Intracrine PTHrP Protects against Serum Starvation- Induced ...

7 downloads 0 Views 346KB Size Report
promotion of transplacental calcium transport, and cartilage development (9, 17–21). The peptide is ...... Carcinogenesis 18:23–29. 33. Falzon M, Du P 2000 ...
0013-7227/02/$15.00/0 Printed in U.S.A.

Endocrinology 143(2):596 – 606 Copyright © 2002 by The Endocrine Society

Intracrine PTHrP Protects against Serum StarvationInduced Apoptosis and Regulates the Cell Cycle in MCF-7 Breast Cancer Cells VERONICA A. TOVAR SEPULVEDA, XIAOLI SHEN,

AND

MIRIAM FALZON

Department of Pharmacology and Toxicology and Sealy Center for Molecular Science, University of Texas Medical Branch, Galveston, Texas 77555 PTHrP is secreted by breast cancer cells in vivo and in vitro. In the breast cancer cell line MCF-7, PTHrP overexpression is associated with increased mitogenesis. We used this cell line to study the mechanism for the proliferative effects of PTHrP. Clonal MCF-7 lines were established overexpressing wildtype PTHrP or PTHrP mutated in the nuclear localization signal (NLS). Mutation of the NLS negated the proliferative effects and nuclear trafficking of PTHrP, indicating that increased mitogenesis is mediated via an intracrine pathway. Cells overexpressing wild-type PTHrP were enriched in G2 ⴙ M stage of the cell cycle compared with cells overexpressing NLS-mutated PTHrP, indicating an intracrine role for PTHrP

in cell cycle regulation. Wild-type PTHrP also protected MCF-7 cells from serum starvation-induced apoptosis. Cells overexpressing wild-type PTHrP showed significantly greater cell survival than cells overexpressing NLS-mutated PTHrP. The ratios of the apoptosis-regulating proteins Bcl-2 to Bax and Bcl-xL to Bax were higher in cells overexpressing wild-type, but not NLS-mutated, PTHrP compared with control cells. These findings suggest that the proliferative effects of PTHrP in breast cancer cells are mediated through regulation of the cell cycle and apoptosis, and that controlling PTHrP production in breast cancer may be therapeutically useful. (Endocrinology 143: 596 – 606, 2002)

P

THrP WAS INITIALLY identified through its role in humoral hypercalcemia of malignancy, one of the most frequent paraneoplastic syndromes and a very common complication in breast cancer patients (1– 6). More recently, the protein was found to be widely distributed in most fetal and adult tissues (7–9). However, PTHrP is not normally present in the circulation (10), suggesting that it may act in an autocrine/paracrine manner. The mature peptide can undergo extensive posttranslational processing to generate secretory forms of PTHrP representing N-terminal, midregion, and C-terminal portions of the molecule (11, 12). Each region exhibits unique biological activities and presumably acts through its own cognate receptor (13, 14). However, only the PTH/PTHrP (PTH 1) receptor that binds PTH, PTHrP, and their N-terminal analogs has been cloned to date (15, 16). Physiological roles attributed to PTHrP include regulation of cell growth and differentiation, smooth muscle relaxation, promotion of transplacental calcium transport, and cartilage development (9, 17–21). The peptide is secreted in a regulated or constitutive fashion depending on the cell type (12). In addition to effects that are mediated via signal transduction cascades initiated at membrane receptors, PTHrP can also function in an intracrine manner after translocation to the nucleus or nucleolus (22–24). The PTHrP molecule contains a midregion nuclear localization sequence (NLS) comprising multibasic clusters in the 88 –106 region, which is similar to nuclear localization signals found in viral and mammalian transcription factors (22–25). Therefore, PTHrP

may function locally in an autocrine/paracrine or intracrine fashion. PTHrP is closely linked to normal mammary gland function (26), with enhanced expression during lactation when the mammary gland is in a proliferative state (27). Studies in transgenic mice have shown the involvement of the protein in branching morphogenesis of the mammary gland, indicating active participation in normal mammary development (28). Conversely, the absence of PTHrP in knockout mice leads to mammary epithelial degeneration (29). Most breast cancer cells secrete higher levels of PTHrP than do normal breast cells (1, 30). In fact, it has been proposed that PTHrP production by breast cancer cells may be one of the key elements instrumental in supporting carcinogenesis (31, 32). Therefore, the effects of PTHrP in breast cancer cells have both biological and clinical significance. PTHrP exerts a mitogenic effect in the breast cancer cell line MCF-7 (33). Here, we report that the increase in cell number is mediated via an intracrine pathway through both cell cycle regulation and inhibition of apoptosis through the Bcl-2 family of proteins. These findings suggest that endogenously produced PTHrP may enhance breast cancer growth via these pathways. Materials and Methods Materials Synthetic human (h) PTHrP-(1–34), hPTHrP-(1– 86), and hPTHrP(107–139) were purchased from Bachem (Torrance, CA). Recombinant hPTHrP-(1–139) was prepared using the IMPACT (intein-mediated purification with an affinity chitin-binding tag) method (New England Biolabs, Inc., Beverly, MA) (34). Peptide stocks (10⫺4 m) were prepared in 10 mm acetic acid. FBS and NuSerum were obtained from Atlanta Biologicals (Norcross, GA) and Collaborative Research (Bedford, MA),

Abbreviations: 7-AAD, 7-Aminoactinomycin D; FITC, fluorescein isothiocyanate; hPTH, human PTH; NCS, newborn calf serum; NLS, nuclear localization signal; PerCP, peridinin chlorophyll protein.

596

Sepulveda et al. • PTHrP and Breast Cancer Cell Proliferation

respectively. Tissue culture supplies were purchased from Life Technologies, Inc. (Gaithersburg, MD). FuGENE 6 and annexin V were obtained from Roche Molecular Biochemicals (Indianapolis, IN). 7Aminoactinomycin D (7-AAD) and valinomycin were obtained from Sigma (St. Louis, MO). Anti-Bax (mouse IgG2b, clone B-9) antibody, anti-Bcl-xL (rabbit IgG, clone H-62) antibody, and isotype normal antirabbit IgG antibody were obtained from Santa Cruz Biotechnology, Inc. (Santa Cruz, CA). Fluorescein isothiocyanate (FITC)-labeled anti-Bcl-2 (hamster IgG, clone 6C8) antibody, FITC-labeled hamster IgG antibody, biotin-labeled antirabbit IgG, and peridinin chlorophyll protein (PerCP)labeled streptavidin were obtained from PharMingen (San Diego, CA). Phycoerythrin-labeled goat antimouse IgG antibody, FITC-labeled goat antimouse IgG, and isotype mouse IgG2b antibody were obtained from Caltag Laboratories, Inc. (Burlingame, CA). Anti-PTHrP mouse monoclonal antibody and the FITC-labeled goat antimouse antibody (Alexa 488) were purchased from Oncogene Research Products (Cambridge, MA) and Molecular Probes, Inc. (Eugene, OR), respectively.

Plasmid constructs The PTHrP construct expressing wild-type PTHrP was constructed by cloning PTHrP cDNA coding for amino acids ⫺5 to ⫹139 (a gift from Dr. W. I. Wood, Genentech, Inc., South San Francisco, CA) in the sense orientation into the expression vector pcDNA3.1⫹ (Invitrogen, San Diego, CA; ⫹ refers to the orientation of the multiple cloning site within the vector, relative to the direction of transcription from the T7 promoter). This construct was used to prepare PTHrP cDNAs mutated over the NLS using a Transformer Site-Directed Mutagenesis Kit (CLONTECH Laboratories, Inc., Palo Alto, CA). The cDNAs encoded either a single deletion (elimination of residues 88 –91 or 102–106) or a double deletion (elimination of residues 88 –91 and 102–106). Mutations were confirmed by DNA sequencing. These constructs as well as the empty vector control pcDNA 3.1⫹ were transfected into MCF-7 cells using FuGENE 6 transfection reagent (Roche Molecular Biochemicals). The DNA template used to prepare the probe for Northern blot analysis was a 231-bp cloned DNA fragment spanning exons 3 and 4 of the PTHrP gene (35, 36).

Cell culture and stable transfection MCF-7 cells were obtained from American Type Culture Collection (Manassas, VA) and grown at 37 C in a humidified 95% O2-5% CO2 atmosphere in DMEM supplemented with 10% FBS and l-glutamine. MCF-7 cells were stably transfected using FuGENE 6 transfection reagent, according to the manufacturer’s specifications (Roche Molecular Biochemicals). Two days after transfection, 600 ␮g/ml G418 (Geneticin, Life Technologies, Inc.) were added, and resistant clones were selected. Single clones of stably transfected cells, isolated by limiting dilution in 96-well plates, were transferred to individual flasks and cultured in medium containing 150 ␮g/ml G418. Individual clones were tested for PTHrP production using an immunoradiometric assay (described below), and transfected mRNAs were detected by Northern blot analysis (also described below). To measure the effects of added peptides on cell cycle regulation and apoptosis, MCF-7 cells were plated in 100-mm dishes in the presence of 10% FBS. To measure apoptosis, cells were transferred to 0.1% NuSerum after 24 h. When they had reached about 30% confluence, the cells were treated with 10⫺8 or 10⫺7 m of the indicated hPTHrP peptides for 3 or 5 d. Control cells received an equivalent volume of 10 mm acetic acid as a vehicle control. When the cells were treated for 5 d, the growth medium was replaced with fresh peptide-containing medium after 3 d. The cells were then analyzed by FACS for their cell cycle or annexin V profile.

Northern blot analysis Total RNA was isolated using RNA STAT-60 (Tel-Test B, Friendswood, TX). RNA gel electrophoresis was performed under standard conditions (37), using 25 ␮g total RNA. The RNA was then blotted onto nitrocellulose (Schleicher & Schuell, Inc., Keene, NH) by capillary action. Probes for hybridization were prepared by the asymmetric PCR (38). The PCR template was a 231-bp cloned DNA fragment spanning exons 3 and 4 of the hPTHrP gene (35, 36). To detect RNA produced by the transfected sense PTHrP, an antisense probe was prepared using the down-

Endocrinology, February 2002, 143(2):596 – 606 597

stream primer from exon 4 of the human PTHrP gene (5⬘-GTTAGGGGACACCTCCGAGGT-3⬘). The blots were prehybridized for 30 min and hybridized for 2 h in Expresshyb (CLONTECH Laboratories, Inc.) at 65 C. After hybridization, the blots were washed twice in 2⫻ SSC/0.05% SDS for 15 min at room temperature and then twice in 0.1⫻ SSC/0.1% SDS at 50 C for 30 min. The washed membranes were exposed to Kodak X-Omat film (Eastman Kodak Co., Rochester, NY) at ⫺70 C with intensifying screens. The ethidium bromide-stained RNA on the nitrocellulose membrane was photographed immediately after transfer. 18S ribosomal RNA signal was used to provide a reference to normalize for equal RNA loading and transfer. The intensities of the bands representing PTHrP and 18S ribosomal RNA were evaluated using the Sigmagel program (Jandel Scientific, San Rafael, CA).

Immunoassay for secreted PTHrP The amount of PTHrP secreted into the culture medium was measured using an immunoradiometric sandwich assay (Nichols Institute Diagnostics, San Juan Capistrano, CA), as previously described (35). Single clones of transfected and control (untransfected) cells were plated in 24-well dishes (5 ⫻ 104 cells/well), and the medium was replaced after 24 h. Conditioned medium was collected after a further 4 d and frozen at ⫺80 C for future use; cell number was determined using a Coulter counter (Hialeah, FL). Unconditioned medium (never exposed to cells) served as a negative control. The assay was carried out according to the manufacturer’s specifications. The detection limit of the assay is 0.7 pmol/liter (39).

Determination of cell number To measure the effects of overexpression of wild-type and NLSmutated PTHrP on cell number, cells were plated into 24-well dishes at 104 cells/well in medium containing 10% FBS (d 0). Cells were then trypsinized and counted on d 1, 2, 4, and 7.

Thymidine incorporation assay For these experiments, cells were plated in 12-well plates at 2 ⫻ 104 cells/well in medium containing 10% FBS. After 3 d, when cells had reached about 70% confluence, they were pulse-treated with [3H]thymidine (0.2 ␮Ci/ml) for 3 h. To determine [3H]thymidine incorporation, the cell monolayer was washed twice with PBS, and nucleic acids in the cell fraction were precipitated with trichloroacetic acid and solubilized with 1 m sodium hydroxide for scintillation counting and protein determination. The protein concentration was measured using the Bio-Rad Laboratories, Inc. assay (Hercules, CA). The results were expressed as counts per ␮g protein.

PTHrP immunofluorescence labeling Cells grown on Lab-Tek chamber slides (Nalge Nunc International, Naperville, IL) were washed with cold PBS, fixed in acetone for 10 min at ⫺20 C, and washed again with PBS. Acetone permeabilizes the cells so they can stain intracellularly. The cells were then incubated in 50 mm ammonium chloride in PBS for 15 min at room temperature and washed with PBS. After blocking nonspecific protein-binding sites by preincubation with 5% milk (Novagen, Madison, WI) for 1 h at room temperature and washing with PBS, the cells were incubated with primary antibody (anti-PTHrP mouse monoclonal IgG, Oncogene Research Products) for 60 min at room temperature. Control cells received no primary antibody. After washing with cold PBS, the cells were treated with goat antimouse secondary antibody (Alexa 488, Molecular Probes) for 1 h at room temperature. The cells were washed again in PBS, mounted using a Prolong Antifade Kit (Molecular Probes), and analyzed using a fluorescence microscope (Olympus Corp., Melville, NY) with a wide-band green filter. The percentage of cells whose nuclei stained positively were counted in a blinded fashion by two observers using computerized histomorphometry (Optimus Corp., Bothell, WA).

Flow cytometry Cell cycle analysis of MCF-7 cells overexpressing wild-type and NLSmutated PTHrP was performed as follows. Cells were plated and grown

598

Endocrinology, February 2002, 143(2):596 – 606

in 100-mm culture dishes in medium containing 10% FBS. When the cells reached about 75% confluence, they were trypsinized and collected by centrifugation. The cell pellets were then suspended in 200 ␮l PBS, and 5 ml ice-cold 75% ethanol were added dropwise to the cell suspension with agitation at 4 C. The cells were collected by centrifugation, and the cell pellets were washed twice with PBS and resuspended in PBS containing 50 ␮g/ml propidium iodide (Sigma) and 50 ␮g/ml deoxyribonuclease-free ribonuclease A (Sigma). The cell suspension was incubated on ice for 1 h in the dark, then analyzed with a FACScan flow cytometer (Becton Dickinson and Co., Franklin Lakes, NJ). Early stage apoptosis was measured using annexin V-Fluos (Roche Molecular Biochemicals) in conjunction with 7-AAD (Sigma) to distinguish among early apoptosis, late apoptosis, and necrosis (40, 41). Cells were plated in T-75 flasks in medium containing 10% FBS and were transferred to medium containing 0.1% NuSerum after 24 h. When the cells had reached about 50% confluence, they were synchronized by treating with aphidicolin (10 ␮g/ml) for 16 h, then cultured in the absence of aphidicolin for a further 24 h. The cells were then trypsinized and counted, a fraction of the cells (0.2 ⫻ 106) was collected by centrifugation, and the pellet was washed twice with PBS. After a further centrifugation, the cell pellet was resuspended in 100 ␮l labeling solution containing 2 ␮l annexin V-FITC labeling reagent in 100 ␮l HEPES buffer [10 mm HEPES/NaOH (pH 7.4), 140 mm NaCl, and 5 mm CaCl2]. For the negative control, the cell pellet was resuspended in HEPES buffer in the absence of annexin V. After a 15-min incubation on ice, the cell suspension was collected by centrifugation and washed twice with PBS containing 1 mm CaCl2. The cell pellet was then resuspended in PBS and 1 mm CaCl2 containing 20 ␮g/ml 7-AAD. Apoptosis was measured by two-color flow cytometry on a FACScan flow cytometer (Becton Dickinson and Co.). Viable cells were electronically gated on forward and side scatter parameters. In addition, viable cells were gated on their ability to exclude the dye 7-AAD (40, 41). The expression of Bcl-2, Bcl-xL, and Bax was measured after aphidicolin synchronization (described above) by staining intracellularly with antibodies to Bcl-2, Bcl-xL, and Bax antibody (42). Briefly, 0.2 ⫻ 106 cells were incubated in 100 ␮l 0.2% saponin buffer [0.2% saponin in PBS containing 2% newborn calf serum (NCS), 5% nonfat dry milk, and 0.02% sodium azide] on ice for 30 min. Primary antibody was then added, and the cells were incubated on ice for a further 30 min. The cells were then washed twice with 1 ml 0.1% saponin buffer (0.1% saponin in PBS containing 2% NCS and 0.02% sodium azide), incubated on ice with secondary antibody for 30 min, and washed twice with 0.1% saponin buffer and once with PBS containing 2% NCS and 0.02% sodium azide. For calculation of the Bcl-xL/Bax ratio, cells were further incubated on ice with PerCP-labeled streptavidin for 30 min, then washed twice with 0.1% saponin buffer. The samples were analyzed on a FACScan flow cytometer (Becton Dickinson and Co.). Viable cells were electronically gated on forward and side scatter parameters. Acquired data were analyzed with CellQuest software (Becton Dickinson and Co.). The ratio of Bcl-2/Bax or Bcl-xL/Bax was calculated as follows: % Bcl-2- or BclxL-positive cells/% Bax-positive cells. Levels of Bcl-2 and Bax were calculated in the same cells by staining with FITC-labeled anti-Bcl-2 (or hamster IgG as isotype control) and anti-Bax (or mouse IgG2b as isotype control) as first antibodies and phycoerythrin-labeled goat antimouse IgG as second antibody. The levels of Bcl-xL and Bax were calculated in the same cells by staining with anti-Bcl-xL (or normal rabbit IgG as isotype control) and anti-Bax (or mouse IgG2b as isotype control) as first antibodies, FITC-labeled goat antimouse IgG and biotin-labeled goat antirabbit Ig as secondary antibodies, and PerCP-labeled streptavidin to bind to biotin.

Nuclear staining assay Cells were grown on Lab-Tek chamber slides (Nalge Nunc International) and at 50% confluence were transferred to medium containing 0.1% NuSerum for 24 h, then treated with the potassium ionophore valinomycin (40 ␮m) for 4 h (43). Control cells were kept in 10% FBS. Apoptosis was induced, and cells were then stained with the DNAbinding fluorescent dye Hoechst 33342 (10 ␮m in PBS) for 10 min. After washing with PBS, the cells were analyzed under a fluorescence microscope (Olympus Corp.) with excitation at UV 360 nm (44).

Sepulveda et al. • PTHrP and Breast Cancer Cell Proliferation

Statistics Numerical data are presented as the mean ⫾ sem. The data were analyzed by ANOVA, followed by a Bonferroni posttest to determine the statistical significance of differences. P ⬍ 0.05 was considered significant. All statistical analysis were performed using Instat Software (GraphPad Software, Inc., San Diego, CA).

Results Establishment and characterization of cell lines overexpressing NLS-mutated PTHrP

We have previously shown that transfection of MCF-7 cells with a sense PTHrP construct increases the proliferation of these cells (33). To determine whether the growth effects of PTHrP in MCF-7 cells require an intact NLS, cells were transfected with PTHrP cDNA constructs mutated at the NLS. Transfected wild-type and NLS-mutated sense PTHrP mRNAs were detected by Northern blot analysis. Transfected sense PTHrP mRNA was only detected in MCF-7 cells transfected with the sense wild-type and NLS mutant PTHrP expression constructs (⬃1.1-kb transcript; Fig. 1), but not in untransfected (Fig. 1) or empty vector-transfected (data not shown) cells. Detection of the endogenous 1.5-kb transcript required much longer exposure of the autoradiograms (33). PTHrP secretion was measured by immunoassay. The amount of PTHrP secreted by control cells was 1.3 ⫾ 0.04 pm/105 cells. PTHrP secretion from the empty vector-transfected cells was not significantly different from that from untransfected cells (Fig. 2). Transfection with the wild-type PTHrP constructs produced a significant increase in PTHrP production compared with empty vector-transfected and untransfected cells (Fig. 2). Similarly, each of the clones overexpressing NLS-mutated PTHrP (⌬88 –91, ⌬102–106, and ⌬88 –91 ⫹ ⌬102–106) secreted significantly higher PTHrP levels (Fig. 2). Secretion from these NLS-mutated PTHrP transfectants was also higher than that from the wild-type PTHrP transfectants. During the process of cloning, secretion from a minimum of four clones for wild-type PTHrP and each of the deletions was tested. Clones expressing NLSmutated PTHrP consistently showed higher levels of PTHrP secretion compared with clones expressing wild-type

FIG. 1. Characterization of MCF-7 cells overexpressing wild-type or NLS-mutated PTHrP. Northern blot analysis of total RNA was carried out to detect transfected transcript from untransfected (parent) cells (P) and from cells transfected with a vector expressing wild-type sense PTHrP mRNA (SN) or with vectors expressing NLS-mutated PTHrP mRNA (⌬1, ⌬88 –91; ⌬2, ⌬102–106; ⌬⌬, ⌬88 –91 ⫹ ⌬102–106). The probe was prepared by asymmetric PCR of a cDNA fragment spanning exons 3 and 4 of the human PTHrP gene, as described in Materials and Methods. The positions of the 5S, 18S, and 28S ribosomal bands are indicated. Top panel, PTHrP mRNA; bottom panel, 18S RNA.

Sepulveda et al. • PTHrP and Breast Cancer Cell Proliferation

Endocrinology, February 2002, 143(2):596 – 606 599

firmed using the [3H]thymidine incorporation assay. After 3 d in culture, [3H]thymidine incorporation of wild-type sense PTHrP transfectants was approximately 1.7-fold that of empty vector transfectants (Fig. 4). This value matches the growth rate obtained by direct cell counting, where, after 4 d in culture, the number of wild-type PTHrP-overexpressing cells was approximately twice that of the empty vector controls (Fig. 3). The [3H]thymidine incorporation rate of the NLS-mutated PTHrP transfectants tended to be lower, but was not significantly different from that of empty vector transfectants (Fig. 4). We cannot at present explain the difference between the growth rate of empty vector-transfected and NLS-mutated PTHrP as measured by direct cell counting (⬃2.5-fold after 4 d in culture; Fig. 3) and by [3H]thymidine FIG. 2. PTHrP secretion by wild-type and NLS mutated PTHrP transfectants. Conditioned medium from MCF-7 cells transfected with a vector expressing wild-type or mutant NLS PTHrP was collected after 4 d in culture; secreted PTHrP was measured by the immunoradiometric assay. P, Untransfected (parent) cells; V, empty vector-transfected cells; SN, cells expressing wild-type PTHrP; ⌬1, ⌬2, and ⌬⌬, cells expressing PTHrP ⌬88 –91, ⌬102–106, and ⌬88 –91 ⫹ ⌬102–106, respectively. Each bar, representing an independent clone, is the mean ⫾ SEM of four wells obtained after subtracting the background value, represented by unconditioned medium (not exposed to cells). Asterisks denote significant differences from untransfected cells (P ⬍ 0.001).

PTHrP. Thus, clones overexpressing wild-type PTHrP secreted between 30 and 50 pm PTHrP/105 cells, whereas clones overexpressing NLS-mutated PTHrP secreted between 80 and 125 pm/105 cells. The empty vector transfectants secreted less than 1.5 pm PTHrP/105 cells. It may be possible that secretion of wild-type PTHrP on the same scale as that for NLS-mutated PTHrP is detrimental to cell viability, or that secretion levels are affected by the inability of NLS-mutated PTHrP to enter the nucleus. Two clones for wild-type PTHrP and each of the NLS mutants (⌬88 –91, ⌬102–106, and ⌬88 –91 ⫹ ⌬102–106) were chosen for further experiments; secretion from these clones is shown in Fig. 2. These clonal lines were used to examine whether the mutated NLS sequence influenced the ability of PTHrP to promote cell growth.

FIG. 3. Proliferation of MCF-7 cells overexpressing wild-type or NLS mutant PTHrP. Cells were plated in 10% FBS at a density of 104 cells/well in 24-well dishes. At the indicated time intervals, cells were trypsinized, and cell numbers were determined using a Coulter counter. Each point is the mean ⫾ SEM of three independent experiments for each of two individual clones (four wells per experiment). Where no error bar is shown, the SEM is smaller than the point. F, Empty vector; E, sense; , ⌬88 –91; f, ⌬102–106; ƒ, ⌬88 –91 ⫹ ⌬102– 106. *, Significantly different from control (empty vector transfectant) at P ⬍ 0.001.

Mutant NLS negates the proliferative effects of PTHrP overexpression

The growth rates of cells stably transfected with cDNAs expressing wild-type or NLS-mutated PTHrP are shown in Fig. 3. The growth rate of MCF-7 clones overexpressing wildtype PTHrP was more rapid than that of empty vector-transfected cells, such that after 4 and 7 d in culture, the cell number of the PTHrP-overexpressing clones was approximately 2 and 3 times that of the empty vector-transfected clones, respectively. On the other hand, the growth rate of the clones overexpressing NLS-mutated PTHrP was about 0.4 times that of empty vector-transfected clones (Fig. 3). NLS deletions still permit the secretion of large concentrations of N-terminally intact PTHrP, which may act to inhibit MCF-7 cell proliferation through an autocrine/paracrine pathway (33), but cannot increase cell proliferation through the intracrine pathway. The differences in the rate of proliferation of clones transfected with wild-type or NLS-mutated PTHrP were con-

FIG. 4. Thymidine incorporation by MCF-7 cells overexpressing wildtype or NLS-mutated PTHrP. Cells were plated in 12-well dishes at 2 ⫻ 104 cells/well in medium containing 10% FBS. After 3 d, cells were pulse-treated with [3H]thymidine for 3 h. Thymidine incorporation was determined as described in Materials and Methods. Each bar is the mean ⫾ SEM of six independent experiments (three experiments for each of two independent clones). P, Parent; V, empty vector; SN, sense; ⌬1, ⌬88 –91; ⌬2, ⌬102–106; ⌬⌬, ⌬88 –91 ⫹ ⌬102–106. *, Significantly different from control (empty vector transfectant) at P ⬍ 0.001.

600

Endocrinology, February 2002, 143(2):596 – 606

incorporation (Fig. 4). The small difference in the [3H]thymidine incorporation rate over 3 d of culture may account for a relatively larger difference in cell number. Mutant NLS inhibits nuclear localization

Immunoflurescence was used to detect the presence of PTHrP in transfected and control MCF-7 cells. Wild-type PTHrP showed strong nuclear and cytoplasmic staining (Fig. 5A). Cytoplasmic staining was present in all cells, whereas nuclear staining was present in approximately 70% of nuclei. Cells transfected with NLS-mutated PTHrP ⌬88 –91 and ⌬88 –91 ⫹ ⌬102–106 still showed strong cytoplasmic staining in all cells, but in this case nuclear staining was present in less than 10% of nuclei (Fig. 5A). The ⌬102–106 mutant showed approximately 35% nuclear staining (Fig. 5A). Cytoplasmic staining was present in all cells. Control (empty vector-transfected) cells showed significantly less intense cytoplasmic staining, and nuclear staining was present in about 25% of nuclei (Fig. 5A). Cells stained in the absence of primary antibody showed no cytoplasmic or nuclear staining (Fig. 5A). Data for nuclear staining in the different clones are summarized in Fig. 5B. Wild-type, but not NLS-mutated, PTHrP transfectants are enriched in G2 ⫹ M

To determine whether the proliferative effects of PTHrP overexpression are accompanied by an effect on the cell cycle profile, this parameter was measured in asynchronously cycling cells grown in 10% FBS. For this purpose, MCF-7 clones overexpressing wild-type or NLS-mutated PTHrP, and control (empty vector-transfected) cells were stained with propidium iodide and analyzed by flow cytometry to determine

Sepulveda et al. • PTHrP and Breast Cancer Cell Proliferation

the distribution of cells in the various stages of the cell cycle. All clones exhibited a major diploid peak (G0/G1). As shown in Fig. 6, A and B, cells overexpressing wild-type PTHrP were enriched in the G2 ⫹ M phase of the cell cycle (32% in G2 ⫹ M), whereas empty vector-transfected cells were largely in G1 (only 4% in G2 ⫹ M). Parental MCF-7 cells were also largely in G1 (only 3.7% in G2 ⫹ M; Fig. 6B). Cells transfected with the NLS mutants ⌬88 –91, ⌬102–106, and ⌬88 –91 ⫹ ⌬102–106) showed the same profile as empty vector-transfected cells (4.2– 6.7% in G2 ⫹ M; Fig. 6, A, B). We also determined the cell cycle profile of parental MCF-7 cells cultured in 10% FBS in the presence of hPTHrP(1–34), -(1–139), or -(107–139) for 3 or 5 d. Under these conditions, there was no difference in the cell cycle profile between treated and untreated cells (data not shown). As PTHrP decreases MCF-7 cell proliferation when acting through the autocrine/paracrine pathway (33), we expected that treatment with exogenous PTHrP analogs would result in fewer cells in G2 ⫹ M. The lack of any significant effect probably can be ascribed to the fact that a very low percentage of cells (3.7%) are normally in G2 ⫹ M in parental MCF-7 cells (data not shown). Wild-type, but not NLS-mutated, PTHrP inhibits apoptosis

Because growth factors and cytokines are known to play important roles in maintaining cell health and survival, serum deprivation has become a widely used strategy to induce apoptosis in a variety of cells (45– 47). In the early stages of apoptosis, phosphatidylserine migrates from the inner part of the plasma membrane to the outer layer, and thus becomes exposed on the external surface of the cell (48). Annexin V is a Ca2⫹-dependent phospholipid-binding pro-

FIG. 5. Immunofluorescence detection of PTHrP in MCF-7 cells overexpressing wild-type or NLS-mutated PTHrP. SN, Wild-type PTHrP transfectants; ⌬1, ⌬88 –91; ⌬2, ⌬102–106; ⌬⌬, ⌬88 –91 ⫹ ⌬102–106; V, empty vector transfectants; C, negative control (absence of primary antibody). A, Cells overexpressing wild-type PTHrP show abundant nuclear and cytoplasmic staining. Cells overexpressing NLS-mutated PTHrP show abundant cytoplasmic, but significantly less nuclear, staining. Cells transfected with empty vector show reduced cytoplasmic and nuclear staining. Magnification, ⫻100. B, The percentage of nuclei that contain PTHrP in the various MCF-7 clones. Each bar represents the mean of 8 slides (4 slides for each of 2 individual clones), and 100 –200 cells were counted per slide in a blinded fashion by 2 observers. Data are presented as the mean ⫾ SEM. *, P ⬍ 0.01.

Sepulveda et al. • PTHrP and Breast Cancer Cell Proliferation

Endocrinology, February 2002, 143(2):596 – 606 601

tein with high affinity for phosphatidylserine. Apoptosis in serum-depleted (0.1% NuSerum) transfected, clonal MCF-7 cells was evaluated using annexin V as a marker of early stage apoptotic cells. This methodology was chosen over DNA fragmentation techniques because a number of reports have concluded that apoptosis in MCF-7 cells is not associated with DNA degradation (49 –51). Annexin V labeling was measured by FACS analysis. When coupled with 7-AAD, this method differentiates between apoptotic and necrotic cells (40, 41). After serum starvation, cells overexpressing wild-type PTHrP showed significantly less apoptosis than control cells (Fig. 7). The level of apoptosis in cells expressing NLS-mutated PTHrP was not significantly different from that in control cells (Fig. 7). Apoptosis in parental MCF-7 cells cultured in serumdepleted medium (0.1% NuSerum for 24, 48, or 72 h) was also measured by FACS analysis in the presence of hPTHrP-(1– 34), -(1–139), or -(107–139). None of the analogs, tested at a dose of 10⫺7 m for each of the three time points, influenced the extent of apoptosis, as determined by FACS analysis (data not shown). Wild-type and NLS-mutated PTHrP-overexpressing cells express different levels of the Bcl-2 family of proteins

The Bcl-2 family of proteins includes central regulators of the apoptotic pathway regardless of the stimulus for apo-

FIG. 6. Cell cycle analysis of wild-type or NLS-mutated PTHrP overexpressing and empty vector-transfected MCF-7 cells. Cells were plated in 10% FBS and at 75% confluence were collected and analyzed by flow cytometry as described in Materials and Methods. SN, Wildtype PTHrP transfectants; V, empty vector transfectants; ⌬1, ⌬88 –91; ⌬2, ⌬102–106; ⌬⌬, ⌬88 –91 ⫹ ⌬102–106. A, Cell cycle profile. B, Percentage of cells in G2 ⫹ M. Each bar is the mean ⫾ SEM of six independent experiments (three experiments for each of two individual clones). *, Significantly different from control (empty vector transfectant) at P ⬍ 0.001.

FIG. 7. Annexin V staining of MCF-7 cells overexpressing wild-type or NLS-mutated PTHrP. Cells were plated in 10% FBS, then transferred to 0.1% NuSerum after 24 h. At 50% confluence, they were synchronized with aphidicolin and analyzed by FACS analysis as described in Materials and Methods. Viable cells were gated based on their forward and side light scatter characteristics. Additional gating of viable cells was based on their ability to exclude 7-AAD. A, Annexin V vs. 7-AAD staining of cells. Percentages refer to the number of annexin V-positive cells in the lower right quadrant. B, Percentage of annexin V-stained cells (from values in the lower right quadrant). SN, Sense; V, empty vector; ⌬1, ⌬88 –91; ⌬2, ⌬102–106; ⌬⌬, ⌬88 –91 ⫹ ⌬102–106. Each bar is the mean ⫾ SEM of six independent experiments (three experiments for each of two individual clones). *, Significantly different from all other values at P ⬍ 0.001.

602

Endocrinology, February 2002, 143(2):596 – 606

Sepulveda et al. • PTHrP and Breast Cancer Cell Proliferation

ptosis. Certain family members (e.g. Bcl-2 and Bcl-xL) act as potent suppressors of apoptosis, whereas others (e.g. Bax and Bak) have opposing functions and promote cell death. The ratio of anti- to proapoptotic molecules determines the response to a death signal (52–54). We performed FACS analysis to determine whether PTHrP alters the expression of these apoptosis-related proteins. Figures 8 and 9 show an increase in Bcl-2/Bax and Bcl-xL/Bax ratios in wild-type PTHrP-overexpressing cells compared with those in NLSmutated PTHrP-overexpressing and vector control cells. Wild-type, but not NLS-mutated, PTHrP inhibits nuclear degradation

We compared the nuclear morphology of MCF-7 cells overexpressing wild-type or NLS-mutated PTHrP after treatment with valinomycin, a potassium ionophore that gives rise to atypical cell death, with chromatin condensation appearing with (43) or without (55) DNA fragmentation. This treatment was chosen because, compared with serum starvation alone, it produced very dramatic differences in terms of nuclear vulnerability between wild-type PTHrP-overexpressing vs. control (empty vector-transfected) and NLSmutated PTHrP-overexpressing cells. Cells cultured in 10%

FIG. 9. Determination of the Bcl-xL/Bax ratio in MCF-7 cells overexpressing wild-type or NLS-mutated PTHrP. Cells were plated in 10% FBS, then transferred to 0.1% NuSerum after 24 h. At 50% confluence, they were synchronized with aphidicolin and analyzed for intracellular Bcl-xL and Bax by FACS analysis as described in Materials and Methods. SN, Sense; V, empty vector; ⌬1, ⌬88 –91; ⌬2, ⌬102–106; ⌬⌬, ⌬88 –91 ⫹ ⌬102–106. A, Dot plot of Bcl-xL vs. Bax staining of cells. B, Bcl-xL/Bax ratio, calculated as described in Materials and Methods. Each bar is the mean ⫾ SEM of six independent experiments (three experiments for each of two individual clones). Where no error bar is shown, the SEM is smaller than the bar line. *, Significantly different from all other values at P ⬍ 0.01.

FIG. 8. Determination of the Bcl-2/Bax ratio in MCF-7 cells overexpressing wild-type or NLS-mutated PTHrP. Cells were plated in 10% FBS, then transferred to 0.1% NuSerum after 24 h. At 50% confluence, they were synchronized with aphidicolin and analyzed for intracellular Bcl-2 and Bax by FACS analysis as described in Materials and Methods. SN, Sense; V, empty vector; ⌬1, ⌬88 –91; ⌬2, ⌬102–106; ⌬⌬, ⌬88 –91 ⫹ ⌬102–106. A, Dot plot of Bcl-2 vs. Bax staining of cells. B, Bcl-2/Bax ratio, calculated as described in Materials and Methods. Each bar is the mean ⫾ SEM of six independent experiments (three experiments for each of two individual clones). *, Significantly different from all other values at P ⬍ 0.01.

serum in the absence of valinomycin show predominantly round or oval nuclei (healthy nuclei; Fig. 10A). The predominant effect of valinomycin treatment in MCF-7 cells overexpressing wild-type PTHrP was nuclear condensation (micronuclei; Fig. 10, A and B). There were no degraded nuclei. In contrast, valinomycin treatment of empty vector-transfected cells predominantly caused nuclear degradation (Fig. 10, A and B). Nuclei from cells overexpressing NLS-mutated PTHrP (⌬88 –91 and ⌬88 –91 ⫹ ⌬102–106) were either condensed or degraded, with very few healthy nuclei (Fig. 10, A and B). There were approximately 30% healthy nuclei in valinomycin-treated cells overexpressing the ⌬102–106 PTHrP mutant, with about 60% micronuclei and the rest degraded nuclei (Fig. 10, A and B). Discussion

Determining the factors involved in promoting breast cancer cell growth is crucial for understanding the pathogenesis of the disease. This study has examined the mechanism by which PTHrP enhances growth of the breast cancer cell line MCF-7. Our main findings are that overexpression of PTHrP in clonal MCF-7 cells stably transfected with PTHrP cDNA

Sepulveda et al. • PTHrP and Breast Cancer Cell Proliferation

FIG. 10. DNA staining with Hoechst 33342 of MCF-7 cells overexpressing wild-type or NLS-mutated PTHrP. U, Control cells (10% serum); all other cells were treated with valinomycin (40 ␮M) for 4 h in the presence of 0.1% NuSerum. SN, Wild-type PTHrP transfectants; V, empty vector transfectants; ⌬1, ⌬88 –91; ⌬2, ⌬102–106; ⌬⌬, ⌬88 –91 ⫹ ⌬102–106 NLS-mutated PTHrP transfectants. A, Arrows point to condensed nuclei (micronuclei). Magnification, ⫻100. B, The percentage of micronuclei and degraded nuclei in the various MCF-7 clones. Each bar represents the mean ⫾ SEM of 8 slides (4 slides for each of 2 individual clones), and 50 –100 cells were counted per slide in a blinded fashion by 2 observers. *, Significantly different from respective vector control at P ⬍ 0.01.

protects against apoptosis induced by serum starvation. PTHrP overexpression also increased [3H]thymidine incorporation and the percentage of cells in the G2 ⫹ M phase of the cell cycle, indicating that the peptide increases the mitotic rate in these cells. Taken together, these effects on apoptosis and cell cycle regulation account for the observed increased growth rate of PTHrP-overexpressing MCF-7 cells (33). Regulation of cell proliferation by PTHrP can be mediated via autocrine/paracrine or intracrine pathways. We have previously shown that exogenously added N-terminal or full-length PTHrP exerts an antimitogenic effect in MCF-7 cells, and that this effect is mediated via the cell surface PTH/PTHrP receptor, which is present and functional in these cells (33). These same exogenously added analogs had no effect on apoptosis, suggesting that the antiapoptotic effects are mediated solely via an intracrine pathway. It is noteworthy, however, that nuclear targeting of PTHrP can be accomplished via protein that is endo-

Endocrinology, February 2002, 143(2):596 – 606 603

cytosed from the cell exterior, even in cells that do not express the PTH/PTHrP receptor (56 –58), providing nuclear targeting of exogenously added PTHrP. However, MCF-7 cells may lack the necessary cellular component(s) to endocytose and translocate PTHrP to the nucleus. In these cells, endogenous PTHrP may normally translocate to the nucleus without ever leaving the cytosol. Such a mechanism has been proposed by workers who propose that alternative translation start sites may bypass the PTHrP N-terminal signal sequence and result in a peptide that cannot enter the endoplasmic reticular space to be secreted, but, having the NLS sequence, can travel directly to the nucleus (56). In addition, we have not tested all available PTHrP analogs [e.g. PTHrP-(38 – 64) and -(87–106)]. The basic question of whether the antiapoptotic effects of PTHrP observed in this study involved a nuclear action was addressed by overexpressing PTHrP with a mutated NLS, an approach often used to answer this question (47, 56 – 60). The results, obtained with MCF-7 cells overexpressing mutant PTHrP, showed unequivocally that deleting either multibasic cluster (⌬88 –91 or ⌬102–106) or both (⌬88 –91 ⫹ ⌬102– 106) in the NLS of PTHrP negated its ability to protect against apoptosis. Mutant PTHrP was also ineffective in preventing against valinomycin-induced nuclear degradation. These findings indicate that an intact NLS is required for the PTHrP-mediated protective effects against apoptosis and suggest that this action requires that PTHrP enter the nucleus. The protective effects of PTHrP on apoptosis and survival have been noted in a number of other systems, including chondrocytes (23), cerebellar granule cells (61), and prostate cancer cells (62). The mechanisms underlying the ability of PTHrP to protect against apoptosis may be mediated via the Bcl-2 family of proteins. The proapoptotic protein Bax binds to the antiapoptotic Bcl-2 and Bcl-xL through interaction of their BH3 domains, forming heterodimers (63, 64). The balance of homodimers of Bcl-2 and Bcl-xL to Bcl-2/Bax and Bcl-xL/Bax heterodimers influences cell fate (53, 54, 65). When Bax is expressed preferentially, Bax homodimers predominate, promoting apoptosis. When Bcl is expressed preferentially, Bcl homodimers form, and apoptosis is inhibited. Therefore, the ratio of expression of Bcl-2 and Bcl-xL to Bax influences cell survival. Overexpression of PTHrP increases the Bcl-2/Bax and Bcl-xL/Bax ratios, suggesting that the protective effect of wild-type PTHrP overexpression against apoptosis may be mediated via this pathway. Whether additional pathways are involved in these PTHrP effects are at present unknown and are currently being addressed. PTHrP overexpression has also been found to modulate Bax and Bcl-xL levels in IEC-6 cells, a nontumor rat intestinal crypt cell line (47). Interestingly, in this cell line, PTHrP overexpression enhanced apoptosis via an intracrine pathway, and Bcl-xL expression was decreased whereas Bax expression was increased (47). Overall, the peptide still increased the proliferation of IEC-6 cells. It is at present not understood how the peptide can exert opposing effects in different cell lines. However, both proliferative and antiproliferative effects have pre-

604

Endocrinology, February 2002, 143(2):596 – 606

viously been attributed to PTHrP, depending on the cell type (33, 56, 59, 66, 67). Deletion of either one or both of these multibasic clusters within the NLS also abolished the proliferative effects of PTHrP in MCF-7 cells. Similar results have been reported in a number of studies, including normal rat small intestinal crypt cell-derived IEC-6 cells (59), vascular smooth muscle cells (56), and COS-1 cells (23). Mutation of the NLS also abolished the effect of PTHrP on the cell cycle, thus supporting an intracrine role for the peptide in cell cycle regulation. Decreased proliferation of the NLS mutants was reflected by a decrease in nuclear localization, as determined by immunofluorescence. Thus, significantly more nuclei stained positively for PTHrP in cells transfected with the wild-type peptide than in cells transfected with NLS mutants or empty vector. Interestingly, more nuclei stained positively for PTHrP in the ⌬102–106 mutant compared with the ⌬88 –91 and the ⌬88 –91 ⫹ ⌬102–106 mutants. The single NLS-mutated PTHrP clones target two different regions of PTHrP. Amino acids 88 –91 are recognized by ␤-importin (58), and therefore the presence of this sequence in the ⌬102– 106 mutant may still permit some trafficking of PTHrP to the nucleus. However, the increased nuclear presence of PTHrP in the ⌬102–106 mutant did not appear to influence the growth rate (as measured by direct cell counts, [3H]thymidine incorporation, and the cell cycle profile) or extent of apoptosis, compared with the ⌬88 –91 and ⌬88 –91 ⫹ ⌬102– 106 mutants. Taken together, these data demonstrate a critical role for nuclear targeting in the antiapoptotic and cell cycle regulatory effects of PTHrP. PTHrP plays an important role in breast cancer. Breast cancer cells secrete high levels of PTHrP, accounting for the high prevalence of humoral hypercalcemia of malignancy in breast cancer patients (68), and there is a positive correlation between PTHrP expression in breast cancer and skeletal metastasis (69 –71). In a mouse model of bone metastases, neutralizing antibodies to PTHrP-(1–34) inhibited the development of breast carcinoma metastases to bone of the human breast carcinoma cells MDA-MB-231, which produce moderate amounts of PTHrP (72). Conversely, in the same mouse model, overexpression of PTHrP by MCF-7 cells, which normally express low levels of PTHrP and do not cause osteolytic metastases, induced marked bone destruction associated with increased osteoclast formation compared with controls (73). These data suggest that tumor production of PTHrP is important for the establishment and progression of osteolytic bone metastases. However, Henderson et al. (74) report that patients presenting with PTHrP-positive breast cancers have a more favorable outcome and have fewer metastases to bone and other sites. This study used an antiPTHrP-(1–14) antibody to detect PTHrP by immunocytochemistry (74). This antibody may not recognize all forms of PTHrP generated by posttranslational processing. In addition, the intracrine form(s) of PTHrP has not all been identified and again may not be recognized by this antibody. It is possible that breast cancer growth and bone metastasis are mediated by specific forms of PTHrP not recognized by anti-PTHrP-(1–14) antibody. Therefore, these findings do not exclude a central role for PTHrP in breast cancer growth and metastasis. The PTHrP-mediated modulation of cell prolif-

Sepulveda et al. • PTHrP and Breast Cancer Cell Proliferation

eration through apoptosis and cell cycle regulatory pathways in breast cancer is of both physiological and clinical relevance. We believe that ours is the first report demonstrating the mechanisms responsible for imparting a growth advantage to PTHrP-expressing breast cancer cells. In summary, our studies show that overexpression of PTHrP in clones of MCF-7 cells stably transfected with PTHrP cDNA significantly increases the proliferation of these cells through both antiapoptotic and cell cycle regulatory effects. The presence of an intact NLS within PTHrP is required for this action. These findings suggest that PTHrP, produced in abundance by breast cancer cells, may increase the proliferation of these cells in an intracrine manner, thereby establishing a positive regulatory loop between protein production and breast tumor burden. Acknowledgments We thank Dr. Rolf Konig for advice with FACS analysis, and Drs. David Konkel, Patricia K. Seitz, Mary L. Thomas, and Cheryl S. Watson for critical reading of the manuscript. Received July 19, 2001. Accepted October 15, 2001. Address all correspondence and requests for reprints to: Dr. Miriam Falzon, Department of Pharmacology and Toxicology and Sealy Center for Molecular Science, University of Texas Medical Branch, 10th and Market Streets, Galveston, Texas 77555. E-mail: [email protected].

References 1. Burtis WJ, Wu T, Bunch C, Wysolmerski JJ, Insogna KL, Weir EC, Broadus AE, Stewart AF 1987 Identification of a novel 17,000 dalton parathyroid hormone-like adenylate cyclase-stimulating protein from a tumor associated with humoral hypercalcemia of malignancy. J Biol Chem 262:7151–7156 2. Moseley JM, Kubota M, Diefenbach-Jagger H, Wettenhall REH, Kemp BE, Suva LJ, Rodda CP, Ebeling PR, Hudson PJ, Zajac JD, Martin TJ 1987 Parathyroid hormone-related protein purified from a human lung cancer cell line. Proc Natl Acad Sci USA 84:5048 –5052 3. Suva LJ, Winslow GA, Wettenhall REH, Hammonds RG, Moseley JM, Diefenbach-Jagger H, Rodda CP, Kemp BE, Rodriguez H, Chen, EY, Hudson, PJ, Martin TJ, Wood WI 1987 A parathyroid hormone-related protein implicated in malignant hypercalcemia: cloning and expression. Science 237:893– 896 4. Broadus AE, Mangin M, Ikeda K, Insogna KL, Weir EC, Burtis WJ, Stewart AF 1988 Humoral hypercalcemia of cancer. N Engl J Med 319:556 –563 5. Strewler GI 2000 Mechanisms of disease: the physiology of parathyroid hormone-related protein. N Engl J Med 342:177–185 6. Stewart AF, Insogna KL, Broadus, AE 1995 Malignancy-associated hypercalcemia. In: DeGroot L, ed. Endocrinology. 3rd ed. Philadelphia: Saunders; 1061–1074 7. Philbrick WM, Wysolmerski JJ, Galbraith S, Holt E, Orloff JJ, Yang KH, Vasavada RC, Weir EC, Broadus AE, Stewart AF 1996 Defining the role of parathyroid hormone-related protein in normal physiology. Physiol Rev 76: 127–173 8. Yang KH, Stewart AF 1996 The PTH-related protein gene and protein products. In: Bilezikian JP, Riasz L, Rodan G, eds. Principles of bone biology. San Diego: Academic Press; 347–376 9. de Papp AE, Yang KH, Soifer NE, Stewart AF 1995 Parathyroid hormonerelated protein as a prohormone: Posttranslational processing and receptor interaction. Endocr Rev Monogr A4:207–210 10. Burtis WJ 1992 Parathyroid hormone-related protein: structure, function and measurement. Clin Chem 38:2171–2183 11. Wu T, Vasavada R, Yang K, Massfelder T, Ganz M, Abbas SK, Care AD, Stewart AF 1996 Structural and physiologic characterization of the mid-region secretory species of parathyroid hormone-related protein. J Biol Chem 271: 24371–24381 12. Plawner LL, Philbrick WM, Burtis WJ, Broadus AE, Stewart AF 1995 Cell type-specific secretion of parathyroid hormone-related protein via the regulated versus the constitutive secretory pathway. J Biol Chem 270:14078 –14084 13. Orloff JJ, Kats Y, Urena P, Schipani E, Vasavada RC, Philbrick WM, Behal A, Abou-Samra AB, Segre GV, Ju¨ppner H 1995 Further evidence for a novel receptor for amino-terminal parathyroid hormone-related protein on keratinocytes and squamous carcinoma cell line. Endocrinology 136:3016 –3023

Sepulveda et al. • PTHrP and Breast Cancer Cell Proliferation

14. Mannstadt M, Ju¨ ppner H, Gardella TJ 1999 Receptors for PTH and PTHrP. Am J Physiol 277:F665–F675 15. Ju¨ ppner H, Abou-Samra A-B, Freeman M, Kong XF, Schipani E, Richards J, Kolakowski Jr LF, Hock J, Potts Jr JT, Kronenberg HM, Segre GV 1991 A G protein-linked receptor for parathyroid hormone and parathyroid hormonerelated peptide. Science 254:1024 –1026 16. Abou-Samra A-B, Ju¨ ppner H, Force T, Freeman MW, Kong X-F, Schipani E, Urena P, Richards J, Bonventre JV, Potts JT, Kronenberg HM, Segre GV 1992 Expression cloning of a common receptor for parathyroid hormone and parathyroid hormone-related peptide from rat osteoblast-like cells: a single receptor stimulates intracellular accumulation of both cAMP and inositol triphosphates and increases intracellular free calcium. Proc Natl Acad Sci USA 89: 2732–2736 17. Wysolmerski JJ, Stewart AF 1998 The physiology of parathyroid hormonerelated protein: an emerging role as a developmental factor. Annu Rev Physiol 60:431– 460 18. Orloff JJ, Ganz MB, Nathanson MH, Moyer MS, Kats Y, Mitnick M, Behal A, Gasalla-Herraiz J, Isales CM 1996 A mid-region parathyroid hormonerelated protein mobilizes cytosolic calcium and stimulates formation of inositol triphosphate in a squamous cell line. Endocrinology 137:5376 –5385 19. Fenton AJ, Kemp BE, Hammonds Jr RG, Mitchelhill K, Moseley JM, Martin TJ, Nicholson GC 1991 A potent inhibitor of osteoclastic bone resorption within a highly conserved pentapeptide region of PTHrP (107–111). Endocrinology 129:3424 –3426 20. Valin A, Garcia-Ocana A, De Miguel F, Sarasa JL, Esbrit P 1997 Antiproliferative effect of the C-terminal fragments of parathyroid hormone-related protein, PTHrP (107–111) and (107–139), on osteoblastic osteosarcoma cells. J Cell Physiol 170:209 –215 21. Amling M, Neff L, Tanaka S, Inoue D, Kuida K, Wei E, Philbrick WM, Broadus AE, Baron R 1997 Bcl-2 lies downstream of parathyroid hormonerelated peptide in a signaling pathway that regulates chondrocyte maturation during skeletal development. J Cell Biol 136:205–213 22. Kaiser SM, Sebag M, Rhim JS, Kremer R, Goltzman D 1994 Antisensemediated inhibition of parathyroid hormone-related peptide production in a keratinocyte cell line impedes differentiation. Mol Endocrinol 8:139 –147 23. Henderson JE, Amikuza N, Warshawsky H, Biasotto D, Lanske BMK, Goltzman D, Karaplis AC 1995 Nucleolar localization of parathyroid hormonerelated peptide enhances survival of chondrocytes under conditions that promote apoptotic cell death. Mol Cell Biol 15:4064 – 4075 24. Lam MHC, Olsen SL, Rankin WA, Ho PWM, Martin TJ, Gillespie MT, Moseley JM 1997 PTHrP and cell division: expression and localization of PTHrP in a keratinocyte cell line (HaCaT) during the cell cycle. J Cell Physiol 173:433– 446 25. Dingwall C, Laskey RA 1991 Nuclear targeting sequences—a consensus? Trends Biochem Sci 16:478 – 481 26. Liapis H, Crouch EC, Grosso LE, Kitazawa S, Wick MR 1993 Expression of parathyroid like protein in normal, proliferative and neoplastic human breast tissues. Am J Pathol 143:1169 –1178 27. Ratcliffe WA 1992 Role of parathyroid hormone-related protein in lactation Endocrinology 37:402– 404 28. Wysolmerski JJ, McCaughern-Carucci JF, Daifotis AG, Broadus AE, Philbrick WM 1995 Overexpression of parathyroid hormone-related protein or parathyroid hormone in transgenic mice impairs branching morphogenesis during mammary gland development. Development 121:3539 –3547 29. Wysolmerski JJ, Philbrick WM, Dunbar ME, Lanske B, Kronenberg H, Broadus AE 1998 Rescue of the parathyroid hormone-related protein knockout mouse demonstrates that parathyroid hormone-related protein is essential for mammary gland development. Development 125: 1285–1294 30. Stewart AF, Wu T, Goumas D, Burtis WJ, Broadus AE 1987 N-terminal amino-acid sequence of two novel tumor-derived adenylate cyclase-stimulating proteins: identification of parathyroid hormone-like and -unlike domains. Biochem Biophys Res Commun 146:672– 678 31. Luparello C, Burtis WJ, Raue F, Birch MA, Gallagher JA 1995 Parathyroid hormone-related peptide and 8701-BC breast cancer cell growth and invasion in vitro: evidence for growth-inhibiting and invasion-promoting effects. Mol Cell Endocrinol 111:225–232 32. Luparello C, Birch MA, Gallagher JA, Burtis WJ 1997 Clonal heterogeneity of the growth and invasive response of a human breast carcinoma cell line to parathyroid hormone-related peptide fragments. Carcinogenesis 18:23–29 33. Falzon M, Du P 2000 Enhanced growth of MCF-7 breast cancer cells overexpressing parathyroid hormone-related peptide. Endocrinology 141: 1882–1892 34. Wu C, Seitz PK, Falzon M 2000 Single-column purification and bio-characterization of recombinant parathyroid hormone-related protein (1–139). Mol Cell Endocrinol 170:163–174 35. Falzon M, Zong J 1998 The noncalcemic vitamin D analogs EB1089 and 22-oxacalcitriol suppress serum-induced parathyroid hormone-related peptide gene expression in a lung cancer cell line. Endocrinology 139:1046 –1053 36. Yasuda T, Banville D, Hendy GN, Goltzman D 1989 Characterization of the human parathyroid hormone-like peptide gene. J Biol Chem 264:7720 –7725

Endocrinology, February 2002, 143(2):596 – 606 605

37. Maniatis T, Fritsch EF, Sambrook J, eds. 1989 Molecular cloning: a laboratory manual. 2nd ed. Cold Spring Harbor: Cold Spring Harbor Laboratory 38. Coen DM 1995 The polymerase chain reaction. In: Ausubel FM, Brent R, Kingston RE, Moore DD, Seidman JG, Smith JA, Struhl K, eds. Current protocols in molecular biology. New York: Wiley & Sons; unit 1553 39. Fraser WD, Robinson J, Lawton R, Durham B, Gallacher SJ, Boyle IT, Beastall GH, Logue FC 1993 Clinical and laboratory studies of a new immunoradiometric assay for parathyroid hormone-related protein. Clin Chem 39: 414 – 419 40. Schmid I, Krall WJ, Uittenbogaart CH, Braun J, Giorgi JV 1992 Dead cell discrimination with 7-amino-actinomycin D in combination with dual color immunoflurescence in single laser flow cytometry. Cytometry 13:204 –208 41. Lecoeur H, Ledru E, Prevost MC, Gougeon ML 1997 Strategies for phenotyping apoptotic periperal human lymphocytes comparing ISNT, annexin-V and 7-AAD cytofluorimetric staining methods. J Immunol Methods 209: 111–123 42. Segal DM 1996 Flow cytometry. In: Coligan JE, Kruisbeek AM, Margulies DH, Shevach EM, Strober W, eds. Current protocols in immunology. New York: Wiley & Sons; unit 5.6 43. Deckers CLP, Lyons AB, Samuel K, Sanserson A, Maddy AH 1993 Alternative pathways of apoptosis induced by methylprednisolone and valinomycin analyzed by flow cytometry. Exp Cell Res 208:362–371 44. Hamatake M, Iguchi K, Ishida R 2000 Zinc induces mixed types of cell death, necrosis, and apoptosis, in Molt-4 cells. J Biochem 128:933–939 45. Wyllie AH 1987 Apoptosis: cell death in tissue regulation. J Pathol 153:313–316 46. Raff MC 1992 Social controls on cell survival and cell death. Nature 356:397– 400 47. Ye Y, Wang C, Du P, Falzon M, Seitz PK, Cooper CW 2001 Overexpression of parathyroid hormone-related protein enhances apoptosis in the rat intestinal cell line, IEC-6. Endocrinology 142:1906 –1914 48. Vermes I, Haanen C, Steffens-Nakken H, Reutelingsperger C 1995 A novel assay for apoptosis. Flow cytometric detection of phosphatidylserine expression on early apoptotic cells using fluorescein labelled Annexin V. J Immunol Methods 184:39 –51 49. Miller T, Beausang LA, Meneghini M, Lidgard G 1993 Death-induced changes to the nuclear matrix: the use of anti-nuclear matrix antibodies to study agents of apoptosis. BioTechniques 15:1042–1047 50. Oberhammer F, Wilson JW, Dive C, Wakeling AE, Walker PR, Sikorska M 1993 Apoptotic death in epithelial cells: cleavage of DNA to 300 and/or 50 kb fragments prior to or in the absence of internucleosomal fragmentation. EMBO J 12:3679 –3684 51. Cohen GM, Sun X-M, Snowden RT, Dinsdale D, Skilleter DN 1992 Key morphological features of apoptosis may occur in the absence of internucleosomal DNA fragmentation. Biochem J 286:331–334 52. Reed JC 1998 Bcl-2 family proteins. Oncogene 17:3225–3236 53. Oltvai Z, Korsmeyer SJ 1994 Checkpoints of dueling dimers foil death wishes. Cell 79:189 –192 54. Oltvai ZN, Milliman CL, Korsmeyer SJ 1993 Bcl-2 heterodimerizes in vivo with a conserved homolog, bax, that accelerates programmed cell death. Cell 74:609 – 619 55. Sun DY, Jiang S, Zheng LM, Ojcius DM, Young JD 1994 Separate metabolic pathways leading to DNA fragmentation and apoptotic chromatin condensation. J Exp Med 179:559 –568 56. Massfelder T, Dann P, Wu TL, Vasavada R, Helwig J-J, Stewart AF 1997 Opposing mitogenic and anti-mitogenic actions of parathyroid hormonerelated protein in vascular smooth muscle cells: a critical role for nuclear targeting. Proc Natl Acad Sci USA 94:13630 –13635 57. Aarts MM, Rix A, Bringhurst R, Henderson JE 1999 The nuclear targeting signal (NTS) of parathyroid hormone related protein mediates endocytosis and nucleolar translocation. J Bone Miner Res 14:1493–1503 58. Lam MHC, Briggs LJ, Hu W, Martin TJ, Gillespie MT, Jans DA 1999 Importin b recognizes parathyroid hormone-related protein with high affinity and mediates its nuclear import in the absence of importin a. J Biol Chem 274:7391– 7398 59. Ye Y, Falzon M, Seitz PK, Cooper CW 2001 Overexpression of parathyroid hormone-related protein promotes cell growth in the rat intestinal cell line IEC-6. Regul Pept 99:169 –174 60. Nguyen MTA, Karaplis, AC 1998 The nucleus: a target site for parathyroid hormone-related peptide (PTHrP) action. J Cell Biochem 70:193–199 61. Brines ML, Ling Z, Broadus AE 1999 Parathyroid hormone-related protein protects against kainic acid excitotoxicity in rat cerebellar granule cells by regulating L-type channel calcium flux. Neurosci Lett 274:13–16 62. Dougherty KM, Blomme EA, Koh AJ, Henderson JE, Pienta KJ, Rosol TJ, McCauley LK 1999 Parathyroid hormone-related protein as a growth regulator of prostate carcinoma. Cancer Res 59:6015– 6022 63. Simonen M, Keller H, Heim J 1997 The BH3 domain of Bax is sufficient for interaction of Bax with itself and with other family members and is required for induction of apoptosis. Eur J Biochem 249:85–91 64. Zha H, Aime-Sempe C, Sato T, Reed JC 1996 Proapoptotic protein Bax heterodimerizes with Bcl-2 and homodimerizes with Bax via a novel domain (BH3) distinct from BH1 and BH2. J Biol Chem 271:7440 –7444 65. Sedlak TW, Oltvai ZN, Wang K, Boise LH, Thompson CB, Korsmeyer SJ 1995

606

66. 67. 68. 69. 70.

Endocrinology, February 2002, 143(2):596 – 606

Multiple Bcl-2 family members demonstrate selective dimerizations with Bax. Proc Natl Acad Sci USA 92:7834 –7838 Kaiser SM, Laneuville P, Bernier SM, Rhim JS, Kremer R, Goltzman D 1992 Enhanced growth of a human keratinocyte cell line induced by antisense RNA for parathyroid hormone-related peptide. J Biol Chem 267:13623–13628 Vasavada RC, Garcia-Ocana A, Massfelder T, Dann P, Stewart AF 1998 Parathyroid hormone-related protein in the pancreatic islet and the cardiovascular system. Recent Prog Horm Res 53:305–340 Hickey RC, Samaan NA, Jackson GL 1981 Hypercalcemia in patients with breast cancer: osseous metastases, hyperplastic parathyroid tissue or pseudoparathyroidism? Arch Surg 116:545–552 Bundred NJ, Walker RA, Ratcliffe WA, Warwick J, Morrison JM, Ratcliffe JG 1992 Parathyroid hormone related protein and skeletal morbidity in breast cancer. Eur J Cancer 28:690 – 692 Kohno N, Kitazawa S, Fukase M, Sakoda Y, Kanbara Y, Furuya Y, Ohashi O, Ishikawa Y, Saitoh Y 1994 The expression of parathyroid hormone-related

Sepulveda et al. • PTHrP and Breast Cancer Cell Proliferation

71. 72.

73. 74.

protein in human breast cancer with skeletal metastases. Surg Today 24: 215–220 Kissin MW, Henderson MA, Danks JA, Hayman JA, Bennett RC, Martin TJ 1993 Parathyroid hormone-related protein in breast cancers of widely varying prognosis. Eur J Surg Oncol 19:134 –142 Guise TA, Yin JJ, Taylor SD, Kumagai Y, Dallas M, Boyce BF, Yoneda T, Mundy GR 1996 Evidence for a causal role of parathyroid hormone-related protein in the pathogenesis of human breast cancer-mediated osteolysis. J Clin Invest 98:1544 –1549 Thomas RJ, Guise TA, Yin JJ, Elliott J, Horwood NJ, Martin TJ, Gillespie MT 1999 Breast cancer cells interact with osteoblasts to support osteoclast formation. Endocrinology 140:4451– 4458 Henderson MA, Danks JA, Moseley JM, Slavin JL, Harris TL, McKinlay MR, Hopper JL, Martin TJ 2001 Parathyroid hormone-related protein production by breast cancers, improved survival, and reduced bone metastases. J Natl Cancer Inst 93:234 –237