{alpha}3ß1 integrin regulates MMP-9 mRNA stability in immortalized keratinocytes: a novel mechanism of integrin-mediated MMP gene expression

Vandana Iyer, Kevin Pumiglia and C. Michael DiPersio*

Center for Cell Biology and Cancer Research, Albany Medical College, MC-165, 47 New Scotland Avenue, Albany, New York 12208, USA

* Author for correspondence (e-mail: dipersm{at}mail.amc.edu)

Accepted 6 January 2005


    Summary
 Top
 Summary
 Introduction
 Materials and Methods
 Results
 Discussion
 References
 
Matrix metalloproteinases facilitate cell migration and tumor invasion through their ability to proteolyse the extracellular matrix. The laminin-binding integrin {alpha}3ß1 is expressed at high levels in squamous cell carcinomas and in normal keratinocytes during cutaneous wound healing. We showed previously that {alpha}3ß1 is required for MMP-9/gelatinase B secretion in immortalized mouse keratinocytes (MK cells) and that this regulation was acquired as part of the immortalized phenotype, suggesting a possible role for {alpha}3ß1 during malignant conversion. In the current study, we identify a novel mechanism whereby {alpha}3ß1 regulates the induction of MMP-9 expression that occurs in response to activation of a MAPK kinase (MEK)/extracellular signal-regulated kinase (ERK) pathway. Inhibition of MEK/ERK signaling in wild-type MK cells with a pharmacological inhibitor, U0126, showed that ERK activation was necessary for high levels of endogenous MMP-9 gene expression and activity of a transfected MMP-9 promoter. Furthermore, activation of MEK/ERK signaling in these cells with an oncogenic mutant of Ras, RasV12, increased both endogenous MMP-9 gene expression and MMP-9 promoter activity. Experiments with {alpha}3ß1-deficient MK cells revealed that {alpha}3ß1 was required for both baseline levels and RasV12-induced levels of MMP-9 mRNA expression. However, {alpha}3ß1 was not required for RasV12-mediated activation of ERK or for ERK-dependent MMP-9 promoter activity. Direct comparison of mRNA turnover in the wild type and {alpha}3-null MK cells identified a requirement for {alpha}3ß1 in stabilization of MMP-9 mRNA transcripts. These results identify a novel function for integrins in promoting mRNA stability as a mechanism to potentiate MAPK-mediated gene expression. They also suggest a role for {alpha}3ß1 in maintaining high levels of MMP-9 mRNA expression in response to oncogenic activation of MEK/ERK signaling pathways.

Key words: {alpha}3ß1 integrin, MMP-9, Keratinocytes, mRNA stability


    Introduction
 Top
 Summary
 Introduction
 Materials and Methods
 Results
 Discussion
 References
 
Remodeling of the extracellular matrix (ECM) during tissue development, wound healing and tumor invasion requires the functions of extracellular proteinases that remodel ECM proteins at the leading edges of invading or migrating cells (Werb, 1997Go). Matrix metalloproteinases (MMPs) are secreted into the extracellular space as zymogens that are functionally activated through self-proteolytic cleavage or through proteolysis by other MMPs or proteinases (Werb, 1997Go). Activated MMPs play important roles in promoting tumor growth, progression and invasion. They can be expressed either directly by tumor cells (Canete-Soler et al., 1994Go; Jiang and Muschel, 2002Go; Pyke et al., 1992Go) or by other cells present in the tumor microenvironment, such as stromal cells or infiltrating inflammatory cells (Coussens et al., 2000Go; Werb, 1997Go).

MMP-9/gelatinase B belongs to the gelatinase subgroup of the MMP family and its expression and activity has been correlated with different stages of carcinoma progression. For example, MMP-9 has been shown to induce the angiogenic switch in a mouse model of pancreatic carcinogenesis (Bergers et al., 2000Go). Other studies have implicated MMP-9 gene activation in promoting tumor invasion and metastasis (Bernhard et al., 1994Go; Kupferman et al., 2000Go). Many potential substrates have been identified for MMP-9 that could contribute to its tumor-promoting functions, including numerous ECM proteins, other MMPs and proteinase inhibitors (Liu et al., 2000Go; Werb, 1997Go).

MMP-9 expression can be regulated at several levels. Although most published studies have focused on transcriptional control of MMP-9 (reviewed by Westermarck and Kahari, 1999Go), there is increasing evidence that its expression can also be regulated at the steps of mRNA stability, translation and protein secretion (Akool et al., 2003Go; Eberhardt et al., 2002Go; Jiang and Muschel, 2002Go; Morini et al., 2000Go; Sehgal and Thompson, 1999Go; Thant et al., 2000Go). The ability to modulate MMP-9 expression at multiple steps through distinct signaling pathways may be particularly important during malignant conversion and metastasis, when tumor cells need to induce or maintain MMP-9 levels in response to changing environmental cues. A number of distinct extracellular stimuli, including growth factors, cytokines and extracellular calcium, have been shown to regulate MMP-9 expression in normal or malignant keratinocytes via the mitogen-activated protein kinases (MAPKs) p38 and/or extracellular signal-regulated kinase (ERK) (Johansson et al., 2000Go; McCawley et al., 1998Go; Mukhopadhyay et al., 2004Go; Westermarck and Kahari, 1999Go; Zeigler et al., 1999Go). In addition, several studies have demonstrated the importance of cell adhesion in regulating the expression of MMP-9 or other MMPs in immortalized epithelial cells or carcinoma-derived cells (DiPersio et al., 2000Go; Lochter et al., 1999Go; Morini et al., 2000Go; Thomas et al., 2001Go).

Integrins are the major receptors for cell adhesion to the ECM and many integrins initiate `outside-in' signal transduction events that modulate cell functions such as gene expression, cell migration and invasion (Hynes, 2002Go). {alpha}3ß1 integrin is expressed at high levels on epithelial cells where it is a major receptor for laminin-5 (LN-5) and other basement membrane laminins (Kreidberg, 2000Go). Knockout studies in mice have revealed important roles for {alpha}3ß1 in maintaining basement membrane integrity during embryonic development of the epidermis and other tissues (DiPersio et al., 1997Go; Kreidberg et al., 1996Go). {alpha}3ß1 is also expressed at high levels in most primary and metastatic tumors (Bartolazzi et al., 1994Go; Natali et al., 1993Go; Patriarca et al., 1998Go) and several recent studies in immortalized epithelial cells or carcinoma-derived cell lines suggest that this integrin has important roles in regulating functions that promote malignant tumor growth and progression, including cell survival, migration, invasion and metastasis (Morini et al., 2000Go; Choma et al., 2004Go; Wang et al., 2004Go; Manohar et al., 2004Go). Some of these {alpha}3ß1-mediated cell functions may involve MMP-9, as {alpha}3ß1 regulates MMP-9 secretion in some immortalized and malignantly transformed epithelial cells (Morini et al., 2000Go; DiPersio et al., 2000Go). We showed previously that {alpha}3ß1-dependent secretion of MMP-9 was acquired as part of the immortalized phenotype in keratinocytes (DiPersio et al., 2000Go), consistent with a role during malignant conversion of squamous cell carcinoma (SCC). Many invasive carcinoma cells also express high levels of LN-5 (Pyke et al., 1995Go) and MMP-9 (Juarez et al., 1993Go; Morini et al., 2000Go; Pyke et al., 1992Go; Westermarck and Kahari, 1999Go), reinforcing a possible link between {alpha}3ß1-mediated adhesion and MMP-9 induction during tumor progression and invasion.

Despite the established importance of integrins in regulating MMP gene expression in immortalized or transformed epithelial cells, the mechanisms of this regulation are still unclear. In the current study, we determined the mechanism whereby {alpha}3ß1 integrin regulates the induction of MMP-9 gene expression in immortalized keratinocytes cultured on LN-5. We showed that ERK activation was necessary for high levels of MMP-9 expression and that {alpha}3ß1 was necessary for ERK-mediated induction of MMP-9 in response to oncogenic RasV12. Importantly, although {alpha}3ß1 was required for high levels of MMP-9 mRNA, it was not required for RasV12-mediated activation of ERK, or for ERK-mediated activation of a transfected MMP-9 promoter construct. However, {alpha}3ß1 was required post-transcriptionally to maintain high levels of MMP-9 mRNA, as mRNA decay experiments indicated that MMP-9 transcripts turn over more rapidly in {alpha}3-null cells than in {alpha}3ß1-expressing cells. These results identify a novel function for {alpha}3ß1 integrin in promoting mRNA stability as a mechanism to potentiate MAPK-mediated gene expression. In addition, our findings suggest an important role for {alpha}3ß1 integrin in augmenting the response of some immortalized or transformed cells to growth factors and other environmental cues, or to activated oncogenes, that stimulate MMP-9 gene expression through MAPK signaling pathways.


    Materials and Methods
 Top
 Summary
 Introduction
 Materials and Methods
 Results
 Discussion
 References
 
MK cell culture and stable transfection with {alpha}3 integrin
The MK+/+ cell line (MK-1.16) and MK–/– cell line (MK-5.4.6) were derived from keratinocytes isolated from wild-type mice or {alpha}3 integrin knockout mice, respectively (DiPersio et al., 2000Go). MK growth medium consisted of Eagle's minimum essential medium (EMEM; BioWhittaker, Walkersville, MD) supplemented with 4% FBS (BioWhittaker) from which Ca2+ had been chelated, 0.05 mM CaCl2, 0.4 µg/ml hydrocortisone (Sigma, St Louis, MO), 5 µ g/ml insulin (Sigma), 10 ng/ml EGF (Invitrogen Corporation, Carlsbad, CA), 2x10–9 M T3 (Sigma), 10 U/ml interferon {gamma} (INF{gamma}; Sigma), 100 U/ml penicillin and 100 µg/ml streptomycin (Invitrogen) and L-glutamine (Invitrogen). MK cell lines were maintained at 33°C, 8% CO2, on tissue culture plates coated with 30 µg/ml denatured rat tail collagen (Cohesion, Palo Alto, CA). For experiments, MK cells were sub-cultured on laminin-5-rich extracellular matrix (LN-5 ECM) prepared from the human squamous cell carcinoma line SCC-25 (Rheinwald and Beckett, 1981Go), as described previously (DiPersio et al., 2000Go). For experiments in Fig. 4, untreated or LN-5 ECM coated tissue culture plates were coated with fibronectin (20 µg/ml; a gift from Livingston Van De Water, Albany Medical College), vitronectin (15 µg/ml; a gift from Paula McKeown-Longo, Albany Medical College), or both fibronectin and vitronectin (20 µg/ml and 15 µg/ml, respectively), then blocked with 10 mg/ml heat-denatured BSA in PBS. For assessment of cell spreading and F-actin, parallel cultures were grown under serum-free conditions for 24 hours on glass coverslips coated with ECM proteins as described above, fixed in 4% paraformaldehyde and stained with TRITC-conjugated phalloidin (Sigma). Images were acquired on an Olympus BX60 microscope using a cooled CCD camera (Sensicam) and the software Slidebook 3.0.



View larger version (99K):
[in this window]
[in a new window]
 
Fig. 4. Culture on fibronectin and/or vitronectin does not restore MMP-9 expression in {alpha}3ß1-deficient MK–/– cells. (A) Total RNA was isolated from MK+/+ cells or MK–/– cells grown on surfaces coated with LN-5 ECM (L), LN-5 ECM plus 20 µg/ml fibronectin (L+F), 20 µg/ml fibronectin (F), 20 µg/ml fibronectin plus 15 µg/ml vitronectin (F+V), or 15 µg/ml vitronectin (V). RT-PCR was performed to compare MMP-9 and ß-actin mRNA levels. (B) Parallel cultures of MK+/+ cells (a, b, e, f) or MK–/– cells (c, d, g, h) were grown on coverslips coated with 20 µg/ml fibronectin (a-d) or 20 µg/ml fibronectin plus 15 µg/ml vitronectin (e-h). F-actin was stained with TRITC-conjugated phalloidin (b, d, f, h); the corresponding phase-contrast image is shown for each field (a, c, e, g). Arrows in d and h indicate spindle-shaped cells. Bar, 20 µm.

 

{alpha}3ß1 integrin expression was restored in MK–/– cells by stable transfection with the plasmid pCMVzeo-h{alpha}3A, which contains the full-length human {alpha}3 cDNA (a generous gift from Martin Hemler, Dana-Farber Cancer Institute, Boston, MA) cloned into the XbaI site proximal to the CMV promoter in pcDNA3.1/Zeo(+) (Invitrogen). MK–/– cells were transfected using lipofectamine reagent (Invitrogen). Pools of zeocin-resistant cells were expanded and assayed for {alpha}3ß1 surface expression by FACS analysis with the anti-human {alpha}3 monoclonal antibody P1B5 (Gibco/BRL, Gaithersburg, MD), as described previously (DiPersio et al., 2000Go).

Western blotting
MK cell lysates were prepared in cell lysis buffer (Cell Signaling Technology, Beverly, MA) and equal amounts of protein from each lysate (either 10 µg or 20 µg) was subject to reducing 10% SDS-PAGE. Proteins were then transferred to nitrocellulose membranes and assayed by western blot. Primary antibodies were used at the following concentrations: rabbit anti-phosphorylated ERK1/2 (Cell Signaling Technology), 1:1000; rabbit anti-ERK1/2 (Promega, Madison, WI), 1:5000; mouse monoclonal anti-HA-tag (Covance Inc., Princeton, NJ), 1:1000. Peroxidase (HRP)-conjugated secondary antibodies were used at the following concentrations: goat anti-rabbit IgG (Cell Signaling Technology), 1:2000; goat anti-rabbit IgG (Pierce, Rockford, IL), 1:15,000; goat anti-mouse IgG (Pierce), 1:15,000. Chemiluminescence was performed with the SuperSignal Kit (Pierce).

Analysis of MMP-9 expression by gelatin zymography
MMP-9 protein expression was assayed essentially as described previously (DiPersio et al., 2000Go). Briefly, MK cells were plated onto 12-well tissue culture plates coated with LN-5 ECM at 1.4x105 cells per well and allowed to attach overnight in MK growth medium. Cultures were then rinsed and maintained in serum-free medium for an additional 36 or 48 hours, as indicated in figure legends, and culture media were collected for zymography. In some experiments, cells were treated with the MEK inhibitor U0126 (10 µm), as described in the figure legends. All zymography experiments were performed in the absence of INF{gamma}, since INF{gamma} has been reported to inhibit MMP-9 expression (Hujanen et al., 1994Go). The same zymography results were obtained whether cells were grown in serum-free medium containing or lacking additional hormonal supplements that are normally present in MK growth medium (see above). MMPs were concentrated from culture media by binding overnight at 4°C to gelatin-agarose beads (Sigma). Agarose beads were recovered by centrifugation, and bound MMPs were eluted in zymography sample buffer (2.25% SDS, 9% glycerol, 45 mM Tris-HCl, pH 6.8, Bromophenol Blue) and resolved by non-reducing SDS-PAGE on 10% polyacrylamide gels impregnated with 1 mg/ml gelatin (Sigma). Gels were processed for zymography as described previously (DiPersio et al., 2000Go).

Analysis of MMP-9 mRNA
For northern blots, MK cells were cultured on LN-5 ECM for 4 days in serum-containing growth medium, and total cellular RNA was isolated using the Purescript RNA isolation kit (Gentra Systems, Minneapolis, MN). 10 µg total RNA was denatured at 55°C for 15 minutes in 1x MOPS, 6.5% formaldehyde and 50% formamide prior to electrophoresis on agarose-formaldehyde gels (1.2% agarose, 1.1% formaldehyde, 1x MOPS). RNA was then transferred to Nytran membranes in 10x SSC (3 M NaCl, 0.3 M sodium citrate, pH 7.0), UV cross-linked and incubated for 3 hours at 42°C in pre-hybridization buffer (50% formamide, 5x Denhardt's solution, 1% SDS, 150 µg/ml salmon sperm DNA, 5x SSC). Blots were then incubated overnight at 42°C in hybridization buffer (pre-hybridization buffer with 2.5x Denhardt's solution and 10% dextran sulfate) containing a 32P-labeled cDNA probe against mouse MMP-9 (a generous gift from Karl Tryggvason, University of Oulu, Finland), washed and exposed to autoradiographic film. Blots were stripped and reprobed with a cDNA probe against A50 ribosomal protein as a control, as its expression does not change in response to cell adhesion (Samarakoon and Higgins, 2002Go).

For RT-PCR, total RNA was isolated as described above and reverse transcribed to produce cDNA template using the first-strand cDNA synthesis kit (Promega). PCR reactions were carried out in 12.5 µl PCR REDTaq ReadyMix (Sigma) with 5 µl cDNA and 0.4 µM of each primer. PCR conditions were 94°C for 60 seconds, 58°C for 90 seconds and 72°C for 90 seconds. Samples were subjected to 28 cycles for MMP-9 and 15 cycles for ß-actin, with the last cycle for each at 72°C for 7 minutes. PCR primers for MMP-9 were as described (Bouloumie et al., 2001Go) and generate a 371 bp product: forward primer, 5'-TGTACCGCTATGGTTACAC-3'; reverse primer, 5'-CCGCGACACCAAACTGGAT-3'. An alternative set of MMP-9 primers was used for certain experiments, as described (Kim et al., 2000Go) and generates a 754 bp product: forward primer, 5'-AGTTTGGTGTCGCGGAGCAC-3'; reverse primer, 5'-TACATGAGCGCTTCCGGCAC-3'. PCR conditions for these primers were similar to those described above, except that the annealing temperature was increased to 64°C for 60 seconds. PCR primers for ß-actin were as described (Li et al., 2002Go) and generate a 450 bp product: forward primer, 5'-AGGGAAATCGTGCGTGACAT-3'; reverse primer, 5'-CATCTGCTGGAAGGTGGACA-3'. PCR products were run on a 1.8% agarose gel, stained with ethidium bromide and visualized using a Bio-Rad Gel-Doc 2000. Signals were quantified using Quantity One software (Bio-Rad, Hercules, CA). The number of PCR cycles was optimized so that signals were within the linear range of detection for both MMP-9 and ß-actin.

For assaying mRNA turnover, cells cultured on LN-5 ECM in serum-free MK growth medium were pre-treated with 10 µg/ml cycloheximide for 16 hours to promote accumulation of MMP-9 mRNA, as described (Huang et al., 2000Go). Cycloheximide was then removed and 4 hours later cells were switched to serum-free MK growth medium containing 10 µg/ml actinomycin D to inhibit transcription. Control experiments demonstrated that this concentration of actinomycin D effectively inhibited transcription from the MMP-9 promoter in MK cells (data not shown). The initial time point was collected 30 minutes after actinomycin D treatment. Subsequent time points were collected as indicated in the figure legend and assayed for MMP-9 mRNA and ß-actin mRNA levels by RT-PCR.

Adenoviral infection of MK cells
MK cells were sub-cultured on LN-5 ECM in growth medium overnight and then infected for 24 hours with adenovirus expressing either HA-tagged RasV12 or ß-galactosidase as a control. RasV12 and ß-galactosidase were cloned into pAdTrack, as described (Meadows et al., 2001Go). Multiplicity of infection (MOI) for each experiment is provided in the figure legends. In some experiments, cells were pre-treated for 3 hours with 10 µM U0126 (Calbiochem, San Diego, CA), or with DMSO as a control, then cultured in serum-free medium for an additional 24-48 hours in the presence or absence of U0126, as indicated; for zymography experiments, U0126 was replenished in the medium every 12 hours. Culture media were assayed by zymography and cell lysates were assayed by western blotting, as described above.

Analysis of the transfected MMP-9 promoter
An MMP-9 promoter/firefly luciferase reporter plasmid was a generous gift from Y. Sasaguri (University of Occupational and Environmental Health, Kitakyushu, Japan) and contained a 1868 bp DNA fragment of the MMP-9 promoter region (–1879 to –12 from the transcription start site) cloned upstream of the luciferase gene in the pGL3 vector (Promega), as described (Shimajiri et al., 1999Go). Uninfected MK cells, or MK cells infected with RasV12 adenovirus (see above), were co-transfected with the MMP-9 promoter/luciferase reporter plasmid and a TK promoter/Renilla luciferase control plasmid (pRLTK, Promega) at a 50:1 ratio using lipofectamine. Following a 5 hour transfection, cells were cultured in MK growth medium without serum (unless otherwise indicated), in the presence or absence of 10 µM U0126 for an additional 24 hours. Cell lysates were then collected and assayed for luciferase expression using the Dual-Luciferase Reporter Assay Kit (Promega). Luciferase expression was measured in a TD-20/20 luminometer (Turner Designs) and expression from the MMP-9 promoter/luciferase plasmid was normalized to that from the control pRLTK plasmid for each sample. Normalized luciferase signals were then plotted as described in the legend to Fig. 5.



View larger version (17K):
[in this window]
[in a new window]
 
Fig. 5. MMP-9 promoter activation by RasV12 is MEK/ERK-dependent but {alpha}3ß1-independent. (A) MMP-9 promoter activity: MK+/+ cells (+/+), MK–/– cells (–/–) and MK–/– cells expressing human {alpha}3 ({alpha}3) were grown on LN-5 ECM and transfected with an MMP-9 promoter/luciferase reporter plasmid. 24 hours later, cells were assayed for luciferase expression as described in the text. Normalized luciferase signals are plotted as the percentage of control levels seen in MK+/+ cells. Data are presented as the mean±s.e.m.; n=3 in two separate experiments. (B) MK+/+ cells (MK+/+), MK–/– cells (MK–/–), or MK–/– cells transfected with {alpha}3 ({alpha}3) grown on LN-5 ECM were infected with adenovirus expressing either HA-tagged RasV12 (Ras) or ß-galactosidase (C); MOI=10. Cells were then transfected with the MMP-9 promoter/luciferase reporter plasmid in the presence or absence of 10 µM U0126, as indicated and assayed for luciferase expression as in A. Normalized luciferase signals are plotted relative to the control levels seen in the absence of U0126, which is set to 1.0 for each cell type (C, –). Data are presented as the mean±s.e.m.; n=3 in two separate experiments.

 


    Results
 Top
 Summary
 Introduction
 Materials and Methods
 Results
 Discussion
 References
 
A MEK/ERK signaling pathway is necessary for MMP-9 expression in immortalized mouse keratinocytes
MMP-9 production is dependent on expression of {alpha}3ß1 integrin in immortalized mouse keratinocytes (MK cells) cultured on laminin 5-rich extracellular matrix (LN-5 ECM) and this mode of regulation was acquired as part of the immortalized phenotype (DiPersio et al., 2000Go). ERKs are phosphorylated and activated by MEKs, and MEK/ERK pathways have been shown to regulate MMP-9 gene expression in a number of cell types (Gum et al., 1997Go; McCawley et al., 1999Go; Zeigler et al., 1999Go; Genersch et al., 2000Go). To determine whether MEK/ERK signaling is required for MMP-9 expression in MK cells, wild-type MK+/+ cells cultured on LN-5 ECM were treated with the MEK-specific inhibitor, U0126, and gelatin zymography was used to assess levels of MMP-9 secreted into the culture medium. Inhibition of MEK caused decreased levels of MMP-9 production by MK+/+ cells (Fig. 1). We showed previously that expression of MMP-9 protein was absent from {alpha}3ß1-deficient MK–/– cells, but was restored in these cells by transfection with the human {alpha}3 subunit (DiPersio et al., 2000Go). This {alpha}3ß1-mediated restoration of MMP-9 expression was also blocked with the MEK inhibitor (data not shown). To confirm that MEK-dependent ERK activation was inhibited effectively by U0126 in these experiments, lysates from the same cells were immunoblotted for the activated forms of ERK1/2 with an antibody specific for phosphorylation on residues Thr202 and Tyr204. Levels of activated ERK were reduced in U0126-treated cells (Fig. 1, pERK blot), whereas levels of total ERK protein were unchanged (Fig. 1, ERK blot). These results demonstrate that signaling through MEK/ERK is necessary for MMP-9 expression in MK cells.



View larger version (40K):
[in this window]
[in a new window]
 
Fig. 1. MEK/ERK signaling is required for MMP-9 expression in MK cells. Wild-type MK+/+ cells were grown on laminin-5-rich extracellular matrix (LN-5 ECM) in serum-free medium for 36 hours in the presence or absence of the MEK inhibitor U0126 (10 µM), as indicated. MMP-9 levels in the culture medium were assayed by gelatin zymography (top panel). Position of the 105 kDa murine pro-MMP-9 (mMMP-9) is indicated on the left. S lane, positions of human MMP-9 (hMMP-9) and MMP-2 (hMMP-2) purified standards are indicated. ERK activation was assayed by immunoblot of the corresponding cell lysates for phospho-ERK (pERK panel). Blots were stripped and reprobed for total ERK (ERK panel).

 

Activation of MEK/ERK signaling by oncogenic RasV12 induces MMP-9 protein levels in an {alpha}3ß1-dependent manner
As activation of ERK is necessary for MMP-9 expression in the MK+/+ cells, we next tested whether ERK activation is sufficient to induce MMP-9 expression in the {alpha}3ß1-deficient MK–/– cells. Ras is a small GTPase that activates the MEK/ERK signaling pathway. Oncogenic mutations in the c-rasHa gene occur frequently in skin tumors (Yuspa, 1994Go) and Ras activation has been shown to induce expression of MMP-9 (Genersch et al., 2000Go; Gum et al., 1996Go). We expressed a constitutively activated mutant of Ras, RasV12, in the MK cells and determined the effects on ERK activation and MMP-9 expression. MK+/+ and MK–/– cells were each infected with an adenovirus expressing either HA-tagged RasV12 or ß-galactosidase as a control. For both MK+/+ cells and MK–/– cells, we did not observe any obvious differences in morphology between RasV12-infected and control-infected cells (data not shown). Gelatin zymography showed that expression of RasV12 in MK+/+ cells induced levels of secreted MMP-9 considerably above basal levels detected in control-infected cells (Fig. 2, mMMP-9, compare lanes 1 and 3). The induction of MMP-9 by RasV12 was MEK/ERK dependent as treatment with U0126 abrogated this response (Fig. 2, mMMP-9, compare lanes 3 and 4). Immunoblots of lysates from the corresponding cell layers confirmed the presence of HA-tagged RasV12 in infected cells (Fig. 2, RasV12, lanes 3 and 4). Immunoblotting for phosphorylated ERK confirmed that RasV12 caused an increase in ERK activation and that this activation was MEK-dependent (Fig. 2, pERK).



View larger version (25K):
[in this window]
[in a new window]
 
Fig. 2. Activation of MEK/ERK signaling by oncogenic RasV12 induces MMP-9 in an {alpha}3ß1-dependent manner. Wild-type MK+/+ cells (MK+/+) or {alpha}3ß1-deficient MK–/– cells (MK–/–) grown on LN-5 ECM were infected with adenovirus expressing either HA-tagged RasV12 (RasV12) or ß-galactosidase (control); MOI=70. Cells were then grown in serum-free medium for an additional 36 hours in the presence or absence of 10 µM U0126, as indicated. Levels of secreted MMP-9 in the culture medium were assayed by gelatin zymography (top panels); murine MMP-9 and human MMP-9 standard are indicated. To confirm RasV12 expression, equal protein from corresponding cell lysates was immunoblotted with anti-HA tag (RasV12 panel). Parallel blots with anti-phospho-ERK confirmed ERK activation in RasV12-infected cells and inhibition of ERK activation by treatment with U0126 (pERK panel).

 

In stark contrast with the results in MK+/+ cells, RasV12 was unable to induce MMP-9 levels in MK–/– cells (Fig. 2, mMMP-9, compare lanes 5 and 7) suggesting that {alpha}3ß1 is required for RasV12-mediated induction of MMP-9. The inability of RasV12 to induce MMP-9 was not due to a defect in ERK activation, since ERK was phosphorylated robustly in RasV12-infected MK–/– cells (Fig. 2, pERK, lane 7). Thus, oncogenic RasV12 was able to activate ERK in both MK–/– cells and MK+/+ cells, but it did not induce MMP-9 secretion in the MK–/– cells, demonstrating that {alpha}3ß1 is required for MMP-9 expression at a point downstream of ERK activation.

{alpha}3ß1 is required for MMP-9 mRNA accumulation
Regulation of MMP-9 expression has been reported to occur at the levels of gene transcription, mRNA stability and protein production/secretion (Akool et al., 2003Go; Eberhardt et al., 2002Go; Jiang and Muschel, 2002Go; Morini et al., 2000Go; Sehgal and Thompson, 1999Go; Thant et al., 2000Go; Westermarck and Kahari, 1999Go). In order to determine whether {alpha}3ß1 regulates MMP-9 mRNA levels, we performed northern blot analysis on total RNA isolated from MK cells cultured on LN-5 ECM in the presence of serum for 4 days, conditions which support similarly high survival of both MK+/+ cells and MK–/– cells (Manohar et al., 2004Go). Two distinct MMP-9 mRNA transcripts were detected in MK+/+ cells, a major transcript of ~3.5 kb and a minor transcript of ~2.7 kb (Fig. 3A, +/+ lane), as described for other cell types (Graubert et al., 1993Go; Tanaka et al., 1993Go). However, MMP-9 transcripts were undetectable in MK–/– cells (Fig. 3A, –/– lane) suggesting that {alpha}3ß1 is required for MMP-9 mRNA expression. Indeed, restoration of {alpha}3ß1 integrin expression in MK–/– cells through stable transfection with the human {alpha}3 subunit (see Fig. 3C) restored MMP-9 mRNA expression (Fig. 3A, –/–, {alpha}3 lane), whereas transfection with a control vector did not (Fig. 3A, –/–, V lane). Using the more sensitive method of RT-PCR, we confirmed that MMP-9 mRNA was detected in MK+/+ cells and in {alpha}3-transfected MK–/– cells, but was barely detectable in {alpha}3ß1-deficient MK–/– cells (Fig. 3B).



View larger version (25K):
[in this window]
[in a new window]
 
Fig. 3. {alpha}3ß1 is required for MMP-9 mRNA expression. (A) MK+/+ cells (+/+ lane), MK–/– cells (–/– lane) and MK–/– cells stably transfected with the human {alpha}3 subunit (–/–, {alpha}3 lane) or the parental expression vector (–/–, V lane) were cultured on LN-5 ECM for 4 days in serum-containing medium. Total RNA was isolated and assayed by northern blotting with cDNA probes for murine MMP-9 (top panel) or A50 as a control (bottom panel). Two distinct MMP-9 mRNA transcripts of 3.5 kb and 2.7 kb are indicated. (B) RT-PCR was performed using RNA collected as in A as a template. MMP-9 mRNA levels (top panel) and ß-actin mRNA levels (bottom panel) are shown for MK+/+ cells (+/+ lane), MK–/– cells (–/– lane) and MK–/– cells transfected with human {alpha}3 ({alpha}3 lane). (C) FACS analysis with the monoclonal antibody P1B5 confirms high levels of {alpha}3ß1 surface expression in MK–/– cells transfected with the human {alpha}3 integrin subunit (gray peak), but not in untransfected MK–/– cells (white peak).

 

We showed previously that {alpha}3-null MK–/– cells cultured on LN-5 ECM fail to spread properly and display reduced actin stress fiber formation (DiPersio et al., 2000Go; Choma et al., 2004Go), suggesting that reduced MMP-9 expression in these cells could be caused by a general spreading defect. Indeed, previous studies have shown that MMP expression can be regulated by changes in cell spreading and actin cytoskeletal organization (Kheradmand et al., 1998Go; Yan et al., 2000Go). To test whether restoring cell spreading to MK–/– cells can induce MMP-9 expression, MK–/– cells were plated either on a mixture of LN-5 ECM plus fibronectin or on purified fibronectin. MK–/– cells were able to spread on both fibronectin alone (Fig. 4B, c and d) and LN-5 ECM plus fibronectin (data not shown). However, neither of these substrates induced MMP-9 expression above the low basal levels seen on LN-5 ECM alone (Fig. 4A). By contrast, MK+/+ cells showed high levels of MMP-9 mRNA expression on either substrate (Fig. 4A); MMP-9 expression on fibronectin alone was probably due to deposition of endogenous LN-5 (DiPersio et al., 2000Go).

Although MK–/– cells were able to spread on fibronectin, the degree of spreading was detectably lower in MK–/– cells (Fig. 4B, c and d) than in MK+/+ cells (Fig. 4B, a and b), with a higher proportion of MK–/– cells showing a spindle-shaped appearance (Fig. 4d, arrow) and/or fewer stress fibers. The expression of other endogenous integrins is similar in the MK+/+ cells and MK–/– cells (DiPersio et al., 2000Go), suggesting that the ability of MK+/+ cells to spread better on fibronectin is most likely due to their ability to adhere to endogenously deposited LN-5 through {alpha}3ß1 (Frank et al., 2004; DiPersio et al., 2000Go). In an attempt to enhance MK–/– cell spreading, we also tested a mixed substrate of fibronectin plus vitronectin. Under these conditions, MK–/– cells showed markedly improved spreading and stress fiber formation that approached that seen in MK+/+ cells (Fig. 4B, e-h), although subtle differences in the number of spindle-shaped cells remained (Fig. 4B, h, arrow). Despite improved spreading, MK–/– cells still showed considerably lower MMP-9 mRNA levels than did MK+/+ cells on fibronectin plus vitronectin (Fig. 4A, F+V). {alpha}3ß1-dependent MMP-9 expression was also seen in MK cells cultured on vitronectin alone (Fig. 4A, V). Taken together, our results indicate that {alpha}3ß1 is required for MMP-9 mRNA expression and that engagement of other endogenous integrins/ECM receptors and subsequent spreading is not sufficient to restore MMP-9 expression in MK–/– cells. Nevertheless, we cannot rule out a requirement for cell spreading, or a role for subtle aspects of cell spreading or cytoskeletal organization, in the {alpha}3ß1-dependent regulation of MMP-9.

{alpha}3ß1 is not required for MEK/ERK-dependent MMP-9 promoter activity
MEK/ERK signaling has been shown to regulate transcriptional activation of the MMP-9 promoter (Westermarck and Kahari, 1999Go). Given that {alpha}3ß1 was required for MEK/ERK-mediated induction of MMP-9 expression, we next tested whether {alpha}3ß1 is also required for MEK/ERK-mediated activation of the MMP-9 promoter. For these experiments, we used an MMP-9 promoter-driven luciferase assay that was shown previously to provide a sensitive measure of transcriptional responsiveness (Shimajiri et al., 1999Go). MK cells cultured on LN-5 ECM were transfected with a reporter plasmid containing the firefly luciferase gene under transcriptional control of a ~1.9 kb MMP-9 promoter fragment that includes the major regulatory elements for responsiveness to growth factors and cytokines (Shimajiri et al., 1999Go; Westermarck and Kahari, 1999Go). Interestingly, MMP-9 promoter activity was reduced only slightly in MK–/– cells compared with MK+/+ cells and restoring {alpha}3ß1 expression in the former cells did not increase promoter activity (Fig. 5A), indicating that activity of the transfected promoter was independent of {alpha}3ß1. Treatment with U0126 reduced MMP-9 promoter activity below basal levels in MK+/+ cells, MK–/– cells and {alpha}3-transfected MK–/– cells (Fig. 5B), indicating that MEK/ERK signaling is necessary for baseline MMP-9 promoter function regardless of {alpha}3ß1 expression. We have consistently observed reduced levels of phosphorylated ERK in MK–/– cells compared with MK+/+ cells or {alpha}3-transfected MK–/– cells, indicating that {alpha}3ß1 is required for full ERK activation in these cells (Manohar et al., 2004Go). Nevertheless, the results presented here show that basal ERK activity in MK–/– cells is both necessary (Fig. 5B) and sufficient (Fig. 5A) for MMP-9 promoter activity.

To determine whether oncogenic Ras can induce MEK/ERK-dependent activation of the MMP-9 promoter, MK cells that had been infected with the RasV12 adenovirus were transfected with the MMP-9 promoter/luciferase reporter plasmid and cultured in the presence or absence of U0126. RasV12 induced the MMP-9 promoter in MK+/+ cells more than threefold over basal levels seen in uninfected cells (Fig. 5B, MK+/+) consistent with the induction of endogenous MMP-9 that was detected in these cells by zymography (Fig. 2). RasV12 also induced the MMP-9 promoter in {alpha}3-transfected MK–/– cells (Fig. 5B, {alpha}3). MMP-9 promoter activity in RasV12-infected cells was reduced below basal levels upon treatment with U0126 (Fig. 5B), showing that RasV12-mediated induction of the MMP-9 promoter is MEK-dependent. These results suggest that RasV12-mediated induction of MMP-9 secretion in MK+/+ cells is due, at least partly, to activation of a MEK/ERK signaling pathway that induces the MMP-9 promoter.

Importantly, even though RasV12 also induced the MMP-9 promoter in {alpha}3 1-deficient MK–/– ß cells (Fig. 5B, MK–/–), there was not a concomitant induction of endogenous MMP-9 secretion by RasV12 in these cells (Fig. 2). These results were confirmed in a separate set of experiments in which transfected MK–/– cells were assayed simultaneously for luciferase expression and for endogenous MMP-9 protein expression (data not shown). Thus, {alpha}3ß1 was not required for induction of the transfected MMP-9 promoter by Ras/MEK/ERK signaling, but it was required for endogenous MMP-9 expression in the same cells. As promoter-driven luciferase assays are an indirect measure of a gene's transcriptional activity, our results do not rule out a role for {alpha}3ß1 in transcriptional regulation of the endogenous MMP-9 gene. Nevertheless, they prompted us to explore the possibility that {alpha}3ß1 has a post-transcriptional role in MMP-9 mRNA accumulation (see below).

{alpha}3ß1 is required for induction of endogenous MMP-9 mRNA by oncogenic RasV12
As RasV12 was able to induce the MMP-9 promoter independently of {alpha}3ß1 (Fig. 5B), but accumulation of endogenous MMP-9 mRNA required {alpha}3ß1 (Fig. 3), we next determined whether the ability of RasV12 to induce steady-state levels of endogenous MMP-9 mRNA was dependent on {alpha}3ß1. MK+/+ cells and MK–/– cells were cultured on LN-5 ECM and infected with adenovirus expressing either HA-tagged RasV12 or ß-galactosidase as a control. MMP-9 and ß-actin mRNA levels were then assayed by RT-PCR (Fig. 6A) and signals for MMP-9 were normalized to those for ß-actin (Fig. 6B). Although RasV12 had a slight inductive effect on MMP-9 mRNA levels in MK–/– cells, substantial induction by RasV12 was dependent on expression of {alpha}3ß1 (Fig. 6B). Taken together, results from Figs 5 and 6 suggest that {alpha}3ß1 may potentiate RasV12-mediated induction of MMP-9 mRNA through a mechanism that is downstream from ERK-dependent promoter activation. One possibility is that {alpha}3ß1 promotes post-transcriptional mRNA stability, and that increased mRNA turnover in {alpha}3-null cells prevents the accumulation of MMP-9 transcripts following transcriptional stimulation by RasV12.



View larger version (17K):
[in this window]
[in a new window]
 
Fig. 6. RasV12-mediated induction of MMP-9 mRNA accumulation is {alpha}3ß1-dependent. (A) MK+/+ and MK–/– cells were infected with adenovirus expressing either ß-galactosidase as a control (– lanes) or RasV12 (+ lanes), then cultured on LN-5 ECM for 48 hours in serum-free medium. Total RNA was isolated and RT-PCR was performed to compare MMP-9 and ß-actin mRNA levels. (B) Signals for MMP-9 mRNA and ß-actin mRNA were quantified; graphs depict relative MMP-9 mRNA levels for each sample, after normalization to ß-actin mRNA. Data are presented as the mean±s.e.m. for three separate experiments; *P<0.05, two-way ANOVA with Newman Keuls post-hoc test.

 

{alpha}3ß1 promotes MMP-9 mRNA stability
To directly compare MMP-9 mRNA stability between {alpha}3-null and {alpha}3-expressing MK cells, we utilized an established approach to inhibit new gene transcription and measure MMP-9 mRNA decay over time (Sehgal and Thompson, 1999Go). Owing to the very low basal level of MMP-9 mRNA in the MK–/– cells, it was necessary to first induce MMP-9 mRNA to detectable levels in these cells so that we could monitor the rate of mRNA turnover. Previous studies of mRNA stability have shown that treatment with the translation inhibitor cycloheximide leads to accumulation of unstable mRNAs in some cells, possibly by inhibiting new synthesis of labile destabilization factors that promote mRNA turnover (reviewed in Guhaniyogi and Brewer, 2001Go). Indeed, in preliminary experiments we observed that MMP-9 mRNA expression in {alpha}3-null MK–/– cells was induced ~tenfold to easily detectable levels by treatment with 10 µg/ml cycloheximide (data not shown). Therefore, cells were pre-treated with cycloheximide to induce accumulation of MMP-9 mRNA in the MK–/– cells. Cells were then cultured in the presence of the transcriptional inhibitor actinomycin D to prevent new mRNA synthesis, allowing us to monitor MMP-9 mRNA decay over a time course as described previously (Sehgal and Thompson, 1999Go). RT-PCR was used to assess mRNA levels for MMP-9, or for ß-actin as a control. MMP-9 mRNA levels remained high for up to 24 hours in MK+/+ cells (Fig. 7A, diamonds). By contrast, by 24 hours MMP-9 mRNA levels dropped to 37% of original levels in MK–/– cells cultured under identical conditions (Fig. 7A, squares), indicating dramatically decreased mRNA stability in these cells. Restoring expression of {alpha}3ß1 in MK–/– cells resulted in increased MMP-9 mRNA levels at each time point, compared with untransfected MK–/– cells (Fig. 7A, open circles). In contrast with MMP-9 mRNA, ß-actin mRNA levels were comparable between {alpha}3ß1-expressing MK cells and {alpha}3-null MK cells at each time point (Fig. 7B), demonstrating that {alpha}3ß1-mediated effects on transcript stability were specific to MMP-9 mRNA. These results identify a novel role for {alpha}3ß1 in promoting MMP-9 mRNA stability and they provide a mechanism whereby the integrin can potentiate the induction of MMP-9 expression in response to oncogenic Ras or other activators of MEK/ERK-mediated gene transcription.



View larger version (20K):
[in this window]
[in a new window]
 
Fig. 7. {alpha}3ß1 promotes stability of MMP-9 mRNA. MK+/+ cells, MK–/– cells, or MK–/– cells transfected with {alpha}3 grown on LN-5-ECM were pre-treated with cycloheximide to promote accumulation of MMP-9 mRNA in MK–/– cells, and then grown under serum-free conditions in the presence of actinomycin D to inhibit new transcription (see text for details). Total RNA was isolated at various time points and RT-PCR was performed to monitor turnover of mRNA for MMP-9 (A) and ß-actin (B). Results are plotted as the percentage of mRNA remaining relative to the starting amounts at 0 hour; diamonds, MK+/+ cells; squares, MK–/– cells; open circles, {alpha}3-transfected MK–/– cells. Data are presented as the mean±s.e.m. for three separate experiments; *P<0.05, one-way ANOVA with Newman Keuls post-hoc test. Gels show results from a representative experiment.

 


    Discussion
 Top
 Summary
 Introduction
 Materials and Methods
 Results
 Discussion
 References
 
Integrins have been shown to play important roles in regulating MMP-9 expression in immortalized epithelial cells and cell lines derived from SCC and breast carcinomas (DiPersio et al., 2000Go; Morini et al., 2000Go; Thomas et al., 2001Go). However, the mechanisms whereby integrins regulate MMP-9 gene expression have been unclear. In the current study, we demonstrated that {alpha}3ß1 integrin is required for accumulation of MMP-9 mRNA following MEK/ERK-mediated induction of the MMP-9 gene in immortalized keratinocytes. We propose a model in which {alpha}3ß1 potentiates MEK/ERK-induced MMP-9 synthesis by regulating mRNA stability downstream of ERK-mediated promoter activation (Fig. 8). Through this mechanism, {alpha}3ß1 potentiates induction of MMP-9 mRNA and protein by activated Ras. Thus, our data identify a novel mechanism whereby integrins can regulate MMP gene induction following the activation of MAPK pathways that stimulate transcription.



View larger version (64K):
[in this window]
[in a new window]
 
Fig. 8. A model for regulation of MMP-9 expression by {alpha}3ß1. MMP-9 gene transcription can be induced by Ras/MEK/ERK signaling in response to growth factors, cytokines, or oncogenic Ras. Our data reveal a novel mechanism wherein {alpha}3ß1 is required for stabilization of MMP-9 mRNA, leading to the accumulation of MMP-9 transcripts and subsequent production of MMP-9 protein in response to ERK-mediated transcriptional induction. In the absence of {alpha}3ß1, transcriptional induction by Ras/MEK/ERK does not lead to substantial accumulation of MMP-9 transcript and increased MMP-9 protein owing to the destabilization and rapid decay of MMP-9 mRNA. The {alpha}3 and ß1 integrin subunits are indicated.

 

The acquisition of {alpha}3ß1-dependent MMP-9 expression by immortalized keratinocytes may reflect an important role for {alpha}3ß1 during malignant progression of SCCs or other tumors, as discussed below. Our findings may also be relevant to other tissue remodeling events in which {alpha}3ß1 has been implicated, such as embryonic tissue development or wound healing (Kreidberg, 2000Go). It seems likely that {alpha}3ß1-mediated regulation of MMP-9 expression is controlled by {alpha}3ß1 binding to LN-5, as it is well established that {alpha}3ß1 is a strong receptor for keratinocyte adhesion to LN-5 (Kreidberg, 2000Go). Alternatively, or in addition, MMP-9 expression may be controlled by functions of {alpha}3ß1 at sites of cell-cell adhesion, as {alpha}3ß1 was recently shown to be a functional component of E-cadherin-mediated adherens junctions in kidney collecting duct epithelial cells (Chattopadhyay et al., 2003Go). Indeed, we recently demonstrated that monolayer cultures of {alpha}3ß1-deficient MK–/– cells show reduced cell-cell interactions (Choma et al., 2004Go) and we have determined that {alpha}3ß1 is physically and functionally associated with E-cadherin at MK cell-cell junctions (N. Myneni and M.D., unpublished). Future experiments will address the relative importance of {alpha}3ß1 functions at cell-ECM adhesions and cell-cell adhesions in the regulation of MMP-9 expression.

Although most previous studies of MMP-9 gene expression have focused on transcriptional control, it has become increasingly evident that post-transcriptional mechanisms play critical roles in the regulation of MMP-9 mRNA levels in response to growth factors and cytokines (Akool et al., 2003Go; Sehgal and Thompson, 1999Go). In the current study we have identified mRNA stabilization as a primary mechanism of {alpha}3ß1-mediated MMP-9 induction in immortalized keratinocytes. Although regulation of mRNA stability is clearly an important post-transcriptional mechanism for the control of gene expression (Sachs, 1993Go), there have been few published studies regarding the roles of integrins and ECM in the modulation of mRNA stability (Feng et al., 1999Go; Retta et al., 2001Go; Xu and Clark, 1996Go) and this mechanism of integrin-mediated gene regulation remains largely unexplored. Regulation of mRNA half-life occurs primarily through conserved U-rich or AU-rich elements, collectively termed AREs, that are usually present in multiple, non-tandem copies within the 3'-untranslated region (3'-UTR) of mRNAs with short or variable half-lives (Brennan and Steitz, 2001Go). AREs interact with specific RNA-binding proteins (RBPs) that control the rate of mRNA decay. Two of the most extensively studied RBPs are HuR, which promotes mRNA stability (Brennan and Steitz, 2001Go) and AUF1/hnRNP D, several isoforms of which promote mRNA decay (Loflin et al., 1999Go). The 3'-UTR of the mouse and rat MMP-9 transcripts each contain several ARE consensus sequences, some of which have been shown to bind HuR and regulate MMP-9 mRNA stability (Graubert et al., 1993Go; Tanaka et al., 1993Go; Akool et al., 2003Go). Future studies should determine whether {alpha}3ß1 integrin promotes MMP-9 mRNA stability by regulating the functions of specific RBPs that bind to these AREs.

Although a number of specific signal transduction pathways have been implicated in the regulation of mRNA stability (Brennan and Steitz, 2001Go), several studies have reported a particularly prominent role for MAPK p38 signaling pathways in ARE-mediated mRNA stability (Montero and Nagamine, 1999Go; Dean et al., 1999Go; Reunanen et al., 2002Go; Winzen et al., 1999Go; Tran et al., 2003Go). Although we have not yet identified the signaling pathways whereby {alpha}3ß1 controls MMP-9 mRNA stability, we recently showed that several signaling proteins are activated by {alpha}3ß1 in MK cells, including focal adhesion kinase (FAK) and the Rho family GTPase Rac1 (Choma et al., 2004Go; Manohar et al., 2004Go). Rac1 is a particularly good candidate for mediating MMP-9 mRNA stability, as a Rac1/MKK3/p38 pathway has been shown to promote stability of the uPA mRNA in invasive breast epithelial cells (Han et al., 2002Go). It remains to be determined whether {alpha}3ß1-mediated activation of Rac1 and/or p38 regulates ARE-mediated stability of MMP-9 mRNA.

Previous studies have demonstrated cooperativity between integrins and growth factor receptors in the regulation of MAPK signaling pathways (Giancotti and Ruoslahti, 1999Go). For example, growth factor-dependent induction of ERK signaling in NIH 3T3 cells is strongly dependent on integrin-mediated cell adhesion (Aplin and Juliano, 1999Go; Renshaw et al., 1997Go). In the latter studies, activation of MEK/ERK signaling was identified as an adhesion-dependent event. Our model in Fig. 8 suggests a distinct, novel mechanism whereby integrins can cooperate with growth factors or other stimuli to induce MEK/ERK-dependent gene expression. Specifically, our results in MK cells using RasV12 as an activator of MEK/ERK showed that {alpha}3ß1 was not required for Ras-mediated ERK activation, but that it was required for MEK/ERK-mediated induction of MMP-9 expression at a point downstream from ERK activation (i.e. mRNA stabilization). Thus, {alpha}3ß1 may regulate the ability of some cells to synthesize and secrete MMP-9 in response to growth factors or cytokines that stimulate MMP-9 gene transcription through MEK/ERK pathways. In addition, our results suggest that {alpha}3ß1 expressed in malignant tumors may potentiate MMP-9 induction in response to MAPK pathways that are activated constitutively by oncogenic Ras or other oncogenes. Indeed, activating mutations in the c-rasHa gene are among the most common initiating mutations in epidermal tumors (Yuspa, 1994Go).

Carcinogenesis is generally accompanied by cellular changes that facilitate tumor growth and cell invasion, including altered ECM synthesis, altered integrin function and increased MMP expression (Johnsen et al., 1998Go; Werb, 1997Go; Westermarck and Kahari, 1999Go). Tumor cells also undergo changes in their capacity to respond to extracellular signals generated from growth factors or ECM, and they are likely to acquire certain signaling pathways to induce or maintain MMP expression. As discussed above, the ability of {alpha}3ß1 integrin to potentiate MEK/ERK-mediated induction of MMP-9 through mRNA stabilization may reflect an important role for this integrin in regulating the ability of some tumor cells to respond to environmental cues or oncogenes that stimulate gene transcription through MEK/ERK signaling pathways. This ability of {alpha}3ß1 to induce MMP-9 expression is acquired as part of the immortalized keratinocyte phenotype (DiPersio et al., 2000Go). Similarly, the ability of TGFß to induce MMP-9 expression in prostate cancer cells is acquired during cellular transformation (Sehgal et al., 1996Go; Sehgal and Thompson, 1999Go). In addition, changes in calcium-mediated regulation of MMP-9 expression occur during malignant transformation of oral keratinocytes (Mukhopadhyay et al., 2004Go). Thus, it is clear that certain signaling pathways that regulate MMP-9 expression are activated or altered during cellular immortalization or malignant progression. Many invasive tumors show increased expression levels of {alpha}3ß1 (Bartolazzi et al., 1994Go; Natali et al., 1993Go; Patriarca et al., 1998Go) and its ECM ligand laminin-5 (Pyke et al., 1995Go) and recent studies in immortalized keratinocytes and tumor-derived cell lines have demonstrated important roles for {alpha}3ß1 in several cell functions that contribute to tumor growth and malignant progression, including migration, invasion, metastasis and survival (Choma et al., 2004Go; Manohar et al., 2004Go; Morini et al., 2000Go; Tsuji et al., 2002Go; Wang et al., 2004Go). As MMP-9 has been implicated in some of these cell functions, our current data suggest that part of the role of {alpha}3ß1 may be to regulate the stability and degradation of MMP-9 mRNA. The acquisition by immortalized cells of {alpha}3ß1-dependent mRNA stability may extend to other mRNA transcripts as well and could represent an important selection step during tumor progression. Therefore, further study of the signal transduction pathways involved in this regulation may lead to the identification of therapeutic targets for inhibiting carcinoma progression and/or metastasis that are specific to cancer cells.


    Acknowledgments
 
We are grateful to John Lamar and Lee Stirling for excellent technical assistance. We also thank Karl Tryggvason, Yasuyuki Sasaguri, Martin Hemler, Livingston Van De Water, Paula McKeown-Longo and Paul Higgins for valuable reagents. We thank Peter Vincent, Andrew Aplin, André Melendez and Livingston Van De Water for valuable discussions and critical reading of the manuscript. This research was supported by a grant from the National Institutes of Health to C.M.D. (R01CA84238), as well as by a grant from the National Institutes of Health to K.P. (R01CA081419). V.I. was supported by a pre-doctoral training grant from the National Heart, Lung and Blood Institute (NIH-T32-HL-07194).


    References
 Top
 Summary
 Introduction
 Materials and Methods
 Results
 Discussion
 References
 

Akool, el-S., Kleinert, H., Hamada, F. M., Abdelwahab, M. H., Forstermann, U., Pfeilschifter, J. and Eberhardt, W. (2003). Nitric oxide increases the decay of matrix metalloproteinase 9 mRNA by inhibiting the expression of mRNA-stabilizing factor HuR. Mol. Cell Biol. 23, 4901-4916.[Abstract/Free Full Text]

Aplin, A. E. and Juliano, R. L. (1999). Integrin and cytoskeletal regulation of growth factor signaling to the MAP kinase pathway. J. Cell Sci. 112, 695-706.[Abstract/Free Full Text]

Bartolazzi, A., Cerboni, C., Nicotra, M. R., Mottolese, M., Bigotti, A. and Natali, P. G. (1994). Transformation and tumor progression are frequently associated with expression of the {alpha}3ß1 heterodimer in solid tumors. Int. J. Cancer 58, 488-491.[Medline]

Bergers, G., Brekken, R., McMahon, G., Vu, T. H., Itoh, T., Tamaki, K., Tanzawa, K., Thorpe, P., Itohara, S., Werb, Z. et al. (2000). Matrix metalloproteinase-9 triggers the angiogenic switch during carcinogenesis. Nat. Cell Biol. 2, 737-744.[CrossRef][Medline]

Bernhard, E. J., Gruber, S. B. and Muschel, R. J. (1994). Direct evidence linking expression of matrix metalloproteinase 9 (92-kDa gelatinase/collagenase) to the metastatic phenotype in transformed rat embryo cells. Proc. Natl. Acad. Sci. USA 91, 4293-4297.[Abstract/Free Full Text]

Bouloumie, A., Sengenes, C., Portolan, G., Galitzky, J. and Lafontan, M. (2001). Adipocyte produces matrix metalloproteinases 2 and 9: involvement in adipose differentiation. Diabetes 50, 2080-2086.[Abstract/Free Full Text]

Brennan, C. M. and Steitz, J. A. (2001). HuR and mRNA stability. Cell. Mol. Life Sci. 58, 266-277.[Medline]

Canete-Soler, R., Litzky, L., Lubensky, I. and Muschel, R. J. (1994). Localization of the 92 kd gelatinase mRNA in squamous cell and adenocarcinomas of the lung using in situ hybridization. Am. J. Pathol. 144, 518-527.[Abstract]

Chattopadhyay, N., Wang, Z., Ashman, L. K., Brady-Kalnay, S. M. and Kreidberg, J. A. (2003). alpha3beta1 integrin-CD151, a component of the cadherin-catenin complex, regulates PTPmu expression and cell-cell adhesion. J. Cell Biol. 163, 1351-1362.[Abstract/Free Full Text]

Choma, D. P., Pumiglia, K. and DiPersio, C. M. (2004). Integrin {alpha}3ß1 directs the stabilization of a polarized lamellipodium in epithelial cells through activation of Rac1. J. Cell Sci. 117, 3947-3959.[Abstract/Free Full Text]

Coussens, L. M., Tinkle, C. L., Hanahan, D. and Werb, Z. (2000). MMP-9 supplied by bone marrow-derived cells contributes to skin carcinogenesis. Cell 103, 481-490.[CrossRef][Medline]

Dean, J. L., Brook, M., Clark, A. R. and Saklatvala, J. (1999). p38 mitogen-activated protein kinase regulates cyclooxygenase-2 mRNA stability and transcription in lipopolysaccharide-treated human monocytes. J. Biol. Chem. 274, 264-269.[Abstract/Free Full Text]

DiPersio, C. M., Hodivala-Dilke, K. M., Jaenisch, R., Kreidberg, J. A. and Hynes, R. O. (1997). {alpha}3ß1 integrin is required for normal development of the epidermal basement membrane. J. Cell Biol. 137, 729-742.[Abstract/Free Full Text]

DiPersio, C. M., Shao, M., di Costanzo, L., Kreidberg, J. A. and Hynes, R. O. (2000). Mouse keratinocytes immortalized with large T antigen acquire {alpha}3ß1 integrin-dependent secretion of MMP-9/gelatinase B. J. Cell Sci. 113, 2909-2921.[Abstract/Free Full Text]

Eberhardt, W., Akool el, S., Rebhan, J., Frank, S., Beck, K. F., Franzen, R., Hamada, F. M. and Pfeilschifter, J. (2002). Inhibition of cytokine-induced matrix metalloproteinase 9 expression by peroxisome proliferator-activated receptor alpha agonists is indirect and due to a NO-mediated reduction of mRNA stability. J. Biol. Chem. 277, 33518-33528.[Abstract/Free Full Text]

Feng, X., Clark, R. A., Galanakis, D. and Tonnesen, M. G. (1999). Fibrin and collagen differentially regulate human dermal microvascular endothelial cell integrins: stabilization of alphav/beta3 mRNA by fibrin1. J. Invest. Dermatol. 113, 913-919.[CrossRef][Medline]

Frank, D. E. and Carter, W. G. (2004). Laminin 5 deposition regulates keratinocyte polarization and persistent migration. J. Cell Sci. 117, 1351-1363.[Abstract/Free Full Text]

Genersch, E., Hayess, K., Neuenfeld, Y. and Haller, H. (2000). Sustained ERK phosphorylation is necessary but not sufficient for MMP-9 regulation in endothelial cells: involvement of Ras-dependent and -independent pathways. J. Cell Sci. 113, 4319-4330.[Abstract/Free Full Text]

Giancotti, F. G. and Ruoslahti, E. (1999). Integrin signaling. Science 285, 1028-1032.[Abstract/Free Full Text]

Graubert, T., Johnston, J. and Berliner, N. (1993). Cloning and expression of the cDNA encoding mouse neutrophil gelatinase: demonstration of coordinate secondary granule protein gene expression during terminal neutrophil maturation. Blood 82, 3192-3197.[Abstract]

Guhaniyogi, J. and Brewer, G. (2001). Regulation of mRNA stability in mammalian cells. Gene 265, 11-23.[CrossRef][Medline]

Gum, R., Lengyel, E., Juarez, J., Chen, J. H., Sato, H., Seiki, M. and Boyd, D. (1996). Stimulation of 92-kDa gelatinase B promoter activity by ras is mitogen-activated protein kinase kinase 1-independent and requires multiple transcription factor binding sites including closely spaced PEA3/ets and AP-1 sequences. J. Biol. Chem. 271, 10672-10680.[Abstract/Free Full Text]

Gum, R., Wang, H., Lengyel, E., Juarez, J. and Boyd, D. (1997). Regulation of 92 kDa type IV collagenase expression by the jun aminoterminal kinase- and the extracellular signal-regulated kinase-dependent signaling cascades. Oncogene 14, 1481-1493.[CrossRef][Medline]

Han, Q., Leng, J., Bian, D., Mahanivong, C., Carpenter, K. A., Pan, Z. K., Han, J. and Huang, S. (2002). Rac1-MKK3-p38-MAPKAPK2 pathway promotes urokinase plasminogen activator mRNA stability in invasive breast cancer cells. J. Biol. Chem. 277, 48379-48385.[Abstract/Free Full Text]

Huang, Z. F., Massey, J. B. and Via, D. P. (2000). Differential regulation of cyclooxygenase-2 (COX-2) mRNA stability by interleukin-1 beta (IL-1 beta) and tumor necrosis factor-alpha (TNF-alpha) in human in vitro differentiated macrophages. Biochem. Pharmacol. 59, 187-194.[CrossRef][Medline]

Hujanen, E. S., Vaisanen, A., Zheng, A., Tryggvason, K. and Turpeenniemi-Hujanen, T. (1994). Modulation of M(r) 72,000 and M(r) 92,000 type-IV collagenase (gelatinase A and B) gene expression by interferons alpha and gamma in human melanoma. Int. J. Cancer 58, 582-586.[Medline]

Hynes, R. O. (2002). Integrins: bidirectional, allosteric signaling machines. Cell 110, 673-687.[CrossRef][Medline]

Jiang, Y. and Muschel, R. J. (2002). Regulation of matrix metalloproteinase-9 (MMP-9) by translational efficiency in murine prostate carcinoma cells. Cancer Res. 62, 1910-1914.[Abstract/Free Full Text]

Johansson, N., Ala-aho, R., Uitto, V., Grenman, R., Fusenig, N. E., Lopez-Otin, C. and Kahari, V. M. (2000). Expression of collagenase-3 (MMP-13) and collagenase-1 (MMP-1) by transformed keratinocytes is dependent on the activity of p38 mitogen-activated protein kinase. J. Cell Sci. 113, 227-235.[Abstract/Free Full Text]

Johnsen, M., Lund, L. R., Romer, J., Almholt, K. and Dano, K. (1998). Cancer invasion and tissue remodeling: common themes in proteolytic matrix degradation. Curr. Opin. Cell Biol. 10, 667-671.[CrossRef][Medline]

Juarez, J., Clayman, G., Nakajima, M., Tanabe, K. K., Saya, H., Nicolson, G. L. and Boyd, D. (1993). Role and regulation of expression of 92-kDa type-IV collagenase (MMP-9) in 2 invasive squamous-cell-carcinoma cell lines of the oral cavity. Int. J. Cancer 55, 10-18.[Medline]

Kheradmand, F., Werner, E., Tremble, P., Symons, M. and Werb, Z. (1998). Role of Rac1 and oxygen radicals in collagenase-1 expression induced by cell shape change. Science 280, 898-902.[Abstract/Free Full Text]

Kim, M. H., Kitson, R. P., Albertsson, P., Nannmark, U., Basse, P. H., Kuppen, P. J., Hokland, M. E. and Goldfarb, R. H. (2000). Secreted and membrane-associated matrix metalloproteinases of IL-2-activated NK cells and their inhibitors. J. Immunol. 164, 5883-5889.[Abstract/Free Full Text]

Kreidberg, J. A. (2000). Functions of {alpha}3ß1 integrin. Curr. Opin. Cell Biol. 12, 548-553.[CrossRef][Medline]

Kreidberg, J. A., Donovan, M. J., Goldstein, S. L., Rennke, H., Shepherd, K., Jones, R. C. and Jaenisch, R. (1996). Alpha 3 beta 1 integrin has a crucial role in kidney and lung organogenesis. Development 122, 3537-3547.[Abstract/Free Full Text]

Kupferman, M. E., Fini, M. E., Muller, W. J., Weber, R., Cheng, Y. and Muschel, R. J. (2000). Matrix metalloproteinase 9 promoter activity is induced coincident with invasion during tumor progression. A. J. Pathol. 157, 1777-1783.

Li, J., Schwimmbeck, P. L., Tschope, C., Leschka, S., Husmann, L., Rutschow, S., Reichenbach, F., Noutsias, M., Kobalz, U., Poller, W. et al. (2002). Collagen degradation in a murine myocarditis model: relevance of matrix metalloproteinase in association with inflammatory induction. Cardiovasc. Res. 56, 235-247.[CrossRef][Medline]

Liu, Z., Zhou, X., Shapiro, S. D., Shipley, J. M., Twining, S. S., Diaz, L. A., Senior, R. M. and Werb, Z. (2000). The serpin alpha1-proteinase inhibitor is a critical substrate for gelatinase B/MMP-9 in vivo. Cell 102, 647-655.[Medline]

Lochter, A., Navre, M., Werb, Z. and Bissell, M. J. (1999). {alpha}1 and {alpha}2 integrins mediate invasive activity of mouse mammary carcinoma cells through regulation of stromelysin-1 expression. Mol. Biol. Cell 10, 271-282.[Abstract/Free Full Text]

Loflin, P., Chen, C. Y. and Shyu, A. B. (1999). Unraveling a cytoplasmic role for hnRNP D in the in vivo mRNA destabilization directed by the AU-rich element. Genes Dev. 13, 1884-1897.[Abstract/Free Full Text]

Manohar, A., Shome, S. G., Lamar, J., Stirling, L., Iyer, V., Pumiglia, K. and DiPersio, C. M. (2004). Alpha3 beta1 integrin promotes keratinocyte cell survival through activation of a MEK/ERK signaling pathway. J. Cell Sci. 117, 4043-4054.[Abstract/Free Full Text]

McCawley, L. J., O'Brien, P. and Hudson, L. G. (1998). Epidermal growth factor (EGF)- and scatter factor/hepatocyte growth factor (SF/HGF)-mediated keratinocyte migration is coincident with induction of matrix metalloproteinase (MMP)-9. J. Cell. Physiol. 176, 255-265.[CrossRef][Medline]

McCawley, L. J., Li, S., Wattenberg, E. V. and Hudson, L. G. (1999). Sustained activation of the mitogen-activated protein kinase pathway. A mechanism underlying receptor tyrosine kinase specificity for matrix metalloproteinase-9 induction and cell migration. J. Biol. Chem. 274, 4347-4353.[Abstract/Free Full Text]

Meadows, K. N., Bryant, P. and Pumiglia, K. (2001). Vascular endothelial growth factor induction of the angiogenic phenotype requires Ras activation. J. Biol. Chem. 276, 49289-49298.[Abstract/Free Full Text]

Montero, L. and Nagamine, Y. (1999). Regulation by p38 mitogen-activated protein kinase of adenylate- and uridylate-rich element-mediated urokinase-type plasminogen activator (uPA) messenger RNA stability and uPA-dependent in vitro cell invasion. Cancer Res. 59, 5286-5293.[Abstract/Free Full Text]

Morini, M., Mottolese, M., Ferrari, N., Ghiorzo, F., Buglioni, S., Mortarini, R., Noonan, D. M., Natali, P. G. and Albini, A. (2000). The alpha 3 beta 1 integrin is associated with mammary carcinoma cell metastasis, invasion, and gelatinase B (MMP-9) activity. Int. J. Cancer 87, 336-342.[CrossRef][Medline]

Mukhopadhyay, S., Munshi, H. G., Kambhampati, S., Sassano, A., Platanias, L. C. and Stack, M. S. (2004). Calcium induced matrix metalloproteinase 9 gene expression is differentially regulated by ERK1/2 and p38 MAPK in oral keratinocytes and oral squamous cell carcinoma. J. Biol. Chem. 279, 33139-33146.[Abstract/Free Full Text]

Natali, P. G., Nicotra, M. R., Bartolazzi, A., Cavaliere, R. and Bigotti, A. (1993). Integrin expression in cutaneous malignant melanoma: association of the alpha 3/beta 1 heterodimer with tumor progression. Int. J. Cancer 54, 68-72.[Medline]

Patriarca, C., Alfano, R. M., Sonnenberg, A., Graziani, D., Cassani, B., de Melker, A., Colombo, P., Languino, L. R., Fornaro, M., Warren, W. H. et al. (1998). Integrin laminin receptor profile of pulmonary squamous cell and adenocarcinomas. Hum. Pathol. 29, 1208-1215.[Medline]

Pyke, C., Ralfkiaer, E., Huhtala, P., Hurskainen, T., Dano, K. and Tryggvason, K. (1992). Localization of messenger RNA for Mr 72,000 and 92,000 type IV collagenases in human skin cancers by in situ hybridization. Cancer Res. 52, 1336-1341.[Abstract]

Pyke, C., Salo, S., Ralfkiaer, E., Romer, J., Dano, K. and Tryggvason, K. (1995). Laminin-5 is a marker of invading cancer cells in some human carcinomas and is coexpressed with the receptor for urokinase plasminogen activator in budding cancer cells in colon adenocarcinomas. Cancer Res. 55, 4132-4139.[Abstract]

Renshaw, M. W., Ren, X. D. and Schwartz, M. A. (1997). Growth factor activation of MAP kinase requires cell adhesion. EMBO J. 16, 5592-5599.[Abstract/Free Full Text]

Retta, S. F., Cassara, G., D'Amato, M., Alessandro, R., Pellegrino, M., Degani, S., de Leo, G., Silengo, L. and Tarone, G. (2001). Cross talk between beta(1) and alpha(V) integrins: beta(1) affects beta(3) mRNA stability. Mol. Biol. Cell 12, 3126-3138.[Abstract/Free Full Text]

Reunanen, N., Li, S. P., Ahonen, M., Foschi, M., Han, J. and Kahari, V. M. (2002). Activation of p38 alpha MAPK enhances collagenase-1 (matrix metalloproteinase (MMP)-1) and stromelysin-1 (MMP-3) expression by mRNA stabilization. J. Biol. Chem. 277, 32360-32368.[Abstract/Free Full Text]

Rheinwald, J. G. and Beckett, M. A. (1981). Tumorigenic keratinocyte lines requiring anchorage and fibroblast support cultures from human squamous cell carcinomas. Cancer Res. 41, 1657-1663.[Abstract]

Sachs, A. B. (1993). Messenger RNA degradation in eukaryotes. Cell 74, 413-421.[Medline]

Samarakoon, R. and Higgins, P. J. (2002). MEK/ERK pathway mediates cell-shape-dependent plasminogen activator inhibitor type 1 gene expression upon drug-induced disruption of the microfilament and microtubule networks. J. Cell Sci. 115, 3093-3103.[Abstract/Free Full Text]

Sehgal, I., Baley, P. A. and Thompson, T. C. (1996). Transforming growth factor beta1 stimulates contrasting responses in metastatic versus primary mouse prostate cancer-derived cell lines in vitro. Cancer Res. 56, 3359-3365.[Abstract]

Sehgal, I. and Thompson, T. C. (1999). Novel regulation of type IV collagenase (matrix metalloproteinase-9 and -2) activities by transforming growth factor-beta1 in human prostate cancer cell lines. Mol. Biol. Cell 10, 407-416.[Abstract/Free Full Text]

Shimajiri, S., Arima, N., Tanimoto, A., Murata, Y., Hamada, T., Wang, K. Y. and Sasaguri, Y. (1999). Shortened microsatellite d(CA)21 sequence down-regulates promoter activity of matrix metalloproteinase 9 gene. FEBS Lett. 455, 70-74.[CrossRef][Medline]

Tanaka, H., Hojo, K., Yoshida, H., Yoshioka, T. and Sugita, K. (1993). Molecular cloning and expression of the mouse 105-kDa gelatinase cDNA. Biochem. Biophy. Res. Commun. 190, 732-740.[CrossRef][Medline]

Thant, A. A., Nawa, A., Kikkawa, F., Ichigotani, Y., Zhang, Y., Sein, T. T., Amin, A. R. and Hamaguchi, M. (2000). Fibronectin activates matrix metalloproteinase-9 secretion via the MEK1-MAPK and the PI3K-Akt pathways in ovarian cancer cells. Clin. Exp. Metastasis 18, 423-428.[CrossRef][Medline]

Thomas, G. J., Lewis, M. P., Hart, I. R., Marshall, J. F. and Speight, P. M. (2001). AlphaVbeta6 integrin promotes invasion of squamous carcinoma cells through up-regulation of matrix metalloproteinase-9. Int. J. Cancer 92, 641-650.[CrossRef][Medline]

Tran, H., Maurer, F. and Nagamine, Y. (2003). Stabilization of urokinase and urokinase receptor mRNAs by HuR is linked to its cytoplasmic accumulation induced by activated mitogen-activated protein kinase-activated protein kinase 2. Mol. Cell. Biol. 23, 7177-7188.[Abstract/Free Full Text]

Tsuji, T., Kawada, Y., Kai-Murozono, M., Komatsu, S., Han, S. A., Takeuchi, K., Mizushima, H., Miyazaki, K. and Irimura, T. (2002). Regulation of melanoma cell migration and invasion by laminin-5 and alpha3beta1 integrin (VLA-3). Clin. Exp. Metastasis 19, 127-134.[CrossRef][Medline]

Wang, H., Fu, W., Im, J. H., Zhou, Z., Santoro, S. A., Iyer, V., DiPersio, C. M., Yu, Q. C., Quaranta, V., Al-Mehdi, A. et al. (2004). Tumor cell {alpha}3ß1 integrin and vascular laminin-5 mediate pulmonary arrest and metastasis. J. Cell Biol. 164, 935-941.[Abstract/Free Full Text]

Werb, Z. (1997). ECM and cell surface proteolysis: regulating cellular ecology. Cell 91, 439-442.[Medline]

Westermarck, J. and Kahari, V.-M. (1999). Regulation of matrix metalloproteinase expression in tumor invasion. FASEB J. 13, 781-792.[Abstract/Free Full Text]

Winzen, R., Kracht, M., Ritter, B., Wilhelm, A., Chen, C. Y., Shyu, A. B., Muller, M., Gaestel, M., Resch, K. and Holtmann, H. (1999). The p38 MAP kinase pathway signals for cytokine-induced mRNA stabilization via MAP kinase-activated protein kinase 2 and an AU-rich region-targeted mechanism. EMBO J. 18, 4969-4980.[Abstract/Free Full Text]

Xu, J. and Clark, R. A. (1996). Extracellular matrix alters PDGF regulation of fibroblast integrins. J. Cell Biol. 132, 239-249.[Abstract]

Yan, L., Moses, M. A., Huang, S. and Ingber, D. E. (2000). Adhesion-dependent control of matrix metalloproteinase-2 activation in human capillary endothelial cells. J. Cell Sci. 113, 3979-3987.[Abstract/Free Full Text]

Yuspa, S. H. (1994). The pathogenesis of squamous cell cancer: lessons learned from studies of skin carcinogenesis–thirty-third G. H. A. Clowes Memorial Award Lecture. Cancer Res. 54, 1178-1189.[Abstract]

Zeigler, M. E., Chi, Y., Schmidt, T. and Varani, J. (1999). Role of ERK and JNK pathways in regulating cell motility and matrix metalloproteinase 9 production in growth factor-stimulated human epidermal keratinocytes. J. Cell. Physiol. 180, 271-284.[CrossRef][Medline]


Related articles in JCS:

Integrins look to mRNA to remodel

JCS 2005 118: 603. [Full Text]  




This Article
Summary
Figures Only
Full Text (PDF)
All Versions of this Article:
jcs.01708v1
118/6/1185    most recent
Alert me when this article is cited
Alert me if a correction is posted
Services
Email this article to a friend
Related articles in JCS
Similar articles in this journal
Similar articles in PubMed
Alert me to new issues of the journal
Download to citation manager
Google Scholar
Articles by Iyer, V.
Articles by DiPersio, C. M.
Articles citing this Article
PubMed
PubMed Citation
Articles by Iyer, V.
Articles by DiPersio, C. M.