Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Mol Cell Biol. 2011 Nov; 31(21): 4270–4285.
PMCID: PMC3209326
PMID: 21896780

Map2k4 Functions as a Tumor Suppressor in Lung Adenocarcinoma and Inhibits Tumor Cell Invasion by Decreasing Peroxisome Proliferator-Activated Receptor γ2 Expression

Abstract

MAP2K4 encodes a dual-specificity kinase (mitogen-activated protein kinase kinase 4, or MKK4) that is mutated in a variety of human malignancies, but the biochemical properties of the mutant kinases and their roles in tumorigenesis have not been fully elucidated. Here we showed that 8 out of 11 cancer-associated MAP2K4 mutations reduce MKK4 protein stability or impair its kinase activity. On the basis of findings from bioinformatic studies on human cancer cell lines with homozygous MAP2K4 loss, we posited that MKK4 functions as a tumor suppressor in lung adenocarcinomas that develop in mice owing to expression of mutant Kras and Tp53. Conditional Map2k4 inactivation in the bronchial epithelium of mice had no discernible effect alone but increased the multiplicity and accelerated the growth of incipient lung neoplasias induced by oncogenic Kras. MKK4 suppressed the invasion and metastasis of Kras-Tp53-mutant lung adenocarcinoma cells. MKK4 deficiency increased peroxisomal proliferator-activated receptor γ2 (PPARγ2) expression through noncanonical MKK4 substrates, and PPARγ2 enhanced tumor cell invasion. We conclude that Map2k4 functions as a tumor suppressor in lung adenocarcinoma and inhibits tumor cell invasion by decreasing PPARγ2 levels.

INTRODUCTION

A growing body of evidence from cellular and murine models indicates that tumor cells must acquire multiple somatic mutations in order to invade locally, disseminate hematogenously, and colonize at distant sites. Genomic sequencing studies have been performed on human tumor tissues in an effort to identify somatic mutations that might act in concert to promote tumorigenesis (11, 16, 39, 40, 44). The relevance of those mutations is, in many cases, obscure, because the mutated genes have no documented roles in cancer, are found at low prevalence in tumors, or have indeterminate effects on the functions of their gene products. Although bioinformatic tools have aided investigators in predicting whether specific mutations are “drivers” or “passengers” in tumor cells (39, 44), determining whether these mutations do, in fact, promote tumorigenesis and how they do so requires biochemical studies on the mutated gene products and biological studies on tumor cells that express them.

Elucidating the genetic mutations that drive the progression of lung cancer has been an area of particular interest, given that lung cancer is the most common cause of cancer-related death in Western countries. Mice that express KrasG12D alleles inducibly, conditionally, or somatically develop lung adenocarcinomas with low invasive and metastatic potential (13, 17, 20, 21, 23). The subjection of Kras-mutant mice to a second oncogenic event leads to lung adenocarcinomas that arise earlier, grow faster, and metastasize widely. Proven cooperative events include inactivating mutations in tumor suppressor genes (Tsc1, Lkb1, or Pten), overexpression of Hif2α, and introduction of Tp53R172H, a mutation that confers metastatic potential to tumors in mice and is found in patients with Li-Fraumeni syndrome and sporadic lung cancer (5, 19, 28, 33, 52). However, the incidence of metastatic disease in these double-mutant mice is, in many cases, far less than 100%, suggesting that additional genetic events are necessary for such an outcome.

Genomic studies to identify somatic mutations in the kinome of human cancer samples found that a total of 11 tumors (3% of the 356 tumors evaluated) had somatic mutations in MAP2K4 (11, 16, 40). These mutations are located primarily in the kinase domain of the MAP2K4 gene; include frameshift, nonsense, and missense mutations; and occur in colorectal cancer, non-small-cell lung cancer, melanoma, and ovarian cancer specimens. MAP2K4 encodes MKK4, a dual-specificity kinase that is activated by environmental stress, cytokines, and peptide growth factors (6). As a component of stress signaling pathways, MKK4 directly phosphorylates c-Jun N-terminal kinase (JNK) and p38 (47). MKK4 and its substrates can reportedly suppress or promote tumorigenesis and metastasis, leading investigators to conclude that MKK4 and its substrates play an important role in tumor progression and can function as either a tumor suppressor or an oncogene (18, 25, 26, 34, 45, 47). However, the spectrum of MAP2K4 somatic mutations has not been fully evaluated in MKK4 biochemical assays to delineate loss or gain of function, and downstream effectors of MKK4 that mediate cellular transformation have not been adequately explored.

Here we performed a comprehensive biochemical analysis of a panel of MAP2K4 somatic mutations reported in cases of human cancers and found that the majority represent loss-of-function mutations, and data mining of human cancer cell lines revealed that homozygous inactivating MAP2K4 mutations are accompanied by TP53 and KRAS mutations. Therefore, we posited that MKK4 functions as a tumor suppressor in lung adenocarcinomas driven by mutant Kras and Tp53 and tested this hypothesis in mice. The findings reported here support this hypothesis and revealed that MKK4 functions as a tumor suppressor partly by decreasing levels of peroxisomal proliferator-activated receptor γ (PPARγ), a nuclear receptor for fatty acids and eicosanoids that plays divergent roles in different tumor models (31, 32, 37, 41, 42).

MATERIALS AND METHODS

Antibodies and reagents.

Polyclonal antibodies against MKK4, p-MKK4, JNK, p-JNK, p38, p-p38, extracellular signal-regulated kinase (ERK), p-ERK, MKK7, tubulin (Cell Signaling Technologies), actin (Sigma-Aldrich), MEKK1, hemagglutinin (HA), and PPARγ2 (Santa Cruz Biotechnologies) were purchased. Sorbitol, MG-132, actinomycin D (Sigma-Aldrich), clasto-lactacystin-β-lactone, SP600125, SB202190 (Calbiochem), and T0070907 (Cayman Chemical) were purchased.

Plasmids and site-directed mutagenesis.

Human MKK4 cDNA (provided by Jia Le Dai, M. D. Anderson Cancer Center) was Flag-tagged at the N terminus and inserted into pLHCX vector (Clontech). MEKK1/pCMV5 (Zhimin Lu, M. D. Anderson Cancer Center), HA-ubiquitin/pcDNA3.1 (Edward Yeh, M. D. Anderson Cancer Center), and HA-JNK2 (APF)/SRα (Bing Su, Yale School of Medicine) were gifts. Murine Pparg2 cDNA (catalog no. 8895) was purchased (Addgene). Vectors expressing mouse Map2k4 (OriGene), Bgn, Vegfc, Satb1, Pparg (SA Biosciences), MAP2K4, and PPARG short hairpin RNAs (shRNAs) were purchased (Open Biosystems). To construct MAP2K4 mutants, a PCR-based site-directed mutagenesis strategy (36) was carried out using Flag-MKK4/pLHCX as a backbone. Site-specific primers are listed in Fig. 1A.

An external file that holds a picture, illustration, etc.
Object name is zmb9991092650001.jpg

MAP2K4 mutations in human cancers and sequence alignment of MKK family members. (A) MAP2K4 mutations. Eleven MAP2K4 mutations found in human cancers analyzed herein are depicted schematically and in tabular form, denoting the cancer types associated with them (which are referenced) and the primers used to create these mutants by site-directed mutagenesis (mutation sites are underlined and in bold). The synthetic MAP2K4 dominant-active S257E; T261D (ED) mutant and dominant-negative K131R (KR) mutant used as controls in this study are indicated. *, nonsense mutation; fs, frameshift mutation. (B) Amino acid sequences in MKK family members were aligned using the ClustalW method. MAP2K4 mutation sites are labeled and boxed. Certain mutation sites (N234, S251, and P326) are well conserved across MKK family members.

Cell culture.

293T, GP-293, and RAW 264.7 (mouse macrophage) cells were maintained in Dulbecco's modification of Eagle's medium (DMEM) (Mediatech) supplemented with 10% fetal bovine serum (FBS) (Sigma-Aldrich). 393P (nonmetastatic) and 344SQ (highly metastatic) murine lung cancer cells, BxPC3 cells, and H2009 cells were cultured in RPMI 1640 (Mediatech) with 10% FBS. Cells were transfected using Lipofectamine with Plus reagent (Invitrogen) or infected with retrovirus produced from GP-293 packaging cells (Clontech). For soft-agar assays, 5 × 104 cells (in 0.3% agar) were seeded into 6-well plates layered with 0.8% agar, and colonies were stained with 0.5 mg/ml nitrotetrazolium blue (Sigma-Aldrich) 21 days later. For the migration and invasion assays, RAW 264.7 cells (1 × 105) were plated in the bottom chamber and 393P cells (1 × 105) were cultured in the upper chamber of 24-well Transwell and Matrigel-coated chambers, respectively (BD Biosciences). After 16 h of incubation, migrating or invading cells were stained with 0.1% crystal violet, photographed, and counted. Cell numbers were counted using a hemocytometer or a Countess automated cell counter (Invitrogen).

Western blotting and in vitro kinase assay.

After 293T cells were transiently transfected with Flag-MKK4 mutants, cells were lysed in 50 mM Tris-HCl (pH 7.4)–150 mM NaCl–1 mM EDTA–1% Triton X-100–protease-phosphatase inhibitors (Sigma-Aldrich). Cell lysates were used for Western blotting or immunoprecipitated with anti-Flag M2 affinity gel (Sigma-Aldrich) for the kinase assay. The immunoprecipitated complex was incubated in kinase assay buffer (20 mM HEPES [pH 7.5], 20 mM β-glycerophosphate, 100 mM MgCl2, 50 mM NaCl, 1 mM dithiothreitol [DTT], 0.1 mM orthovanadate, 1 μM okadaic acid, 50 μM ATP) with glutathione S-transferase–JNK1 (GST-JNK1) (K55M mutant; Carna Biosciences) and [γ-32P]ATP (MP Biomedicals) at 30°C for 30 min. After sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE), the gel was dried and exposed to X-ray film to determine the phosphorylation of JNK1.

35S-Met labeling.

Flag-MKK4-transfected 293T cells were incubated with Met/Cys-free DMEM (Invitrogen) for 30 min and then labeled with 35S-Met (Tran35S-Label; MP Biomedicals) for 45 min. After incubation using normal growth medium for the indicated times, cell lysates were prepared and immunoprecipitated with anti-Flag M2 affinity gel. The immunoprecipitated complex was separated using SDS-PAGE, and 35S-labeled MKK4 was detected using autoradiography.

Ubiquitination assay.

293T cells were transfected with Flag-MKK4 and HA-Ub and lysed. Lysates were subjected to immunoprecipitation with anti-Flag M2 affinity gel. After SDS-PAGE, polyubiquitinated MKK4 was detected by Western blotting with anti-HA antibody.

Animal husbandry.

Before their initiation, all mouse studies were submitted to, and approved by, the Institutional Animal Care and Use Committee (IACUC) at the University of Texas M. D. Anderson Cancer Center. Mice received care and were euthanized according to the standards set forth by the IACUC. KrasG12D, MKK4L/L, and Scgb1a1Cre strains were interbred to evaluate the effects of oncogenic Kras and MKK4 loss of function in the lung as described previously (19, 46). Syngeneic (129/Sv) mice (n = 10 or 15 per group) were injected subcutaneously in the right flank with 393P murine lung adenocarcinoma cells (106 cells per mouse) that had been stably transfected with MKK4-shRNA or scrambled control retroviral vectors. Mice were monitored daily for tumor growth, sacrificed at 8 weeks, and subjected to necropsy to isolate primary tumors and sites of metastasis, which were confirmed histologically by analysis of hematoxylin- and eosin-stained, formalin-fixed tissues.

Affymetrix gene expression profiling.

Total RNA was extracted from 393P-scr and 393P-shZ cells by the use of a RiboPure kit (Ambion) and then hybridized to an Affymetrix GeneChip mouse genome 430 2.0 array (Asuragen). Data processing, determination of differentially expressed genes, and gene ontology term enrichment analysis were carried out essentially as described previously (10).

RT-PCR.

Total RNA was isolated from Flag-MKK4-transfected 293T cells by the use of TRIzol (Invitrogen) according to manufacturer's protocol. After reverse transcription, PCR was performed to detect Flag-tagged MKK4 mRNA with forward (5′-GGATTACAAGGATGACGACGA-3′) and reverse (5′-CTGCCATTATTTGCCCACTT-3′) primers. Quantitative reverse transcriptase PCR (RT-PCR) assays were performed using a SYBR green-based system (Applied Biosystems) as previously described (8). Values were normalized on the basis of ribosomal protein L32 values. Primer sequences are available upon request.

RNA polymerase II-chromatin immunoprecipitation (ChIP) assay.

Cells were cross-linked with 1% formaldehyde and then incubated in lysis buffer (50 mM Tris-HCl [pH 8.1], 1% SDS, 10 mM EDTA, protease inhibitor cocktail) on ice for 10 min. After sonication (Cole-Parmer GEX-130 sonicator) (50% power; 20 cycles of pulse on for 10 s and pulse off for 10 s), samples were immunoprecipitated with anti-RNA polymerase II antibody (Millipore) or anti-mouse IgG. DNA was eluted and purified with PCR purification kit (Qiagen), and then quantitative PCR was carried out with specific primers to amplify the promoter regions of Pparg2 and Gapdh to normalize for differences in loading.

Luciferase assays.

393P and 344SQ cells were seeded on 24-well plates (1 × 105 cells/well) 1 day before transfection and then transfected with 500 ng of pGL3-basic (Promega) or PPRE (PPAR-responsive element)/pGL3-basic (from David Mangelsdorf, University of Texas Southwestern Medical School). After 48 h, luciferase activity was measured using a dual-luciferase reporter assay system (Promega).

Statistical analysis.

With the exception of Affymetrix expression arrays, data were analyzed using Student's t test, Spearman's correlation test, and Fisher's exact test.

RESULTS

Using computational tools, we examined the structural consequences of 11 MAP2K4 somatic mutations (point mutations and deletions) that are representative of the mutations reported thus far (11, 16, 40). The locations of mutations within the kinase domain are illustrated graphically in Fig. 1A. These mutations involve sites that are conserved among MKK family members (Fig. 1B). Mapping of the mutation sites on solved MKK4 crystal structures (35) revealed that several mutations involve residues in critical domain structures, including the ATP-binding pocket (N234I), activation loop (S251N), and αC-helix (R134Q and Q142L) (Fig. 2A). The predicted effects of these mutations on kinase function were further examined by using PolyPhen (2) and SIFT (38), which examine the functional consequence of a mutation on the basis of sequence conservation and physical properties of amino acids. Both of these tools classified several of the MAP2K4 mutations (S251N, N234I, P326L, V321M, and R154W) as functionally “damaging” and the others (R134Q, A279T, and Q142L) as “tolerated” (Fig. 2B). A third tool, CanPredict (24), which is designed to distinguish cancer-associated mutations from genetic polymorphisms and rare mutations, classified all of the “damaging” mutations and one “tolerated” mutation (Q142L) as “likely cancer-associated” and the remaining tolerated mutations (R134Q and A279T) as “likely non-cancer-associated” (Fig. 2B).

An external file that holds a picture, illustration, etc.
Object name is zmb9991092650002.jpg

MAP2K4 mutants exhibit aberrant kinase activities. (A) MKK4 mutation sites were superimposed on an MKK4 X-ray crystal structure (Protein Data Bank; 3ALO) using PyMOL (http://www.pymol.org). Mutated amino acids and their locations are indicated with arrows and labeled in red. (B) Prediction algorithms were used to predict whether MAP2K4 missense mutations affect kinase activity (PolyPhen and SIFT) and are associated with cancer (CanPredict). (C) 293T cells were transfected with Flag-tagged MKK4 mutants and then immunoprecipitated (IP) with Flag antibody. In vitro kinase assays were performed using a kinase-dead mutant form (K55M) of JNK1 as a substrate. As controls, the synthetic MAP2K4 dominant-active S257E; T261D (ED) and dominant-negative K131R (KR) mutants were used. KA, kinase activity; WB, Western blot; total WB, WB of total lysate. (D) Quantification of kinase assays. The activity of each mutant was normalized to that of the wild type (WT), which was set at a value of 1. Graph values are presented as means ± standard deviations (SD) (n = 3), and P values were calculated with Student's t test (* or #, P < 0.01). (E) Response of mutants to stress. 293T cells were transfected with Flag-MKK4 alone and treated with UV (250 J/m2 for 30 min) or with sorbitol (0.5 M for 3 h). Kinase assays performed as described above. (F) Response of mutants to upstream signal. 293T cells were cotransfected with Flag-MKK4 mutants and MEKK1 (an upstream kinase of MKK4), and then in vitro kinase assays were performed.

The same 11 MAP2K4 somatic mutants were created by performing site-directed mutagenesis using human MAP2K4 cDNA. The mutants were expressed in 293T cells, immunoprecipitated, and subjected to in vitro kinase assays using as a substrate a mutant JNK1 (JNK1-K55M) that is incapable of autophosphorylation. Relative to wild-type (WT) MKK4, eight of the mutations (V321M, P326L, R154W, S251N, N234I, W310*, R304*, and I295fs*23) were loss-of-function mutations, one (Q142L) was a gain-of-function mutation with activity similar to that of a synthetic constitutively active mutant (ED), and two (R134Q and A279T) were mutations with activities similar to that of wild-type MKK4 (Fig. 2C and D).

Given the importance of MKK4 in stress-activated pathways (47), we examined the effect of UV irradiation and sorbitol on kinase activity, focusing first on the missense mutants. The results of activation of Q142L, R134Q, and A279T were similar to those seen with the wild type, whereas R154W, P326L, S251N, and N234I had sharply attenuated responses (Fig. 2E). To determine whether the most inactive mutants (P326L, S251N, and N234I) were also refractory to an MKK4 kinase, these mutants were cotransfected with MEKK1 in 293T cells and subjected to MKK4 kinase assays. Indeed, P326L and S251N had attenuated responses to MEKK1, and N234I was not activated at all (Fig. 2F). This was not related to an inability to bind to JNK, as determined on the basis of findings from immunoprecipitation and Western blot assays (data not shown). Furthermore, when expressed in a green fluorescent protein (GFP)-tagged construct and transfected into PT67 cells, the mutants localized within the nucleus and cytoplasm in a pattern similar to that seen with wild-type MKK4 (data not shown), suggesting that their loss of function was not related to aberrant trafficking.

In 293T cells, the C-terminal truncation mutant proteins (W310*, R304*, and I295fs*23) were present at sharply reduced levels relative to those of the wild-type and missense mutant proteins (Fig. 3A). Findings from semiquantitative RT-PCR analysis performed on RNA samples from the transfectants (wild type and C-terminally truncated mutants) revealed no differences in the levels of exogenous MKK4 mRNA (Fig. 3B), raising the possibility that these mutants had reduced protein stability. Indeed, 35S-methionine labeling studies revealed the C-terminally truncated mutants to have a shorter half-life than the wild-type protein (Fig. 3C and D). Furthermore, these mutants (W310*, R304*, and I295fs*23) were highly ubiquitinated (Fig. 3E), and the degradation of one of them, W310*, was attenuated by treatment with the proteasome inhibitor MG132 or clasto-lactacystin-β-lactone (Fig. 3F), suggesting that its degradation was mediated through ubiquitin-dependent pathways in a manner similar to that seen with wild-type MKK4 (3). Deletion of 200 C-terminal amino acids from wild-type MKK4 enhanced MKK4 ubiquitination (Fig. 3G), indicating that the C-terminal domains are required for protein stability.

An external file that holds a picture, illustration, etc.
Object name is zmb9991092650003.jpg

C-terminally truncated MAP2K4 mutants are rapidly degraded. (A) 293T cells were transfected with MAP2K4 mutants and subjected to Western blotting using anti-MKK4 or antitubulin. The protein levels of truncated forms (ΔC-term) were very low compared with those of full-length forms. (B) Reverse transcriptase PCR of Flag-tagged MKK4 mRNA. There was no difference in mRNA expression levels between wild-type and C-terminally truncated MKK4. GAPDH, glyceraldehyde-3-phosphate dehydrogenase. (C) 35S-Met labeling to measure protein stability, demonstrating the extinction of the truncated mutant forms by 4 h after the pulse. (D) Quantification of data in panel C by densitometric analysis. (E) In vivo ubiquitination assay. Immunoprecipitation and Western blotting of transfectants revealed that truncated forms were more highly ubiquitinated than were full-length forms. A total of 2-fold more lysates of the truncated mutants were loaded to compensate for their rapid degradation. (F) Proteasome inhibitors MG132 (50 μM) and β-lactone (clasto-lactacystin-β-lactone) (5 μM) attenuated the degradation of C-terminally truncated mutants based on 35S-Met labeling analyses. (G) Deletion of 300 (300del) but not 100 (100del) or 200 (200del) C-terminal amino acids abrogated ubiquitination of MKK4. MKK4 C-terminal deletion constructs were transiently transfected into 293T cells, immunoprecipitated using anti-Flag antibodies, and subjected to Western blotting. The numbers of N-terminal amino acids remaining in the deletion constructs are indicated on right side of the panel.

We searched existing databases and found 17 (2.2% of 746 analyzed) human cancer cell lines that contain different inactivating MAP2K4 mutations (16 with homozygous deletions, C-terminal truncations, or point mutations and 1 with a heterozygous point mutation [Table 1 ]) and comprise a variety of epithelial tumor types, including lung cancer (4 cell lines). Therefore, we posited that MAP2K4 functions as a tumor suppressor in lung cancer and tested this hypothesis by inactivating Map2k4 in the bronchial epithelium of mice in the presence or absence of a potent oncogenic driver, KrasG12D. Mice were interbred that have a Map2k4 allele with LoxP sites surrounding exon 6 within the kinase domain (MKK4L/L) (46), a Kras allele with a G12D mutation under the control of a Lox-Stop-Lox transcriptional cassette (KrasG12D) (20), and a Scgb1a1 allele in which the Cre recombinase has been knocked in and that is expressed specifically in Clara cells (19). Cohorts were generated that did not express Cre (WT; n = 40), expressed KrasG12D alone (KrasG12D; n = 34), were MKK4 deficient (MKK4L/L; n = 38), or expressed KrasG12D and were MKK4 deficient (KrasG12D) (MKK4L/L; n = 34). Mice were sacrificed either at the first sign of morbidity for Kaplan-Meier analysis or at 7 to 10 months of age for examination of tumor histologic features. KrasG12D mice in this age range are expected to have incipient lung neoplasias of various histologic descriptions and rare lung adenocarcinomas (19).

Table 1.

MAP2K4 mutant cell lines from COSMIC (Catalogue of Somatic Mutations in Cancer; http://www.sanger.ac.uk/genetics/CGP/cosmic)

Cell lineMAP2K4 mutationbPrimary tissue typeHistologyZygosityOther mutation(s)
TP53CDKN2RB1BRAFSMAD4KRASPTEN
647-V1-?_115+?delUrinary tractTransitional-cell carcinomaHomomutamut
AsPC-11-?_393+?delPancreasCarcinomaHomomutmutmut
BxPC-31041_1200del160PancreasCarcinomaHomomutmutmut
DU-44751-?_1200+?delBreastDuctal carcinomaHomomutmut
EFM-191-?_1200+?delBreastDuctal carcinomaHomomutmut
HCC2998460C>T (R154W)Large intestineAdenocarcinomaHomomutmut
KLE1_218del218EndometriumAdenocarcinomaHomomut
MZ7-mel752G>A (S251N)SkinMalignant melanomaHeteromutmut
NCI-H146814_891del78LungSmall-cell carcinomaHomomutmut
NCI-H23421075A>T (K359*)LungAdenocarcinomaHomomut
NCI-H2405814_1200del387LungAdenocarcinomaHomomutmutmutmut
NCI-H7161-?_115+?delLarge intestineAdenocarcinomaHomomut
NCI-H774219-?_1200+?delLungSmall-cell carcinomaHomomutmutmut
NMC-G11-?_115+?delCentral nervous systemGliomaHomomutmut
OAW-281-?_1200+?delOvaryCarcinomaHomo
SW403219_891del673Large intestineAdenocarcinomaHomomutmutmut
SW620219_891del673Large intestineAdenocarcinomaHomomutmutmut
amut, mutation.
b*, nonsense mutation.

Based on Kaplan-Meier analysis of the two cohorts, Map2k4 inactivation accelerated mortality (P = 0.0186; log rank test) (Fig. 4A). Lung tumors were present exclusively in the two cohorts that expressed KrasG12D (examples illustrated in Fig. 4D). The conditional Kras allele was recombined in all of the tumors analyzed (Fig. 4B), and the conditional Map2k4 allele was recombined in tumors exclusively from the KrasG12D;MKK4L/L cohort (Fig. 4C). Although it induced no lung histologic abnormalities alone, Map2k4 inactivation increased the frequency and accelerated the growth of incipient lung neoplasias driven by mutant Kras (Fig. 4D and E). Relative to lung tumors in KrasG12D mice, tumors in KrasG12D; MKK4L/L mice were more numerous (P = 0.030; two-tailed Student's t test) and involved a greater surface area per lung (P = 0.019; two-tailed Student's t test) (Fig. 4F and G). None of the tumor-bearing mice had distant metastatic disease. Microscopic analysis of lung tissues demonstrated a variety of histopathologic changes in both cohorts that were more numerous in the KrasG12D; MKK4L/L cohort (P = 0.048; two-tailed Fisher's exact test) (Table 2). In fact, 5 (31%) of 16 KrasG12D mice had no lung histologic abnormalities at all, whereas all of the KrasG12D; MKK4L/L mice had abnormalities. We conclude that Map2k4 inactivation potentiated the development of incipient lung neoplasias initiated by oncogenic Kras, accelerated the growth of these tumors, and caused a more rapid demise of tumor-bearing mice.

An external file that holds a picture, illustration, etc.
Object name is zmb9991092650004.jpg

MKK4 inactivation enhances lung tumorigenesis driven by mutant Kras. (A) Map2k4 inactivation shortened mouse survival times. The graph presents the results of Kaplan-Meier survival analysis of mice with the indicated genotypes. P values for comparison of the results obtained with KrasG12D/+ and KrasG12D/+; MKK4L/L groups were calculated with the log rank (Mantel-Cox) test. (B) Recombination of KrasLSL allele specifically in lung tumors. KrasWT, wild-type Kras allele (∼622 bp); KrasLSL, germ line conditional Kras allele (500 bp); KrasG12D, recombined conditional Kras allele (∼650 bp). The PCR amplification strategy employed has been previously reported (20). (C) PCR confirmation of floxed allele recombination. A diagrammatic illustration of the PCR amplification strategy (top) used to detect the wild-type Map2k4 allele (MKK4+) and germ line allele (MKK4L) and recombined (MKK4) floxed Map2k4 allele by the use of forward (F) and reverse (R) primers is shown. The left gel demonstrates recombination of MKK4L specifically in lung tissues of KrasG12D/+; MKK4L/L mice. The right gel demonstrates recombination of the MKK4L allele specifically in lung tumors. The PCR amplification strategy used has been previously reported (46). (D) Histologic sections of lung tissues. Lung tissue sections from each cohort were stained with hematoxylin and eosin and photographed (main images). Magnifications of adenomas are shown in the insets. (E) Examples of lung tissues obtained at necropsy. Lungs and heart (H) were removed at necropsy. Photographs illustrate representative lung tissues that are wild type (WT) or have inactivated Map2k4 (MKK4L/L) or express KrasG12D or both. Lung tumors (arrows) are indicated. (F and G) Map2k4 inactivation promoted lung tumor growth. The numbers of lung tumors visible on the pleural surface were determined at necropsy and are expressed as the mean value per lung per cohort (F). Areas of tumor surface in lung tissue sections were measured using ImageJ (http://rsbweb.nih.gov/ij) and are expressed as the ratio of the sum of all tumor surface areas to lung surface area per cohort (G). Mean values, long horizontal lines; standard deviations, short horizontal lines.

Table 2.

Lung histopathology of KrasG12D and KrasG12D; MKK4L/L mice

HistologyaNo. of mice with indicated condition/total no. of mice
KrasG12DKrasG12D; MKK4L/Lb
Normal5/160/13*
AAH5/166/13
APBP11/1611/13
BAC-like lesion3/163/13
Adenoma9/1613/13*
Adenocarcinoma1/163/13
aAAH, atypical adenomatous hyperplasia; APBP, atypical papillary bronchiolar proliferation; BAC, bronchioloalveolar carcinoma.
bP < 0.05 (KrasG12D versus KrasG12D; MKK4L/L [two-tailed Fisher's exact test]).

To identify other oncogenic driver mutations in human cancer cell lines that have homozygous MAP2K4 loss, we mined the catalogue of COSMIC (Catalogue of Somatic Mutations in Cancer; http://www.sanger.ac.uk/cosmic) (Table 1). Coexisting mutations identified in the 17 cell lines with MAP2K4 mutations included TP53 (13 cell lines), CDKN2A (6 cell lines), RB1 (4 cell lines), BRAF1 (4 cell lines), SMAD4 (4 cell lines), KRAS (4 cell lines), and PTEN (1 cell line). Thus, MAP2K4 mutations coincided frequently with mutations in TP53 and its upstream activator CDKN2A. In four cell lines, KRAS mutations coincided with mutations in MAP2K4 and TP53. On the basis of these findings, we posited that MAP2K4 functions as a tumor suppressor in lung adenocarcinomas driven by mutant KRAS and TP53 and tested this hypothesis in lung adenocarcinoma cell lines derived from KrasG12D;p53R172HΔG mice, which develop metastatic lung adenocarcinomas owing to expression of mutant Kras and p53 (52). In syngeneic mice, these cell lines have different characteristics with respect to tumorigenicity and metastatic potential despite their uniform expression of mutant Kras and Tp53 (15), suggesting that additional genetic events are required for malignant progression.

In a panel of 13 murine lung adenocarcinoma cell lines, none was deficient in MKK4 expression (data not shown), but this cell line sample size was underpowered given that homozygous MAP2K4 deletions occur in only 3% of TP53-mutant human cancer cell lines (4). Therefore, we undertook a synthetic approach to examine the role of MKK4 in these cell lines. MKK4 (wild type, kinase dead, or constitutively active) was overexpressed in 344SQ (Fig. 5A), a highly invasive and metastatic lung adenocarcinoma cell line derived from KrasLA1;p53R172HΔG mice. Conversely, MKK4 was knocked down in 393P, a lung adenocarcinoma cell line derived from KrasLA1;p53R172HΔG mice that has low invasive activity in vitro and undetectable metastatic potential in syngeneic mice despite the presence of Kras and Tp53 mutations (15). Two 393P clones (393P-shY and 393P-shZ) were generated that were deficient with respect to MKK4 (Fig. 5D). Forced expression of wild-type or constitutively active (ED mutant) MKK4 minimally suppressed 344SQ cell proliferation (Fig. 5B) and caused a 60 to 70% suppression of cell invasion, whereas kinase-dead MKK4 (KR or N234I) did not have these effects (Fig. 5C). Conversely, MKK4-deficient 393P cells proliferated slightly faster than controls did at days 2 and 3 but not at later time points (Fig. 5E and F) and exhibited increased cell migration (3- to 4-fold) (Fig. 5G) and invasion (3- to 5-fold) (Fig. 5H) but no differences in colony formation in soft agar (data not shown). Similarly, MKK4 knockdown in a human lung cancer cell line (H2009) that has mutations in both KRAS (G12A) and TP53 (R273L) increased cell growth (Fig. 5J) and invasion (Fig. 5K).

An external file that holds a picture, illustration, etc.
Object name is zmb9991092650005.jpg

MKK4 depletion enhances tumor cell migration and invasion. (A) Flag-tagged MKK4 (wild-type or mutant) vectors (vec) were stably transfected into 344SQ cells. (B) MKK4-transfected 344SQ cells were counted every day for 5 days under normal culture conditions. Wild-type MKK4 had no effect, but a constitutively active (ED) mutant slightly inhibited cellular growth. Results represent the mean values (±SD) obtained with quadruplicate wells. *, P < 0.05; **, P < 0.01 (compared with vector). (C) RAW 264.7- and MKK4-transfected 344SQ cells were seeded into the lower and upper chambers, respectively, of Matrigel-coated chambers for invasion assays. Invasive cells were photographed (images) and counted (bar graph). Results represent the mean values (±SD) obtained with triplicate wells. (D) 393P cells were transfected with each of four distinct MKK4 shRNAs (A, B, Y, or Z) and subjected to Western blotting, which was quantified densitometrically relative to the results obtained with a scrambled control, which was set at a value of 1. Two clones with the most prominent MKK4 depletion (shY and shZ) were selected for further experiments. (E and F) 393P-scr, shY, and shZ cells were seeded into 12-well plates, and cell numbers were counted every day for 5 days under normal culture conditions (E) or after 2 days under serum-free conditions (0%) (F), revealing a slight increase in cellular proliferation under normal conditions at days 2 and 3 but not at later time points. *, P < 0.05 compared with scr results. (G and H) MKK4 depletion enhanced 393P cell migration and invasion. RAW 264.7 and 393P cells were seeded into the lower and upper chambers, respectively, of Transwell plates for migration assays (G) and Matrigel-coated chambers for invasion assays (H) in the presence of mitomycin C (1 μg/ml). Cells were photographed (images) and counted (bar graphs). Results represent mean values (±SD) obtained with triplicate wells. (I) H2009 human lung cancer cells were stably transfected with each of five distinct human MKK4 shRNAs (sh1 to sh5). The two clones with the most prominent MKK4 depletion (sh3 and sh4) were selected for further experiments. (J) H2009 transfectants (scr, sh3, and sh4) were counted every day for 5 days under normal culture conditions, which revealed increased proliferation in clone sh3. Results represent mean values (±SD) obtained with triplicate wells. *, P < 0.05; **, P < 0.01 (compared with scr results). (K) H2009 transfectants were seeded into Matrigel-coated chambers for invasion assays. Invasive cells were photographed (images) and counted (bar graph). Results represent mean values (±SD) obtained with triplicate wells.

To examine whether tumor cell invasion is dependent on the presence of MKK4 kinase, 393P-shZ cells were stably transfected with wild-type Map2k4 or mutants with defined kinase activity (reduced, unchanged, or increased [Fig. 2D]) (Fig. 6A), and the transfectants were subjected to invasion assays. Adding back wild-type MKK4 suppressed 393P-shZ cell invasion (Fig. 6B and C), and the mutant MKK4 expression constructs reduced invasion in direct proportion to the kinase activities of the mutants (R = −0.070; P = 0.013 by one-tailed Spearman's rank correlation) (Fig. 6D), suggesting that MKK4 kinase activity was required. Similarly, forced expression of wild-type MKK4 in a human pancreatic cancer cell line with homozygous MAP2K4 loss (BxPC3) (Fig. 6E) had no detectable effect on cell proliferation in monolayer culture but suppressed invasion by 40% (Fig. 6F and G). However, in contrast to 393P cells, BxPC3 cells transfected with kinase-dead MKK4 (S251N) exhibited small but statistically significant increases in cell proliferation and invasion (Fig. 6F and G), suggesting a dominant-negative effect of kinase-dead MAP2K4 that is specific with respect to cell type.

An external file that holds a picture, illustration, etc.
Object name is zmb9991092650006.jpg

MKK4 reintroduction abrogates invasive activity. (A) Flag-tagged MKK4 (wild-type or mutant) vectors were stably transfected into 393P-shZ cells. (B and C) 393P-shZ-MKK4 cells and RAW 264.7 cells were seeded into the upper and lower chambers, respectively, of Matrigel-coated wells, and then invasive cells were photographed (B) and counted (C). Results represent mean values (±SD) obtained with triplicate wells. *, P < 0.01 compared to empty vector. (D) Graph denoting the relationship between MKK4 kinase activity and invasion activity. According to linear regression analysis, most MKK4 mutants were within 95% confidence intervals (dotted curves; R2 = 0.6737). (E) Flag-tagged MKK4 (wild-type or S251N mutant) vectors were stably transfected into BxPC3 cells. (F) Cell proliferation of MKK4-transfected BxPC3 cells was measured by 3-(4,5-dimethyl-2-thiazolyl)-2,5-diphenyl-2H-tetrazolium bromide (MTT) assays daily for 4 days. Wild-type MKK4 had no effect, whereas the S251N mutant increased cellular growth. Results represent mean values (±SD) obtained with quadruplicate wells. *, P < 0.01 compared with vector. (G) Wild-type MKK4 (WT) suppressed BxPC3 cell invasion in Matrigel-coated chambers, whereas the S251N mutant slightly increased invasion. Invasive cells were photographed and counted (bar graph). Results are expressed as mean values (±SD) obtained with triplicate wells. (H) Inhibitors of JNK (SP) or p38 (SB) blocked the phosphorylation of downstream targets but did not increase 393P cell invasiveactivity. 393P cells were stimulated with sorbitol (0.5 M for 3 h) to induce MAP kinase activation in the presence of SP600125 (SP; 10 μM), SB202190 (SB; 10 μM), or vehicle (dimethyl sulfoxide [DMSO]) and subjected to Western blot analysis (gels in left panel) to detect phosphorylated (p) or total proteins. 393P cells were seeded into Matrigel-coated chambers for invasion assays and treated with inhibitors or vehicle (DMSO). Invasive cells were photographed (images) and counted (bar graph). Results represent mean values (±SD) obtained with triplicate wells.

The tumorigenicity and metastatic activity of MKK4-deficient 393P cells were examined by subcutaneous injection into syngeneic, immunocompetent mice. Eight weeks following injection, necropsies were performed to determine the number of mice in each cohort bearing primary tumors and lung metastases. Although the levels of tumor incidence in mice injected subcutaneously with 393P-scr (9 of 15 mice), 393P-shY (8 of 9), and 393P-shZ (11 of 15) were similar, lung metastases in mice bearing 393P-shY tumors (5 of 8 mice) were significantly more frequent (P = 0.013; one-sided chi-square test) than they were in mice bearing 393P-scr tumors (1 of 9 mice), and a similar trend (P = 0.021) was observed in mice bearing 393P-shZ tumors (6 of 11 mice) (Table 3). 393P-shZ tumors reconstituted with wild-type MKK4 generated lung metastases in fewer mice (1 of 5) than did nonreconstituted tumors (6 of 8; P = 0.027) or kinase-dead S251N-reconstituted tumors (7 of 10, P = 0.034), and a similar trend was observed with kinase-dead N234I-reconstituted tumors (5 of 10; P = 0.13) (Table 3). Collectively, these findings suggest that MKK4 kinase activity suppressed the invasion and metastasis of Kras-Tp53-mutant lung adenocarcinoma cells.

Table 3.

Effect of MKK4 depletion on 393P cell tumorigenicity and metastasisa

Cell categoryNo. of mice with palpable tumors/total no. of injected miceNo. of mice with visible lung metastases/total no. of mice bearing primary tumors
393P-scr9/151/9
393P-shY8/95/8b
393P-shZ11/156/11c
393P-shZ + vector8/106/8
393P-shZ + WT5/101/5d
393P-shZ + S251N10/107/10e
393P-shZ + N234I10/105/10f
aMice were injected subcutaneously with tumor cells, sacrificed at 8 weeks, and subjected to necropsy to isolate primary and metastatic tumors.
bP = 0.0134 (scr versus shY [one-sided chi-square test]).
cP = 0.0214 (scr versus shZ).
dP = 0.0265 (shZ + vec versus shZ + WT).
eP = 0.0336 (shZ + WT versus shZ + S251N).
fP = 0.1318 (shZ + WT versus shZ + N234I).

We next sought to identify MKK4 substrates and transcriptional targets that regulate tumor cell invasion. Surprisingly, basal and sorbitol-induced phosphorylation of the two canonical substrates of MKK4 (JNK and p38) was only minimally reduced in MKK4-deficient cells (data not shown), a result that is potentially related to compensatory effects of other JNK and p38 kinases expressed in these cells (i.e., MKK7). Treatment of 393P-shZ cells with inhibitors of JNK (SP600125) or p38 (SB202190) decreased substrate phosphorylation but did not significantly change invasion (Fig. 6H). Collectively, these findings suggest that invasion was dependent on the presence of MKK4 kinase but was mediated through novel substrates. We next turned our attention to MKK4 transcriptional targets. Triplicate RNA samples from 393P-shZ and 393P-scr cells were transcriptionally profiled. Comparison of the resulting profiles demonstrated that, relative to controls, 393P-shZ cells differentially expressed 449 genes (319 up- and 130 downregulated; P < 0.01 [>1.5-fold change]) (Fig. 7 A), and 33 were confirmed by quantitative RT-PCR analysis (Fig. 7B). One of the genes with the most prominently increased (5.0-fold) expression was PPARγ, a transcription factor that functions as a nuclear receptor for fatty acids and eicosanoids and can reportedly promote or inhibit tumorigenesis (31, 32, 37, 41, 42). Pparg encodes two protein isoforms (γ1 and γ2), the mRNA levels of which increased 1.5- and 6.3-fold, respectively, following MKK4 knockdown in 393P cells (Fig. 7C), and a substantial increase in PPARγ2 protein induced by MKK4 depletion was confirmed by Western blot analysis (Fig. 7D). Actinomycin D treatment of 393P-shZ cells abrogated the increase in Pparg2 (Fig. 7E), and chromatin immunoprecipitation assays revealed increased binding of RNA polymerase II to the Pparg2 gene promoter in 393P-shZ cells (Fig. 7F), suggesting that MKK4 regulated Pparg2 expression through transcriptional mechanisms. Relative to their levels in 393P-scr cells, both spliced and nonspliced Pparg2 mRNA levels increased in 393P-shZ cells (Fig. 7G), arguing against posttranscriptional regulation of Pparg2 through mRNA splicing.

An external file that holds a picture, illustration, etc.
Object name is zmb9991092650007.jpg

Expression profiling of 393P-shZ and 393P-scr cells. (A) Expression data matrix (“heat map”) of 574 RNA transcripts (i.e., Affymetrix probe sets, representing 449 unique genes) differentially expressed between 393P-shZ (MKK4-shRNA) and 393P-scr (control). Each row represents a gene, each column a profiled sample. The relative abundance of each gene in each group is represented using a yellow-blue color scale (blue represents low expression; yellow represents high expression). To the right of the data matrix, red, green, blue, and yellow bars denote the corresponding annotations of genes in the matrix presented using selected Gene Ontology (GO) terms. Specific genes of particular interest are indicated. (B) Quantitative RT-PCR (Q-PCR) confirmation of genes that were differentially expressed in the Affymetrix profiles. Results were normalized on the basis of ribosomal protein L32 levels. Mean values (±SD) were computed from triplicate samples and are expressed as the ratio of 393P-shZ to 393P-scr. *, P < 0.01 compared with scrambled shRNA transfectant. (C) Q-PCR analyses of PPARγ1 and PPARγ2 mRNAs in 393P-scr and 393P-shZ cells. Data represent mean values (±SD) obtained with triplicate samples. (D) Western blot of PPARγ2 in 393P-scr and 393P-shZ cells. Actin was included as a loading control. (E) Q-PCR analysis of PPARγ2 mRNA in 393P-scr and 393P-shZ cells treated with actinomycin D (Act.D; 2 μg/ml) for the indicated times. Results were normalized on the basis of ribosomal protein L32 mRNA values and are expressed as mean values (±SD) obtained with triplicate wells. (F) RNA polymerase II chromatin immunoprecipitation assays on 393P-scr and 393P-shY cells. Chromatin-RNA polymerase II complexes were immunoprecipitated, and then promoter regions of Pparg2 and Gapdh were amplified by quantitative PCR. Results were normalized on the basis of Gapdh values and are expressed as mean values (±SD) obtained with triplicate samples. (G) Q-PCR analysis of nonspliced (pre-PPARγ2) and spliced (PPARγ2) mRNA performed using PPARγ2-specific primers to amplify exon-intron and exon-exon products, respectively. Results were normalized on the basis of L32 mRNA values and are expressed as mean values (±SD) obtained with triplicate samples. (H) Q-PCR analyses of PPARγ2 mRNA in 393P-shZ cells and BxPC3 cells stably transfected with wild type (WT) or kinase-dead mutant (S251N) MKK4.

JNK phosphorylates PPARγ and inhibits PPARγ transcriptional activity (1). Therefore, we examined whether the increase in Pparg2 expression was dependent on the presence of MKK4 kinase and required JNK. Following reexpression of (wild-type or S251N) MKK4 in 393P-shZ cells, Pparg2 expression decreased in wild-type but not S251N transfectants (Fig. 7H). Similarly, Pparg2 levels decreased following reintroduction of wild-type but not kinase-dead MKK4 into human BxPC3 pancreatic cancer cells that had a homozygous MAP2K4 deletion (Fig. 7H). However, treatment of 393P cells with SP600125 or SB202190 did not increase the levels of expression of Pparg2 or Pgc-1α, a PPARγ target gene (19) (data not shown). Thus, MKK4 suppressed PPARγ expression through a kinase-dependent mechanism involving novel MKK4 substrates.

PPARγ has numerous transcriptional targets that mediate diverse biological functions. To examine the extent to which MKK4 deficiency activated a PPARγ-dependent transcriptional program, we mined a database that lists genes regulated by treatment with a PPARγ agonist in human lung cancer cells (Gene Expression Omnibus GSE7035) and found that, of the 1,343 genes that exhibited increased expression, 35 overlapped with genes with increased expression in 393P-shZ cells (Table 4), which exceeded the expected overlap due to chance (23 genes; P < 0.01 [Fisher's exact test]). Furthermore, MKK4 knockdown in 393P cells increased the levels of Pgc-1α (Fig. 8 A). Thus, MKK4 depletion activated a PPARγ-dependent transcriptional program.

An external file that holds a picture, illustration, etc.
Object name is zmb9991092650008.jpg

MKK4 suppresses tumor cell invasion through PPARγ. (A) Quantitative RT-PCR (Q-PCR) analysis of PGC-1α mRNA in 393P-scr and 393P-shZ cells. Data represent mean values (±SD) obtained with triplicate samples. (B) Q-PCR analyses of PPARγ2 (left) and PGC-1α (middle) and reporter assays (right) in cells transiently transfected with a PPAR-responsive element (PPRE) luciferase reporter plasmid. Results were normalized on the basis of renilla luciferase activity values and are expressed as mean values (±SD) obtained with triplicate samples. (C and D) Q-PCR analyses of PPARγ2 mRNA in 344SQ cells (C) or 393P-shZ cells (D) stably transfected with shRNAs against PPARγ (shP-A or shP-B) or scrambled control (scr). Results were normalized on the basis of L32 mRNA values and are expressed as mean values (±SD) obtained with triplicate samples. (E and F) Invasion assays on 344SQ cells (E) or 393P-shZ cells (F) stably transfected with PPARγ or scrambled shRNA. Invasive cells were photographed (images) and quantified (bar graphs). Results are expressed as mean values (±SD) obtained with triplicate wells. (G) Invasion assays on 393P-shZ cells treated with a PPARγ-selective antagonist (T0070907) or vehicle (DMSO). Invaded cells were photographed (images) and quantified (bar graph). Vehicle-treated cells are indicated as a “0” dose. Cellular toxicity after 24 h of treatment of 393P-shZ cells with T0070907 was examined by MTT assay (line graph). Vehicle-treated cells are indicated as a “0” dose. Results are expressed as mean values (±SD) obtained with triplicate wells. (H) Invasion assays on 393P cells stably transfected with PPARγ cDNA or empty (vec) expression vectors. Invasive cells were photographed (images) and quantified (left bar graph). Results are expressed as mean values (±SD) obtained with triplicate wells. Q-PCR analysis confirmed exogenous PPARγ mRNA expression (right bar graph). Results were normalized on the basis of L32 mRNA values and are expressed as mean values (±SD) obtained with triplicate samples. (I) Q-PCR analysis of human pancreatic cancer cells (BxPC3) stably transfected with one of four distinct shRNAs against PPARγ (shP1 to shP4) or scrambled control (scr). Results were normalized on the basis of L32 mRNA values and are expressed as mean values (±SD) obtained with triplicate samples. (J) Invasion assays on BxPC3 cells stably transfected with PPARγ shRNAs (shP3 and shP4) or scrambled (scr) shRNA. Invasive cells were photographed (images) and quantified (bar graphs). Results are expressed as mean values (±SD) obtained with triplicate wells.

Table 4.

Genes induced both by MKK4 knockdown and by the PPARγ activator rosiglitazone

Entrez no. (human)DesignationDescription
1363CPECarboxypeptidase E
1435CSF1Colony-stimulating factor 1 (macrophage)
1462VCANVersican
1896EDAEctodysplasin A
2260FGFR1Fibroblast growth factor receptor 1
22905EPN2Epsin 2
23321TRIM2Tripartite motif-containing 2
2335FN1Fibronectin 1
2444FRKFyn-related kinase
2555GABRA2Gamma-aminobutyric acid (GABA) A receptor, alpha 2
27330RPS6KA6Ribosomal protein S6 kinase, 90 kDa, polypeptide 6
2770GNAI1Guanine nucleotide binding protein (G protein), alpha inhibiting activity polypeptide 1
2892GRIA3Glutamate receptor, ionotrophic, AMPA 3
30001ERO1LERO1-like (Saccharomyces cerevisiae)
3006HIST1H1CHistone cluster 1, H1c
3172HNF4AHepatocyte nuclear factor 4, alpha
3572IL6STInterleukin 6 signal transducer (gp130, oncostatin M receptor)
4070TACSTD2Tumor-associated calcium signal transducer 2
4254KITLGKIT ligand
4613MYCNv-myc myelocytomatosis virus-related oncogene, neuroblastoma derived (avian)
5087PBX1Pre-B-cell leukemia homeobox 1
51196PLCE1Phospholipase C, epsilon 1
51299NRN1Neuritin 1
51635DHRS7Dehydrogenase/reductase (SDR family) member 7
5569PKIAProtein kinase (cyclic AMP dependent, catalytic) inhibitor alpha
5950RBP4Retinol binding protein 4, plasma
64116SLC39A8Solute carrier family 39 (zinc transporter), member 8
6595SMARCA2SWI/SNF-related, matrix-associated, actin-dependent regulator of chromatin, subfamily a, member 2
6664SOX11SRY (sex-determining region Y)-box 11
7026NR2F2Nuclear receptor subfamily 2, group F, member 2
7404UTYUbiquitously transcribed tetratricopeptide repeat gene, Y-linked
8829NRP1Neuropilin 1
8878SQSTM1Sequestosome 1
950SCARB2Scavenger receptor class B, member 2
952CD38CD38 molecule

Relative to 393P cells, 344SQ cells are more invasive (15) and exhibit higher Pparg2 and Pgc-1α expression and higher PPAR transcriptional activity in reporter assays (Fig. 8B). We posited that PPARγ promotes invasion and tested this hypothesis by introducing one of two distinct murine PPARγ shRNA constructs (shP-A or shP-B) into 393P-shZ cells and 344SQ cells, which resulted in basal PPARγ levels that were decreased by 48% in 344SQ cells (Fig. 8C) and abrogation of the MKK4 shRNA-induced increase in Pparg2 in 393P-shZ cells (Fig. 8D). As a comparison to the effects of PPARγ knockdown, shRNAs against other genes that exhibited increased expression with MKK4 depletion (biglycan, vascular endothelial growth factor c, and special AT-rich sequence-binding protein 1) were introduced into 393P-shZ cells. PPARγ knockdown decreased the invasion of 344SQ cells (Fig. 8E) and 393P-shZ cells (Fig. 8F), whereas the other shRNAs did not abrogate 393P-shZ cell invasion (data not shown). Treatment with T0070907, a PPARγ antagonist (29), blocked 393P-shZ cell invasion without affecting cell viability (Fig. 8G). Conversely, forced expression of PPARγ enhanced 393P cell invasion (Fig. 8H). To determine whether PPARγ promotes invasion in human cancer cells that have homozygous MAP2K4 loss, BxPC3 cells were stably transfected with one of four distinct human PPARγ shRNAs (shP1 to shP4) (Fig. 8I). Relative to scrambled shRNA controls, PPARγ-depleted BxPC3 cells (shP3 and shP4) exhibited reduced invasion (Fig. 8J).

DISCUSSION

Numerous studies have demonstrated MAP2K4 somatic mutations in cancer cells (18, 26, 34, 45, 47, 50), but the spectrum of MAP2K4 somatic mutations has not been fully evaluated in MKK4 biochemical assays to delineate loss or gain of function, and downstream effectors of MKK4 that mediate cellular transformation have not been adequately explored. Here we systematically characterized the biochemical properties of a large panel of cancer-associated MAP2K4 mutations, examined the consequences of Map2k4 inactivation in genetically engineered mouse models of human lung cancer driven by mutant Kras alone or in combination with mutant Tp53, and profiled transcriptional changes induced by MKK4 depletion in Kras/Tp53-mutant lung adenocarcinoma cells. Findings from these studies revealed that Map2k4 functions as a tumor suppressor in lung adenocarcinoma and inhibits tumor cell invasion by decreasing PPARγ2 expression.

Map2k4 inactivation increased the multiplicity and accelerated the growth of incipient lung neoplasias and promoted the invasion and metastasis of Kras/Tp53-mutant lung adenocarcinoma cells, leading us to conclude that MKK4 exerts tumor suppressor activity at both early and late stages of lung tumorigenesis. These findings add to a growing list of somatic mutations that cooperate with mutant Kras to promote lung adenocarcinoma development in mice (19, 22, 28, 33, 52). However, Map2k4 inactivation mitigates carcinogen-induced skin tumors in the same mouse strain (MKK4L/L) used in this study (12), which may reflect tissue-specific roles of MKK4, differential signaling by mutant Ras family members (Kras and Hras in lung and skin carcinomas, respectively), or other factors.

Biochemical analysis of mutant MKK4 revealed that the MAP2K4 somatic mutations inactivated MKK4 through diverse mechanisms. The missense mutations targeted motifs required for kinase activation, including the ATP-binding pocket (N234I) and activation loop (S251N), and nonsense and frameshift mutations resulted in rapid degradation of MKK4 through the ubiquitin-proteasome complex, which was presumably engaged due to misfolding of the C-terminally truncated proteins, as has been reported for other mutant tumor suppressors (14, 43, 48). Although the biological relevance of degrading an inactive kinase is not immediately obvious, one potential consequence is the removal of proteins that would otherwise have a dominant-negative effect. The C-terminally truncated mutants lack a domain required for interactions with an upstream kinase, MEKK1, creating kinase-dead proteins that can compete with wild-type MKK4 for binding to MKK4 substrates and other proteins within the larger mitogen-activated protein (MAP) kinase complex. Given that these mutants are unstable, such a dominant-negative effect is unlikely.

Several findings reported here suggest that the tumor suppressor activity of MKK4 is kinase dependent. MAP2K4 somatic mutations cluster within the kinase domain, and most of the mutants had reduced kinase activity. Furthermore, the ability of MAP2K4 mutants to attenuate invasion correlated positively with their residual kinase activities. On the basis of these findings, we expected that canonical MKK4 substrates would function as key mediators of tumor cell invasion and metastasis. However, basal and sorbitol-induced JNK and p38 phosphorylation did not decrease in MKK4-depleted cells, and chemical antagonists of JNK (SP600125) and p38 (SB202190) did not enhance 393P cell invasion. These findings were surprising, given that MKKs are viewed as kinases dedicated primarily to the regulation of MAP kinases (9). The only noncanonical MKK4 substrate reported thus far is RXR, which MKK4 phosphorylates on Tyr249 in the ligand-binding domain, thereby inhibiting RXR transactivation (30). However, RXR knockdown did not decrease PPARγ expression levels in 393P-shZ cells (data not shown), arguing against RXR as the requisite MKK4 substrate. These findings warrant efforts to discover additional substrates through which MKK4 mediates tumor suppression.

To gain insight into candidate mediators of MKK4, we performed gene expression profiling studies and identified 449 genes from a broad range of Gene Ontology categories that increased or decreased in expression in MKK4-depleted tumor cells. Some of these genes play critical roles in the extracellular matrix (Bgn and Vcam1), lymphangiogenesis (Vegfc), inflammation and response to inflammatory cytokines (Il13ra1, Csf1, Ccr9, Cxcl11, and Cxcl7), transcriptional regulation (Satb1, Smarca2, and Id2), cell polarity (Inadl, Crb3, and Pard6b), and epithelium-to-mesenchyme transition (Ncam1, Vim, and Snail1). Although many of these genes could have contributed to the enhanced invasive and metastatic activity, we chose to investigate the role of PPARγ, owing to the degree to which its expression increased in MKK4-deficient cells and its reported role in other tumor models (7, 27, 37, 41). We found that PPARγ promoted tumor cell invasion and increased in abundance through an MKK4 kinase-dependent pathway that did not involve canonical MKK4 substrates. Collectively, these findings reinforce a growing body of evidence indicating that MAP2K4 somatic mutations contribute to tumor progression in a variety of epithelial tumor types (18, 4951) and suggest that MKK4 exerts its tumor suppressor properties through noncanonical substrates that have diverse transcriptional targets.

ACKNOWLEDGMENTS

We thank Dror Berel for technical assistance.

Y.-H.A., Y.Y., D.L.G., C.J.C., F.Y., I.I.W., W.L., N.T., C.A.A., J.R., E.J.G., and C.T. performed these experiments, and J.M.K. conceived and supervised the work.

This work was supported by NIH grant R01 CA105155.

Footnotes

Published ahead of print on 6 September 2011.

REFERENCES

1. Adams M., Reginato M. J., Shao D., Lazar M. A., Chatterjee V. K. 1997. Transcriptional activation by peroxisome proliferator-activated receptor gamma is inhibited by phosphorylation at a consensus mitogen-activated protein kinase site. J. Biol. Chem. 272:5128–5132 [PubMed] [Google Scholar]
2. Adzhubei I. A., et al. 2010. A method and server for predicting damaging missense mutations. Nat. Methods 7:248–249 [PMC free article] [PubMed] [Google Scholar]
3. Ahn Y. H., Kurie J. M. 2009. MKK4/SEK1 is negatively regulated through a feedback loop involving the E3 ubiquitin ligase itch. J. Biol. Chem. 284:29399–29404 [PMC free article] [PubMed] [Google Scholar]
4. Bignell G. R., et al. 2010. Signatures of mutation and selection in the cancer genome. Nature 463:893–898 [PMC free article] [PubMed] [Google Scholar]
5. Carretero J., et al. 2010. Integrative genomic and proteomic analyses identify targets for Lkb1-deficient metastatic lung tumors. Cancer Cell 17:547–559 [PMC free article] [PubMed] [Google Scholar]
6. Chang L., Karin M. 2001. Mammalian MAP kinase signalling cascades. Nature 410:37–40 [PubMed] [Google Scholar]
7. Chen L., et al. 2006. Peroxisome proliferator-activated receptor gamma promotes epithelial to mesenchymal transformation by Rho GTPase-dependent activation of ERK1/2. J. Biol. Chem. 281:24575–24587 [PubMed] [Google Scholar]
8. Choi K., Creighton C. J., Stivers D., Fujimoto N., Kurie J. M. 2007. Transcriptional profiling of non-small cell lung cancer cells with activating EGFR somatic mutations. PLoS One 2:e1226. [PMC free article] [PubMed] [Google Scholar]
9. Cobb M. H., Goldsmith E. J. 1995. How MAP kinases are regulated. J. Biol. Chem. 270:14843–14846 [PubMed] [Google Scholar]
10. Creighton C. J., et al. 2008. Insulin-like growth factor-I activates gene transcription programs strongly associated with poor breast cancer prognosis. J. Clin. Oncol. 26:4078–4085 [PMC free article] [PubMed] [Google Scholar]
11. Davies H., et al. 2005. Somatic mutations of the protein kinase gene family in human lung cancer. Cancer Res. 65:7591–7595 [PubMed] [Google Scholar]
12. Finegan K. G., Tournier C. 2010. The mitogen-activated protein kinase kinase 4 has a pro-oncogenic role in skin cancer. Cancer Res. 70:5797–5806 [PMC free article] [PubMed] [Google Scholar]
13. Fisher G. H., et al. 2001. Induction and apoptotic regression of lung adenocarcinomas by regulation of a K-Ras transgene in the presence and absence of tumor suppressor genes. Genes Dev. 15:3249–3262 [PMC free article] [PubMed] [Google Scholar]
14. Gautreau A., et al. 2002. Mutant products of the NF2 tumor suppressor gene are degraded by the ubiquitin-proteasome pathway. J. Biol. Chem. 277:31279–31282 [PubMed] [Google Scholar]
15. Gibbons D. L., et al. 2009. Contextual extracellular cues promote tumor cell EMT and metastasis by regulating miR-200 family expression. Genes Dev. 23:2140–2151 [PMC free article] [PubMed] [Google Scholar]
16. Greenman C., et al. 2007. Patterns of somatic mutation in human cancer genomes. Nature 446:153–158 [PMC free article] [PubMed] [Google Scholar]
17. Guerra C., et al. 2003. Tumor induction by an endogenous K-ras oncogene is highly dependent on cellular context. Cancer Cell 4:111–120 [PubMed] [Google Scholar]
18. Hickson J. A., et al. 2006. The p38 kinases MKK4 and MKK6 suppress metastatic colonization in human ovarian carcinoma. Cancer Res. 66:2264–2270 [PubMed] [Google Scholar]
19. Iwanaga K., et al. 2008. Pten inactivation accelerates oncogenic K-ras-initiated tumorigenesis in a mouse model of lung cancer. Cancer Res. 68:1119–1127 [PMC free article] [PubMed] [Google Scholar]
20. Jackson E. L., et al. 2001. Analysis of lung tumor initiation and progression using conditional expression of oncogenic K-ras. Genes Dev. 15:3243–3248 [PMC free article] [PubMed] [Google Scholar]
21. Ji H., et al. 2006. K-ras activation generates an inflammatory response in lung tumors. Oncogene 25:2105–2112 [PubMed] [Google Scholar]
22. Ji H., et al. 2007. LKB1 modulates lung cancer differentiation and metastasis. Nature 448:807–810 [PubMed] [Google Scholar]
23. Johnson L., et al. 2001. Somatic activation of the K-ras oncogene causes early onset lung cancer in mice. Nature 410:1111–1116 [PubMed] [Google Scholar]
24. Kaminker J. S., et al. 2007. Distinguishing cancer-associated missense mutations from common polymorphisms. Cancer Res. 67:465–473 [PubMed] [Google Scholar]
25. Kennedy N. J., et al. 2003. Suppression of Ras-stimulated transformation by the JNK signal transduction pathway. Genes Dev. 17:629–637 [PMC free article] [PubMed] [Google Scholar]
26. Khatlani T. S., et al. 2007. c-Jun N-terminal kinase is activated in non-small-cell lung cancer and promotes neoplastic transformation in human bronchial epithelial cells. Oncogene 26:2658–2666 [PubMed] [Google Scholar]
27. Kim K. R., et al. 2007. A peroxisome proliferator-activated receptor gamma antagonist induces vimentin cleavage and inhibits invasion in high-grade hepatocellular carcinoma. Oncol. Rep. 18:825–832 [PubMed] [Google Scholar]
28. Kim W. Y., et al. 2009. HIF2alpha cooperates with RAS to promote lung tumorigenesis in mice. J. Clin. Invest. 119:2160–2170 [PMC free article] [PubMed] [Google Scholar]
29. Lee G., et al. 2002. T0070907, a selective ligand for peroxisome proliferator-activated receptor gamma, functions as an antagonist of biochemical and cellular activities. J. Biol. Chem. 277:19649–19657 [PubMed] [Google Scholar]
30. Lee H. Y., et al. 2000. Stress pathway activation induces phosphorylation of retinoid X receptor. J. Biol. Chem. 275:32193–32199 [PubMed] [Google Scholar]
31. Lefebvre A. M., et al. 1998. Activation of the peroxisome proliferator-activated receptor gamma promotes the development of colon tumors in C57BL/6J-APCMin/+ mice. Nat. Med. 4:1053–1057 [PubMed] [Google Scholar]
32. Lehrke M., Lazar M. A. 2005. The many faces of PPARgamma. Cell 123:993–999 [PubMed] [Google Scholar]
33. Liang M. C., et al. 2010. TSC1 loss synergizes with KRAS activation in lung cancer development in the mouse and confers rapamycin sensitivity. Oncogene 29:1588–1597 [PMC free article] [PubMed] [Google Scholar]
34. Lotan T., et al. 2008. c-Jun NH2-terminal kinase activating kinase 1/mitogen-activated protein kinase kinase 4-mediated inhibition of SKOV3ip. 1 ovarian cancer metastasis involves growth arrest and p21 up-regulation. Cancer Res. 68:2166–2175 [PMC free article] [PubMed] [Google Scholar]
35. Matsumoto T., et al. 2010. Crystal structures of MKK4 kinase domain reveal that substrate peptide binds to an allosteric site and induces an auto-inhibition state. Biochem. Biophys. Res. Commun. 400:369–373 [PubMed] [Google Scholar]
36. Meza R., Nunez-Valdez M. E., Sanchez J., Bravo A. 1996. Isolation of Cry1Ab protein mutants of Bacillus thuringiensis by a highly efficient PCR site-directed mutagenesis system. FEMS Microbiol. Lett. 145:333–339 [PubMed] [Google Scholar]
37. Nakajima A., et al. 2008. Inhibition of peroxisome proliferator-activated receptor gamma activity suppresses pancreatic cancer cell motility. Cancer Sci. 99:1892–1900 [PMC free article] [PubMed] [Google Scholar]
38. Ng P. C., Henikoff S. 2003. SIFT: predicting amino acid changes that affect protein function. Nucleic Acids Res. 31:3812–3814 [PMC free article] [PubMed] [Google Scholar]
39. Parmigiani G., et al. 2009. Design and analysis issues in genome-wide somatic mutation studies of cancer. Genomics 93:17–21 [PMC free article] [PubMed] [Google Scholar]
40. Parsons D. W., et al. 2005. Colorectal cancer: mutations in a signalling pathway. Nature 436:792. [PubMed] [Google Scholar]
41. Saez E., et al. 1998. Activators of the nuclear receptor PPARgamma enhance colon polyp formation. Nat. Med. 4:1058–1061 [PubMed] [Google Scholar]
42. Sarraf P., et al. 1999. Loss-of-function mutations in PPAR gamma associated with human colon cancer. Mol. Cell 3:799–804 [PubMed] [Google Scholar]
43. Shimizu H., et al. 2006. Destabilizing missense mutations in the tumour suppressor protein p53 enhance its ubiquitination in vitro and in vivo. Biochem. J. 397:355–367 [PMC free article] [PubMed] [Google Scholar]
44. Torkamani A., Verkhivker G., Schork N. J. 2009. Cancer driver mutations in protein kinase genes. Cancer Lett. 281:117–127 [PMC free article] [PubMed] [Google Scholar]
45. Wang L., Pan Y., Dai J. L. 2004. Evidence of MKK4 pro-oncogenic activity in breast and pancreatic tumors. Oncogene 23:5978–5985 [PubMed] [Google Scholar]
46. Wang X., et al. 2007. Targeted deletion of the mitogen-activated protein kinase kinase 4 gene in the nervous system causes severe brain developmental defects and premature death. Mol. Cell. Biol. 27:7935–7946 [PMC free article] [PubMed] [Google Scholar]
47. Whitmarsh A. J., Davis R. J. 2007. Role of mitogen-activated protein kinase kinase 4 in cancer. Oncogene 26:3172–3184 [PubMed] [Google Scholar]
48. Xu J., Attisano L. 2000. Mutations in the tumor suppressors Smad2 and Smad4 inactivate transforming growth factor beta signaling by targeting Smads to the ubiquitin-proteasome pathway. Proc. Natl. Acad. Sci. U. S. A. 97:4820–4825 [PMC free article] [PubMed] [Google Scholar]
49. Yamada S. D., et al. 2002. Mitogen-activated protein kinase kinase 4 (MKK4) acts as a metastasis suppressor gene in human ovarian carcinoma. Cancer Res. 62:6717–6723 [PubMed] [Google Scholar]
50. Yeasmin S., et al. 2011. Loss of MKK4 expression in ovarian cancer: a potential role for the epithelial to mesenchymal transition. Int. J. Cancer 128:94–104 [PubMed] [Google Scholar]
51. Yoshida B. A., et al. 1999. Mitogen-activated protein kinase kinase 4/stress-activated protein/Erk kinase 1 (MKK4/SEK1), a prostate cancer metastasis suppressor gene encoded by human chromosome 17. Cancer Res. 59:5483–5487 [PubMed] [Google Scholar]
52. Zheng S., El-Naggar A. K., Kim E. S., Kurie J. M., Lozano G. 2007. A genetic mouse model for metastatic lung cancer with gender differences in survival. Oncogene 26:6896–6904 [PubMed] [Google Scholar]

Articles from Molecular and Cellular Biology are provided here courtesy of Taylor & Francis

-