation with reduced breast tumor levels of tamoxifen andisomerization of trans-4-hydroxytamoxifen.

J Natl Cancer Inst

1991

;

83

:

1477

-82.

(29)

Johnston SR, Haynes BP, Smith IE, Jarman M, Sacks NP, EbbsSR, et al. Acquired tamoxifen resistance in human breast cancer and reduced intra-tumoral drugconcentration.

Lancet

1993

;

342

:

1521

-2.

(30)

Daffada AA, Johnston SR, Smith IE, Detre S, King N, DowsettM. Exon 5 deletion variant ER messenger RNA expression in relation to tamoxifen resistance andprogesterone receptor/pS2 status in human breast cancer.

Cancer Res

1995

;

55

:

288

-93.

(31)

Fuqua SA, Wolf DM. Molecular aspects of estrogen receptorvariants in breast cancer.

Breast Cancer Res Treat

1995

;

35

:

233

-41.

(32)

Murphy LC, Hilsenbeck SG, Dotzlaw H, Fuqua SA.Relationship of clone 4 estrogen receptor variant messenger RNA expression to some knownprognostic variables in human breast cancer.

Clin Cancer Res

1995

;

1

:

155

-9.

(33)

Clarke R, Dickson RB, Lippman ME. Hormonal aspects ofbreast cancer. Growth factors, drugs and stromal interactions.

Crit Rev Oncol Hematol

1992

;

12

:

1

-23.

(34)

Dorssers LC, Van Agthoven T, Sieuwerts AM. Geneticmechanisms involved in progression to hormone independence of human breast cancer. In: BernsPM, Romijn JC, Schroeder FH, editors. Mechanisms of progression to hormone-independentgrowth of breast and prostatic cancer. Carnforth (U.K.): Parthenon Publishing Group Ltd; 1991.p. 169-82.

(35)

Clarke R, Thompson EW, Leonessa F, Lippman J, McGarveyM, Frandsen TL, et al. Hormone resistance, invasiveness, and metastatic potential in breastcancer.

Breast Cancer Res Treat

1993

;

24

:

227

-39.

(36)

Dorssers LC, Van Agthoven T, Dekker A, Van Agthoven TL,Kok EM. Induction of antiestrogen resistance in human breast cancer cells by random insertionalmutagenesis using defective retroviruses: identification of bcar-1, a common integration site.

Mol Endocrinol

1993

;

7

:

870

-8.

(37)

Dorssers LC, Veldscholte J. Identification of a novelbreast-cancer-anti-estrogen-resistance (BCAR2) locus by cell-fusion-mediated gene transfer inhuman breast-cancer cells.

Int J Cancer

1997

;

72

:

700

-5.

(38)

Van Agthoven T, Van Agthoven TL, Portengen H, Foekens JA,Dorssers LC. Ectopic expression of epidermal growth factor receptors induces hormoneindependence in ZR-75-1 human breast cancer cells.

Cancer Res

1992

;

52

:

5082

-8.

(39)

Miller AD, Rosman GJ. Improved retroviral vectors for genetransfer and expression.

BioTechniques

1989

;

7

:

980

-90.

(40)

Van den Ouweland AM, Breuer ML, Steenbergh PH, SchalkenJA, Bloemers HP, Van de Ven WJ. Comparative analysis of the human and feline c-sisproto-oncogenes: identification of 5′ human c-sis coding sequences that are nothom*ologous to the transforming gene of simian sarcoma virus.

Biochim Biophys Acta

1985

;

825

:

140

-7.

(41)

Sambrook J, Fritsch EF, Maniatis T. Molecular cloning: alaboratory manual. New York (NY): Cold Spring Harbor Press; 1989.

(42)

Van Agthoven T, Van Agthoven TL, Dekker A, Van der SpekPJ, Vreede L, Dorssers LC. Identification of BCAR3 by a random search for genes involved inantiestrogen resistance of human breast cancer cells.

EMBO J

1998

;

17

:

2799

-808.

(43)

Van Agthoven T, Van Agthoven TL, Dekker A, Foekens JA,Dorssers LC. Induction of estrogen independence of ZR-75-1 human breast cancer cells byepigenetic alterations.

Mol Endocrinol

1994

;

8

:

1474

-83.

(44)

Sander EE, van Delft S, ten Klooster JP, Reid T, van derKammen RA, Michiels F, et al. Matrix-dependent Tiam1/Rac signaling in epithelial cells promoteseither cell-cell adhesion or cell migration and is regulated by phosphatidylinositol 3-kinase.

J Cell Biol

1998

;

143

:

1385

-98.

(45)

van der Flier S, Brinkman A, Look MP, Kok EM, Meijer-vanGelder ME, Klijn GM, et al. Bcar1/p130Cas protein and primary breast cancer: prognosis andresponse to tamoxifen treatment.

J Natl Cancer Inst

2000

;

92

:

000

-000.

(46)

Sakai R, Iwamatsu A, Hirano N, Ogawa S, Tanaka T, Mano H,et al. A novel signaling molecule, p130, forms stable complexes in vivo with v-Crk andv-Src in a tyrosine phosphorylation-dependent manner.

EMBO J

1994

;

13

:

3748

-56.

(47)

Kozak M. An analysis of vertebrate mRNA sequences:intimations of translational control.

J Cell Biol

1991

;

115

:

887

-903.

(48)

Nakamoto T, Sakai R, Ozawa K, Yazaki Y, Hirai H. Directbinding of C-terminal region of p130Cas to SH2 and SH3 domains of Src kinase.

J Biol Chem

1996

;

271

:

8959

-65.

(49)

Polte TR, Hanks SK. Interaction between focal adhesion kinaseand Crk-associated tyrosine kinase substrate p130Cas.

Proc Natl Acad Sci US A

1995

;

92

:

10678

-82.

(50)

Callen DF, Lane SA, Kozman H, Kremmidiotis G, WhitmoreSA, Lowenstein M, et al. Integration of transcript and genetic maps of chromosome 16 at near-1-Mb resolution: demonstration of a “hot spot” for recombination at 16p12.

Genomics

1995

;

29

:

503

-11.

(51)

Schlaepfer DD, Broome MA, Hunter T. Fibronectin-stimulatedsignaling from a focal adhesion kinase-c-Src complex: involvement of the Grb2, p130cas, and Nck adaptor proteins.

Mol Cell Biol

1997

;

17

:

1702

-13.

(52)

Liu F, Sells MA, Chernoff J. Transformation suppression byprotein tyrosine phosphatase 1B requires a functional SH3 ligand.

Mol Cell Biol

1998

;

18

:

250

-9.

(53)

Blaukat A, Ivankovic-Dikic I, Gronroos E, Dolfi F, Tokiwa G,Vuori K, et al. Adaptor proteins grb2 and crk couple pyk2 with activation of specificmitogen-activated protein kinase cascades.

J Biol Chem

1999

;

274

:

14893

-901.

(54)

Lu Y, Brush J, Stewart TA. NSP1 defines a novel family ofadaptor proteins linking integrin and tyrosine kinase receptors to the c-Jun N-terminalkinase/stress-activated protein kinase signaling pathway.

J Biol Chem

1999

;

274

:

10047

-52.

(55)

Jonkers J, Berns A. Retroviral insertional mutagenesis as astrategy to identify cancer genes.

Biochim Biophys Acta Rev Cancer

1996

;

1287

:

29

-57.

(56)

Sakai R, Iwamatsu A, Hirano N, Ogawa S, Tanaka T, Nishida J,et al. Characterization, partial purification, and peptide sequencing of p130, the mainphosphoprotein associated with v-Crk oncoprotein.

J Biol Chem

1994

;

269

:

32740

-6.

(57)

Auvinen M, Paasinen-Sohns A, Hirai H, Andersson LC, HolttaE. Ornithine decarboxylase- and ras-induced cell transformations: reversal by protein tyrosinekinase inhibitors and role of pp130CAS.

Mol Cell Biol

1995

;

15

:

6513

-25.

(58)

Honda H, Oda H, Nakamoto T, Honda Z, Sakai R, Suzuki T, etal. Cardiovascular anomaly, impaired actin bundling and resistance to Src- induced transformationin mice lacking p130Cas.

Nat Genet

1998

;

19

:

361

-5.

(59)

Harte MT, Hildebrand JD, Burnham MR, Bouton AH, ParsonsJT. p130Cas, a substrate associated with v-Src and v-Crk, localizes to focaladhesions and binds to focal adhesion kinase.

J Biol Chem

1996

;

271

:

13649

-55.

(60)

Nakamoto T, Sakai R, Honda H, Ogawa S, Ueno H, Suzuki T,et al. Requirements for localization of p130CAS to focal adhesions.

Mol CellBiol

1997

;

17

:

3884

-97.

(61)

Nojima Y, Morino N, Mimura T, Hamasaki K, Furuya H, SakaiR, et al. Integrin-mediated cell adhesion promotes tyrosine phosphorylation of p130Cas, a Src hom*ology 3-containing molecule having multiple Src hom*ology 2-binding motifs.

J Biol Chem

1995

;

270

:

15398

-402.

(62)

Vuori K, Hirai H, Aizawa S, Ruoslahti E. Induction of p130cas signaling complex formation upon integrin-mediated cell adhesion: a role for Srcfamily kinases.

Mol Cell Biol

1996

;

16

:

2606

-13.

(63)

Howe A, Aplin AE, Alahari SK, Juliano RL. Integrin signalingand cell growth control.

Curr Opin Cell Biol

1998

;

10

:

220

-31.

(64)

Schlaepfer DD, Hunter T. Integrin signalling and tyrosinephosphorylation: just the FAKs?

Trends Cell Biol

1998

;

8

:

151

-7.

(65)

Ojaniemi M, Vuori K. Epidermal growth factor modulatestyrosine phosphorylation of p130Cas. Involvement of phosphatidylinositol3′-kinase and actin cytoskeleton.

J Biol Chem

1997

;

272

:

25993

-8.

(66)

Rozengurt EV. Gastrointestinal peptide signaling throughtyrosine phosphorylation of focal adhesion proteins.

Am J Physiol

1998

;

275

:

G177

-82.

(67)

Sorokin A, Reed E. Insulin stimulates the tyrosinedephosphorylation of docking protein p130cas (Crk-associated substrate), promoting the switchof the adaptor protein crk from p130cas to newly phosphorylated insulin receptor substrate-1.

Biochem J

1998

;

334

:

595

-600.

(68)

Murakami H, Iwash*ta T, Asai N, Iwata Y, Narumiya S,Takahashi M. Rho-dependent and -independent tyrosine phosphorylation of focal adhesion kinase,paxillin and p130Cas mediated by Ret kinase.

Oncogene

1999

;

18

:

1975

-82.

(69)

Manie SN, Beck AR, Astier A, Law SF, Canty T, Hirai H, et al.Involvement of p130Cas and p105HEF1, a novel Cas-like dockingprotein, in a cytoskeleton-dependent signaling pathway initiated by ligation of integrin or antigenreceptor on human B cells.

J Biol Chem

1997

;

272

:

4230

-6.

(70)

Cary LA, Han DC, Polte TR, Hanks SK, Guan JL. Identificationof p130Cas as a mediator of focal adhesion kinase-promoted cell migration.

JCell Biol

1998

;

140

:

211

-21.

(71)

Klemke RL, Leng J, Molander R, Brooks PC, Vuori K, ChereshDA. CAS/Crk coupling serves as a “molecular switch” for induction of cellmigration.

J Cell Biol

1998

;

140

:

961

-72.

(72)

Tamura M, Gu J, Takino T, Yamada KM. Tumor suppressorPTEN inhibition of cell invasion, migration, and growth: differential involvement of focal adhesionkinase and p130Cas.

Cancer Res

1999

;

59

:

442

-9.

(73)

Black DS, Bliska JB. Identification of p130Cas as asubstrate of Yersinia YopH (Yop51), a bacterial protein tyrosine phosphatase thattranslocates into mammalian cells and targets focal adhesions.

EMBO J

1997

;

16

:

2730

-44.

(74)

Persson C, Carballeira N, Wolf-Watz H, Fallman M. The PTPaseYopH inhibits uptake of Yersinia, tyrosine phosphorylation of p130Cas andFAK, and the associated accumulation of these proteins in peripheral focal adhesions.

EMBO J

1997

;

16

:

2307

-18.

(75)

Pawson T, Scott JD. Signaling through scaffold, anchoring, andadaptor proteins.

Science

1997

;

278

:

2075

-80.

(76)

Law SF, Estojak J, Wang BL, Mysliwiec T, Kruh G, GolemisEA. Human enhancer of filamentation 1, a novel p130cas-like dockingprotein, associates with focal adhesion kinase and induces pseudohyphal growth in Saccharomyces cerevisiae.

Mol Cell Biol

1996

;

16

:

3327

-37.

(77)

Minegishi M, Tachibana K, Sato T, Iwata S, Nojima Y,Morimoto C. Structure and function of Cas-L, a 105-kD Crk-associated substrate-related proteinthat is involved in β1 integrin-mediated signaling in lymphocytes.

J Exp Med

1996

;

184

:

1365

-75.

(78)

Ishino M, Ohba T, Sasaki H, Sasaki T. Molecular cloning of acDNA encoding a phosphoprotein, Efs, which contains a Src hom*ology 3 domain and associateswith Fyn.

Oncogene

1995

;

11

:

2331

-8.

(79)

Alexandropoulos K, Baltimore D. Coordinate activation of c-Srcby SH3- and SH2-binding sites on a novel, p130Cas-related protein, Sin.

Genes Dev

1996

;

10

:

1341

-55.

(80)

Law SF, Zhang YZ, Klein-Szanto AJ, Golemis EA. Cellcycle-regulated processing of HEF1 to multiple protein forms differentially targeted to multiplesubcellular compartments.

Mol Cell Biol

1998

;

18

:

3540

-51.

(81)

Astier A, Avraham H, Manie SN, Groopman J, Canty T,Avraham S, et al. The related adhesion focal tyrosine kinase is tyrosine-phosphorylated afterβ1-integrin stimulation in B cells and binds to p130cas.

J Biol Chem

1997

;

272

:

228

-32.

(82)

Garton AJ, Burnham MR, Bouton AH, Tonks NK. Associationof PTP-PEST with the SH3 domain of p130cas; a novel mechanism of proteintyrosine phosphatase substrate recognition.

Oncogene

1997

;

15

:

877

-85.

(83)

Ohba T, Ishino M, Aoto H, Sasaki T. Interaction of twoproline-rich sequences of cell adhesion kinase beta with SH3 domains of p130Cas-related proteinsand a GTPase-activating protein, Graf.

Biochem J

1998

;

330

:

1249

-54.

(84)

Burnham MR, Harte MT, Richardson A, Parsons JT, BoutonAH. The identification of p130cas-binding proteins and their role in cellulartransformation.

Oncogene

1996

;

12

:

2467

-72.

(85)

Khwaja A, Hallberg B, Warne PH, Downward J. Networks ofinteraction of p120cbl and p130cas with Crk and Grb2 adaptorproteins.

Oncogene

1996

;

12

:

2491

-8.

Oxford University Press

Topic:

  • phenotype
  • cell proliferation
  • estrogen
  • cell fusion procedure
  • cell lines
  • dna, complementary
  • estrogen antagonists
  • estrogen receptor modulators
  • exons
  • genes
  • genome
  • libraries
  • retroviridae
  • tamoxifen
  • transfection
  • rats
  • breast cancer
  • breast cancer cells
  • transfer technique

Issue Section:

Articles

Download all slides

Advertisem*nt

Citations

Views

1,594

Altmetric

More metrics information

Metrics

Total Views 1,594

1,220 Pageviews

374 PDF Downloads

Since 2/1/2017

Month: Total Views:
February 2017 15
March 2017 1
April 2017 2
May 2017 7
June 2017 2
August 2017 3
September 2017 2
October 2017 3
November 2017 5
December 2017 18
January 2018 25
February 2018 17
March 2018 24
April 2018 65
May 2018 121
June 2018 24
July 2018 31
August 2018 29
September 2018 18
October 2018 14
November 2018 17
December 2018 25
January 2019 19
February 2019 21
March 2019 23
April 2019 42
May 2019 37
June 2019 39
July 2019 38
August 2019 21
September 2019 19
October 2019 10
November 2019 13
December 2019 5
January 2020 12
February 2020 15
March 2020 18
April 2020 21
May 2020 17
June 2020 21
July 2020 26
August 2020 17
September 2020 12
October 2020 16
November 2020 15
December 2020 11
January 2021 15
February 2021 13
March 2021 23
April 2021 14
May 2021 11
June 2021 10
July 2021 10
August 2021 3
September 2021 23
October 2021 26
November 2021 23
December 2021 18
January 2022 14
February 2022 16
March 2022 13
April 2022 20
May 2022 10
June 2022 9
July 2022 22
August 2022 17
September 2022 28
October 2022 8
November 2022 8
December 2022 17
January 2023 13
February 2023 6
March 2023 14
April 2023 14
May 2023 8
June 2023 13
July 2023 3
August 2023 15
September 2023 9
October 2023 13
November 2023 21
December 2023 28
January 2024 18
February 2024 21
March 2024 25
April 2024 18
May 2024 21
June 2024 7

Citations

Powered by Dimensions

124 Web of Science

Altmetrics

×

Email alerts

Article activity alert

Advance article alerts

New issue alert

Receive exclusive offers and updates from Oxford Academic

Citing articles via

Google Scholar

  • Latest

  • Most Read

  • Most Cited

The state of Cancer-Focused community outreach and engagement (COE): reflections of black COE directors
RE: Plea to stop calling for new clinical trials to test metformin in cancer
Resource requirements to accelerate clinical applications of next generation sequencing and radiomics: Workshop commentary and review
RET overexpression leads to increased brain metastatic competency in luminal breast cancer
RE: Spatial intratumor heterogeneity of programmed death-ligand 1 expression predicts poor prognosis in resected non-small cell lung cancer

More from Oxford Academic

Medicine and Health

Books

Journals

Advertisem*nt

BCAR1, a Human hom*ologue of the Adapter Protein p130Cas, and Antiestrogen
Resistance in Breast Cancer Cells (2024)
Top Articles
Latest Posts
Article information

Author: Saturnina Altenwerth DVM

Last Updated:

Views: 6178

Rating: 4.3 / 5 (44 voted)

Reviews: 83% of readers found this page helpful

Author information

Name: Saturnina Altenwerth DVM

Birthday: 1992-08-21

Address: Apt. 237 662 Haag Mills, East Verenaport, MO 57071-5493

Phone: +331850833384

Job: District Real-Estate Architect

Hobby: Skateboarding, Taxidermy, Air sports, Painting, Knife making, Letterboxing, Inline skating

Introduction: My name is Saturnina Altenwerth DVM, I am a witty, perfect, combative, beautiful, determined, fancy, determined person who loves writing and wants to share my knowledge and understanding with you.