Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase C–Potentiated, p53-Mediated Apoptosis

download Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase C–Potentiated, p53-Mediated Apoptosis

of 12

Transcript of Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase C–Potentiated, p53-Mediated Apoptosis

  • 8/12/2019 Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase CPotentiated, p53-Mediated Apoptosis

    1/12

    Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase

    CPotentiated, p53-Mediated Apoptosis

    Eric Meyer, Jean-Yves Vollmer, Raymonde Bovey, and Ivan Stamenkovic

    Division of Experimental Pathology, Institut Universitaire de Pathologie, Universite de Lausanne, Lausanne, Switzerland

    Abstract

    p53, a major sensor of DNA damage, is a transcription factor

    that, depending on its phosphorylation status, regulates the

    cell cycle, DNA repair, or apoptosis. The protein kinase C

    (PKC) family of isozymes is also implicated in cell cycle and

    programmed cell death (PCD) control and has recently been

    shown to influence p53 function. Using three human colon

    adenocarcinoma cell lines SW480, EB-1, and HCT116 that

    either lack p53 function and were engineered to express

    inducible wild-type p53 (wt p53), or that constitutively express

    wt p53, we show that phorbol estermediated PKC activation

    potentiates p53-induced PCD. Despite the effectiveness of

    PKC/p53 synergy in inducing SW480 tumor cell death,however, a fraction of the cells invariably survive. To address

    the putative mechanisms that underlie resistance to PKC/p53-

    induced cell death, we generated a phorbol 12-myristate 13-

    acetate/p53resistant SW480 subline and compared the gene

    expression profile of resistant and parental cells by DNA

    microarray analysis. The results of these experiments show

    that PKC/p53-resistant cells express a higher level of several

    matrix metalloproteinases (MMP), including MMP-9, MMP-10,

    and MMP-12, and corresponding real-time PCR assays

    indicate that p53 is a negative regulator of MMP-9 gene

    expression. Using MMP inhibitors and MMP-specific small

    interfering RNA, we show that MMP function confers

    protection from PKC/p53-induced apoptosis and identify theprotective MMPs as MMP-9 and MMP-10. Taken together,

    these observations provide evidence that MMPs are implicat-

    ed in tumor cell resistance to the synergistic proapoptotic

    effect of PKC and p53. (Cancer Res 2005; 65(10): 4261-72)

    Introduction

    Resistance to apoptosis is a hallmark of cancer. Although

    numerous molecular events can enable cells to survive in the face

    of proapoptotic stimuli, the full repertoire of mechanisms used by

    cancer cells to resist programmed cell death has yet to be

    established. In the physiologic setting, the transcription factor p53

    plays a major role in regulating cell survival, in addition to several

    other cellular processes that include cell cycle checkpoint control(1), DNA repair and senescence (2, 3). It is therefore not surprising

    that p53 is among the most commonly affected molecules in

    cancer, >50% of malignant human tumors bearing mutations in the

    TP53gene. p53 is activated in response to a multitude of signals,

    including oncogene activation, hypoxia (4), and DNA damage (5),

    and when required, induces a cell cycle checkpoint blockade and/

    or cell death through modification of the expression level of several

    target genes. p53 target genes include, among many others, Bax(6),

    Mdm2 (7), Apaf1 (8), Fas (9), p21WAF1 , and DR5 (10). The

    consequences of the induction of these target genes are a function

    of the cellular context.

    Recently, a new mode of action has been proposed for the

    proapoptotic activity of p53 (11). According to this view, p53 may

    function in a transcription-independent manner and may induce

    cell death by modifying mitochondrial membrane permeability,

    thereby activating the intrinsic apoptotic pathway. Regulation of

    p53 activity occurs at several levels, including post-translational

    modification (phosphorylation and acetylation), induction of the

    Mdm2 protein, which in turn induces ubiquitin-mediated p53degradation (12), and intracellular localization. To act as a

    transcription factor, p53 must translocate from the cytosol to the

    nucleus. In contrast, its transcription-independent activity

    requires its presence at the mitochondrial surface where it can

    associate with the mitochondrial regulators of apoptosis Bax and

    Bcl-XL (13). Localization is regulated in part by Mdm2, which

    sequesters p53 and forces its translocation back toward the

    cytosol, augmenting its availability for mitochondrial association

    (14, 15).

    Recent evidence suggests that protein kinase C (PKC) isozymes

    are involved in the regulation of the p53 DNA-binding activity by

    directly phosphorylating p53 on the Ser378 residue (16). It has also

    been proposed that PKCy regulates p53 expression at thetranscriptional level and contributes to the accumulation of the

    p53 protein (17). PKC family members are serine/threonine

    protein kinases that are divided into three subfamilies depending

    on their structure and mode of activation. The classic PKC

    subfamily includes the a, hI, hII, and g isotypes, which are

    activated by diacylglycerol and phosphatidylserine in a calcium-

    dependent manner. The atypical PKCs, which constitute the

    second subfamily, include the ~ and Eisotypes that are activated

    by phosphatidylserine alone. Finally, the novel PKC subfamily is

    composed of the y, q, D, and u, isotypes, which are activated by

    diacylglycerol and phosphatidylserine in a calcium-independent

    manner. These kinases play a role in multiple cellular functions,

    one of them being apoptosis. However, the precise role played byeach of the different isotypes in programmed cell death (PCD)

    remains to be fully elucidated. For example, the novel PKC, PKCy,

    is considered proapoptotic and is cleaved by caspase-3 (18, 19),

    one of the effector caspases whose activation is common to

    several apoptotic pathways. The resulting cleaved fragment is

    proapoptotic and initiates an apoptotic amplification loop by

    targeting its own activator, caspase-3. On the other hand, the full-

    size PKCy can mediate apoptosis at the mitochondrial level

    (20, 21). In contrast, the classic PKCs, PKCa, PKChI, and PKChIIare believed to exert antiapoptotic activity. PKCa, for example, is

    able to phosphorylate and stabilize the antiapoptotic molecule

    Bcl-2 (22). Several other reports show data suggesting that classic

    Note: E. Meyer and J-Y. Vollmer contributed equally to this work.Requests for reprints: Ivan Stamenkovic, Institute of Pathology, Centre

    Hospitalier Universitaire Vaudois, Rue du Bugnon 25, 1011 Lausanne, Switzerland.Phone: 121-314-7136; Fax: 121-314-7110; E-mail: [email protected] American Association for Cancer Research.

    www.aacrjournals.org 4261 Cancer Res 2005; 65: (10). May 15, 2005

    Research Article

  • 8/12/2019 Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase CPotentiated, p53-Mediated Apoptosis

    2/12

  • 8/12/2019 Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase CPotentiated, p53-Mediated Apoptosis

    3/12

    sample in binding buffer [10 mmol/L HEPES (pH 7.4), 150 mmol/L NaCl,

    2.5 mmol/L CaCl2, 1 mmol/L MgCl2, 4% bovine serum albumin] and

    incubated at room temperature in the dark for 15 minutes, centrifuged, and

    resuspended in 0.5 mL binding buffer supplemented with 15 AL

    streptavidin-FITC (eBioscience, San Diego, CA; 15 Ag/mL) and 10 AL PI

    (Sigma, 1 Ag/mL). Cells were kept at 4jC in the dark and analyzed

    immediately by flow cytometry.

    Terminal deoxynucleotidyl transferasemediated nick-end labeling

    staining.Terminal deoxynucleotidyl transferasemediated nick-end label-

    ing (TUNEL) assays were done using the In situ Cell Death Detection Kit,AP from Roche Diagnostics (Rotkreuz, Switzerland), according to the

    manufacturers instructions. Briefly, cells were grown on glass chamber

    slides. Following stimulation, the slides were washed in PBS, air-dried, and

    fixed in 4% paraformaldehyde in PBS (pH 7,4). After a second washing

    step, cells were permeabilized in 0.1% Triton X-100 with 0.1% sodium

    citrate. Samples were labeled by adding fluorescein-dUTP conjugate and

    terminal deoxynucleotidyl transferase and incubating for 1 hour at 37jC

    in the dark. After several washes in PBS, the cells were examined under a

    fluorescence microscope. The signal was converted using an AP-

    conjugated anti-fluorescein antibody. Phosphatase activity was detected

    by immersion of the slides in AP substratecontaining staining solution

    [nitroblue tetrazolium/5-bromo-4-chloro-3-indolyl phosphate mix in 0.1

    mol/L Tris (pH 9,5), 0.05 mol/L MgCl2, and 0.1 mol/L NaCl] for 45 minutes

    in the dark. The reaction was stopped by addition of double distilled

    water.

    Mitochondrial transmembrane potential (#Wm) analysis. Briefly,

    DMSO- or PMA/tetracycline-stimulated cells (5 105 cells per sample) werewashed twice with cold PBS containing 1% serum. For each sample, two

    tubes were prepared, one containing cells stained with the carbocyanine dye

    3,3V-dihexyloxacarbocyanine iodide (Molecular Probes, Eugene, OR) and the

    other containing unstained cells that served as negative controls for

    fluorescence quantification. After 15 minutes of incubation at 37jC, the

    samples were washed in PBS containing 1% serum and resuspended in a

    final volume of 200 AL PBS/1% FBS. Samples were kept at 4jC in the dark

    and analyzed immediately by flow cytometry.

    Crystal violet staining.DMSO-, PMA-, 5-FU-, or 5-FU/PMA-treated cells

    were washed with PBS. Cells were stained with 0.5% Cr ystal violet solution

    (25% methanol) for 15 minutes. After two washes with distilled water, cells

    were lysed in 1% SDS and absorbance at 620 nm was measured. All 620-nmabsorbance values were expressed as a percentage of values obtained from

    lysates of DMSO-treated cells, giving the percentage of cell viability. Cell

    death rate was calculated by subtracting this percentage from 1.

    Western BlottingSDS-PAGE protein electrophoresis and Western blotting were done

    according to standard procedures. For immunodetection of the human p53

    protein, we used the DO-1 mouse monoclonal IgG2a antibody (Santa Cruz

    Biotechnology, Santa Cruz, CA). For caspase-3/Yama/Apopain/CPP32, we

    used the Ab-1 mouse monoclonal IgG1 antibody (Oncogene Research

    Products), which recognizes both the zymogen and the cleaved form of the

    caspase-3. For humanh-actin, we used the AC-74 mouse monoclonal IgG2aantibody (Sigma). The secondary antibody was a horseradish peroxidase

    conjugated antimouse IgG and the signal was revealed using the Enhanced

    Chemiluminescence PLUS substrate solutions (Amersham Biosciences.,Otelfingen, Switzerland).

    mRNA InterferenceFor specific gene silencing, we used the small interfering RNA (siRNA)

    technology. Cells grown to 30% to 40% confluence in 24-well culture plates

    were transfected with siRNA solution consisting of oligofectamine reagent

    and the Opti-MEM (both from Invitrogen). Following transfection, cells

    were maintained in DMEM/10% FCS without antibiotics for 48 hours. We

    used the following siRNA: control, sense sequence 5V-AAAGGAAACUG-

    GAAAAAUGtt-3Vand antisense sequence 5V-CAUUUUUCCAGUUUCCUUUtt-3V;

    antiMMP-9, sense sequence 5V-CAUCACCUAUUGGAUCCAAtt-3V and

    antisense sequence 5V-UUGGAUCCAAUAGGUGAUGtt-3V; antiMMP-10,

    sense sequence 5V-GGAGGACUCCAACAAGGAUtt-3Vand antisense sequence

    5V-AUCCUUGUUGGAGUCCUCCtc-3V; antiMMP-12, sense sequence 5V-

    GGAAACAUUAUAUCACCUAtt-3V and antisense sequence 5V-UAGGUGAU-

    AUAAUGUUUCCtc-3V.

    All siRNAs were synthesized by Ambion, Inc. (Huntingdon, United

    Kingdom).

    Gelatin ZymographyAfter cell stimulation, the serum-free supernatant was concentrated with

    centricon column with a 30-kDa cutoff value (Millipore, Volketswil,

    Switzerland). The concentrated solution was mixed with Laemmli sample

    buffer, without h-mercaptoethanol, and loaded onto a gelatin (1 mg/mL)containing 10% SDS-PAGE gel. Following the completion of electrophoresis,

    the gel was washed in Triton X-100 solution and the proteases were

    activated overnight at 37jC in 5 mmol/L CaCl2and 50 mmol/L Tris (pH 8).

    The gel was then stained with Coomassie blue.

    Results

    To address the mechanisms that underlie tumor cell resistance to

    apoptosis and to elucidate the effect of PKC activation on p53-

    induced proapoptotic signaling in human cancer cells, we selected

    the SW480 cell line established from a primary human colon

    adenocarcinoma. These cells express a p53 protein containing two

    distinct mutations: G273A (ref. 23; converting an arginine to a

    histidineresidue)and C309T (converting a prolineto a serineresidue;ref. 27), which result in p53 inactivation. To study the effect of wt p53

    overexpression in SW480cells, we applied a tetracycline-regulated wt

    p53 expression system (28). PKC activation was achieved by

    exogenous administration of the phorbol ester PMA, which is a

    major activator of PKCs as well as of other signaling pathways,

    including mitogen-activated protein kinases and nuclear factor nh.SMF1 cells overexpress wild-type human p53 under the

    control of tetracycline (Tet-ON system) and initiate a celldeath programme accelerated by phorbol 12-myristate 13-acetate stimulation. SMF1 cells were generated by genetically

    modifying SW480 cells to stably express the tetracycline repressor

    protein allowing conditional expression of wt p53. Expression of

    exogenous p53 was achieved by addition of tetracycline to the

    culture medium (Fig. 1A). DMSO and PMA stimulation (Fig. 1A,lanes 1 and 3, respectively) had no effect on endogenous p53

    expression in SMF1 cell lysates, whereas lysates from SMF1 cells

    stimulated with either tetracycline alone or tetracycline and PMA

    show expression of both the endogenous and the wild-type

    exogenous protein (Fig. 1A, lanes 2 and 4, respectively). Twenty-

    four hours following induction of expression, p53-dependent cell

    death was almost negligible (mean 1%, Fig. 1B). Stimulation with

    PMA alone was not accompanied by significant induction of cell

    death, but costimulation with tetracycline and PMA resulted in 73%

    cell death at 24 hours (Fig. 1B). Observations by light microscopy

    indicated the presence of detectable cell death 12 hours following

    combined p53 induction and PMA stimulation in SMF1 cells (data

    not shown). Thus, PMA treatment sensitized SMF1 cells to p53-dependent cell death. The notion that this effect was p53 dependent

    and not related to some property of the tetracycline molecule itself

    is supported by the observation that PMA-induced sensitization did

    not occur in tetracycline-treated parental SW480 cells that do not

    express wt p53 (Fig. 1B). In addition, tetracycline stimulation of

    SW480 cells stably transfected with tetracycline repressor and the

    p53-independent gene PLSCR1did not result in a higher ratio of cell

    death than in the parental SW480 cells (data not shown).

    To ensure that the observed sensitization of p53-mediated cell

    death by PMA was not a cell typespecific event, we tested the

    effect of p53 induction in the presence and absence of PMA in the

    EB-1 colon carcinoma cell line (24) engineered to express wt p53

    MMP-9 and MMP-10 Inhibit p53/PKC-Induced Apoptosis

    www.aacrjournals.org 4263 Cancer Res 2005; 65: (10). May 15, 2005

  • 8/12/2019 Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase CPotentiated, p53-Mediated Apoptosis

    4/12

    gene under the control of the metallothionein MT-1 promoter.

    Unlike SMF1 cells, which contain mutated p53, EB-1 cells lack

    endogenous p53 expression. Treatment of EB-1 cells with ZnCl2led to a robust induction of p53 (Fig. 1C) accompanied by 20%

    cell death at 24 hours, whereas costimulation with ZnCl2 and

    PMA resulted in 40% cell death at the same time point (Fig. 1C).

    To determine whether the observed effect of p53/PMA

    costimulation occurs in a more physiologic context, we stimulated

    HCT116 colon carcinoma cells, which constitutively express wt p53

    (HCT+/+) , and their p53/ derivatives (HCT/) with the

    anticancer drug 5-FU, a known p53 inducer. Treatment of HCT+/+

    cells with 5-FU augmented endogenous p53 expression and

    induced 34% cell death at 24 hours (Fig. 1D). Consistent with the

    observations using SMF1 and EB-1 cells, 5-FUinduced cell death

    was potentiated by costimulation with PMA (Fig. 1D). 5-FU

    cytotoxicity was due, in part, to p53-independent events, as 5-FUtreatment of HCT p53/ cells resulted in 15% to 19% cell death

    (Fig. 1D). Importantly, however, PMA costimulation did not

    enhance 5-FUmediated death of HCT/ cells (Fig. 1D).Cell death observed in phorbol 12-myristate 13-acetate/

    tetracycline stimulated cells shows features of apoptosis,

    including DNA cleavage, phosphatidylserine externalization,

    and mitochondrial transmembrane potential#Wmalteration.

    Differentiation between apoptotic and necrotic cell death is

    based on several morphologic and functional criteria. We

    therefore applied a series of approaches to determine the type

    of cell death that was induced by the combination of p53 and

    PMA. Analysis of genomic DNA from dying cells revealed a DNA

    ladder (data not shown) which is a hallmark of apoptosis. p53/

    PMA costimulation was observed to induce several other

    apoptosis-specific molecular and cellular features. Thus, TUNEL

    staining (Fig. 2A) showed the presence of nick-containing DNA

    in the PMA/tetracycline-stimulated cells, which occurs in most

    forms of apoptosis as a result of endonuclease activation.

    Annexin V-biotin conjugate reactivity revealed externalization of

    phosphatidylserine (29, 30), another characteristic feature of

    programmed cell death (Fig. 2B). Similar phosphatidyl serine

    externalization was observed in EB-1 and HCT+/+ cells in

    response to ZnCl2 and 5-FU stimulation, respectively (data not

    shown). Finally, using a mitochondrial membrane carbocyanine

    dye (Fig. 2C), we could show that p53/PMA altered the

    mitochondrial transmembrane potential DWm (31, 32), in SMF1

    cells, suggesting that activation of the intrinsic mitochondrialapoptotic pathway may contribute to cell death induced by the

    combined effect of p53 and PMA.Selection of S1R6 cells that are partially resistant to p53/

    phorbol 12-myristate 13-acetate costimulation induced apo-

    ptosis.Because all three cell types displayed a similar response to

    p53/PMA costimulation, we selected SMF1 cells for subsequent

    experiments. Although most of the SMF1 cells treated with

    tetracycline and PMA underwent apoptosis, a small fraction

    (

  • 8/12/2019 Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase CPotentiated, p53-Mediated Apoptosis

    5/12

    tetracycline/PMA treatment and selection. The resulting S1R6 cell

    line displayed markedly different sensitivity to PMA/p53 compared

    with the parental SMF1 cells (Fig. 2). Thus, in the TUNEL assay

    (Fig. 2A), PMA/tetracycline stimulation induced only weak stainingin S1R6 cells, whereas the majority of the SMF1 cells were strongly

    positive (DMSO treated SMF1 and S1R6 cells, used as controls,

    were TUNEL negative). Similarly, phosphatidylserine externaliza-

    tion in S1R6 was weaker than in SMF1 (18% compared with 45%,

    Fig. 2B). In addition, whereas there was little relevant mitochon-

    drial transmembrane potential difference between DMSO- and

    PMA/tetracycline-treated S1R6 cells (DMSO: 14% and PMA/

    tetracycline: 16%), a major difference was observed in SMF1 cells

    (DMSO: 9% and PMA/tetracycline: 49%, Fig. 2C). A cell death time

    course (Fig. 2D) indicates that at 24 and 48 hours, the cell death

    observed in S1R6 represents 21% and 46%, respectively, of that

    induced in SMF1 cells.

    Phorbol 12-myristate 13-acetate/tetracyclineinduced apo-

    ptosis is caspase dependent. Most of the apoptotic signaling

    pathways described thus far are mediated by the caspase family of

    cysteine-proteases (33). The activated forms of these enzymescleave diverse targets that are important for several key cellular

    functions, including DNA repair and cell structure maintenance,

    and are also able to activate other caspases. We therefore

    addressed their implication in PMA/p53-induced apoptosis.

    Caspase-3 (Apopain/YAMA/CPP32), a member of the effector

    caspase subfamily, was observed to be activated by PMA/p53, as

    shown by the cleavage of the 32-kDa zymogen (Fig. 3). The active

    17-kDa fragment was detectable only in the PMA/tetracycline-

    stimulated SMF1 cell lysates after 24 hours (Fig. 3A). Cleavage was

    not detected in SMF1 cell lysates at 8 hours or in S1R6 cell lysates

    at any time. The 116-kDa form of poly(ADP-ribose) polymerase

    (PARP), a nuclear protein implicated in DNA repair and known to

    Figure 2. S1R6 cells display augmented resistance to PMA/tetracycline-induced apoptosis. A, TUNEL staining of SMF1 and S1R6 cells treated with DMSO or0.16Amol/L PMA + 2Amol/L tetracycline (PT) for 24 hours. B, Annexin V staining of phosphatidylserine externalization. SMF1 (black column) and S1R6 (white column)cells were stimulated with DMSO (D) or PMA + tetracycline (P + T) for 24 hours.Columns, mean of triplicate experiments;bars, SD.C, mitochondrial transmembranepotential analysis by flow cytometry. SMF1 and S1R6 cells were treated with DMSO or PMA + tetracycline (PT) for 20 hours. Solid arrows, peaks correspondingto cells with normal mitochondria;hatched arrows, peaks corresponding to cells with damaged mitochondria. D, sub-G1apoptotic DNA quantification by flow cytometry.SMF1 cells (n) or S1R6 cells (.) were stimulated with PMA + tetracycline for 24 or 48 hours. Points, mean of three different experiments; bars, SD.

    MMP-9 and MMP-10 Inhibit p53/PKC-Induced Apoptosis

    www.aacrjournals.org 4265 Cancer Res 2005; 65: (10). May 15, 2005

  • 8/12/2019 Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase CPotentiated, p53-Mediated Apoptosis

    6/12

    be a target of at least caspase-3 and caspase-7 was also found to be

    cleaved in SMF1 cells after PMA + tetracycline stimulation (data

    not shown). Finally, using the generic caspase inhibitor, z-VAD-fmk

    (34), which inactivates most members of the family, we observedcomplete inhibition of PMA/tetracycline-induced apoptosis in S1R6

    cells but incomplete inhibition in SMF1 cells (Fig. 3B).

    Because activation of Bax, which leads to the release of

    cytochrome c from mitochondria and caspase activation, is a key

    mechanism underlying p53-mediated cell death, we tested the

    effect of 5-FU in the presence and absence of PMA on HCT 116

    Bax/ cells. Cell death induced by 5-FU and 5-FU/PMA was lower

    in HCT Bax/ cells than in HCT Bax+/ counterparts (Fig. 3C).

    The observed cell death in HCT Bax/ cells most likely cor-

    responds to p53-independent cytotoxicity of 5-FU consistent with

    the notion that Bax is implicated in apoptosis induced by p53 and

    p53/PMA.

    Phorbol 12-myristate 13-acetate sensitization to p53-induced apoptosis is protein kinase C mediated. Because the

    phorbol ester PMA is known to activate several members of the

    PKC family in addition to other signaling pathways, we attempted

    to identify the PKC family members implicated in the observed

    PMA/p53-induced apoptosis using specific PKCs inhibitors,

    including rottlerin (35), a novel PKC subfamily inhibitor, Go6976

    (36, 37), a specific inhibitor of the classic PKCs a and h, and bis-

    indolylmaleimide I (38), a broad classic PKC inhibitor. When

    pretreated with either classic PKC inhibitor, both SMF1 and S1R6

    cells displayed a marked reduction in cell death following PMA/

    tetracycline stimulation (Fig. 4). In contrast, the novel PKC

    inhibitor rottlerin enhanced PMA/tetracycline sensitivity of both

    cell lines. These observations suggest that the potentiation of p53-

    induced cell death by PMA in SW480 cells is mediated by PKCs of

    the classic subfamily and antagonized by PKCs of the novel

    subfamily.S1R6 expresses higher levels of matrix metalloproteinases

    after phorbol 12-myristate 13-acetate/tetracycline stimula-

    tion. To address the mechanisms responsible for the observed

    resistance of S1R6 cells to PKC/p53-induced apoptosis, we did a

    microarray gene expression analysis of SMF1 and S1R6 cells.

    First, we compared gene expression in DMSO- and PMA/

    tetracycline-stimulated SMF1 cells. Among 474 genes that

    displayed a >3-fold induction in cells treated with PMA/

    tetracycline, were several p53 target genes, PKC-related genes,

    and genes implicated in apoptosis (data not shown). P53 target

    genes included PIG11, CDKN1A (encoding p21), and MDM2. The

    gene encoding p53 itself, TP53, was up-regulated, and as

    expected, genes encoding PKC family members, including PRKCAand PRKCZ, and genes encoding proapoptotic proteins,

    including BIK, CASP10, and APAF1 were also induced by PMA/

    p53 (data not shown).

    A second round of microarray analysis was done to compare gene

    expression in SMF1- versus S1R6-PKC/p53stimulated cells. Ninety-

    seven genes were found to display a >2 fold induction in S1R6 cells

    (Table 1 and data not shown). At least two groups of genes of

    potential interest were up-regulated. The first includes genes

    encoding proteins involved in cell adhesion and cell-cell commu-

    nication. The second includes genes encoding matrix metal-

    loproteinases (MMP), and more specifically, MMP-9, MMP-10, and

    MMP-12. Because of their implication in the regulation of cell

    Figure 3. PMA/tetracycline-dependent apoptosis is mediated by caspase-3. A, Western blot analysis of SMF1 and S1R6 cell lysates using antihuman caspase-3antibody. SMF1 and S1R6 cells were stimulated for 8 or 24 hours with DMSO ( D) or 0.16Amol/L PMA + 2Amol/L tetracycline (PT). Representative of two independentexperiments.B, sub-G1apoptotic DNA quantification by flow cytometry. SMF1 cells (white columns) and S1R6 cells (black columns) were stimulated for 15 hours withDMSO or 0.16 Amol/L PMA + 2 Amol/L tetracycline. z-VAD-fmk was added 3 hours before stimulation at a concentration of 100 Amol/L. Representative of threeindependent experiments. C, crystal violet staining: HCT 116 Bax+/ (black columns) or HCT 116 Bax/ (white columns) cells were stimulated with DMSO (D), PMA(P), 50 Ag/mL 5-FU, or 50 Ag/mL 5-FU + PMA (5-FU + P) for 24 hours. Columns, mean of experiments performed in triplicate; bars, SD.

    Cancer Research

    Cancer Res 2005; 65: (10). May 15, 2005 4266 www.aacrjournals.org

  • 8/12/2019 Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase CPotentiated, p53-Mediated Apoptosis

    7/12

    survival, we chose to address the functional role played by the

    MMPs in response to PMA/p53 stimulation.Matrix metalloproteinase activity plays a role in protein

    kinase C/p53-induced apoptosis. To test the potential role of

    MMP-mediated proteolysis in colon carcinoma cell response to

    PMA/p53 signals, we first blocked MMP proteolytic activity using

    two broad spectrum MMP inhibitors: the metal chelating agent

    o-phenanthroline (PHE), a molecule known to interfere with the

    MMP-zinc ion fixation (39), and GM6001 (40), a hydroxamic acid

    isomer known to be a potent collagenase inhibitor. Figure 5Aillustrates the effect of pretreatment of SMF1 or S1R6 cells with

    these drugs. PMA/tetracycline-induced apoptosis was augmented

    in both cell lines following PHE compared with DMSO treatment

    (64 % a nd 36% for S MF1 and 53% an d 2 4% for S1R 6,

    respectively). Similarly, treatment with GM6001 increased the

    fraction of cells undergoing apoptosis in response to PMA/

    tetracycline to 75% and 35% among SMF1 and S1R6 cells,

    respectively. Thus, MMP activity seems to provide the tumor

    cells with partial protection against apoptosis induced by p53/

    PKC costimulation. To make sure that these effects were not

    specific to the DNA degradation marker of apoptosis, we did

    Annexin V binding analysis and found that both inhibitors affect

    phosphatidylserine externalization in a similar way (data not

    shown).Matrix metalloproteinase 9 and matrix metalloproteinase

    10specific small interfering RNAs sensitize S1R6 cells

    towards phorbol 12-myristate 13-acetate/tetracyclinein-

    duced cell death. To identify the MMPs that confer protection

    to S1R6 cells from PKC/p53-induced cell death, we did siRNA-

    transient transfection experiments (41) to specifically silence MMP-

    9, MMP-10, and MMP-12 transcripts. Each siRNA induced areduction of the corresponding mRNA expression as detected by

    real-time PCR analysis (Fig. 5B-D). MMP-9 and MMP-10 gene

    silencing was observed to be followed by corresponding cell

    sensitization to PMA/tetracycline-induced cell death (Fig. 5Band C). MMP-9 and MMP-10specific siRNAs induced 257% and

    161% sensitization, respectively. In contrast, MMP-12 siRNA failed

    to modify S1R6 cell sensitivity to PKC/p53-induced apoptosis

    (Fig. 5D). Thus, consistent with the results obtained using chemical

    MMP inhibitors, MMP-9 and MMp-10 expression may play

    a protective role in colon cancer against PKC/p53-induced

    apoptosis.

    This notion is supported by the observation that sensitization ofEB-1 cells, which express MMP-9, to ZnCl2/PMA-mediatedapoptosis could also be achieved by MMP-9specific siRNA, witha 30% increase in cell death (data not shown). By contrast, HCT+/+

    cells, which are negative for MMP-9 expression, as assessed bygelatin zymography, displayed no sensitization to p53-mediatedcell death following transfection with MMP-9 siRNA (data notshown).

    p53 down-regulates expression of matrix metalloproteinase

    9 at the transcriptional level. p53 is known to induce the

    expression of several MMPs, but its effect on MMP-9 has not been

    fully explored. To determine the effect of p53 on MMP-9

    expression, we quantified the mRNA expression of the gene

    encoding MMP-9 by real-time PCR in response to p53 induction.

    Surprisingly, p53 was observed to be a negative regulator of the

    Figure 4. PMA sensitization to p53-induced apoptosis is PKC mediated. Sub-G1apoptotic DNA quantification by flow cytometry. A, SMF1 cells were stimulatedfor 24 hours with DMSO or 0.16 Amol/L PMA + 2 Amol/L tetracycline. PKCinhibitors bisindolylmaleimide I (1 Amol/L), Go6976 (1 Amol/L), or rottlerin(25 Amol/L) were added 3 hours before stimulation (white columns, no

    pretreatment with PKC inhibitors; black columns, pretreatment with Go6976;dotted columns, pretreatment with bisindolylmaleimide I; hatched columns,pretreatment with rottlerin). B, S1R6 cells were treated as in ( A). Representativeof three independent experiments. Table 1.S1R6 cells display augmented MMP expression

    compared with SMF1 cells

    Cluster Overexpression Gene Description

    MMPs 4 MMP-9 MMP-9

    (gelatinase B)

    4 MMP-10 MMP-10

    (stromelysin 2)

    3 MMP-12 MMP-12

    (macrophage

    elastase)

    Cell adhesion 8 CLDN7

    Claudin 75 CDH1 E-cadherin

    4 OCLN Occludin

    2 CLDN1 Claudin1

    2 CLDN11 Claudin 11

    NOTE: SMF1 and S1R6 cells were treated for 8 h with 0.16 Amol/L

    PMA + 2 Amol/L tetracyclin, and mRNA was extracted and amplified

    for microarray analysis. Clusters of genes that were upregulated in

    PMA/tetracyclin-stimulated S1R6 cells compared with PMA/tetracy-

    clin-treated SMF1 cells. The fold increase was rounded off to the

    nearest whole number.

    MMP-9 and MMP-10 Inhibit p53/PKC-Induced Apoptosis

    www.aacrjournals.org 4267 Cancer Res 2005; 65: (10). May 15, 2005

  • 8/12/2019 Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase CPotentiated, p53-Mediated Apoptosis

    8/12

    MMP-9 gene. Figure 6A shows relative quantification using the

    cyclophilin A mRNA as endogenous control. As expected, PMA

    induced MMP-9 (15.6-fold), but the addition of tetracycline

    together with PMA reduced this induction to 6.07-fold. It is

    important to note that the mRNA was extracted from PMA +

    tetracycline stimulated SMF1 cells before the onset of cell death.To eliminate the possibility of an effect due to intrinsic properties

    of the tetracycline molecule (independent of p53 induction),

    we did the same experiment in the SW480 parental cells. Weobserved a similar induction by PMA, but the addition of

    tetracycline failed to down-regulate MMP-9 (data no shown).These observations suggest that p53 is a negative regulator ofMMP-9 mRNA expression. Accordingly, MMP-9 protein levels, as

    assessed by gelatin zymography, were decreased by p53 (Fig. 6B).

    Discussion

    Inactivation of p53-dependent cell death by mutation or

    deletion of the TP53 gene is one of the key events incarcinogenesis. Nevertheless, many malignant tumors retain p53expression and function, and an understanding of the mecha-

    nisms that potentiate p53-mediated tumor cell death as well asthose that tumor cells deploy to survive in the face of p53signals may have important implications in cancer treatment.

    One of the principal observations of the present work is that

    PMA treatment of three different colon carcinoma cell lines

    expressing inducible exogenous or constitutive endogenous p53

    accelerates and enhances cell death induced by p53 expression

    alone. Importantly, cell death triggered as a result of 5-FU

    induced endogenous p53 expression, which reflects a more

    physi ologic situation than tetracycline- or ZnCl2-mediated

    induction of exogenously introduced p53, was augmented in

    the presence of PMA.

    Several criteria support the notion that the observed cell death

    was apoptotic. First, phosphatidylserine, normally present in theinner layer of the cytoplasmic membrane, was externalized in all

    three cell types. Second, SMF1 cells were positive for TUNEL

    staining. Third, chromatolysis and DNA fragmentation occurred in

    a nonstochastic manner. Fourth, caspases were involved in this

    process as shown by the cleavage of PARP and caspase-3. However,

    z-VAD-fmk was not able to completely abrogate SMF1 cell death

    even at a high concentration (100 Amol/L), suggesting that the cell

    death we observed is predominantly but not exclusively caspase

    dependent. Interestingly, mitochondria were also affected as shown

    by the modification of the mitochondrial transmembrane potential,

    suggesting that the intrinsic mitochondrial apoptotic pathway is

    also involved here. As might be expected, the use of HCT116 cells

    Figure 5. PKC/p53-induced apoptosis is inhibited by MMP inhibitors.A, sub-G1 apoptotic DNA quantification by flow cytometry. SMF1 andS1R6 cells were stimulated for 24 hours with 0.16 Amol/L PMA + 2 Amol/Ltetracycline. No inhibitors (white columns) or MMP inhibitors, 100 Amol/LPHE (black columns), 100Amol/L GM (dotted columns), and 100Amol/L GM+

    (hatched columns) were added 3 hours before stimulation. Representativeof three independent experiments. B, sub-G1apoptotic DNA quantificationby flow cytometry (% specific cell death) as a function of MMP expression.S1R6 cells were transfected with control siRNA (V) or MMP-9specificsiRNA (M) and 48 hours later, stimulated with 0.16 Amol/L PMA + 2 Amol/Ltetracycline for 24 hours. Expression was determined by quantification ofMMP-9 mRNA by real-time PCR 48 hours post-transfection in S1R6cells (relative expression with respect to control siRNA transfected cells).C and D, experiment was performed as in (B) but using MMP-10 siRNA(C) and MMP-12 siRNA (D). Representative of three independent experiments.

    Cancer Research

    Cancer Res 2005; 65: (10). May 15, 2005 4268 www.aacrjournals.org

  • 8/12/2019 Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase CPotentiated, p53-Mediated Apoptosis

    9/12

    with homozygous deletion of Bax showed that Bax is involved in

    p53/PMA-mediated apoptosis.

    The mechanism that underlies PMA-mediated potentiation of

    p53-induced cell death is consistent with PKC involvement.

    Phorbol esters activate the PKC family whose members

    phosphorylate and thereby regulate the activity of numerous

    cellular substrates, possibly including p53 itself. p53 activity is

    regulated, at least in part, by phosphorylation at several levels.

    First, p53 stability is increased by the activity of several kinases,

    including ATM, ATR, Chk1, and Chk2 (42), which directlyphosphorylate p53 and/or its negative regulator MDM2 (43).

    Second, by phosphorylating sites at its NH2 terminus, kinases can

    inactivate a nuclear export signal and thereby regulate the cellular

    localization of p53 (44). Third, the activity itself of p53 is prone to

    regulation by kinases. Thus, phosphorylation of p53 has been

    shown to control both its DNA binding ability and its

    transcriptional activity. In addition, p53 can be phosphorylated

    by the c-Jun NH2-terminal protein kinases, known to be activated

    by PMA-induced PKC activity (45), and by the CAK (46) complex

    (CDK7, cyclin H, and p36MAT1). Changes in p53 phosphorylation

    may therefore provide at least one mechanism to explain the

    PKC-mediated sensitization of p53-dependent cell death observed

    in this study.Members of the PKC family have been shown to interact with

    p53 and modulate p53-induced apoptosis and cell growth arrest.

    The classic PKCs are able to induce p53 translocation to the

    nucleus and the G1cell cycle control checkpoint (47) and PKCyhas

    been shown to induce an accumulation of p53 (17). Moreover, p53

    contains a PKC binding site in the central oligomerization domain

    and can be directly phosphorylated by PKCs at Ser378 and by PKCa

    and PKC~ at Ser371 (48). In addition, a recent study has shown

    activation by cyclin-dependent kinase 2/cyclin A and PKC of the

    p53-specific DNA sequence binding ability (16).

    Because all novel PKCs and classic PKCs contain a phorbol ester

    binding motif, it would seem reasonable to expect that either or

    both of these subfamilies are implicated in the potentiation of p53-

    mediated cell death. Consistent with this notion, both classic PKC

    and novel PKC inhibitors affected p53-dependent apoptosis but in

    opposite directions. Thus, bis-indolylmaleimide I, which inhibits

    PKCa, PKChI, PKChII, PJCg, PKCy, and PKCq isozymes, was the

    strongest inhibitor of PMA-dependent cell death potentiation, with

    Go6976, which inhibits PKCa and PKChI isozymes, displaying a

    lesser degree of inhibition. Based on these findings, it would seem

    that the observed PMA-associated cell death enhancement wasmediated by the PKCa, PKChI, PKChII, PKCg, and PKCqisozymes.

    In contrast, PKCyand PKCuseemed to have an antiapoptotic effect

    in p53-stimulated SW480 cells.

    PKC targets independent of the p53 may contribute to the

    observed synergistic proapoptotic effect of PKC activation and p53

    induction (Fig. 7). The function of PKCy, which is both a substrate

    and an activator of caspase-3, has been suggested to be linked to

    the mitochondrial release of cytochrome c during apoptosis.

    Consistent with these observations, the phorbol ester 12-O-

    tetradecanoylphorbol-13-acetate induces the release of cytochrome

    c and activation of caspase-3. Our observation that inhibition

    of PKCy by rottlerin sensitized colon carcinoma cells to p53 in

    our in vitro system may therefore be somewhat surprising andwould seem to contradict the evidence suggesting that PKCy is

    proapoptotic. However, rottlerin has been suggested to inhibit

    the intrinsic cell death pathway but to potentiate extrinsic cell

    death pathways (49). Thus, depending on the type of proapoptotic

    stimulus, rottlerin may be expected to exert a stimulatory or

    inhibitory effect. In the case of our model system, the observed

    effect of rottlerin is consistent with the possibility that even if there

    was activation of the intrinsic pathway, as suggested by changes

    in the mitochondrial transmembrane potential, the extrinsic pro-

    apoptotic pathway was likely to be dominant.

    Selection of SMF1 cells that were resistant to PMA/p53-induced

    cell death (termed S1R6), allowed the identification of potential

    mechanisms that underlie the resistance. S1R6 cells were observed

    to retain tetracycline-regulated p53 expression, such that theirresistance could not be attributed to loss of p53 induction.

    However, comparison of the gene expression profiles of S1R6 and

    SMF1 cells showed up-regulation of a relatively small number of

    genes among which two groups seemed of immediate interest:

    MMPs and cell adhesion molecules.

    Members of the MMP family, also called matrixins, are cysteine

    proteases with zinc iondependent proteolytic activity (reviewed in

    refs. 5052). Thus far, >20 MMPs have been identified in mammals

    and are currently grouped according to structural features (51). The

    majority of MMPs are secreted into the extracellular matrix (ECM),

    a subset being membrane bound by virtue of a transmembrane or

    PI-linked domain (51). All are produced as inactive zymogens and

    are activated by proteolysis. MMPs were originally described asbeing primarily involved in the proteolysis of ECM components

    and therefore believed to primarily mediate tissue remodeling.

    However, it is becoming increasingly clear that these enzymes play

    a critical role in a broad range of physiologic and pathologic

    processes. Thus, MMPs can cleave and regulate the activity of a

    host of growth factors, chemokines, and cytokines as well as cell

    surface adhesion receptors and proteoglycans involved in intercel-

    lular communication and cell migration that are directly implicat-

    ed in tumor progression and metastasis (5052).

    Functional analysis in the present study showed that

    expression of MMP-9, and to a lesser extent that of MMP-10

    but not MMP-12, confer partial resistance to PMA/p53-mediated

    Figure 6. p53 is a negative regulator of MMP-9 mRNA and protein expression.A, relative quantification of MMP-9 mRNA expression by real-time PCR. SMF1cells were stimulated with DMSO (D), 0.16Amol/L PMA (P), 2 Amol/Ltetracycline (T), or PMA + tetracycline (PT) for 24 hours. The 2DDCT value

    represents MMP-9 expression relative to that in DMSO treated cells.B, MMP-9activity as illustrated by gelatin zymography. SMF1 cells were treated asindicated for 24 hours. Columns, mean of experiments performed in triplicate;bars, SD.

    MMP-9 and MMP-10 Inhibit p53/PKC-Induced Apoptosis

    www.aacrjournals.org 4269 Cancer Res 2005; 65: (10). May 15, 2005

  • 8/12/2019 Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase CPotentiated, p53-Mediated Apoptosis

    10/12

    cell death. A potential issue in our approach is the use of

    tetracycline to induce p53, because tetracycline and its derivative

    doxycyline have been shown to inhibit MMP activity and MMP-2

    gene expression. However, inhibition occurs with an IC50 of 30 to

    300 Amol/L, whereas we used a concentration of 2 Amol/L, well

    below the reported inhibitory levels. Accordingly, gelatin

    zymography assays showed no difference between lysate and

    supernatant MMP activity from cells treated or not with

    tetracycline. The presence of tetracycline therefore seems to

    have no effect on the activity of MMP-9 in our system at the

    concentration used.

    Similar to PKC, MMP activity may have both proapoptotic and

    antiapoptotic effects. Proapoptotic effects can be due to the

    proteolytic degradation of ECM proteins, such as laminin, that

    serve as ligands of the integrin class of cell surface adhesionreceptors (5052). In the absence of ligand, specific subclasses of

    integrins can no longer trigger signals necessary for normal

    epithelial and endothelial cell survival, resulting in a form of

    apoptosis known as anoikis. Conversely, ECM degradation can

    result in the release and augmented bioactivity of potent survival

    factors, such as insulin-like growth factor I (IGF-I) that may

    directly promote tumor cell resistance to proapoptotic signals and

    favor tumor growth. MMP-mediated cleavage of the integralmembrane ligand of the death-inducing receptor Fas (FasL) can

    result in increased or reduced apoptosis depending on the

    physiologic context (5052). The proapoptotic or antiapoptotic

    effect of any given MMP may therefore depend on the local cellular

    production and relative availability of proteolytically released and

    activated death-inducing and/or growth and survival factors.There are several potential mechanisms whereby MMP-9 may

    promote tumor cell resistance to proapoptotic stimuli. MMP-9

    can release and activate transforming growth factor h (TGF-h; ref.

    50), cleave several chemokines and activate other MMPs including

    MMP-1 (4951). TGF-h can promote survival in transformed cells,

    as can the chemokine IL-8, whose function is potentiated by

    MMP-9mediated proteolysis (5052). Activation of MMP-1 leads

    to the release of IGF-I from ECM-sequestered IGF binding

    proteins (5052). However, the full range of MMP substrates is far

    from elucidated, and it is quite possible that proteolytic

    activation or inactivation of other regulators of cell survival

    may be implicated in the observed antiapoptotic effect of MMP-9

    expression. High expression of MMP-9 is often associated withpoor prognosis in cancer patients and, in at least some tumor

    types, correlates with increased angiogenesis and metastatic

    proclivity (5052).

    Interestingly, regulation of MMP expression is reported to be

    subjected, at least in part, to p53 control. Thus, p53 has been

    shown to induce transcriptional activation of the MMP-2 (53) gene

    promoter and to attenuate the production of MMP-1 (54) and

    MMP-13 (55). Consistent with earlier work by others (53), we have

    found that MMP-2 expression is enhanced by p53 (data not shown).

    In the present study, we observed that p53 represses the human

    MMP-9gene, which in and of itself may contribute to p53-induced

    cell death. It is also interesting to note that the two MMPs (MMP-2

    Figure 7. Schematic representation of the effect of PKC isoforms and MMP-9 on p53-mediated cell death. PKCs directly affect p53 phosphorylation, localization,and expression at the protein level. Classic PKCs such as the a and h isoforms synergize with p53 in inducing apoptosis, by directly influencing p53 activity and/orpromoting nonp53-mediated proapoptotic events, whereas PKCy has an antagonistic effect. In addition to inducing intracellular proapoptotic pathways, p53 inhibitstranscription of MMP-9, which in its active form, promotes cell survival by proteolytically activating and/or inhibiting extracellular survival and proapoptotic factors,respectively. These factors may function in an autocrine and/or paracrine manner.

    Cancer Research

    Cancer Res 2005; 65: (10). May 15, 2005 4270 www.aacrjournals.org

  • 8/12/2019 Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase CPotentiated, p53-Mediated Apoptosis

    11/12

    and MMP-9) that form the gelatinase subfamily of MMPs, based on

    structural similarities and shared substrate specificity (5052), are

    differentially regulated by p53.

    The second group of genes of potential interest found to be

    overexpressed in S1R6 cells encode proteins involved in cell-cellcommunication and/or adhesion. Several are components of tight

    junctions (zonula occludens) and contribute to the epithelial

    barrier function (claudin and occludin family of proteins),

    whereas E-cadherin is part of the zonula adherens junctions thatgenerate signals sensed by p53 (56). Although further work will berequired to address their role, it is conceivable that elevated

    expression of these proteins contributes to resistance to PKC/p53-

    dependent apoptosis.

    Taken together, our observations provide evidence that PKCs are

    able to potentiate p53-induced apoptosis and that expression ofMMP-9 and MMP-10 confers partial protection against the proapop-

    totic PKC/p53 signals. Understanding of the precise mechanism of

    PKC/p53 synergy and identification of the relevant PKC targets mayhelp develop therapeutic strategies that, in combination with selective

    MMP inhibitors, could augment the effectiveness of tumor cell

    elimination.

    Acknowledgments

    Received 8/16/2004; revised 12/15/2004; accepted 1/4/2005.

    Grant support: National Center for Competence in Research Molecular Oncology,Swiss National Scientific Foundation grant 31-65090.01, and Oncosuisse grant 1267-08-2002.

    The costs of publication of this article were defrayed in part by the payment of pagecharges. This article must therefore be hereby marked advertisementin accordancewith 18 U.S.C. Section 1734 solely to indicate this fact.

    We thank Josiane Wyniger and Otto Hagenbuchle for providing the cDNAmicroarrays, Bert Vogelstein for the HCT cell lines, Phil Shaw for EB-1 cells, RichardIggo for the wt human p53 cDNA, and Nathalie Mayran for her advice regarding siRNAexperiments.

    References

    1. Schwartz D, Rotter V. p53-dependent cell cyclecontrol: response to genotoxic stress. Semin CancerBiol 1998;8:32536.

    2. Hofseth LJ, Hussain SP, Harris CC. p53: 25 years afterits discovery. Trends Pharmacol Sci 2004;25:17781.

    3. Oren M. Decision making by p53: life, death andcancer. Cell Death Differ 2003;10:43142.

    4. Koumenis C, Alarcon R, Hammond E, et al. Regulationof p53 by hypoxia: dissociation of transcriptionalrepression and apoptosis from p53-dependent trans-activation. Mol Cell Biol 2001;21:1297310.

    5. Selivanova G, Wiman KG. p53: a cell cycle regulatoractivated by DNA damage. Adv Cancer Res 1995;66:14380.

    6. Miyashita T, Reed JC. Tumor suppressor p53 is a directtranscriptional activator of the human bax gene. Cell1995;80:2939.

    7. Juven T, Barak Y, Zauberman A, George DL, Oren M.Wild type p53 can mediate sequence-specific trans-

    activation of an internal promoter within the mdm2gene. Oncogene 1993;8:34116.

    8. Moroni MC, Hickman ES, Denchi EL, et al. Apaf-1 is atranscriptional target for E2F and p53. Nat Cell Biol2001;3:5528.

    9. Owen-Schaub LB, Zhang W, Cusack JC, et al. Wild-typehuman p53 and a temperature-sensitive mutant induceFas/APO-1 expression. Mol Cell Biol 1995;15:303240.

    10. Takimoto R, El-Deiry WS. Wild-type p53 trans-activates the KILLER/DR5 gene through an intronicsequence-specific DNA-binding site. Oncogene 2000;19:173543.

    11. Leu JI, Dumont P, Hafey M, Murphy ME, George DL.Mitochondrial p53 activates Bak and causes disrup-tion of a Bak-Mcl1 complex. Nat Cell Biol 2004;6:44350.

    12. Michael D, Oren M. The p53-Mdm2 module and theubiquitin system. Semin Cancer Biol 2003;13:4958.

    13. Chipuk JE, Kuwana T, Bouchier-Hayes L, et al. Directactivation of Bax by p53 mediates mitochondrialmembrane permeabilization and apoptosis. Science 2004;303:10104.

    14. Momand J, Zambetti GP, Olson DC, George D, LevineAJ. The mdm-2 oncogene product forms a complex withthe p53 protein and inhibits p53-mediated trans-activation. Cell 1992;69:123745.

    15. Geyer RK, Yu ZK, Maki CG. The MDM2 RING-fingerdomain is required to promote p53 nuclear export. NatCell Biol 2000;2:56973.

    16. Pospisilova S, Brazda V, Kucharikova K, et al.Activation of the DNA-binding ability of latent p53protein by protein kinase C is abolished by proteinkinase CK2. Biochem J 2004;378:93947.

    17. Abbas T, White D, Hui L, Yoshida K, Foster DA,Bargonetti J. Inhibition of human p53 basal transcrip-

    tion by down-regulation of protein kinase Cy. J BiolChem 2004;279:99707.

    18. Koriyama H, Kouchi Z, Umeda T, et al. Proteolyticactivation of protein kinase C y and epsilon by caspase-3 in U937 cells during chemotherapeutic agent-inducedapoptosis. Cell Signal 1999;11:8318.

    19. Leverrier S, Vallentin A, Joubert D. Positive feedbackof protein kinase C proteolytic activation duringapoptosis. Biochem J 2002;368:90513.

    20. Li L, Lorenzo PS, Bogi K, Blumberg PM, Yuspa SH.Protein kinase C y targets mitochondria, altersmitochondrial membrane potential, and induces apo-ptosis in normal and neoplastic keratinocytes whenoverexpressed by an adenoviral vector. Mol Cell Biol1999;19:854758.

    21. Brodie C, Blumberg PM. Regulation of cell apoptosisby protein kinase C y . Apoptosis 2003;8:1927.

    22. Ruvolo PP, Deng X, Carr BK, May WS. A functionalrole for mitochondrial protein kinase Ca in Bcl2phosphorylation and suppression of apoptosis. J Biol

    Chem 1998;273:2543642.23. Rodrigues NR, Rowan A, Smith ME, et al. p53

    mutations in colorectal cancer. Proc Natl Acad SciU S A 1990;19:75559.

    24. Shaw P, Bovey R, Tardy S, Sahli R, Sordat B, Costa J.Induction of apoptosis by wild-type p53 in a humancolon tumor-derived cell line. Proc Natl Acad Sci U S A1992;89:44959.

    25. The R engine was used with the SMA library(Statistic for MicroArray, Speeds laboratory, Universityof Berkeley, CA) and the BRAJU library (Bengtsson H.,University of Lund, Sweden).

    26. Sanceau J, Truchet S, Bauvois B. Matrix metal-loproteinase-9 silencing by RNA interference triggersthe migratory-adhesive switch in Ewings sarcoma cells.J Biol Chem 2003;278:3653746.

    27. Abarzua P, LoSardo JE, Gubler ML, Neri A. Microin-jection of monoclonal antibody PAb421 into human

    SW480 colorectal carcinoma cells restores the tran-scription activation function to mutant p53. Cancer Res1995;55:34904.

    28. Gossen M, Bujard H. Tight control of geneexpression in mammalian cells by tetracycline-responsive promoters. Proc Natl Acad Sci U S A 1992;89:554751.

    29. Martin SJ, Reutelingsperger CP, McGahon AJ, et al.Early redistribution of plasma membrane phosphati-dylserine is a general feature of apoptosis regardlessof the initiating stimulus: inhibition by overexpres-sion of Bcl-2 and Abl. J Exp Med 1995;182:154556.

    30. Boersma AW, Nooter K, Oostrum RG, Stoter G.Quantification of apoptotic cells with fluoresceinisothiocyanate-labeled Annexin V in Chinese hamsterovary cell cultures treated with cisplatin. Cytometry1996;24:12330.

    31. Quillet-Mary A, Jaffrezou JP, Mansat V, Bordier C,Naval J, Laurent G. Implication of mitochondrialhydrogen peroxide generation in ceramide-inducedapoptosis. J Biol Chem 1997;272:2138895.

    32. Zamzami N, Marchetti P, Castedo M, et al. Sequentialreduction of mitochondrial transmembrane potentialand generation of reactive oxygen species in earlyprogrammed cell death. J Exp Med 1995;182:36777.

    33. Nunez G, Benedict MA, Hu Y, Inohara N. Caspases:the proteases of the apoptotic pathway. Oncogene1998;17:323745.

    34. Slee EA, Zhu H, Chow SC, MacFarlane M, NicholsonDW, Cohen GM. Benzyloxycarbonyl-Val-Ala-Asp (OMe)fluoromethylketone (Z-VAD.FMK) inhibits apoptosis byblocking the processing of CPP32. Biochem J 1996;315:214.

    35. Gschwendt M, Muller HJ, Kielbassa K, et al. Rottlerin,a novel protein kinase inhibitor. Biochem Biophys ResCommun 1994;199:938.

    36. Qatsha KA, Rudolph C, Marme D, Schachtele C, May

    WS. Go 6976, a selective inhibitor of protein kinase C,is a potent antagonist of human immunodeficiencyvirus 1 inducti on from latent /low-l evel-pro ducingreservoir cells in vitro. Proc Natl Acad Sci U S A1993;90:46748.

    37. Martiny-Baron G, Kazanietz MG, Mischak H, et al.Selective inhibition of protein kinase C isozymes bythe indolocarbazole Go 6976. J Biol Chem 1993;268:91947.

    38. Toullec D, Pianetti P, Coste H, et al. The bisindo-lylmaleimide GF 109203X is a potent and selectiveinhibitor of protein kinase C. J Biol Chem 1991;266:1577181.

    39. Uhm JH, Dooley NP, Villemure JG, Yong VW. Gliomainvasion in vitro : regulation by matrix metalloprotease-2 and protein kinase C. Clin Exp Metastasis 1996;14:42133.

    40. Grobelny D, Poncz L, Galardy RE. Inhibition of

    human skin fibroblast collagenase, thermolysin, andPseudomonas aeruginosa elastase by peptide hydroxa-mic acids. Biochemistry 1992;31:71524.

    41. Elbashir SM, Harborth J, Lendeckel W, Yalcin A,Weber K, Tuschl T. Duplexes of 21-nucleotide RNAsmediate RNA interference in cultured mammalian cells.Nature 2001;411:4948.

    42. Caspari T. How to activate p53. Curr Biol 2000;10:R3157.

    43. Khosravi R, Maya R, Gottlieb T, Oren M, Shiloh Y,Shkedy D. Rapid ATM-dependent phosphorylation ofMDM2 precedes p53 accumulation in response toDNA damage. Proc Natl Acad Sci U S A 1999;96:149737.

    44. Zhang Y, Xiong Y. A p53 amino-terminal nuclearexport signal inhibited by DNA damage-inducedphosphorylation. Science 2001;292:19105.

    MMP-9 and MMP-10 Inhibit p53/PKC-Induced Apoptosis

    www.aacrjournals.org 4271 Cancer Res 2005; 65: (10). May 15, 2005

  • 8/12/2019 Matrix Metalloproteinases 9 and 10 Inhibit Protein Kinase CPotentiated, p53-Mediated Apoptosis

    12/12

    45. Hu MC, Qiu WR, Wang YP. JNK1, JNK2 and JNK3 arep53 N-terminal serine 34 kinases. Oncogene 1997;15:227787.

    46. Ko LJ, Shieh SY, Chen X, et al. p53 is phosphorylatedby CDK7-cyclin H in a p36MAT1-dependent manner.Mol Cell Biol 1997;17:72209.

    47. Scotto C, Delphin C, Deloulme JC, Baudier J.Concerted regulation of wild-type p53 nuclear accumu-lation and activation by S100B and calcium-dependentprotein kinase C. Mol Cell Biol 1999;19:716880.

    48. Youmell M, Park SJ, Basu S, Price BD. Regulation of

    the p53 protein by protein kinase C a and proteinkinase C ~. Biochem Biophys Res Commun 1998;245:5148.

    49. Basu A, Miura A. Differential regulation of extrinsicand intrinsic cell death pathways by protein kinase C.Int J Mol Med 2002;10:5415.

    50. Stamenkovic I. Extracellular matrix remodelling: therole of matrix metalloproteinases. J Pathol 2003;200:44864.

    51. Egelblad M, Werb Z. New functions for the matrixmetalloproteinases in cancer progression. Nat RevCancer 2002;2:16174.

    52. McCawley LJ, Matrisian LM. Matrix metalloprotei-nases: theyre not just for matrix anymore! Curr Opin

    Cell Biol 2001;13:53440.53. Bian J, Sun Y. Transcriptional activation by p53 of thehuman type IV collagenase (gelatinase A or matrix

    metalloproteinase 2) promoter. Mol Cell Biol 1997;17:63308.

    54. Sun Y, Sun Y, Wenger L, Rutter JL, Brinckerhoff CE,Cheung HS. p53 down-regulates human matrix metal-loproteinase-1 (collagenase-1) gene expression. J BiolChem 1999;274:1153540.

    55. Ala-aho R, Grenman R, Seth P, Kahari VM. Adenoviraldelivery of p53 gene suppresses expression of collage-nase-3 (MMP-13) in squamous carcinoma cells. Onco-gene 2002;21:118795.

    56. Yamaguchi M, Hirose F, Inoue YH, et al. Genetic

    link between p53 and genes required for formationof the zonula adherens junction. Cancer Sci 2004;95:43641.

    Cancer Research

    Cancer Res 2005; 65: (10). May 15, 2005 4272 www.aacrjournals.org