Glucocorticoids are keystone drugs in the treatment of childhood acute lymphoblastic leukemia (ALL). To get more insight in signal transduction pathways involved in glucocorticoid-induced apoptosis, Affymetrix U133A GeneChips were used to identify transcriptionally regulated genes on 3 and 8 hours of prednisolone exposure in leukemic cells of 13 children as compared with nonexposed cells. Following 3 hours of exposure no significant changes in gene expression could be identified. Following 8 hours of exposure, 51 genes were differentially expressed (P < .001 and false discovery rate < 10%) with 39 genes being up-regulated (median, 2.4-fold) and 12 genes were down-regulated (median, 1.7-fold). Twenty-one of those genes have not been identified before to be transcriptionally regulated by prednisolone. Two of the 3 most highly up-regulated genes were tumor suppressor genes, that is, thioredoxin-interacting protein (TXNIP; 3.7-fold) and zinc finger and BTB domain containing 16 (ZBTB16; 8.8-fold). About 50% of the differentially expressed genes were functionally categorized in 3 major routes, namely MAPK pathways (9 genes), NF-κB signaling (11 genes), and carbohydrate metabolism (5 genes). Biologic characterization of these genes and pathways might elucidate the action of glucocorticoids in ALL cells, possibly suggesting causes of glucocorticoid resistance and new potential targets for therapy.

Glucocorticoids such as prednisone and dexamethasone have been used extensively in the treatment of childhood acute lymphoblastic leukemia (ALL) for many years. The in vivo and in vitro prednisone response, as determined with a tetrazolium-based (MTT) toxicity assay, have been shown to correlate with each other and long-term clinical outcome in children with ALL.1–3 

The classically proposed way of glucocorticoid action is that glucocorticoids bind to the intracellular glucocorticoid receptor (GR). The glucocorticoid-GR complex then translocates to the nucleus, where it binds to glucocorticoid-responsive elements (GREs), resulting in the transcriptional activation of glucocorticoid-responsive genes.4  Alternatively, the glucocorticoid-GR complex can directly bind to transcription factors such as nuclear factor-κB (NF-κB) or activator protein-1 (AP-1), resulting in so-called transrepression complexes. These complexes disrupt the transcriptional regulation of genes that are normally affected by these transcription factors.5  Depending on the cell type, transcriptional activation and repression result in immunosuppression, stress response, or induction of apoptosis.

Interestingly, glucocorticoids only induce apoptosis in lymphoid cells such as ALL, multiple myelomas, malignant lymphomas, and thymocytes, not in other tissues. Despite the major impact of glucocorticoid resistance on clinical outcome, knowledge about the signal transduction pathways leading to glucocorticoid-induced apoptosis in ALL cells is limited.6,7  Several microarray studies have been performed to determine glucocorticoid-regulated genes in leukemia.8–12  However, the cell lines (ie, immortalized cells) used in these studies do not represent an ideal model to study mechanisms involved in survival and apoptosis of primary ALL cells. Recently, the first microarray study was published in which, besides cell lines, in vivo prednisone-exposed patient ALL samples were analyzed.13  In the present study, we used freshly isolated leukemic cells of pediatric patients with ALL at initial diagnosis to identify which genes are transcriptionally regulated on in vitro prednisolone exposure.

The study has been approved by the medical ethical committee of the Erasmus Medical Center Rotterdam, The Netherlands.

Patients

The study population consisted of 13 patients diagnosed with precursor-B ALL or T-ALL. Pretreatment bone marrow or peripheral blood samples were obtained after written informed consent, in accordance with the Declaration of Helsinki, from the patients and/or their legal guardians. The mononuclear cell fraction was separated by Lymphoprep density gradient centrifugation (density, 1.077 g/mL; Nycomed Pharma, Oslo, Norway). When necessary, nonleukemic cells were depleted by immunomagnetic beads to purify the samples to more than 90% of leukemic cells.

Prednisolone exposure

Leukemic cells were incubated in RPMI 1640 medium (Dutch modification without l-glutamine) supplemented with 2 mM l-glutamine, 5 μg/mL insulin, 5 μg/mL transferrin, 5 ng/mL sodium selenite, 20% heat-inactivated fetal calf serum, 100 IU/mL penicillin, 100 μg/mL streptomycin, 0.125 μg/mL fungizone, 200 μg/mL gentamycin with and without 250 μg/mL prednisolone. After 3 and 8 hours of incubation, 20 × 106 cells were removed from the culture for RNA isolation.

RNA extraction, labeling, and hybridization

Total RNA was extracted using the Trizol method (Gibco BRL, Life Technologies, Breda, The Netherlands) according to the protocol provided by the manufacturer with minor modifications.14  RNA integrity was determined using the Agilent 2100 Bioanalyzer (Agilent, Palo Alto, CA). RNA (5-15 μg) was used for subsequent production of biotinylated antisense cRNA, as described before.15  Samples with less than 10 μg labeled cRNA were excluded. Labeled cRNA was hybridized to the U133A GeneChip oligonucleotide microarray (Affymetrix, Santa Clara, CA) according to the protocol provided by the manufacturer.

Statistics

Raw gene expression values were calculated using Affymetrix Microarray Suite version 5.0.16  Data were normalized using the variance stabilization procedure (VSN) as proposed by Huber et al.17  Data distribution of unexposed controls suggested a γ- or log-normal distribution. Therefore, a generalized linear model with γ error distribution and identity link function was used to describe the effect of exposure (prednisolone exposed or nonexposed) on gene expression levels. This model was fitted to each gene separately, and the effect of exposure was evaluated via the corresponding ANOVA P values. Differentially expressed genes with respect to exposure were selected by controlling the false discovery rate (FDR; by ANOVA for each gene separately), using the procedure by Benjamini and Hochberg.18  A P value less than .001 and a FDR (adapted for multiple testing) less than 10% was considered statistically significant. Normalization and subsequent analysis were performed using R 1.9.1,19  also making use of the Bioconductor packages VSN and Multtest (www.Bioconductor.org). The fold up- or down-regulation was calculated using the formula: e (vsn value pred sample − vsn value control sample).

The TELiS database20,21  was used to search for transcription factor binding motifs (TFBMs) which were overrepresented or underrepresented in the genes differentially expressed on prednisolone exposure, using a promoter size of −1000 to + 200 and a stringency setting of 0.9.

Diagnostic samples of 13 patients with ALL were exposed to prednisolone for 3 and 8 hours. Prednisolone exposed and nonexposed (control sample, culture medium only) leukemic cells of the same patient were analyzed pairwise per patient to correct for the effect of culture in time. Paired samples could be successfully analyzed for 9 of 13 patients at 3 hours and for 10 of 13 patients at 8 hours of exposure time.

After 3 hours of prednisolone exposure, no differentially expressed probe sets were identified at P less than .001 and FDR less than 10%. However, 2 genes were differentially expressed with P less than .005 and FDR less than 20%, that is, ZFP36L2 (zinc finger protein 36, C3H type-like 2) and DSIPI (delta sleep-inducing peptide, immunoreactor; alias GILZ). Eight hours of prednisolone exposure revealed differential expression of 57 probe sets (51 unique genes and 2 expressed sequence tags [ESTs]) at P less than .001 and FDR less than 10%. As shown in Table 1, 44 probe sets (39 genes) were up-regulated (median, 2.4-fold; 25th-75th percentile, 1.8- to 3.1-fold) and 13 probe sets (12 genes) were down-regulated (median, 1.7-fold; 25th-75th percentile, 1.5- to 2-fold). Analysis at lower significance level (P < .001 and FDR < 20%) revealed that 144 probe sets were differentially expressed. In general, we observed that the same direction of prednisolone-induced changes in gene expression (down- or up-regulation) was found in the majority of patients, in all 570 observations (57 probe sets, 10 patient samples) only 6 observations (1%) were opposite from the effect seen in the remaining cases (see Table 1).

Table 1

Prednisolone-induced changes in gene expression in pediatric ALL

Probe IDAccession no.Gene name*DescriptionNo. of patient samplesFold changeP
Up-regulated genes 
    204560_at NM_004117 FKBP5 FK506-binding protein 5 10 35.4 < .001 
    205883_at NM_006006 ZBTB16 (PLZF) Zinc finger and BTB domain containing 16 10 8.8 < .001 
    201008_s_at AI439556 TXNIP (VDUP) Thioredoxin-interacting protein 10 4.4 < .001 
    201009_s_at AA812232 TXNIP (VDUP) Thioredoxin-interacting protein 10 3.0 < .001 
    221756_at AL540260 LIMK2 LIM domain kinase 2 10 4.1 < .001 
    212158_at AL577322 SDC2 Syndecan 2 10 4.0 < .001 
    204698_at U88964 ISG20 Interferon-stimulated exonuclease gene 20 kDa 10 3.7 < .001 
    33304_at NM_002201 ISG20 Interferon-stimulated exonuclease gene 20 kDa 10 3.4 < .001 
    201369_s_at U07802 ZFP36L2 (ERF2) Zinc finger protein 36, C3H type-like 2 10 3.6 < .001 
    201368_at NM_006887 ZFP36L2 (ERF2) Zinc finger protein 36, C3H type-like 2 10 2.6 < .001 
    208078_s_at NMP_030751 SNF1LK SNF1-like kinase 10 3.5 < .001 
    208763_s_at AL110191 DSIPI (TSC22D3, GILZ) Delta sleep-inducing peptide, immunoreactor 10 3.3 < .001 
    202670_at AI571419 MAP2K1 (MEK1) Mitogen-activated protein kinase kinase 1 10 3.1 < .001 
    203542_s_at NM_001206 KLF9 (BTEB1) Kruppel-like factor 9 10 3.1 < .001 
    203543_s_at AI690205 KLF9 (BTEB1) Kruppel-like factor 9 10 3.1 < .001 
    203574_at NM_005384 NFIL3 Nuclear factor, interleukin 3 regulated 10 3.0 < .001 
    209185_s_at AF073310 IRS2 Insulin receptor substrate 2 10 2.8 < .001 
    215890_at X61094 GM2A GM2 ganglioside activator 10 2.6 < .001 
    203973_s_at NM_005195 CEBPD CCAAT/enhancer-binding protein (C/EBP), delta 10 2.5 < .001 
    213792_s_at AA485908 INSR Insulin receptor 2.5 < .001 
    212242_at AL565074 TUBA1 Tubulin, alpha 1 (testis specific) 10 2.4 < .001 
    201041_s_at NM_004417 DUSP1 (MKP1) Dual-specificity phosphatase 1 10 2.4 < .001 
    212188_at AI718937 KCTD12 Potassium channel tetramerization domain containing 12 10 2.3 < .001 
    212192_at AA551075 KCTD12 Potassium channel tetramerization domain containing 12 2.3 < .001 
    218638_s_at NM_012445 SPON2 Spondin 2 10 2.3 < .001 
    207996_s_at NM_004338 C18orf1 Chromosome 18 open reading frame 1 10 2.2 < .001 
    204618_s_at NM_005254 GABPB2 GA-binding protein transcription factor, beta subunit 2 2.2 < .001 
    210001_s_at AB005043 SOCS1 Suppressor of cytokine signaling 1 10 2.2 < .001 
    200921_s_at NM_001731 BTG1 B-cell translocation gene 1, antiproliferative 10 2.2 < .001 
    202643_s_at AI738896 TNFAIP3 (A20) Tumor necrosis factor, alpha-induced protein 3 10 2.1 < .001 
    201037_at NM_002627 PFKP Phosphofructokinase, platelet 10 2.0 < .001 
    207945_s_at NM_001893 CSNK1D Casein kinase 1, delta 10 1.8 < .001 
    201739_at NM_005627 SGK Serum/glucocorticoid-regulated kinase 10 1.8 < .001 
    203819_s_at AU160004 IMP-3 IGF-II mRNA-binding protein 3 1.7 < .001 
    221563_at N36770 DUSP10 (MKP5) Dual-specificity phosphatase 10 1.5 < .001 
    215977_x_at X68285 GK Glycerol kinase 10 1.5 < .001 
    213310_at AI613483 EIF2C2 Eukaryotic translation initiation factor 2C, 2 10 1.4 < .001 
    215046_at AL133053 FLJ23861 EST 10 1.4 < .001 
    217356_s_at S81916 PGK1 Phosphoglycerate kinase 1 10 1.4 < .001 
    211926_s_at AI827941 MYH9 Myosin, heavy polypeptide 9, nonmuscle 10 1.3 < .001 
    218761_at NM_017610 RNF111 (ARK) Ring finger protein 111 10 1.3 < .001 
    217795_s_at W74580 THEM43 Transmembrane protein 43 10 1.3 < .001 
    201859_at NM_002727 PRG1 Proteoglycan 1 10 1.3 < .001 
    218528_s_at NM_022781 RNF38 Ring finger protein 38 10 1.3 < .001 
Down-regulated genes 
    205749_at NM_000499 CYP1A1 Cytochrome P450, family 1, subfamily A, polypeptide 1 10 −2.0 < .001 
    209969_s_at BC002704 STAT1 Signal transducer and activator of transcription 1 10 −2.0 < .001 
    219066_at NM_021823 MDS018 EST 10 −2.0 < .001 
    205013_s_at NM_000675 ADORA2A Adenosine A2a receptor 10 −2.0 < .001 
    205006_s_at NM_004808 NMT2 N-myristoyltransferase 2 10 −1.7 < .001 
    203612_at NM_004053 BYSL Bystin-like 10 −1.7 < .001 
    219665_at NM_024815 NUDT18 Nudix-type motif 18 10 −1.7 < .001 
    221933_at AI338338 NLGN4X Neuroligin 4, X-linked −1.7 < .001 
    204070_at NM_004585 RARRES3 Retinoic acid receptor responder (tazarotene induced) 3 10 −1.7 < .001 
    211430_s_at M87789 IGHG1 Immunoglobulin heavy constant gamma 1 10 −1.4 < .001 
    218046_s_at NM_016065 MRPS16 Mitochondrial ribosomal protein S16 10 −1.4 < .001 
    219344_at NM_018344 SLC29A3 (ENT3) Solute carrier family 29, member 3 10 −1.4 < .001 
    203814_s_at NM_000904 NQO2 NAD(P)H dehydrogenase, quinone 2 10 −1.4 < .001 
Probe IDAccession no.Gene name*DescriptionNo. of patient samplesFold changeP
Up-regulated genes 
    204560_at NM_004117 FKBP5 FK506-binding protein 5 10 35.4 < .001 
    205883_at NM_006006 ZBTB16 (PLZF) Zinc finger and BTB domain containing 16 10 8.8 < .001 
    201008_s_at AI439556 TXNIP (VDUP) Thioredoxin-interacting protein 10 4.4 < .001 
    201009_s_at AA812232 TXNIP (VDUP) Thioredoxin-interacting protein 10 3.0 < .001 
    221756_at AL540260 LIMK2 LIM domain kinase 2 10 4.1 < .001 
    212158_at AL577322 SDC2 Syndecan 2 10 4.0 < .001 
    204698_at U88964 ISG20 Interferon-stimulated exonuclease gene 20 kDa 10 3.7 < .001 
    33304_at NM_002201 ISG20 Interferon-stimulated exonuclease gene 20 kDa 10 3.4 < .001 
    201369_s_at U07802 ZFP36L2 (ERF2) Zinc finger protein 36, C3H type-like 2 10 3.6 < .001 
    201368_at NM_006887 ZFP36L2 (ERF2) Zinc finger protein 36, C3H type-like 2 10 2.6 < .001 
    208078_s_at NMP_030751 SNF1LK SNF1-like kinase 10 3.5 < .001 
    208763_s_at AL110191 DSIPI (TSC22D3, GILZ) Delta sleep-inducing peptide, immunoreactor 10 3.3 < .001 
    202670_at AI571419 MAP2K1 (MEK1) Mitogen-activated protein kinase kinase 1 10 3.1 < .001 
    203542_s_at NM_001206 KLF9 (BTEB1) Kruppel-like factor 9 10 3.1 < .001 
    203543_s_at AI690205 KLF9 (BTEB1) Kruppel-like factor 9 10 3.1 < .001 
    203574_at NM_005384 NFIL3 Nuclear factor, interleukin 3 regulated 10 3.0 < .001 
    209185_s_at AF073310 IRS2 Insulin receptor substrate 2 10 2.8 < .001 
    215890_at X61094 GM2A GM2 ganglioside activator 10 2.6 < .001 
    203973_s_at NM_005195 CEBPD CCAAT/enhancer-binding protein (C/EBP), delta 10 2.5 < .001 
    213792_s_at AA485908 INSR Insulin receptor 2.5 < .001 
    212242_at AL565074 TUBA1 Tubulin, alpha 1 (testis specific) 10 2.4 < .001 
    201041_s_at NM_004417 DUSP1 (MKP1) Dual-specificity phosphatase 1 10 2.4 < .001 
    212188_at AI718937 KCTD12 Potassium channel tetramerization domain containing 12 10 2.3 < .001 
    212192_at AA551075 KCTD12 Potassium channel tetramerization domain containing 12 2.3 < .001 
    218638_s_at NM_012445 SPON2 Spondin 2 10 2.3 < .001 
    207996_s_at NM_004338 C18orf1 Chromosome 18 open reading frame 1 10 2.2 < .001 
    204618_s_at NM_005254 GABPB2 GA-binding protein transcription factor, beta subunit 2 2.2 < .001 
    210001_s_at AB005043 SOCS1 Suppressor of cytokine signaling 1 10 2.2 < .001 
    200921_s_at NM_001731 BTG1 B-cell translocation gene 1, antiproliferative 10 2.2 < .001 
    202643_s_at AI738896 TNFAIP3 (A20) Tumor necrosis factor, alpha-induced protein 3 10 2.1 < .001 
    201037_at NM_002627 PFKP Phosphofructokinase, platelet 10 2.0 < .001 
    207945_s_at NM_001893 CSNK1D Casein kinase 1, delta 10 1.8 < .001 
    201739_at NM_005627 SGK Serum/glucocorticoid-regulated kinase 10 1.8 < .001 
    203819_s_at AU160004 IMP-3 IGF-II mRNA-binding protein 3 1.7 < .001 
    221563_at N36770 DUSP10 (MKP5) Dual-specificity phosphatase 10 1.5 < .001 
    215977_x_at X68285 GK Glycerol kinase 10 1.5 < .001 
    213310_at AI613483 EIF2C2 Eukaryotic translation initiation factor 2C, 2 10 1.4 < .001 
    215046_at AL133053 FLJ23861 EST 10 1.4 < .001 
    217356_s_at S81916 PGK1 Phosphoglycerate kinase 1 10 1.4 < .001 
    211926_s_at AI827941 MYH9 Myosin, heavy polypeptide 9, nonmuscle 10 1.3 < .001 
    218761_at NM_017610 RNF111 (ARK) Ring finger protein 111 10 1.3 < .001 
    217795_s_at W74580 THEM43 Transmembrane protein 43 10 1.3 < .001 
    201859_at NM_002727 PRG1 Proteoglycan 1 10 1.3 < .001 
    218528_s_at NM_022781 RNF38 Ring finger protein 38 10 1.3 < .001 
Down-regulated genes 
    205749_at NM_000499 CYP1A1 Cytochrome P450, family 1, subfamily A, polypeptide 1 10 −2.0 < .001 
    209969_s_at BC002704 STAT1 Signal transducer and activator of transcription 1 10 −2.0 < .001 
    219066_at NM_021823 MDS018 EST 10 −2.0 < .001 
    205013_s_at NM_000675 ADORA2A Adenosine A2a receptor 10 −2.0 < .001 
    205006_s_at NM_004808 NMT2 N-myristoyltransferase 2 10 −1.7 < .001 
    203612_at NM_004053 BYSL Bystin-like 10 −1.7 < .001 
    219665_at NM_024815 NUDT18 Nudix-type motif 18 10 −1.7 < .001 
    221933_at AI338338 NLGN4X Neuroligin 4, X-linked −1.7 < .001 
    204070_at NM_004585 RARRES3 Retinoic acid receptor responder (tazarotene induced) 3 10 −1.7 < .001 
    211430_s_at M87789 IGHG1 Immunoglobulin heavy constant gamma 1 10 −1.4 < .001 
    218046_s_at NM_016065 MRPS16 Mitochondrial ribosomal protein S16 10 −1.4 < .001 
    219344_at NM_018344 SLC29A3 (ENT3) Solute carrier family 29, member 3 10 −1.4 < .001 
    203814_s_at NM_000904 NQO2 NAD(P)H dehydrogenase, quinone 2 10 −1.4 < .001 

After 8 hours of prednisolone exposure, 57 probe sets (51 unique genes and 2 ESTs) were differentially expressed at P < .001 and FDR < 10%. Aliases between brackets.

*

Human genome nomenclature.

Number of up-regulated patient samples for up-regulated genes and number of down-regulated patient samples for down-regulated gene.

The median fold change reflects the change in expression of genes after 8 hours of prednisolone exposure compared with culture medium exposed control cells.

As shown in Figure 1A, the top 10 probe sets that were found to be significantly up-regulated at P less than .005 and FDR less than 10% after 8 hours of prednisolone exposure were also up-regulated after 3 hours of prednisolone exposure, albeit at a lower significance (P < .05). Seven of the 10 most significantly down-regulated genes at 8 hours of prednisolone exposure were also down-regulated after 3 hours of prednisolone exposure but with a much lower significance (P < .35) (Figure 1B).

Figure 1

Comparison between 3 and 8 hours on prednisolone-induced changes in gene expression in pediatric ALL. The fold up-regulation (A) and down-regulation (B) of the top 10 probe sets that were affected after 8 hours of prednisolone exposure are shown at both 3 (□) and 8 (▪) hours of exposure. The median fold change is given.

Figure 1

Comparison between 3 and 8 hours on prednisolone-induced changes in gene expression in pediatric ALL. The fold up-regulation (A) and down-regulation (B) of the top 10 probe sets that were affected after 8 hours of prednisolone exposure are shown at both 3 (□) and 8 (▪) hours of exposure. The median fold change is given.

Close modal

Table 2 shows a summary of literature on previously identified prednisolone-responsive genes in leukemic cell lines (23 genes) and nonleukemic cells (7 genes) that were also found in the present study to be affected after 8 hours of prednisolone exposure in primary ALL cells. Five genes were found to be up-regulated in our study but down-regulated in array studies that included leukemic cell lines DUSP1 (previously reported to be up-regulated in mast cells and fibroblasts), DUSP10, PGK1, EIF2C2, and PFKP. CYP1A1 was found to be the most prominently down-regulated gene in primary ALL cells. In contrast, this gene was found to be up-regulated in human aorta endothelial cells.

Table 2

Genes differentially expressed on 8 hours of prednisolone exposure in this study and previously reported in the literature

GeneUp- or down-regulation
Cell lines studiedReferences
Child ALL*Lymphoid cell linesNonleukemic cell lines
ZFP36L2 Up Up — 697 Yoshida et al11  
FKBP5 Up Up — 697, Jurkat, CEM Obexer et al,9  Planey et al,10  Yoshida et al11  
LIMK2 Up Up — CEM Medh et al,40  Webb et al45  
NFIL3 Up Up — CEM Medh et al,40  Webb et al45  
ISG20 Up Up — 697 Yoshida et al11  
TXNIP Up Up — 697, CEM Tonko et al,8  Planey et al,10  Medh et al,40  Webb et al45  
C18ORF1 Up Up — CEM Tonko et al,8  Webb et al45  
BTEB1 Up Up — 697 Planey et al10  
DUSP1 (MKP-1) Up Down§ — 697 Planey et al10  
DSIP1 (GILZ) Up Up — 697, CEM Tonko et al,8  Planey et al10  
BTG1 Up Up — WEHI7.2, S49.A2, CEM, 697 Yoshida et al,11  Wang et al,12  Medh et al,40  Webb et al45  
ZBTB16 Up Up — CEM Tonko et al8  
SNF1LK Up Up — 697 Yoshida et al11  
TUBA1 Up Up — CEM Medh et al,40  Webb et al45  
DUSP10 (MKP-5) Up Down — 697 Planey et al10  
PGKI Up Down — CEM Tonko et al8  
SGK Up Up — WEHI7.2, S49.A2 Wang et al12  
PRG1 Up Up — CEM, 697 Planey et al,10  Medh et al,40  Webb et al45  
EIF2C2 Up Down — CEM Medh et al,39  Leong et al44  
PFKP Up Down — WEHI7.2, S49.A2 Wang et al12  
BYSL Down Down — CEM Tsai et al4  
SOCS1 Up Up — CEM, 697 Tonko et al,8  Yoshida et al,11  Medh et al,40  Webb et al45  
MAP2K1 Up — Up Human bone marrow stromal cells (TM5) Jeon et al46  
SDC2 Up — Up Human glomerular epithelial cells Kasinath et al47  
CEBPD Up — Up Different rat and rabbit cell types Ramji and Foka48  
GK Up — Up Rat adipocytes Taylor et al49  
STAT1 Down — Down Human peripheral blood mononuclear cells Hu et al50  
INSR Up — Up Human promonocytic cells Leal et al51  
CYP1A1 Down — Up Human aorta endothelial cells Celander et al52  
GeneUp- or down-regulation
Cell lines studiedReferences
Child ALL*Lymphoid cell linesNonleukemic cell lines
ZFP36L2 Up Up — 697 Yoshida et al11  
FKBP5 Up Up — 697, Jurkat, CEM Obexer et al,9  Planey et al,10  Yoshida et al11  
LIMK2 Up Up — CEM Medh et al,40  Webb et al45  
NFIL3 Up Up — CEM Medh et al,40  Webb et al45  
ISG20 Up Up — 697 Yoshida et al11  
TXNIP Up Up — 697, CEM Tonko et al,8  Planey et al,10  Medh et al,40  Webb et al45  
C18ORF1 Up Up — CEM Tonko et al,8  Webb et al45  
BTEB1 Up Up — 697 Planey et al10  
DUSP1 (MKP-1) Up Down§ — 697 Planey et al10  
DSIP1 (GILZ) Up Up — 697, CEM Tonko et al,8  Planey et al10  
BTG1 Up Up — WEHI7.2, S49.A2, CEM, 697 Yoshida et al,11  Wang et al,12  Medh et al,40  Webb et al45  
ZBTB16 Up Up — CEM Tonko et al8  
SNF1LK Up Up — 697 Yoshida et al11  
TUBA1 Up Up — CEM Medh et al,40  Webb et al45  
DUSP10 (MKP-5) Up Down — 697 Planey et al10  
PGKI Up Down — CEM Tonko et al8  
SGK Up Up — WEHI7.2, S49.A2 Wang et al12  
PRG1 Up Up — CEM, 697 Planey et al,10  Medh et al,40  Webb et al45  
EIF2C2 Up Down — CEM Medh et al,39  Leong et al44  
PFKP Up Down — WEHI7.2, S49.A2 Wang et al12  
BYSL Down Down — CEM Tsai et al4  
SOCS1 Up Up — CEM, 697 Tonko et al,8  Yoshida et al,11  Medh et al,40  Webb et al45  
MAP2K1 Up — Up Human bone marrow stromal cells (TM5) Jeon et al46  
SDC2 Up — Up Human glomerular epithelial cells Kasinath et al47  
CEBPD Up — Up Different rat and rabbit cell types Ramji and Foka48  
GK Up — Up Rat adipocytes Taylor et al49  
STAT1 Down — Down Human peripheral blood mononuclear cells Hu et al50  
INSR Up — Up Human promonocytic cells Leal et al51  
CYP1A1 Down — Up Human aorta endothelial cells Celander et al52  

— indicates not applicable.

*

From this study.

From the literature.

697 indicates human pre-B leukemia cell line; Jurkat, T-lineage leukemic cell line; CEM, T-lineage leukemic cell line; WEH17.2, T-cell lymphoma cell line; S49A2, T-cell lymphoma cell line.

§

Up-regulation reported in nonleukemic cell lines.53 

Forty of the 51 genes (57 probe sets) differentially expressed on prednisolone exposure had an annotation in the Gene Ontology database. The representation of these 40 probe sets in each of the functional categories did not statistically differ from the total of U133A Gene Chip probe sets that are annotated by Gene Ontology. Besides analysis through the Gene Ontology database,22  we also analyzed the literature for putative functions of the 51 genes (Table 3). Among the 51 prednisolone responsive genes, 2 of the 3 most highly up-regulated genes are the putative tumor suppressor genes TXNIP (thioredoxin-interacting protein, alias VDUP1; 3.7-fold) and ZBTB16 (zinc finger and BTB domain containing 16, alias PLZF; 8.8-fold). Besides these 2 tumor suppressor genes involved in cell-cycle regulation, approximately 50% of the prednisolone-responsive genes could be assigned to 3 major pathways, the MAPK pathways (9 genes), NF-κB signaling pathways of gene transcription (11 genes), and carbohydrate metabolism (5 genes).

Table 3

Biological functions of prednisolone-responsive genes in pediatric ALL

FunctionGenes
Proapoptotic ADORA2A,*NFIL3, ZBTB16, STAT1,*ISG20, TXN1P, BTG1, RARRES3,*SOCS1, DSIP1 (GILZ) 
Proliferation ADORA2A,*NFIL3, ZBTB16, ZFP36L2, MAP2K1, CEBPD, DUSP1 (MKP-1), NMT2,*RNF111, DUSP10 (MKP-5), TNFAIP3, SGK, IMP-3 
Metabolism CYP1A1,*GM2A,IRS2,§GK,§INSR,§PGK1,§PFKP,§SGK§ 
Other FKBP5, LIMK2, MYH9, SLC29A3, CEBPD, KLF9, NQO2,*SPON2, CSNK1D, SNF1LK, IGHG1,*TUBA1, NLGN4X,*PRG1, EIF2C2, GABPB2, MRPS16* 
Unknown KCTD12, C18ORF1, SDC2, RNF38, BYSL,*THEM43, NUDT18* 
FunctionGenes
Proapoptotic ADORA2A,*NFIL3, ZBTB16, STAT1,*ISG20, TXN1P, BTG1, RARRES3,*SOCS1, DSIP1 (GILZ) 
Proliferation ADORA2A,*NFIL3, ZBTB16, ZFP36L2, MAP2K1, CEBPD, DUSP1 (MKP-1), NMT2,*RNF111, DUSP10 (MKP-5), TNFAIP3, SGK, IMP-3 
Metabolism CYP1A1,*GM2A,IRS2,§GK,§INSR,§PGK1,§PFKP,§SGK§ 
Other FKBP5, LIMK2, MYH9, SLC29A3, CEBPD, KLF9, NQO2,*SPON2, CSNK1D, SNF1LK, IGHG1,*TUBA1, NLGN4X,*PRG1, EIF2C2, GABPB2, MRPS16* 
Unknown KCTD12, C18ORF1, SDC2, RNF38, BYSL,*THEM43, NUDT18* 

The function of the different genes was studied in the literature (PubMed search). Some genes have multiple functions, depending on cell type, stage in cell cycle.

*

Genes down-regulated on glucocorticoid exposure.

Drug-metabolizing enzyme.

Glycosphingolipid metabolism.

§

Carbohydrate metabolism

Coordinated up- and down-regulation of multiple genes on prednisolone exposure may depend on the presence of specific transcription factor binding motifs (TFBMs) in the promoter regions of genes. Forty-five of the 51 genes were annotated in the TELiS database,21  of which 35 genes were up-regulated and 10 genes were down-regulated on prednisolone exposure. Seventeen TFBMs were found to be overrepresented or underrepresented in the up-regulated genes compared with nonregulated genes (P < .01 and a FDR < 14%). Of these 17 TFBMs, 11 were overrepresented and 6 were underrepresented (Table 4). cAMP-responsive element-binding protein (CREB) is the most often represented TFBM in the up-regulated genes. We did not find a significant different frequency of GRE motifs in these 45 up-regulated genes as compared with nonregulated genes. The number of 12 down-regulated genes was too small to allow for a meaningful analysis.

Table 4

Transcription factor-binding motifs in prednisolone-responsive genes

Transcription factorTBFM matrixFold difference*P
Overrepresented 
    cAMP-responsive element-binding protein V$CREB_Q2 2.8 < .001 
    CRE-binding protein 1 V$CREBP1_Q2 2.8 .006 
    cAMP-responsive element-binding protein V$CREB_02 2.5 < .001 
    Activator protein 2 V$AP2_Q6 2.5 < .001 
    cAMP-responsive element-binding protein V$CREB_01 2.5 .003 
    cAMP-responsive element-binding protein V$CREB_Q4 2.2 .006 
    Stimulating protein 1 V$SP1_Q6 2.1 < .001 
    GC box elements V$GC_01 1.9 .002 
    MZF1 V$MZF1_01 1.6 <.001 
    Stimulating protein 1 V$SP1_01 1.6 .004 
    Activator protein 4 V$AP4_Q5 1.4 .006 
Underrepresented 
    YY1 (yin and yang 1) V$YY1_01 −2.5 < .001 
    Octamer factor 1 V$OCT1_03 −2.0 .002 
    GATA-binding factor 3 V$GATA3_01 −1.4 .007 
    GATA-binding factor 1 V$GATA1_01 −1.3 .002 
    GATA-binding factor 2 V$GATA2_01 −1.3 .004 
    Cap signal for transcription initiation V$CAP_01 −1.1 < .001 
Transcription factorTBFM matrixFold difference*P
Overrepresented 
    cAMP-responsive element-binding protein V$CREB_Q2 2.8 < .001 
    CRE-binding protein 1 V$CREBP1_Q2 2.8 .006 
    cAMP-responsive element-binding protein V$CREB_02 2.5 < .001 
    Activator protein 2 V$AP2_Q6 2.5 < .001 
    cAMP-responsive element-binding protein V$CREB_01 2.5 .003 
    cAMP-responsive element-binding protein V$CREB_Q4 2.2 .006 
    Stimulating protein 1 V$SP1_Q6 2.1 < .001 
    GC box elements V$GC_01 1.9 .002 
    MZF1 V$MZF1_01 1.6 <.001 
    Stimulating protein 1 V$SP1_01 1.6 .004 
    Activator protein 4 V$AP4_Q5 1.4 .006 
Underrepresented 
    YY1 (yin and yang 1) V$YY1_01 −2.5 < .001 
    Octamer factor 1 V$OCT1_03 −2.0 .002 
    GATA-binding factor 3 V$GATA3_01 −1.4 .007 
    GATA-binding factor 1 V$GATA1_01 −1.3 .002 
    GATA-binding factor 2 V$GATA2_01 −1.3 .004 
    Cap signal for transcription initiation V$CAP_01 −1.1 < .001 

Transcription factor binding motifs (TFBMs) overrepresented and underrepresented in 39 genes up-regulated on 8 hours of prednisolone exposure.

*

Fold difference in the frequency of specified TFBMs in up-regulated genes compared with the frequency observed in genes that are not differentially expressed on prednisolone exposure.

Although this study contained both in vitro prednisolone-sensitive and -resistant cases, the sample size in each subgroup was too small for a statistically relevant analysis.

Despite the clinical importance of glucocorticoids in the treatment of ALL, the genes which are transcriptionally regulated on glucocorticoid exposure in pediatric ALL and the specific (in)activation of pathways leading to glucocorticoid-induced apoptosis are unknown.

In the present study, 3 hours of prednisolone exposure did not sufficiently alter the level of gene expression to be able to (statistically) discriminate prednisolone-responsive genes in pediatric ALL. In contrast to these primary cells, in leukemic cell lines significant changes in gene expression can be observed already after 3 hours of prednisolone exposure.8,9  The same phenomenon that leukemic cell lines respond faster to a drug than corresponding primary cells was found for L-asparaginase.23  Besides the faster response, different genes were also found to be affected by these drugs in the leukemic cell lines. These studies emphasize that leukemic cell lines behave differently to drugs than primary cells and hence may not be suitable models for pharmacodynamic studies.

Exposure of pediatric ALL cells for 8 hours to prednisolone affected the expression of 57 probe sets (51 genes); 44 probe sets were up-regulated and 13 probe sets were down-regulated. FKBP5 (FK506 binding protein 5) was the most significantly up-regulated gene (35-fold). Its product, FKBP51, functions as co-chaperone molecule of the glucocorticoid receptor that affects the transport of the glucocorticoid receptor into the nucleus. The expression of this gene has been reported before to be highly glucocorticoid inducible in cell lines and primary patient cells.13,24,25  The most significantly down-regulated gene was CYP1A1 (2-fold) which is involved in drug metabolism and detoxification. Although we found only 57 probe sets significantly regulated on 8 hours of prednisolone exposure, more prednisolone-responsive genes can be identified if less-stringent cut-off levels are used for the P value and FDR. For example, at P less than .001 and FDR less than 20%, 144 probe sets are regulated on 8 hours of prednisolone exposure.

Interestingly, a gene expression profiling study has been published recently in which glucocorticoid-responsive genes were studied in both leukemic cell lines and primary cells of pediatric ALL.13  In that study other statistical considerations were made than in our study, such as a minimal change in gene expression of 1.6-fold in at least 6 of 13 studied patients. In our study, 99% of the observations were in correspondence with each other, that is, either up- or down-regulation of gene expression on prednisolone exposure in all tested cases. Despite differences in methodology, 5 of 28 identified responsive genes found in patients that were treated in vivo with prednisone (FKBP5, SOCS1, ZFP36L2, SNF1LK, and ZBTB16) were also found in our study using in vitro–exposed leukemic cells of children with ALL.

Signal transduction pathways possibly involved in glucocorticoid-induced apoptosis

Both TXNIP and ZBTB16 were found to be up-regulated following 8 hours of prednisolone exposure in the present study. TXNIP has been described as a tumor suppressor protein that induces a cell-cycle arrest on formation of a transcriptional repressor complex with ZBTB16.26  TXNIP prevents thioredoxin-mediated apoptosis signal-regulating kinase 1 (ASK1) ubiquitination and degradation27,28  and inhibits the thioredoxin radical scavenging function.29  Thus, TXNIP can act as a proapoptotic regulator. These data support a role for these genes in the induction of apoptosis on prednisolone exposure in childhood ALL, as has been suggested recently by Wang et al.30 

MAP kinase pathways.

In the group of 51 prednisolone-regulated genes, 9 genes were associated with the 3 mitogen-activated protein (MAP) kinase pathways (ie, ERK, JNK and p38 MAPK). These MAP kinase pathways are involved in cell survival (ERK and JNK) and cell death (p38 MAPK) and have been reported to play critical roles in the pathogenesis of various hematologic malignancies.31,32  Miller et al33  recently showed that pharmacologic inhibition of ERK and JNK enhanced glucocorticoid-induced apoptosis, whereas inhibition of p38 MAPK activity opposed glucocorticoid-induced apoptosis in lymphoid cells. Four genes (DUSP1, DUSP10, DSIPI [alias GILZ], and SGK) that we found to be induced on prednisolone exposure are negative regulators of the MAP kinase pathways. In the literature, overexpression of DSIPI has been shown to promote apoptosis in thymocytes.34  Overexpression of DUSP1 in a precursor-B ALL cell line did not alter glucocorticoid sensitivity,35  suggesting that DUSP1 may not be essential for mediating the toxic effect of glucocorticoid in ALL cells. SGK is related to the PI3K/AKT pathway (which is linked to the ERK pathway) and has been reported to be a survival molecule.36 

Because we found both positive and negative regulators of the MAPK pathways among the prednisolone-regulated genes, the net effect of these genes on cell survival needs to be addressed in functional studies in childhood ALL.

NF-κB.

There are 2 classically proposed ways of an inhibitory effect of glucocorticoids on NF-κB function. First, the glucocorticoid-GR complex may interact with NF-κB directly, thereby opposing its function. Second, glucocorticoids up-regulate IκBα (inhibitor of NF-κB α), which negatively regulates NF-κB. However, we did not find a significant up-regulation of IKBA in our study (P = .002; FDR, 30%). Other postulated mechanisms of glucocorticoid-mediated inhibition of NF-κB are the glucocorticoid-induced up-regulation of DSIPI and TNFAIP3.37,38  Both genes were identified in our study as prednisolone-responsive genes (3.35-fold and 2.12-fold, respectively; P < .001; FDR < 10%). This observation suggests that the last 2 mechanisms might be more relevant for NF-κB inhibition (and hence induction of apoptosis) than the 2 classically proposed models in childhood ALL.

Carbohydrate metabolism.

Five genes directly involved in carbohydrate metabolism are up-regulated on prednisolone exposure: IRS2, INSR, PFKP, GK, and PGK1. Activity of these 5 genes results in higher ATP levels in the cell because of a higher glucose uptake of the cell (IRS2 and INSR), a higher rate of glyconeogenesis (GK), and glycolysis (PFKP and PGK1). Interestingly, in an earlier study looking for the baseline expression of genes determining glucocorticoid resistance in primary, untreated ALL cells, we found GLUT3 and GAPDH, 2 genes involved in carbohydrate metabolism, to be overexpressed in prednisolone-resistant ALL cells.16  Moreover, the glycolytic rate of prednisolone-resistant leukemic cell lines was higher as compared with sensitive cell lines and inhibition of glycolysis by 2-deoxy-D-glucose sensitized resistant cells to prednisolone, whereas no effect on sensitive cells was found (A. Holleman, manuscript in preparation). Taken together, these studies strongly suggest an important role for carbohydrate metabolism in glucocorticoid-induced apoptosis.

Transcription factor binding motifs

The TELiS database20,21  was used to study which TFBMs were overrepresented or underrepresented in the 39 up-regulated genes (Table 4). Four TFBMs representing cAMP-responsive element-binding protein (CREB) were overrepresented in the promoter regions of the 35 up-regulated genes as compared with nonregulated genes. CREB is a ubiquitous transcription factor involved in cell proliferation and survival. Interestingly, the interaction between glucocorticoids and CREB has been reported before.39  Forskolin was shown to increase cellular cAMP levels and to promote the phosphorylation of CREB. In combination with dexamethasone, forskolin synergistically induced apoptosis in the glucocorticoid-resistant CEM-C1 lymphoid cell line, suggesting a role for CREB in glucocorticoid response.

Surprisingly, the glucocorticoid-responsive element (GRE) sequence was not overrepresented in the promoter regions of the 35 up-regulated genes, which is in line with a previous report on glucocorticoid-induced genes in 3 ALL cell lines.40  However, at least 3 of the 51 genes found to be regulated on prednisolone exposure contain GREs in the promoter regions, namely FKBP5,25,41 DSIPI,42  and SGK.43,44  The fact that the GRE TFBMs were not overexpressed in the up-regulated genes in our study might be the result of the delicate positioning of GRE-like sequences in glucocorticoid-responding genes that are not recognized in the TELiS database. Another explanation might be that some genes are regulated “directly” by glucocorticoids binding to their GREs, whereas other genes are regulated more “indirectly” by other transcription factors, which in turn are regulated by glucocorticoids.

In conclusion, we found 51 prednisolone-responsive genes in leukemic cells taken from children at initial diagnosis of ALL. Further functional research may identify which genes/pathways are essential for the glucocorticoid responsiveness of cells. Of the 51 identified glucocorticoid-responsive genes, 50% can be linked to 3 pathways, that is, cell proliferation and survival, NF-κB signaling, and glucose metabolism. Two of the up-regulated genes are tumor suppressor genes: TXNIP and ZBTB16, which is possibly related to the induction of apoptosis by glucocorticoids. Knowledge on the pathways leading to glucocorticoid-induced apoptosis is essential to develop more targeted therapy and ways to modulate glucocorticoid resistance in pediatric ALL.

Contribution: W.J.E.T. designed and performed the research, collected and analyzed the data, and wrote the paper; M.L.d.B., J.P.P.M., and R.P. designed the research and analyzed the data; R.X.M., S.S., and P.J.v.d.S. analyzed the data; S.E.S. designed the research; and S.A.A. designed and performed the research and analyzed the data.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

Correspondence: M. L. den Boer, Erasmus MC-Sophia Children's Hospital, Department of Pediatric Oncology/Hematology, Dr Molewaterplein 60, 3015 GJ, Rotterdam, The Netherlands; e-mail m.l.denboer@erasmusmc.nl.

The publication costs of this article were defrayed in part by page charge payment. Therefore, and solely to indicate this fact, this article is hereby marked “advertisement” in accordance with 18 USC section 1734.

This work was supported by the Center of Medical Systems Biology (CMSB) established by The Netherlands Genomics Initiative/Netherlands Organisation for Scientific Research (NGI/NWO) (R.X.M.).

1
Lauten M, Stanulla M, Zimmermann M, Welte K, Riehm H, Schrappe M. Clinical outcome of patients with childhood acute lymphoblastic leukaemia and an initial leukaemic blood blast count of less than 1000 per microliter.
Klin Padiatr
2001
;
213
:
169
–174.
2
Kaspers GJ, Pieters R, Van Zantwijk CH, Van Wering ER, Van Der Does-Van Den Berg A, Veerman AJ. Prednisolone resistance in childhood acute lymphoblastic leukemia: vitro-vivo correlations and cross-resistance to other drugs.
Blood
1998
;
92
:
259
–266.
3
Hongo T, Yajima S, Sakurai M, Horikoshi Y, Hanada R. In vitro drug sensitivity testing can predict induction failure and early relapse of childhood acute lymphoblastic leukemia.
Blood
1997
;
89
:
2959
–2965.
4
Tsai SY, Carlstedt-Duke J, Weigel NL, et al. Molecular interactions of steroid hormone receptor with its enhancer element: evidence for receptor dimer formation.
Cell
1988
;
55
:
361
–369.
5
De Bosscher K, Vanden Berghe W, Haegeman G. The interplay between the glucocorticoid receptor and nuclear factor-kappaB or activator protein-1: molecular mechanisms for gene repression.
Endocr Rev
2003
;
24
:
488
–522.
6
Tissing WJ, Meijerink JP, den Boer ML, Pieters R. Molecular determinants of glucocorticoid sensitivity and resistance in acute lymphoblastic leukemia.
Leukemia
2003
;
17
:
17
–25.
7
Schmidt S, Rainer J, Ploner C, Presul E, Riml S, Kofler R. Glucocorticoid-induced apoptosis and glucocorticoid resistance: molecular mechanisms and clinical relevance.
Cell Death Differ
2004
;
11
:suppl 1,
S45
–S55.
8
Tonko M, Ausserlechner MJ, Bernhard D, Helmberg A, Kofler R. Gene expression profiles of proliferating vs. G1/G0 arrested human leukemia cells suggest a mechanism for glucocorticoid-induced apoptosis.
FASEB J
2001
;
15
:
693
–699.
9
Obexer P, Certa U, Kofler R, Helmberg A. Expression profiling of glucocorticoid-treated T-ALL cell lines: rapid repression of multiple genes involved in RNA-, protein- and nucleotide synthesis.
Oncogene
2001
;
20
:
4324
–4336.
10
Planey SL, Abrams MT, Robertson NM, Litwack G. Role of apical caspases and glucocorticoid-regulated genes in glucocorticoid-induced apoptosis of pre-B leukemic cells.
Cancer Res
2003
;
63
:
172
–178.
11
Yoshida NL, Miyashita T, U M, et al. Analysis of gene expression patterns during glucocorticoid-induced apoptosis using oligonucleotide arrays.
Biochem Biophys Res Commun
2002
;
293
:
1254
–1261.
12
Wang Z, Malone MH, He H, McColl KS, Distelhorst CW. Microarray analysis uncovers the induction of the proapoptotic BH3-only protein Bim in multiple models of glucocorticoid-induced apoptosis.
J Biol Chem
2003
;
278
:
23861
–23867.
13
Schmidt S, Rainer J, Riml S, et al. Identification of glucocorticoid-responsive genes in children with acute lymphoblastic leukemia.
Blood
2006
;
107
:
2061
–2069.
14
Stam RW, den Boer ML, Meijerink JP, et al. Differential mRNA expression of Ara-C-metabolizing enzymes explains Ara-C sensitivity in MLL gene-rearranged infant acute lymphoblastic leukemia.
Blood
2003
;
101
:
1270
–1276.
15
Armstrong SA, Staunton JE, Silverman LB, et al. MLL translocations specify a distinct gene expression profile that distinguishes a unique leukemia.
Nat Genet
2002
;
30
:
41
–47.
16
Holleman A, Cheok MH, den Boer ML, et al. Gene-expression patterns in drug-resistant acute lymphoblastic leukemia cells and response to treatment.
N Engl J Med
2004
;
351
:
533
–542.
17
Huber W, von Heydebreck A, Sultmann H, Poustka A, Vingron M. Variance stabilization applied to microarray data calibration and to the quantification of differential expression.
Bioinformatics
2002
;
18
:suppl 1,
S96
–S104.
18
Benjamini Y and Hochberg Y. Controlling the false discovery rate—a practical and powerful approach to multiple testing.
J Roy Stat Soc B
1995
;
57
:
289
–300.
19
R development core team. R: a language and environment for statistical computing. R foundation for statistical computing
2005
;Vienna, Austria http://www.r-project.org Accessed October 11, 2004.
20
Cole SW, Yan W, Galic Z, Arevalo J, Zack JA. Expression-based monitoring of transcription factor activity: the TELiS database.
Bioinformatics
2005
;
21
:
803
–810.
21
Cole S. UC Regents. TELiS: the transcription element listening system. http://www.telis.ucla.edu/index.htm Accessed March 23, 2005.
22
Gene Ontology Consortium. Gene ontology database. http://geneontology.org Accessed October 30, 2004.
23
Fine BM, Kaspers GJ, Ho M, Loonen AH, Boxer LM. A genome-wide view of the in vitro response to L-asparaginase in acute lymphoblastic leukemia.
Cancer Re
2005
;
65
:
291
–299.
24
Vermeer H, Hendriks-Stegeman BI, van der Burg B, van Buul-Offers SC, Jansen M. Glucocorticoid-induced increase in lymphocytic FKBP51 messenger ribonucleic acid expression: a potential marker for glucocorticoid sensitivity, potency, and bioavailability.
J Clin Endocrinol Metab
2003
;
88
:
277
–284.
25
U M, Shen L, Oshida T, Miyauchi J, Yamada M, Miyashita T. Identification of novel direct transcriptional targets of glucocorticoid receptor.
Leukemia
2004
;
18
:
1850
–1856.
26
Han SH, Jeon JH, Ju HR, et al. VDUP1 upregulated by TGF-beta1 and 1,25-dihydroxyvitamin D3 inhibits tumor cell growth by blocking cell-cycle progression.
Oncogene
2003
;
22
:
4035
–4046.
27
Liu Y and Min W. Thioredoxin promotes ASK1 ubiquitination and degradation to inhibit ASK1-mediated apoptosis in a redox activity-independent manner.
Circ Res
2002
;
90
:
1259
–1266.
28
Junn E, Han SH, Im JY, et al. Vitamin D3 up-regulated protein 1 mediates oxidative strss via suppressing the thioredoxin function.
J. Immunol
2000
;
164
:
6287
–6295.
29
Schulze PC, Yoshioka J, Takahashi T, He Z, King GL, Lee RT. Hyperglycemia promotes oxidative stress through inhibition of thioredoxin function by thioredoxin-interacting protein.
J Biol Chem
2004
;
279
:
30369
–30374.
30
Wang Z, Rong YP, Malone MH, Davis MC, Zhong F, Distelhorst CW. Thioredoxin-interacting protein (txnip) is a glucocorticoid-regulated primary response gene involved in mediating glucocorticoid-induced apoptosis.
Oncogene
2006
;
25
:
1910
–1913.
31
Platanias LC. Map kinase signaling pathways and hematologic malignancies.
Blood
2003
;
101
:
4667
–4679.
32
Ravandi F, Talpaz M, Kantarjian H, Estrov Z. Cellular signaling pathways: new targets in leukaemia therapy.
Br J Haematol
2002
;
116
:
57
–77.
33
Miller AL, Webb MS, Copik AJ, et al. p38 MAP kinase is a key mediator in glucocorticoid-induced apoptosis of lymphoid cells: correlation between p38 MAPK activation and site-specific phosphorylation of the human glucocorticoid receptor at serine 211.
Mol Endocrinol
2005
;
19
:
1569
–1583.
34
Delfino DV, Agostini M, Spinicelli S, Vito P, Riccardi C. Decrease of Bcl-xL and augmentation of thymoctye apoptosis in GILZ overexpressing transgenic mice.
Blood
2004
;
104
:
4134
–4141.
35
Abrahams MT, Robertson NM, Litwack G, Wickstrom E. Evaluation of glucocorticoid sensitivity in 697 pre-B acute lymphoblastic leukemia cells after overexpression or silencing of MAP kinase phosphatase-1.
J Cancer Res Clin Oncol
2005
;
131
:
347
–354.
36
Mikosz CA, Brickley DR, Sharkey MS, Moran TW, Conzen SD. Glucocorticoid receptor-mediated protection from apoptosis is associated with induction of the serine/threonine survival kinase gene, sgk-1.
J Biol Chem
2001
;
276
:
16649
–16654.
37
Ayroldi E, Migliorati G, Bruscoli S, et al. Modulation of T-cell activation by the glucocorticoid-induced leucine zipper factor via inhibition of nuclear factor kappaB.
Blood
2001
;
98
:
743
–753.
38
Chen F. Endogenous inhibitors of nuclear factor-kappaB, an opportunity for cancer control.
Cancer Res
2004
;
64
:
8135
–8138.
39
Medh RD, Saeed MF, Johnson BH, Thompson EB. Resistance of human leukemic CEM-C1 cells is overcome by synergism between glucocorticoid and protein kinase A pathways: correlation with c-Myc suppression.
Cancer Res
1998
;
58
:
3684
–3693.
40
Medh RD, Webb MS, Miller AL, et al. Gene expression profile of human lymphoid CEM cells sensitive and resistant to glucocorticoid-evoked apoptosis.
Genomics
2003
;
81
:
543
–555.
41
Hubler TR and Scammell JG. Intronic hormone response elements mediate regulation of FKBP5 by progestins and glucocorticoids.
Cell Stress Chaperones
2004
;
9
:
243
–252.
42
Asselin-Labat ML, David M, Biola-Vidamment A, et al. GILZ, a new target for the transcription factor FoxO3, protects T lymphocytes from interleukin-2 withdrawal-induced apoptosis.
Blood
2004
;
104
:
215
–223.
43
Naray-Fejes-Toth A, Fejes-Toth G, Volk KA, Stokes JB. SGK is a primary glucocorticoid-induced gene in the human.
J Steroid Biochem Mol Biol
2000
;
75
:
51
–56.
44
Leong ML, Maiyar AC, Kim B, O'Keeffe BA, Firestone GL. Expression of the serum- and glucocorticoid-inducible protein kinase, SGK, is a cell survival response to multiple types of environmental stress stimuli in mammary epithelial cells.
J Biol Chem
2003
;
278
:
5871
–5882.
45
Webb MS, Miller AL, Johnson BH, et al. Gene networks in glucocorticoid-evoked apoptosis of leukemic cells.
J Steroid Biochem Mol Biol
2003
;
85
:
183
–193.
46
Jeon JW, Lee SJ, Kim JB, et al. Cellular proliferative effect of dexamethasone in immortalized trabecular meshwork cell (TM5) line.
Yonsei Med J
2003
;
44
:
299
–306.
47
Kasinath BS, Singh AK, Kanwar YS, Lewis EJ. Dexamethasone increases heparan sulfate proteoglycan core protein content of glomerular epithelial cells.
J Lab Clin Med
1990
;
115
:
196
–202.
48
Ramji DP and Foka P. CCAAT/enhancer-binding proteins: structure, function and regulation.
Biochem J
2002
;
365
:
561
–575.
49
Taylor WM, Goldrick RB, Ishikawa T. Glycerokinase in rat and human adipose tissue: response to hormonal and dietary stimuli.
Horm Metab Res
1979
;
11
:
280284
.
50
Hu X, Li WP, Meng C, Ivashkiv LB. Inhibition of IFN-gamma signaling by glucocorticoids.
J Immunol
2003
;
170
:
4833
–4839.
51
Leal MA, Aller P, Calle C. Effect of dexamethasone on insulin receptor mRNA levels, RNA stability and isotype RNA pattern in U-937 human promonocytic cells.
J Endocrinol Invest
1996
;
19
:
530
–534.
52
Celander M, Weisbrod R, Stegeman JJ. Glucocorticoid potentiation of cytochrome p4501A1 induction by 2,3,7,8-tetrachlorodibenzo-p-dioxin in porcine and human endothelial cells in culture.
Biochem Biophys Res Commun
1997
;
232
:
749
–753.
53
Hermoso MA, Matsuguchi T, Smoak K, Cidlowski JA. Glucocorticoids and tumor necrosis factor alpha cooperatively regulate toll-like receptor 2 gene expression.
Mol Cell Biol
2004
;
24
:
4743
–4756.
Sign in via your Institution