The utility of the folate receptor (FR) type α, in a broad range of targeted therapies and as a diagnostic serum marker in cancer, is confounded by its variable tumor expression levels. FR-α, its mRNA and its promoter activity were coordinately up-regulated by the glucocorticoid receptor (GR) agonist, dexamethasone. Optimal promoter activation which occurred at <50 nmol/L dexamethasone was inhibited by the GR antagonist, RU486, and was enhanced by coactivators, supporting GR mediation of the dexamethasone effect. The dexamethasone response of the FR-α promoter progressed even after dexamethasone was withdrawn, but this delayed effect required prior de novo protein synthesis indicating an indirect regulation. The dexamethasone effect was mediated by the G/C-rich (Sp1 binding) element in the core P4 promoter and was optimal in the proper initiator context without associated changes in the complement of major Sp family proteins. Histone deacetylase (HDAC) inhibitors potentiated dexamethasone induction of FR-α independent of changes in GR levels. Dexamethasone/HDAC inhibitor treatment did not cause de novo FR-α expression in a variety of receptor-negative cells. In a murine HeLa cell tumor xenograft model, dexamethasone treatment increased both tumor-associated and serum FR-α. The results support the concept of increasing FR-α expression selectively in the receptor-positive tumors by brief treatment with a nontoxic dose of a GR agonist, alone or in combination with a well-tolerated HDAC inhibitor, to increase the efficacy of various FR-α–dependent therapeutic and diagnostic applications. They also offer a new paradigm for cancer diagnosis and combination therapy that includes altering a marker or a target protein expression using general transcription modulators.

In recent years, the glycosyl-phosphatidylinositol–anchored folate receptor (FR) type α has served as a model target for tumor-specific delivery of a broad range of pharmacologic and immunologic experimental therapies, for the following reasons: (a) FR-α is expressed in several cancers such as non–mucinous adenocarcinomas of the ovary and uterus, malignant pleural mesothelioma, testicular choriocarcinoma, ependymal brain tumors, nonfunctioning pituitary adenoma, and variably in breast, colon, and renal carcinoma (reviewed in ref. 1); (b) FR-α expression in proliferating normal tissues (reviewed in ref. 1) is restricted to the luminal surface of certain epithelial cells, where it is inaccessible to the circulation, whereas the receptor expressed in tumors is accessible via the circulation. FR-α–targeted low molecular weight agents that may filter through the glomerulus and bind to the receptor in proximal kidney tubules seem to be transcytosed and reabsorbed, avoiding nephrotoxicity (2); (c) other FR isoforms are either expressed in a nonfunctional manner in mature hematopoietic cells (FR-β; refs. 3, 4) or poorly expressed and constitutively secreted (FR-γ/γ′; refs. 5, 6); and (d) FR-α quantitatively recycles between the cell surface and intracellular compartments (reviewed in ref. 7), effectively internalizing receptor-bound folate/antifolate compounds and folate conjugates (8, 9). Various FR-α–targeted therapeutics (reviewed in refs. 1017) and imaging agents (reviewed in ref. 18) have shown promise in preclinical models and in early clinical trials. These agents include radiopharmaceutical and cytotoxic conjugates of folate including prodrugs, prodrug-activating enzymes, nanoparticles, and liposomal drugs as well as potent novel antifolates that are dependent on FR-α for cellular uptake. The FR-α–targeted immunologic therapies include bifunctional antibodies and antibody-interleukin chimeras, peptide and DNA vaccines, and more innovative agents such as dual-specific T cells and folate-hapten conjugates. A portion of the FR-α expressed on the cell surface is released in a soluble form by the combined action of a membrane-associated protease and glycosyl-phosphatidylinositol–specific phospholipase (1922). Soluble FR-α is low or undetected in normal human sera, and therefore, the protein shed into the circulation is a potential serum marker for FR-α–positive tumors (23).

Even though major subtypes of malignant tissues show consistent patterns of FR-α expression, there is a considerable variability and heterogeneity in the tumor expression levels of the receptor covering a range of almost two orders of magnitude (24, 25). The successful experimental FR-targeted therapies in animal models have used xenografts of human tumor cells (e.g., KB cells) that express the receptor uniformly and at levels closer to the high end of this range, underscoring the importance of developing molecular methods of up-regulating the FR gene selectively in malignant cells. Increased FR-α expression by the tumors may be expected to enhance the efficacy of the receptor-targeted therapies and whole-body imaging, and increase the levels of soluble FR-α for early detection as a diagnostic serum marker.

The FR-α gene has seven exons and six introns with multiple transcripts resulting from the use of alternative promoters as well as alternative splicing involving exons 1 to 4 (26, 27). The FR-α gene contains two promoters, named P1 and P4, located upstream of exons 1 and 4, respectively. Transcripts generated by both promoters encode identical proteins, but the P4 promoter activity seems to be predominant in malignant cells (28) and further, P1 promoter-driven transcripts seem to be translated several-fold less efficiently than the P4 promoter-driven transcript (29). The basal TATA-less P4 promoter activity is initiated by a cluster of three G/C-rich sequences that are noncanonical Sp1 binding sites, each of which contributes to promoter activity (27).

We have previously reported that the FR-α gene is directly and negatively regulated by the estrogen receptor (28). Here, we report that the FR-α gene is indirectly and positively regulated at the transcriptional (P4 promoter) level by the glucocorticoid receptor (GR) agonist, dexamethasone, and that this profound regulation is further potentiated by inhibiting histone deacetylase (HDAC). The selectivity of this regulation for FR-α–positive tissues, the innocuous nature of the modulating agents and the ubiquitous expression of GR present a potential means of greatly improving the effectiveness of all available FR-α–targeted therapeutic and diagnostic strategies by the inclusion of GR modulators. The findings also illustrate the potential utility of general transcription modulators in optimizing the expression of genes encoding marker proteins and drug targets that are selectively expressed in tumor tissues. These considerations provided the impetus for the present study of the nature and mechanism of GR regulation of FR-α in vitro and for examining the effect of GR modulation on tumor and serum levels of the protein in vivo.

Chemicals and reagents. DMEM, RPMI, and penicillin/streptomycin/l-glutamine stock mix were purchased from Life Technologies, Inc. (Carlsbad, CA). Fetal bovine serum (FBS) was purchased from Irvine Scientific (Santa Ana, CA). FuGENE 6 was purchased from Roche Diagnostics (Indianapolis, IN), luciferase assay reagents from Promega (Madison, WI) dexamethasone from Sigma (St. Louis, MO), dexamethasone and placebo pellets from Innovative Research of America (Sarasota, FL), trichostatin A, valproic acid, and cycloheximide from Sigma. Affinity-purified rabbit anti-human Sp1, anti-human Sp3, anti-human Sp4 antibodies, and rabbit polyclonal IgG against GR were purchased from Santa Cruz Biotechnology (Santa Cruz, CA). Mouse anti-α tubulin clone B-5-1-2 antibody was from Sigma. Vent DNA polymerase was purchased from New England Biolabs (Beverly, MA), and custom oligonucleotide primers from Life Technologies. The reagents for real-time reverse transcription-PCR were from Applied Biosystems (Branchburg, NJ).

DNA constructs and expression plasmids. Construct design made use of either natural restriction sites or restriction sites created by the addition of appropriate restriction sites to upstream and downstream PCR primers. The PCR products were first digested at both ends with the appropriate restriction enzymes and cloned into the PGL3-basic plasmid (Promega) or subcloned into the FR-α-promoter construct (−3,394 to +33 nt, relative to the transcription initiation site at +1 nt) in the PGL3 basic plasmid. The FR-α −3,394 to −47 nt/SV40(GC)6 and the FR-α-3,394 to −32 nt/SV40(INR) constructs are described elsewhere (28). The 5′ deletion constructs of the FR-α promoter, i.e., −272 to +33 nt, −116 to +33 nt, and −49 to +33 nt were all constructed by PCR using the appropriate primers and subcloned at MluI (upstream) and XhoI (downstream) sites in the pGL3 basic plasmid. In addition the G/C-rich sequence, −49 to −35 nt within the FR-α promoter (−49 to +33 nt)-luciferase construct was replaced by a TATA-box element (5′-AATAATTAA-3′) using PCR. The recombinant plasmids were amplified in E. coli strain XL1Blue and purified using the Qiagen plasmid kit (Qiagen, Chatsworth, CA). The entire cloned DNA sequence was verified by sequencing.

The expression plasmids for hSRC-1, hSRC-2, and hpCAF, the corresponding vector plasmid pCR 3.1 and the GRE2e1b promoter-luciferase plasmid were provided by Dr. Brian Rowan (Medical College of Ohio, OH). The expression plasmids for Sp1 and Sp3 were provided by Dr. Sumudra Periyasamy (Medical College of Ohio, OH). The expression plasmid for Sp4 was provided by Dr. Guntram Suske (Institut for Molekularbiologie und Tumorforschung Philipps-Universitat Marburg).

Cell culture and transfection. All of the cell lines were purchased from American Type Culture Collection (Rockville, MD) except for NB4 and KCL-22 cells, which were provided by Dr. Philip Koeffler (University of California, Los Angeles, CA) and Ishikawa cells provided by Dr. Brian Rowan (Medical College of Ohio, OH). Cells growing as monolayers were grown in 10 cm tissue culture plates at 37°C in 5% CO2 in the appropriate cell culture media supplemented with FBS (10%), 100 units/mL penicillin, 100 μg/mL streptomycin, and 2 mmol/L l-glutamine. Caki-1, HeLa, MG63, MCF-7, JEG, JAR, Ishikawa, SKOV-3, and SVG cells were routinely cultured in DMEM. NB4, KCL-22, K562, and KG-1 cells were grown in RPMI 1640. For treatment with various agents (dexamethasone, valproic acid, and trichostatin A) and for transfection, cells were grown in phenol red–free media supplemented with charcoal-stripped FBS (5% v/v), penicillin (100 units/mL), streptomycin (100 mg/mL), l-glutamine (2 mmol/L), insulin (2 μg/mL), and transferrin (40 μg/mL).

Transfections with various constructs were carried out in six-well plates (Corning, New York, NY) using FuGENE 6 (Roche Diagnostics), according to the manufacturer's suggested protocol. The amount of plasmid DNAs used for the transfections varied as indicated in the appropriate figure legends. Uniformity in transfection efficiencies was ascertained from measurements of β-galactosidase activity after cotransfection with an expression plasmid for this enzyme.

Preparation of cell lysates and luciferase assay. Cells in each well of a six-well tissue culture plate were washed once with PBS (pH 7.5; 2 mmol/L KH2PO4, 2.7 mmol/L KCl, 10 mmol/L Na2HPO4, 137 mmol/L NaCl) and lysed in 400 μL of reporter lysis buffer provided with the luciferase assay system (Promega). The samples were centrifuged at 12,000 × g for 2 minutes at room temperature. The supernatant was assayed for luciferase activity in a luminometer (Lumat LB9501, Berthold, Wildbad, Germany) using the luciferase substrate from Promega.

Real-time reverse transcription-PCR analysis. Total RNA for real-time reverse transcription-PCR from various cell lines was prepared using the RNeasy Mini kit purchased from Qiagen. Real-time reverse transcription-PCR was used to measure endogenous mRNAs for FR-α as well as glyceraldehyde-3-phosphate dehydrogenase (GAPDH) in the same samples. The reverse transcription step was carried out following standard procedures. Essentially, 200 ng of total RNA was mixed with random hexamer primers (5 × 10−4 absorbance units/μL), RNase inhibitor (1 unit/μL), Moloney murine leukemia virus reverse transcriptase (5 units/μL), and deoxynucleoside triphosphates (1.0 mmol/L each) in reverse transcriptase buffer [50 mmol/L potassium chloride and 10 mmol/L Tris-HCl (pH 8.3)]. The 10 μL reaction mixture was first incubated at 25°C for 10 minutes, then at 42°C for 15 minutes, and finally at 99°C for 6.5 minutes. The subsequent real-time PCR step for FR-α was carried out in the presence of 12.5 μL of PCR Mastermix (Applied Biosystems), 0.5 μL each of the forward primer (AAGTGCCGAGTGGGAGCT) and reverse primer (CATTGCACAGAACAGTGGGTG), and 0.5 μL of the TaqMan probe (6FAM-CCTGCCAACCTTTCCATTTCTACTTCCCC-TAMRA). The primers and the TaqMan probe for the control GAPDH gene were purchased from Applied Biosystems. The PCR conditions were 2 minutes at 50°C, then 10 minutes at 95°C, followed by 40 cycles of 15 seconds each at 95°C and finally 1 minute at 60°C. Fluorescence data generated were monitored and recorded by the Gene Amp 5700 sequence detection system (Applied Biosystems). All samples were set-up in triplicate and normalized to GAPDH values.

Measurement of cell surface binding of pteroyl lysine-fluorescein. The fluorescent folate analogue, pteroyl lysine-fluorescein, was kindly provided by Dr. John Hynes. HeLa cells grown in six-well plates and subjected to appropriate treatments were washed with PBS, detached from the plate by treatment with 2 mmol/L EDTA and resuspended in cold PBS. The cells were then washed briefly with isotonic acid buffer [10 mmol/L sodium acetate buffer (pH 3.5)/150 mmol/L NaCl] to remove endogenous bound folate, washed again in PBS and resuspended in PBS containing pteroyl lysine-fluorescein (10 nmol/L) and incubated on ice for 30 minutes with intermittent gentle agitation. The fluorescence on the surface of cells due to pteroyl lysine-fluorescein binding was measured in an EPICS Elite cytometer (Beckman Coulter, Fullerton, CA). Background fluorescence on the cell surface due to nonspecific binding of pteroyl lysine-fluorescein was determined by preincubating the cells with unlabeled folic acid (1 μmol/L) for 10 minutes before the addition of pteroyl lysine-fluorescein.

[3H]Folic acid binding assay for serum folate receptor. Twenty microliters of mouse serum was diluted into 0.5 mL of assay buffer [10 mmol/L sodium phosphate buffer (pH 7.5)/150 mmol/L NaCl/1% Triton X-100] in a 1.5 mL Eppendorf tube. To each assay tube, 2 pmol of [3H]folic acid (Moraveck, Brea, CA) was added and after incubation for 1 hour at 37°C, the protein-bound [3H]folic acid was measured by a charcoal-binding method as described (6). Nonspecific binding of [3H]folic acid was determined by carrying out the assay as above but after preincubating the diluted serum in the assay tube for 10 minutes with a 100-fold excess of unlabeled folic acid (200 pmol).

Phosphatidylinositol-specific phospholipase C treatment. HeLa cells grown in six-well plates and subjected to the appropriate treatments were treated with phosphatidylinositol-specific phospholipase C (0.1 units/mL) by adding the enzyme directly into the culture medium followed by incubation for 3 hours at 37°C. The cell lysates, prepared by lysis in PBS containing 1% Triton X-100 were subjected to Western blot analysis.

Preparation of nuclear extracts. HeLa cells subjected to the appropriate treatments were washed twice with PBS, scraped off the six-well plates, snap-frozen in liquid nitrogen, and stored at −80°C until the next step. Nuclear extracts were prepared as described (28), except that the cytoplasmic fractions were retained after lysis of cells for subsequent Western blot analysis. The nuclear extracts were desalted using G-25 Sephadex columns (Roche Diagnostics) following the supplier's protocol. The protein concentrations were determined by the Bradford assay (Bio-Rad, Hercules, CA).

Western blot analysis. Protein samples (10-20 μg) were resolved by electrophoresis on 8% SDS-PAGE gels and electrophoretically transferred to nitrocellulose filters. The blots were probed with the appropriate primary rabbit antibodies followed by goat anti-rabbit IgG conjugated to horseradish peroxidase and visualized using the enhanced chemiluminescence method. The same membrane was then similarly reprobed with a primary mouse anti-α tubulin antibody and secondary goat anti-mouse IgG conjugated to horseradish peroxidase and also subjected to Coomassie blue staining to ensure uniform sample loading.

Murine tumor xenograft model. Fox chase out-bred SCID female mice (29-35 days old) purchased from Charles River Laboratories were maintained under controlled conditions and fed with folate-free rodent chow ad libitum during the course of the experiments. After a period of acclimation, the mice were injected with 5 × 106 HeLa cells s.c. into the interscapular region. Dexamethasone pellets (0.001 mg/pellet for a 21-day release schedule) or placebo pellets were implanted s.c. when the tumor became palpable and grew to a diameter of approximately 0.5 cm. Five days after implanting the pellets, the mice were euthanized and the blood and tumor tissue collected. The tumors were snap-frozen in liquid nitrogen and stored at −80°C. The frozen tumor tissue was ground using mortar and pestle, lysed in PBS containing 1% Triton X-100, and centrifuged for separation of insoluble cell debris. The supernatant was used for Western blot analysis.

Effect of dexamethasone on FR-α expression in HeLa cells. Treatment of HeLa cells with dexamethasone (100 nmol/L) resulted in a progressive increase in the expression of both endogenous FR-α mRNA, measured by real-time PCR and FR-α protein, detected by Western blot using a FR-specific rabbit antiserum (Fig. 1A). This induction of FR-α began between 24 and 48 hours, and reached up to 7-fold elevation at 96 hours (Fig. 1A). In HeLa cells transfected with a full-length FR-α promoter-luciferase reporter construct (−3,394 to +33 nt), dexamethasone caused a dose-dependent increase in the promoter activity, reaching optimal activity between 5 and 50 nmol/L dexamethasone (Fig. 1B) and a corresponding dexamethasone dose-dependent increase in endogenous FR-α expression (Fig. 1B). These results provide evidence for positive regulation of FR-α by dexamethasone at the transcriptional level.

Figure 1.

Induction of FR-α by dexamethasone. A, HeLa cells were treated with either vehicle alone or dexamethasone (100 nmol/L) for the indicated periods, at the end of which, cells harvested from one triplicate set of wells for purification of total RNA and another set for preparation of cell lysates for Western blot. mRNA for FR-α and GAPDH were measured in the RNA samples by real-time reverse transcription-PCR (top). The Western blots were probed with rabbit anti-FR antibody (bottom). The same blots were re-probed with a mouse antitubulin antibody (bottom). B, HeLa cells grown in six-well plates were transfected with the full length FR-α promoter-luciferase reporter plasmid (0.5 μg) followed by treatment either with vehicle alone or with the indicated concentrations of dexamethasone for 96 hours. The harvested cells were divided into two portions, one of which was used to measure luciferase activity (top) and the other subjected to Western blot analysis using rabbit anti-FR antibody (bottom). C, HeLa cells in triplicate six-well plates were treated with the indicated concentrations of dexamethasone for 96 hours, at the end of which, they were incubated for a further 3 hours, either in the absence or in the presence of phosphatidylinositol-specific phospholipase C. The cell lysates were then subjected to Western blot analysis using rabbit anti-FR antibody. Uniformity of sample loading in the blots was confirmed by Coomassie blue staining. D, HeLa cells in six-well plates were treated with either vehicle or dexamethasone for 96 hours, at the end of which, the cells were harvested for flow cytometric analysis of the binding of pteroyl lysine-fluorescein. In negative controls, the cells were preincubated with a 1,000-fold excess of unlabeled folic acid. The fluorescence shift in the negative control samples was <1% of the signal due to unblocked FR.

Figure 1.

Induction of FR-α by dexamethasone. A, HeLa cells were treated with either vehicle alone or dexamethasone (100 nmol/L) for the indicated periods, at the end of which, cells harvested from one triplicate set of wells for purification of total RNA and another set for preparation of cell lysates for Western blot. mRNA for FR-α and GAPDH were measured in the RNA samples by real-time reverse transcription-PCR (top). The Western blots were probed with rabbit anti-FR antibody (bottom). The same blots were re-probed with a mouse antitubulin antibody (bottom). B, HeLa cells grown in six-well plates were transfected with the full length FR-α promoter-luciferase reporter plasmid (0.5 μg) followed by treatment either with vehicle alone or with the indicated concentrations of dexamethasone for 96 hours. The harvested cells were divided into two portions, one of which was used to measure luciferase activity (top) and the other subjected to Western blot analysis using rabbit anti-FR antibody (bottom). C, HeLa cells in triplicate six-well plates were treated with the indicated concentrations of dexamethasone for 96 hours, at the end of which, they were incubated for a further 3 hours, either in the absence or in the presence of phosphatidylinositol-specific phospholipase C. The cell lysates were then subjected to Western blot analysis using rabbit anti-FR antibody. Uniformity of sample loading in the blots was confirmed by Coomassie blue staining. D, HeLa cells in six-well plates were treated with either vehicle or dexamethasone for 96 hours, at the end of which, the cells were harvested for flow cytometric analysis of the binding of pteroyl lysine-fluorescein. In negative controls, the cells were preincubated with a 1,000-fold excess of unlabeled folic acid. The fluorescence shift in the negative control samples was <1% of the signal due to unblocked FR.

Close modal

The induction of FR-α in HeLa cells by dexamethasone did not reflect a global increase in gene expression, because under these conditions, the expression level of the GAPDH gene (Fig. 1A) as well as tubulin and Sp family proteins (Figs. 1A and 5B, discussed in a later section) were unaltered. The typical plasma membrane localization and glycosyl-phosphatidylinositol membrane anchor attachment known for FR-α was confirmed for the receptor synthesized de novo following dexamethasone treatment, because as seen with the Western blot, the dexamethasone-induced FR was quantitatively cleaved from the cell surface upon treatment with phosphatidylinositol-specific phospholipase C, which is a diagnostic characteristic of proteins with a glycosyl-phosphatidylinositol plasma membrane anchor (Fig. 1C). Furthermore, the dexamethasone-induced FR-α protein retained its ability to bind ligand on the cell surface, evident from an increase in the binding of the fluorescent folate analogue, pteroyl lysine-fluorescein on the surface of the treated cells (Fig. 1D). Thus, the subcellular localization and function of FR-α was unaltered by dexamethasone induction.

GR ligand-specificity of FR-α induction. RU-486, a specific antagonist of GR that competes with dexamethasone for binding to GR, inhibited the induction of FR-α promoter-luciferase reporter in transfected HeLa cells in a dose-dependent manner (Fig. 2A). RU-486 similarly inhibited dexamethasone activation of the control GRE2e1b promoter-luciferase reporter construct in transfected HeLa cells (Fig. 2B). The GRE2e1b promoter contains a glucocorticoid response element (GRE) and is a classical target of dexamethasone activation through GR. Under similar conditions, the dose response of inhibition of the FR-α promoter activity by RU-486 paralleled that for the GRE2e1b promoter, suggesting that the regulation of the FR-α promoter by dexamethasone is, at some level, mediated by GR.

Figure 2.

Effect of RU486 on dexamethasone induction of promoter activity. HeLa cells were transfected with either FR-α promoter-luciferase plasmid (A) or with a control GRE2e1b promoter-luciferase plasmid (B), and grown for a further 72 hours either in the absence or in the presence of dexamethasone (100 nmol/L) in combination with the indicated concentrations of RU-486. The cells were then harvested to assay luciferase activity in the lysates. The fold induction of promoter activity is plotted in each case, considering the promoter activity of the vehicle-treated control as unity.

Figure 2.

Effect of RU486 on dexamethasone induction of promoter activity. HeLa cells were transfected with either FR-α promoter-luciferase plasmid (A) or with a control GRE2e1b promoter-luciferase plasmid (B), and grown for a further 72 hours either in the absence or in the presence of dexamethasone (100 nmol/L) in combination with the indicated concentrations of RU-486. The cells were then harvested to assay luciferase activity in the lysates. The fold induction of promoter activity is plotted in each case, considering the promoter activity of the vehicle-treated control as unity.

Close modal

Response time and reversibility of dexamethasone induction of the FR-α promoter. HeLa cells were transiently transfected with either GRE2e1b promoter-luciferase or FR-α promoter-luciferase and treatment of the cells with either dexamethasone or vehicle begun at 12 hours post-transfection (Fig. 3). In the transiently transfected cells, close to maximal activation of the GRE2e1b promoter-luciferase by dexamethasone occurred at 3 hours, reached the maximum value at 6 hours and declined between 24 and 48 hours (Fig. 3A). In contrast, dexamethasone activation of transiently transfected FR-α promoter-luciferase was relatively low at 3 hours and progressed gradually, reaching its maximum value at 24 hours and was sustained up to 48 hours (Fig. 3B). Furthermore, when dexamethasone was withdrawn at 6, 12, or 24 hours, the activity of the GRE2e1b promoter measured at 48 hours was greatly reduced compared with the values measured at 6, 12, and 24 hours, respectively (Fig. 3A). In contrast, after withdrawal of dexamethasone at 6, 12, or 24 hours, the activity of the FR-α promoter further increased as seen at 48 hours (Fig. 3B). This observation explains why contrary to the GRE2e1b promoter, the activity of the FR-α promoter was sustained in the later stage of transient transfection (Fig. 3A and B). This delayed activation of the FR-α promoter by dexamethasone indicates that the dexamethasone response is likely mediated indirectly through a product(s) of dexamethasone/GR action on an upstream target gene(s) of dexamethasone and that the mode of action of dexamethasone on the FR-α promoter is thus likely fundamentally different from its activation of a GRE-driven promoter.

Figure 3.

Time course and reversibility of dexamethasone induction of FR-α promoter activity and the effect of cycloheximide. HeLa cells were transfected with either the control GRE2e1b promoter-luciferase plasmid (A) or FR-α promoter-luciferase plasmid (B). Twelve hours post-transfection, the cells were treated for the indicated periods with either dexamethasone (1 μmol/L) or vehicle alone and harvested after the indicated periods for measurement of luciferase activity in the cell lysates. In cases where dexamethasone treatment was terminated before harvest, dexamethasone was removed by aspirating the media and washing the cells with dexamethasone-free media before replenishing the cells with media that did not contain dexamethasone. C, HeLa cells were pretreated with either cycloheximide (10 μmol/L) or vehicle for 2 hours followed by the introduction of dexamethasone (1 μmol/L) or vehicle as indicated. Twelve hours later, dexamethasone/cycloheximide was removed by aspirating the media and washing the cells with dexamethasone/cycloheximide–free media before replenishing the cells with media that did not contain dexamethasone/cycloheximide. The cells were harvested 60 hours later and total RNA extracted for quantification of FR-α mRNA by real-time reverse transcription-PCR.

Figure 3.

Time course and reversibility of dexamethasone induction of FR-α promoter activity and the effect of cycloheximide. HeLa cells were transfected with either the control GRE2e1b promoter-luciferase plasmid (A) or FR-α promoter-luciferase plasmid (B). Twelve hours post-transfection, the cells were treated for the indicated periods with either dexamethasone (1 μmol/L) or vehicle alone and harvested after the indicated periods for measurement of luciferase activity in the cell lysates. In cases where dexamethasone treatment was terminated before harvest, dexamethasone was removed by aspirating the media and washing the cells with dexamethasone-free media before replenishing the cells with media that did not contain dexamethasone. C, HeLa cells were pretreated with either cycloheximide (10 μmol/L) or vehicle for 2 hours followed by the introduction of dexamethasone (1 μmol/L) or vehicle as indicated. Twelve hours later, dexamethasone/cycloheximide was removed by aspirating the media and washing the cells with dexamethasone/cycloheximide–free media before replenishing the cells with media that did not contain dexamethasone/cycloheximide. The cells were harvested 60 hours later and total RNA extracted for quantification of FR-α mRNA by real-time reverse transcription-PCR.

Close modal

Effect of cycloheximide on dexamethasone induction of the FR-α promoter. The possibility that the action of dexamethasone on the FR-α promoter requires intermediate synthesis of a protein factor(s) was tested using cycloheximide to inhibit de novo protein synthesis during the early stage (0-12 hours) of dexamethasone treatment (Fig. 3C). The delayed induction of FR-α promoter activity observed at 72 hours after only a 12-hour treatment with dexamethasone was abrogated when cycloheximide was included during the dexamethasone treatment (Fig. 3C). This result indicates that in dexamethasone induction of the FR-α promoter, an early action of dexamethasone is to induce de novo synthesis of some other protein(s).

Effect of coactivators on FR-α promoter activity and its induction by dexamethasone. The effect of the nuclear receptor/GR coactivators, SRC-1, SRC-2, and pCAF on promoter activity as well as its activation by dexamethasone was measured in HeLa cells cotransfected with expression plasmids for the individual coactivators and FR-α promoter-luciferase (Table 1). Each of the coactivators enhanced the basal FR-α promoter activity. However, each of the coactivators also potentiated dexamethasone induction of the FR-α promoter. This result lends further support to GR mediation of the dexamethasone effect on FR-α gene expression and the view that this regulation, albeit indirect, is transcriptional.

Table 1.

Effect of coactivators ± dexamethasone on FR-α promoter activity

Coregulator*Promoter activity
VehicleDexamethasone
None 1.0 ± 0.0 5.7 ± 0.2 
SRC-1 2.9 ± 0.9 13.6 ± 0.6 
SRC-2 3.2 ± 0.2 11.3 ± 0.5 
PCAF 4.6 ± 0.7 11.5 ± 0.8 
Coregulator*Promoter activity
VehicleDexamethasone
None 1.0 ± 0.0 5.7 ± 0.2 
SRC-1 2.9 ± 0.9 13.6 ± 0.6 
SRC-2 3.2 ± 0.2 11.3 ± 0.5 
PCAF 4.6 ± 0.7 11.5 ± 0.8 

NOTE: The promoter activity was determined by measuring luciferase reporter activity. The values are expressed as the ratios to that for the vehicle-treated control in the absence of cotransfected coregulator.

*

HeLa cells (106) were transfected with FR-α promoter-luciferase (0.5 μg plasmid) and cotransfected with an expression plasmid for SRC-1, SRC-2, or pCAF or with the empty plasmid (0.5 μg plasmid).

The transfected cells were treated with either vehicle or dexamethasone (1 μmol/L) for a period of 48 hours post-transfection before harvesting them to measure luciferase activity.

Identification of the target site of dexamethasone action in the FR-α promoter. The FR-α promoter-luciferase reporter construct used in the above studies included the FR-α gene sequence, −3,394 to +33 nt, spanning both the P1 and the P4 promoters. 5′ deleted versions of this promoter construct, i.e., the −272 to +33 nt and the −116 to +33 nt fragments of the promoter, retained the dexamethasone responsiveness of the full-length construct in transfected HeLa cells (Fig. 4A). The time course of the dexamethasone response was also unaltered for the truncated promoter fragments (Fig. 4A). The only functional cis elements known to occur between the initiator sequence and −116 nt are the G/C-rich Sp1 binding sites of the P4 promoter, indicated in Fig. 4B. A further 5′ deletion of FR-α promoter luciferase in which all of the promoter sequence upstream of the most proximal Sp1 element was deleted retained the dexamethasone response (Fig. 4C). Because the single Sp1 element in the FR-α (−49 to +33 nt) construct is necessary for basal promoter activity (27), a possible role for this G/C-rich element in mediating the dexamethasone effect was tested by replacing this sequence (−49 to −35 nt) with a TATA-box element (AATAATTAA) to retain basal promoter activity (Fig. 4C). The chimeric promoter was unresponsive to dexamethasone treatment (Fig. 4C), implicating the Sp1 element as a mediator of the dexamethasone effect.

Figure 4.

Mapping the target site of dexamethasone action in the FR-α promoter. A, HeLa cells were transfected with FR-α promoter-luciferase reporter constructs containing either the entire P1/P4 promoter region (−3,394 to +33 nt) or its 5′ deleted fragments (−272 to +33 nt or −116 to +33 nt). Twelve hours post-transfection, the cells were treated with dexamethasone (1 μmol/L) for the indicated periods before harvesting for luciferase assay. Promoter activity is plotted as fold increase over that in cells that were treated with vehicle alone. B, DNA sequence of the FR-α promoter fragment −116 to +33 nt. The numbers represent the positions of the nucleotides relative to the position of the transcription start site (+1 nt). The Sp1 elements are indicated in boldface letters. C, HeLa cells were transfected with a 5′ deleted FR-α promoter-luciferase reporter constructs containing the promoter fragment −49 to +33 nt. A chimeric promoter construct in which a TATA-box element was attached upstream of the FR-α promoter fragment −35 to +33 nt was also used in the transfection. Twelve hours post-transfection, the cells were treated with dexamethasone (1 μmol/L) for the indicated periods before harvesting for luciferase assay. Promoter activity is plotted as fold increase over that in cells that were treated with vehicle alone. D, HeLa cells were transfected with FR-α or SV40 promoter-luciferase constructs containing the indicated promoter fragments of FR-α, SV40, or FR-α/SV40 promoter chimeras. Twelve hours post-transfection, the cells were treated with dexamethasone (1 μmol/L) for the indicated periods before harvesting for luciferase assay. Promoter activity is plotted as fold increase over that in cells that were treated with vehicle alone. The pGL3 plasmid (Promega) was used for the SV40 promoter-luciferase contained in it. The construct, FR-α −3,394 to −32 nt/SV40(INR) is a chimera in which the initiator region of the SV40 early promoter is substituted for that of the P4 promoter. The construct FR-α −3,394 to −47 nt/SV40(GC)6 is a chimera in which the cluster of six Sp1 elements of the SV40 early promoter is substituted for Sp1 elements of the P4 promoter.

Figure 4.

Mapping the target site of dexamethasone action in the FR-α promoter. A, HeLa cells were transfected with FR-α promoter-luciferase reporter constructs containing either the entire P1/P4 promoter region (−3,394 to +33 nt) or its 5′ deleted fragments (−272 to +33 nt or −116 to +33 nt). Twelve hours post-transfection, the cells were treated with dexamethasone (1 μmol/L) for the indicated periods before harvesting for luciferase assay. Promoter activity is plotted as fold increase over that in cells that were treated with vehicle alone. B, DNA sequence of the FR-α promoter fragment −116 to +33 nt. The numbers represent the positions of the nucleotides relative to the position of the transcription start site (+1 nt). The Sp1 elements are indicated in boldface letters. C, HeLa cells were transfected with a 5′ deleted FR-α promoter-luciferase reporter constructs containing the promoter fragment −49 to +33 nt. A chimeric promoter construct in which a TATA-box element was attached upstream of the FR-α promoter fragment −35 to +33 nt was also used in the transfection. Twelve hours post-transfection, the cells were treated with dexamethasone (1 μmol/L) for the indicated periods before harvesting for luciferase assay. Promoter activity is plotted as fold increase over that in cells that were treated with vehicle alone. D, HeLa cells were transfected with FR-α or SV40 promoter-luciferase constructs containing the indicated promoter fragments of FR-α, SV40, or FR-α/SV40 promoter chimeras. Twelve hours post-transfection, the cells were treated with dexamethasone (1 μmol/L) for the indicated periods before harvesting for luciferase assay. Promoter activity is plotted as fold increase over that in cells that were treated with vehicle alone. The pGL3 plasmid (Promega) was used for the SV40 promoter-luciferase contained in it. The construct, FR-α −3,394 to −32 nt/SV40(INR) is a chimera in which the initiator region of the SV40 early promoter is substituted for that of the P4 promoter. The construct FR-α −3,394 to −47 nt/SV40(GC)6 is a chimera in which the cluster of six Sp1 elements of the SV40 early promoter is substituted for Sp1 elements of the P4 promoter.

Close modal

To determine whether another Sp1-dependent promoter, similar to the FR-α promoter, would respond to dexamethasone in a similar manner as the FR-α P4 promoter, the effect of dexamethasone was tested on the Sp1-dependent SV40 promoter-luciferase reporter. In transfected HeLa cells dexamethasone enhanced the activity of the SV40 promoter, with a time course that was similar to that for the FR-α promoter (Fig. 4D). It may be noted that the G/C-rich region of the SV40 promoter is a stronger Sp1 element than that of the FR-α P4 promoter because it contains six repeat Sp1 elements. To determine the relative roles of the initiator (and the flanking) region and the Sp1 elements in the dexamethasone response of the P4 promoter, either the P4 initiator region (−28 to +33 nt) or the entire G/C-rich region of the P4 promoter in the full length FR-α promoter-luciferase construct was replaced by the corresponding regions of the SV40 promoter (Fig. 4D). Both the chimeric promoter constructs were responsive to dexamethasone with a time course similar to that of the FR-α promoter; however, the magnitude of the dexamethasone response was much greater for the FR-α promoter chimera containing the G/C-rich region of the SV40 promoter (Fig. 4D). This suggests that whereas a G/C-rich sequence element mediates and may determine the magnitude of the delayed dexamethasone response of a promoter, for optimal response to occur, there is a preferred initiator context.

The action of Sp family proteins on the FR-α promoter and the effect of dexamethasone on their expression levels. A well-known mechanism of gene regulation through G/C-rich cis elements involves changes in differential expression and transcriptional activities of their cognate trans factors, i.e., Sp family proteins. To test this possibility, the action of major Sp family members including Sp1, Sp3, and Sp4 on the FR-α promoter activity was tested by cotransfection of HeLa cells with FR-α promoter-luciferase and expression plasmids for the Sp proteins, individually and in combination. All of the Sp proteins were equipotent activators of the FR-α promoter (Fig. 5A). Furthermore, dexamethasone treatment of HeLa cells (up to 96 hours) did not result in any obvious substantive changes in the expression levels of endogenous Sp1, Sp3, or Sp4 or their apparent molecular weights, under conditions in which endogenous FR-α was up-regulated (Fig. 5B). These results exclude changes in the relative expression levels or phosphorylation levels of Sp1, Sp3, or Sp4 as a mechanism mediating the dexamethasone effect on the FR-α gene and suggest that dexamethasone regulates the interaction(s) of some other transcription factor(s) with the core transcription initiation complex of the P4 promoter.

Figure 5.

Transactivation of the FR-α promoter by Sp family proteins and the effect of dexamethasone on their expression levels. A, HeLa cells were cotransfected with FR-α promoter-luciferase (0.5 μg) and an expression plasmid for Sp1, Sp3, or Sp4 alone or in combination as indicated. As a negative control, cells were cotransfected with the pRc vector. The cells were harvested for luciferase assay 72 hours post-transfection. B, HeLa cells were grown for 96 hours in the presence of vehicle alone or with the addition of dexamethasone (1 μmol/L) for either the entire 96 hours or for the last 24 hours. The cells were then harvested for preparation of nuclear extracts and cytoplasmic fractions. The cytoplasmic fractions were probed by Western blot with a rabbit anti-FR-α antibody or with a mouse antibody to tubulin (bottom). The nuclear extracts were probed by Western blot using antibodies specific for Sp1, Sp3, or Sp4.

Figure 5.

Transactivation of the FR-α promoter by Sp family proteins and the effect of dexamethasone on their expression levels. A, HeLa cells were cotransfected with FR-α promoter-luciferase (0.5 μg) and an expression plasmid for Sp1, Sp3, or Sp4 alone or in combination as indicated. As a negative control, cells were cotransfected with the pRc vector. The cells were harvested for luciferase assay 72 hours post-transfection. B, HeLa cells were grown for 96 hours in the presence of vehicle alone or with the addition of dexamethasone (1 μmol/L) for either the entire 96 hours or for the last 24 hours. The cells were then harvested for preparation of nuclear extracts and cytoplasmic fractions. The cytoplasmic fractions were probed by Western blot with a rabbit anti-FR-α antibody or with a mouse antibody to tubulin (bottom). The nuclear extracts were probed by Western blot using antibodies specific for Sp1, Sp3, or Sp4.

Close modal

The action of inhibitors of histone deacetylase on dexamethasone induction of FR-α gene expression. Because the transcriptional activity of nuclear receptors entails modulation of histone acetylation, it was of interest to examine the effects of HDAC inhibitors on dexamethasone induction of FR-α gene transcription. The well-tolerated drug valproic acid and the well-characterized laboratory reagent, trichostatin A, were chosen as the HDAC inhibitors for these experiments. In HeLa cells transfected with FR-α promoter-luciferase, both valproic acid (Fig. 6A) and trichostatin A (Fig. 6B) independently increased the promoter activity to some extent within their pharmacologic dose ranges but they both greatly potentiated dexamethasone induction of the FR-α promoter. Valproic acid also potentiated dexamethasone induction of the endogenous FR-α in HeLa cells, in a dose-dependent manner (Fig. 6C). Finally, the potentiation of the dexamethasone induction of the transfected FR-α promoter-luciferase in HeLa cells by valproic acid occurred both during the early (0-24 hours) and later (24-72 hours) stages of dexamethasone treatment (Fig. 6D). Under the conditions of the above treatments, the HDAC inhibitors did not affect the viability or the growth of HeLa cells (data not shown).

Figure 6.

Effect of HDAC inhibitors on FR-α gene regulation by dexamethasone. A, HeLa cells were transfected with FR-α promoter-luciferase and treated for 72 hours with different concentrations of valproic acid either in the absence or in the presence of dexamethasone (100 nmol/L) and then harvested for luciferase assay. B, HeLa cells were transfected with FR-α promoter-luciferase and treated for 72 hours with different concentrations of trichostatin A, either in the absence or in the presence of dexamethasone (100 nmol/L) and then harvested for luciferase assay. C, HeLa cells that were either untreated or treated with dexamethasone (100 nmol/L) and/or different concentrations of valproic acid for 96 hours were harvested and the cell lysates probed by Western blot using either a rabbit anti-FR antibody or a mouse antibody to tubulin. D, HeLa cells were transfected with FR-α promoter-luciferase and then treated for 72 hours with vehicle alone or with dexamethasone (100 nmol/L). Valproic acid (1 mmol/L) was introduced for the entire 72 hours, the first 24 hours, or for the period between 24 and 72 hours of dexamethasone treatment. The cells were harvested for luciferase assay at the end of 72 hours.

Figure 6.

Effect of HDAC inhibitors on FR-α gene regulation by dexamethasone. A, HeLa cells were transfected with FR-α promoter-luciferase and treated for 72 hours with different concentrations of valproic acid either in the absence or in the presence of dexamethasone (100 nmol/L) and then harvested for luciferase assay. B, HeLa cells were transfected with FR-α promoter-luciferase and treated for 72 hours with different concentrations of trichostatin A, either in the absence or in the presence of dexamethasone (100 nmol/L) and then harvested for luciferase assay. C, HeLa cells that were either untreated or treated with dexamethasone (100 nmol/L) and/or different concentrations of valproic acid for 96 hours were harvested and the cell lysates probed by Western blot using either a rabbit anti-FR antibody or a mouse antibody to tubulin. D, HeLa cells were transfected with FR-α promoter-luciferase and then treated for 72 hours with vehicle alone or with dexamethasone (100 nmol/L). Valproic acid (1 mmol/L) was introduced for the entire 72 hours, the first 24 hours, or for the period between 24 and 72 hours of dexamethasone treatment. The cells were harvested for luciferase assay at the end of 72 hours.

Close modal

Effect of dexamethasone treatment and histone deacetylase inhibition on endogenous FR-α gene expression in FR-α–positive versus FR-α–negative cell lines. In order to test whether dexamethasone increased FR-α gene expression in other FR-α–positive cell lines and to examine whether dexamethasone could alter the tissue expression pattern of FR-α by producing de novo expression of the receptor in FR-α–negative cells, a variety of cell types were treated with dexamethasone, trichostatin A, or a combination of dexamethasone and trichostatin A. Cells in which FR-α mRNA was determined by real-time reverse transcription-PCR to be <1/1,000 of that in HeLa cells and in which the FR-α protein was undetectable by Western blot were considered to be FR-α–negative. In human hematopoietic cells (KG-1, Kcl-22, K-562, and NB-4), fibroblasts (MG-63 and SVG), and epithelial (Caki-1) cell lines that were FR-α–negative, there was no detectable increase in the receptor mRNA expression upon dexamethasone/trichostatin A treatments (Table 2); however, in JAR, Ishikawa, and SKOV-3 cells that are FR-α–positive, dexamethasone, alone or in combination with trichostatin A, increased FR-α expression (Table 2).

Table 2.

Effect of dexamethasone ± trichostatin A on endogenous FR expression in various cell lines

Cell lineCell typeEndogenous FR-αIncrease in FR-α expression*
VehicleDexamethasoneTrichostatin ADexamethasone + trichostatin A
KG-1 Acute myelogeneous leukemia Negative  − − − 
Kcl-22 Myeloblastic leukemia Negative − − − − 
K-562 Erythroleukemia Negative − − − − 
NB-4 Acute promyelocytic leukemia Negative − − − − 
MG-63 Osteosarcoma Negative − − − − 
SVG Transformed fibroblast Negative − − − − 
Caki-1 Kidney carcinoma Negative − − − − 
JAR Choriocarcinoma Positive − − − +§ 
Ishikawa Uterine adenocarcinoma Positive − 
Skov-3 Ovarian carcinoma Positive − 
Cell lineCell typeEndogenous FR-αIncrease in FR-α expression*
VehicleDexamethasoneTrichostatin ADexamethasone + trichostatin A
KG-1 Acute myelogeneous leukemia Negative  − − − 
Kcl-22 Myeloblastic leukemia Negative − − − − 
K-562 Erythroleukemia Negative − − − − 
NB-4 Acute promyelocytic leukemia Negative − − − − 
MG-63 Osteosarcoma Negative − − − − 
SVG Transformed fibroblast Negative − − − − 
Caki-1 Kidney carcinoma Negative − − − − 
JAR Choriocarcinoma Positive − − − +§ 
Ishikawa Uterine adenocarcinoma Positive − 
Skov-3 Ovarian carcinoma Positive − 
*

Treatment of the cells with dexamethasone (0.1 μmol/L) and/or trichostatin A (25 ng/mL) was carried out for 96 hours.

FR-α mRNA levels in the FR-α–negative cell lines were at least 1,000-fold less than that in HeLa cells as assessed by real-time reverse transcription-PCR and the protein could not be detected by Western blot.

The “−” sign indicates that there was no increase in FR-α mRNA as assessed by real-time reverse transcription-PCR.

§

The “+” sign indicates that there was an increase in FR-α mRNA as assessed by real-time reverse transcription-PCR.

Effect of dexamethasone/trichostatin A on GR expression. GR is known to be down-regulated in cells treated with dexamethasone. The effect of dexamethasone/trichostatin A on GR expression in HeLa cells was tested in order to explore the possibility that modulation of GR expression plays a role in the potentiation of dexamethasone induction of the FR-α promoter by HDAC inhibition (Fig. 7). The Western blot data in Fig. 7 shows that over a period of 96 hours of dexamethasone treatment, GR was progressively down-regulated by dexamethasone, alone, as well as by dexamethasone in the presence of trichostatin A. Although trichostatin A by itself did not alter the GR level, it did seem to decrease the extent of down-regulation of GR by dexamethasone. However, this did not seem to be a significant factor in the synergy produced by trichostatin A at least in the early stage (up to 48 hours) because the GR levels were comparable between the two treatments up to 48 hours.

Figure 7.

Effect of dexamethasone and trichostatin A on GR expression. HeLa cells were treated with vehicle, dexamethasone (1 μmol/L) or trichostatin A (4 ng/mL) or with a combination of dexamethasone and trichostatin A for the indicated periods. The cells were harvested and the cell lysates subjected to Western blot analysis using antibody to GR, FR, or tubulin.

Figure 7.

Effect of dexamethasone and trichostatin A on GR expression. HeLa cells were treated with vehicle, dexamethasone (1 μmol/L) or trichostatin A (4 ng/mL) or with a combination of dexamethasone and trichostatin A for the indicated periods. The cells were harvested and the cell lysates subjected to Western blot analysis using antibody to GR, FR, or tubulin.

Close modal

Up-regulation of tumor and serum FR-α by dexamethasone in a murine tumor xenograft model. A murine tumor xenograft model was used to test whether the in vitro observations of FR-α induction by dexamethasone would extend to the regulation of the FR-α gene in the physiologic milieu. In Fig. 8A, two groups of three SCID female mice bearing s.c. HeLa cell tumors (∼0.5 cm diameter) were tested. Either low-dose slow-release dexamethasone pellets (to achieve a circulating concentration of 0.24 μmol/L dexamethasone) or placebo pellets were implanted s.c. in the mice for a duration of 5 days before sacrificing the mice to harvest the tumors. As expected, the dexamethasone treatment did not cause a significant difference between the treated and placebo groups in terms of body weight and activity (data not shown). Dexamethasone treatment caused a substantial increase in FR-α protein in the tumors as seen by Western blot analysis of the tumor cell lysate (Fig. 8A).

Figure 8.

Effect of dexamethasone on tumor and serum FR-α levels in a murine tumor xenograft model. A, two groups of three SCID female mice were inoculated s.c. at a single site with HeLa cells (5 × 106). When the tumors had grown to approximately 0.5 cm diameter, the mice were implanted s.c. with slow-release dexamethasone pellets or placebo pellets. After 5 days, the mice were sacrificed and cell lysates were prepared from the tumors and probed by Western blot using rabbit anti-FR antibody or a mouse antibody to tubulin. B, groups of five mice bearing s.c. HeLa cell tumors or without the tumors were implanted s.c. with slow-release dexamethasone pellets or placebo pellets. After 5 days, the mice were sacrificed and the blood collected. The serum folate binding protein (serum FR) was measured by a [3H]folic acid binding assay. The normal groups represent mice that did not bear tumors. The data show mean ± SE for each group. Statistical differences between groups were determined using Student's t test. The level of significance for the difference between the normal groups and the placebo-treated tumor group was P < 0.1 and between the dexamethasone and placebo groups for the tumor-bearing mice was P < 0.01.

Figure 8.

Effect of dexamethasone on tumor and serum FR-α levels in a murine tumor xenograft model. A, two groups of three SCID female mice were inoculated s.c. at a single site with HeLa cells (5 × 106). When the tumors had grown to approximately 0.5 cm diameter, the mice were implanted s.c. with slow-release dexamethasone pellets or placebo pellets. After 5 days, the mice were sacrificed and cell lysates were prepared from the tumors and probed by Western blot using rabbit anti-FR antibody or a mouse antibody to tubulin. B, groups of five mice bearing s.c. HeLa cell tumors or without the tumors were implanted s.c. with slow-release dexamethasone pellets or placebo pellets. After 5 days, the mice were sacrificed and the blood collected. The serum folate binding protein (serum FR) was measured by a [3H]folic acid binding assay. The normal groups represent mice that did not bear tumors. The data show mean ± SE for each group. Statistical differences between groups were determined using Student's t test. The level of significance for the difference between the normal groups and the placebo-treated tumor group was P < 0.1 and between the dexamethasone and placebo groups for the tumor-bearing mice was P < 0.01.

Close modal

In Fig. 8B, the effect of dexamethasone on the level of soluble FR in the serum was tested in the mouse HeLa cell tumor xenograft model and compared with control mice that did not bear the tumor. Groups of five mice were used in this experiment and serum FR was estimated by using [3H]folic acid binding assay. S.c. inoculation of HeLa cells and treatment with dexamethasone or placebo pellets were carried out as described above for Fig. 8B and at the end of the treatment, blood samples were collected from the sacrificed animals. In mice bearing the HeLa cell tumors and treated with placebo, serum FR was slightly elevated compared with the mice that did not bear tumors (P < 0.1; Fig. 8B). Administration of dexamethasone to the tumor-bearing mice further increased their serum FR substantially (P < 0.01). These results show that dexamethasone induction of FR-α in a solid tumor in vivo will cause an increase in serum FR and that such a response to dexamethasone may be considered to be indicative of the presence of a FR-α-rich tumor.

In view of the large number of preclinical and clinical studies that show the considerable potential for the utility of FR as a target for tumor-selective delivery of a broad range of experimental therapies, there is a pressing need to address the problem of the variable and frequently limiting expression of FR in the target tumors. Recently published (28, 30, 31) and unpublished studies in our laboratory have shown that the FR gene family is regulated by nuclear receptors. The specific regulatory mechanisms, however, are quite varied but none involve classical response elements. Thus, ER acts by directly interacting with the proximal P4 promoter of the FR-α gene to repress it and this repressive effect is enhanced by estrogen; antiestrogens will bind to ER and de-repress FR-α transcription (28). Retinoid compounds act through each of the three retinoic acid receptors (α, β, and γ) in distinct ways, directly interacting with the FR-β gene to up-regulate its expression (30, 31). Other unpublished studies indicate the positive regulation of FR-α by the androgen receptor by direct interaction with proteins bound to an enhancer element in the FR-α gene as well as regulation of FR-α by the progesterone receptor. In this context, the regulation of FR-α by GR, as reported here, is particularly interesting because, unlike other steroid hormone receptors, GR is almost ubiquitously expressed (32). Furthermore, the many clinical applications of dexamethasone and the observation that variations in endogenous cortisol levels do not impact the physiologic effects of dexamethasone (33) also indicate that gene activation by endogenous glucocorticoids is suboptimal.

The results of this study clearly show that dexamethasone up-regulates the expression of the endogenous FR-α gene in cell lines in which the gene is transcriptionally active but not in a variety of cell types that are FR-α–negative; under these conditions, other active genes such as those encoding GAPDH, tubulin, and Sp family proteins were not regulated by dexamethasone. These observations are consistent with the nature of FR regulation by other nuclear receptors and are further supported by the lack of de novo FR expression in various FR-negative tissues in mice following dexamethasone treatment (data not shown). The dexamethasone-induced FR-α retained the desired characteristics of the receptor as a tumor target and a releasable tumor marker, i.e., it's anchoring to the membrane by glycosyl-phosphatidylinositol and its ability to bind ligand. A substantial induction of FR-α by dexamethasone was also observed in a murine tumor xenograft model, both within the tumor and in the serum, confirming the relevance of this regulation to the physiologic setting.

We undertook a study of the molecular processes involved in the action of GR on the FR-α gene, in the context of current knowledge of nuclear receptor functions, to provide a rational approach to designing and optimizing the use of GR ligands in FR-α targeting and FR-dependent diagnostics and to help to understand, anticipate and address associated problems. FR-α up-regulation in response to dexamethasone treatment was relatively slow and progressed even after dexamethasone was withdrawn as early as 6 hours of treatment. This suggests that dexamethasone does not act directly on the FR-α gene and that an early sequence of events involving dexamethasone action on some other target(s) precedes and is necessary for FR-α up-regulation. Further evidence that the expression of an intermediary protein factor in response to dexamethasone treatment is required to mediate the action of dexamethasone on the FR-α promoter is provided by the observation that blocking de novo protein synthesis in the early lag phase of dexamethasone action abolished the delayed activation of the promoter by dexamethasone. The dexamethasone action must be mediated by GR based on (a) the dexamethasone dose-dependence; (b) the inhibition of its action by the GR antagonist, RU486, and (c) potentiation of the dexamethasone effect by coactivators. Whereas the identity of the critical immediate (direct) target(s) of dexamethasone/GR action is unclear, the evidence points to the transcriptional nature of these early regulatory events as opposed to the more recently discovered nongenomic actions of steroid receptors (34, 35). This is evident from the GR coactivator dependence of the dexamethasone effect as well as potentiation of dexamethasone induction of FR-α by HDAC inhibition in the early (within 24 hours) as well as late phases (after 24 hours) of dexamethasone action.

Among the many protein modifications that occur as intermediate steps in transcriptional activation by nuclear receptors, histone acetylation is not only a key event for nuclear receptor function but is reversible, acutely regulated and may be specifically modulated by drugs that have negligible toxicity (3638). All of the GR coactivators that were shown in this study to synergize with dexamethasone to enhance FR-α promoter activity, i.e., SRC-1 (NcoA-1), SRC-2 (GRIP1/TIF2/NcoA-2), and pCAF directly or indirectly promote histone acetylation (39). Nuclear receptor corepressors generate a transcriptionally repressed state by recruiting class II HDACs (4043). Indeed, short chain fatty acids have been shown to sensitize cells to steroid hormones in vitro and in vivo both by activating the mitogen-activated protein kinase pathway and by inhibiting HDAC (44). Therefore, it was logical to test the effect of HDAC inhibitors on dexamethasone induction of FR-α. The short chain fatty acid, valproic acid (an antiepileptic and antineoplastic drug), and the hydroxamic acid, trichostatin A (a well characterized laboratory reagent), are both inhibitors of class I and II HDACs (36, 45). Valproic acid and trichostatin A both potentiated dexamethasone induction of FR-α at the transcriptional level, in their pharmacologically effective millimolar and nanomolar concentration ranges, respectively (36, 45). Despite the broad role of histone acetylation in gene regulation, HDAC inhibitors do not have global effects on gene expression. It has been established that these inhibitors alter the expression of only ∼2% of actively transcribed genes and that most of them have either minimal toxicity or no toxicity at their effective doses (36). The profound effects of HDAC inhibitors on dexamethasone induction of FR-α may be used to advantage in the receptor-targeted therapies and FR-dependent diagnostics in view of the fact that a variety of HDAC inhibitors that have acceptable toxicity profiles, which produce a sustained increase in the level of acetylated histones within hours (45), and that act on HDACs functionally associated with nuclear receptors are currently available (36, 4648).

The results also show that the ultimate downstream site of action of dexamethasone in the FR-α gene is the proximal P4 promoter, more specifically, the G/C-rich Sp1 elements and the initiator and flanking region. Devoid of other regulatory elements, this portion of the FR-α promoter region represents the essential elements of a basal TATA-less promoter. The results of this study showed that a (Sp1 binding) G/C-rich region is essential for the dexamethasone response. The magnitude of this response not only correlated with the strength of the Sp1 element but also depended on the context of the initiator region. A common mechanism of gene regulation through G/C-rich elements involves differential levels and effects of Sp family proteins (49), but such a mechanism for the action of dexamethasone was ruled out in HeLa cells because the major Sp family proteins, Sp1, Sp3, and Sp4 all regulated the FR-α promoter in a similar manner and further, dexamethasone did not alter their expression levels or their apparent molecular weights that are influenced by their phosphorylation state. It thus seems that the ultimate action of dexamethasone in relation to FR-α regulation involves modulation of some component(s) of the transcription initiation complex, whose exact composition is dictated by both the G/C-rich and initiator regions. Such a supposition is reasonable in light of the emerging view that the recruitment of several components of the preinitiation complex is dictated by the basal promoter context (50).

The FR-α gene thus belongs to a class of indirect target genes of dexamethasone/GR that lack a GRE. Based on the foregoing studies, it is reasonable to anticipate that in individuals bearing an FR-α–positive tumor, brief treatment with the combination of innocuous doses of a GR agonist such as dexamethasone and a HDAC inhibitor will boost FR-α expression in the tumor and will, in addition, increase the serum level of FR. Such a treatment should greatly improve the outcome of FR-targeted therapies. GR modulation of FR-α expression also offers a means of using circulating FR as a serum marker to detect and follow the treatment response of major subtypes of ovarian, endometrial and other cancers. Indeed, elevation of serum FR in response to dexamethasone/HDAC inhibitor treatment may by itself serve as a diagnostic test for the presence of a FR-positive tumor. Following detection of the induced FR-α in the serum during initial screening, the FR-positive tumors may be detected by FR-targeted whole-body imaging. Similar principles of using general transcription modulators to induce or even down-regulate gene expression, selectively in tumors, concomitant with the administration of therapeutic agents whose action is dependent on the expression level of those genes, may be of value as a general concept in combination therapies. Such an approach is also applicable to the discovery of new tumor/serum markers.

Note: T. Tran, A. Shatnamwi, and X. Zheng contributed equally to this work.

The findings in this report are covered by pending patents.

Grant support: NIH grants CA 80183 and CA 103964 (to M. Ratnam), and NIH Institutional Pre-Doctoral NRSA grant CA 79450 (to T. Tran).

The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked advertisement in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

1
Elnakat H, Ratnam M. Distribution, functionality and gene regulation of folate receptor isoforms: implications in targeted therapy.
Adv Drug Deliv Rev
2004
;
56
:
1067
–84.
2
Forster MD, Raynaud F, Wood N, et al. CB300945, a new α-folate receptor targeted thymidylate synthase inhibitor [abstract #4618]. Proc Am Assoc
Cancer Res
2004
;
45
:
1066
.
3
Pan XQ, Zheng X, Shi G, Wang H, Ratnam M, Lee RJ. Strategy for the treatment of acute myelogenous leukemia based on folate receptor β-targeted liposomal doxorubicin combined with receptor induction using all-trans retinoic acid.
Blood
2002
;
100
:
594
–602.
4
Ross JF, Wang H, Behm FG, et al. Folate receptor type β is a neutrophilic lineage marker and is differentially expressed in myeloid leukemia.
Cancer
1999
;
85
:
348
–57.
5
Shen F, Ross JF, Wang X, Ratnam M. Identification of a novel folate receptor, a truncated receptor, and receptor type β in hematopoietic cells: cDNA cloning, expression, immunoreactivity, and tissue specificity.
Biochemistry
1994
;
33
:
1209
–15.
6
Shen F, Wu M, Ross JF, Miller D, Ratnam M. Folate receptor type γ is primarily a secretory protein due to lack of an efficient signal for glycosylphosphatidylinositol modification: protein characterization and cell type specificity.
Biochemistry
1995
;
34
:
5660
–5.
7
Kamen BA, Smith AK. A review of folate receptor α cycling and 5-methyltetrahydrofolate accumulation with an emphasis on cell models in vitro.
Adv Drug Deliv Rev
2004
;
56
:
1085
–97.
8
Lu Y, Low PS. Folate-mediated delivery of macromolecular anticancer therapeutic agents.
Adv Drug Deliv Rev
2002
;
54
:
675
–93.
9
Theti DS, Bavetsias V, Skelton LA, et al. Selective delivery of CB300638, a cyclopenta[g]quinazoline-based thymidylate synthase inhibitor into human tumor cell lines overexpressing the α-isoform of the folate receptor.
Cancer Res
2003
;
63
:
3612
–8.
10
Leamon CP, Low PS. Folate-mediated targeting: from diagnostics to drug and gene delivery.
Drug Discov Today
2001
;
6
:
44
–51.
11
Ratnam M, Hao H, Zheng X, et al. Receptor induction and targeted drug delivery: a new antileukaemia strategy.
Expert Opin Biol Ther
2003
;
3
:
563
–74.
12
Jakman AL, Theti DS, Gibbs DD. Antifolates targeted specifically to the folate receptor.
Adv Drug Deliv Rev
2004
;
56
:
1111
–25.
13
Leamon CP, Reddy JA. Folate-targeted chemotherapy.
Adv Drug Deliv Rev
2004
;
56
:
1127
–41.
14
Gabizon A, Shmeeda H, Horowitz AT, Zalipsky S. Tumor cell targeting of liposome-entrapped drugs with phospholipid-anchored folic acid-PEG conjugates.
Adv Drug Deliv Rev
2004
;
56
:
1177
–92.
15
Lu Y, Sega E, Leamon CP, Low PS. Folate receptor-targeted immunotherapy of cancer: mechanism and therapeutic potential.
Adv Drug Deliv Rev
2004
;
56
:
1161
–76.
16
Roy EJ, Gawlick U, Orr BA, Kranz DM. Folate-mediated targeting of T-cell to tumors.
Adv Drug Deliv Rev
2004
;
56
:
1219
–31.
17
Zhao XB, Lee RJ. Tumor-selective targeted delivery of genes and antisense oligodeoxyribonucleotides via the folate receptor.
Adv Drug Deliv Rev
2004
;
56
:
1193
–204.
18
Ke C-Y, Mathias CJ, Gerrn MA. Folate receptor-targeted radionuclide imaging agents.
Adv Drug Deliv Rev
2004
;
56
:
1143
–60.
19
Antony AC, Verma RS, Unune AR, LaRosa JA. Identification of a Mg2+-dependent protease in human placenta which cleaves hydrophobic folate-binding proteins to hydrophilic forms.
J Biol Chem
1989
;
264
:
1911
–4.
20
Elwood PC, Deutsch JC, Kolhouse JF. The conversion of the human membrane-associated folate binding protein (folate receptor) to the soluble folate binding protein by a membrane-associated metalloprotease.
J Biol Chem
1991
;
266
:
2346
–53.
21
Luhrs CA, Slomiany BL. A human membrane-associated folate binding protein is anchored by a glycosyl-phosphatidylinositol tail.
J Biol Chem
1989
;
264
:
21446
–9.
22
Yang XY, Mackins JY, Li QJ, Antony AC. Isolation and characterization of a folate receptor-directed metalloprotease from human placenta.
J Biol Chem
1996
;
271
:
11493
–9.
23
Mantovani LT, Miotti S, Menard S, et al. Folate binding protein distribution in normal tissues and biological fluids from ovarian carcinoma patients as detected by the monoclonal antibodies MOv18 and MOv19.
Eur J Cancer
1994
;
30A
:
363
–9.
24
Toffoli G, Cernigoi C, Russo A, Gallo A, Bagnoli M, Boiocchi M. Overexpression of folate binding protein in ovarian cancers.
Int J Cancer
1997
;
74
:
193
–8.
25
Wu M, Gunning W, Ratnam M. Expression of folate receptor type α in relation to cell type, malignancy, and differentiation in ovary, uterus, and cervix.
Cancer Epidemiol Biomarkers Prev
1999
;
8
:
775
–82.
26
Elwood PC, Nachmanoff K, Saikawa Y, et al. The divergent 5′ termini of the α human folate receptor (hFR) mRNAs originate from two tissue-specific promoters and alternative splicing: characterization of the α hFR gene structure.
Biochemistry
1997
;
36
:
1467
–78.
27
Saikawa Y, Price K, Hance KW, Chen TY, Elwood PC. Structural and functional analysis of the human KB cell folate receptor gene P4 promoter: cooperation of three clustered Sp1-binding sites with initiator region for basal promoter activity.
Biochemistry
1995
;
34
:
9951
–61.
28
Kelley KM, Rowan BG, Ratnam M. Modulation of the folate receptor α gene by the estrogen receptor: mechanism and implications in tumor targeting.
Cancer Res
2003
;
63
:
2820
–8.
29
Roberts SJ, Chung KN, Nachmanoff K, Elwood PC. Tissue-specific promoters of the α human folate receptor gene yield transcripts with divergent 5′ leader sequences and different translational efficiencies.
Biochem J
1997
;
326
:
439
–47.
30
Hao H, Qi H, Ratnam M. Modulation of the folate receptor type β gene by coordinate actions of retinoic acid receptors at activator Sp1/ets and repressor AP-1 sites.
Blood
2003
;
101
:
4551
–60.
31
Wang H, Zheng X, Behm FG, Ratnam M. Differentiation-independent retinoid induction of folate receptor type β, a potential tumor target in myeloid leukemia.
Blood
2000
;
96
:
3529
–36.
32
Adcock IM, Caramori G. Cross-talk between pro-inflammatory transcription factors and glucocorticoids.
Immunol Cell Biol
2001
;
79
:
376
–84.
33
DeRijk RH, Schaaf M, De Kloet ER. Glucocorticoid receptor variants: clinical implications.
J Steroid Biochem Mol Biol
2002
;
81
:
103
–22.
34
Boonyaratanakornkit V, Scott MP, Ribon V, et al. Progesterone receptor contains a proline-rich motif that directly interacts with SH3 domains and activates c-Src family tyrosine kinases.
Mol Cell
2001
;
8
:
269
–80.
35
Simoncini T, Hafezi-Moghadam A, Brazil DP, Ley K, Chin WW, Liao JK. Interaction of oestrogen receptor with the regulatory subunit of phosphatidylinositol-3-OH kinase.
Nature
2000
;
407
:
538
–41.
36
Marks PA, Richon VM, Rifkin RA. Histone deacetylase inhibitors: inducers of differentiation or apoptosis of transformed cells.
J Natl Cancer Inst
2000
;
92
:
1210
–6.
37
Sterner DE, Berger SL. Acetylation of histones and transcription-related factors.
Microbiol Mol Biol Rev
2000
;
64
:
435
–59.
38
Struhl K. Histone acetylation and transcriptional regulatory mechanisms.
Genes Dev
1998
;
12
:
599
–606.
39
Robyr D, Wolffe AP, Wahli W. Nuclear hormone receptor coregulators in action: diversity for shared tasks.
Mol Endocrinol
2000
;
14
:
329
–47.
40
Kao HY, Downes M, Ordentlich P, Evans RM. Isolation of a novel histone deacetylase reveals that class I and class II deacetylases promote SMRT-mediated repression.
Genes Dev
2000
;
14
:
55
–66.
41
Li J, Wang J, Wang J, et al. Both corepressor proteins SMRT and N-CoR exist in large protein complexes containing HDAC3.
EMBO J
2000
;
19
:
4342
–50.
42
Underhill C, Qutob MS, Yee SP, Torchia J. A novel nuclear receptor corepressor complex, N-CoR, contains components of the mammalian SWI/SNF complex and the corepressor KAP-1.
J Biol Chem
2000
;
275
:
40463
–70.
43
Wen YD, Perissi V, Staszewski LM, et al. The histone deacetylase-3 complex contains nuclear receptor corepressors.
Proc Natl Acad Sci U S A
2000
;
97
:
7202
–7.
44
Jansen MS, Nagel SC, Miranda PJ, Lobenhofer EK, Afshari CA, McDonnell DP. Short-chain fatty acids enhance nuclear receptor activity through mitogen-activated protein kinase activation and histone deacetylase inhibition.
Proc Natl Acad Sci U S A
2004
;
101
:
7199
–204.
45
Marks PA, Richon VM, Breslow R, Rifkind RA. Histone deacetylase inhibitors as new cancer drugs.
Curr Opin Oncol
2001
;
13
:
477
–83.
46
Gottlicher M, Minucci S, Zhu P, et al. Valproic acid defines a novel class of HDAC inhibitors inducing differentiation of transformed cells.
EMBO J
2001
;
20
:
6969
–78.
47
Jung M. Inhibitors of histone deacetylase as new anticancer agents.
Curr Med Chem
2001
;
8
:
1505
–11.
48
Kramer OH, Gottlicher M, Heinzel T. Histone deacetylase as a therapeutic target.
Trends Endocrinol Metab
2001
;
12
:
294
–300.
49
Apt D, Watts RM, Suske G, Bernard HU. High Sp1/Sp3 ratios in epithelial cells during epithelial differentiation and cellular transformation correlate with the activation of the HPV-16 promoter.
Virology
1996
;
224
:
281
–91.
50
Smale ST, Kadonaga JT. The RNA polymerase II core promoter.
Annu Rev Biochem
2003
;
72
:
449
–79.