Skip to main content

Main menu

  • Home
  • Articles
    • Current Issue
    • Next in The JI
    • Archive
    • Brief Reviews
    • Pillars of Immunology
    • Translating Immunology
    • Most Read
    • Top Downloads
    • Annual Meeting Abstracts
  • COVID-19/SARS/MERS Articles
  • Info
    • About the Journal
    • For Authors
    • Journal Policies
    • Influence Statement
    • For Advertisers
  • Editors
  • Submit
    • Submit a Manuscript
    • Instructions for Authors
    • Journal Policies
  • Subscribe
    • Journal Subscriptions
    • Email Alerts
    • RSS Feeds
    • ImmunoCasts
  • More
    • Most Read
    • Most Cited
    • ImmunoCasts
    • AAI Disclaimer
    • Feedback
    • Help
    • Accessibility Statement
  • Other Publications
    • American Association of Immunologists
    • ImmunoHorizons

User menu

  • Subscribe
  • Log in

Search

  • Advanced search
The Journal of Immunology
  • Other Publications
    • American Association of Immunologists
    • ImmunoHorizons
  • Subscribe
  • Log in
The Journal of Immunology

Advanced Search

  • Home
  • Articles
    • Current Issue
    • Next in The JI
    • Archive
    • Brief Reviews
    • Pillars of Immunology
    • Translating Immunology
    • Most Read
    • Top Downloads
    • Annual Meeting Abstracts
  • COVID-19/SARS/MERS Articles
  • Info
    • About the Journal
    • For Authors
    • Journal Policies
    • Influence Statement
    • For Advertisers
  • Editors
  • Submit
    • Submit a Manuscript
    • Instructions for Authors
    • Journal Policies
  • Subscribe
    • Journal Subscriptions
    • Email Alerts
    • RSS Feeds
    • ImmunoCasts
  • More
    • Most Read
    • Most Cited
    • ImmunoCasts
    • AAI Disclaimer
    • Feedback
    • Help
    • Accessibility Statement
  • Follow The Journal of Immunology on Twitter
  • Follow The Journal of Immunology on RSS

The Role of High-Mobility Group I(Y) Proteins in Expression of IL-2 and T Cell Proliferation

S. Roy Himes, Raymond Reeves, Joanne Attema, Mark Nissen, Ying Li and M. Frances Shannon
J Immunol March 15, 2000, 164 (6) 3157-3168; DOI: https://doi.org/10.4049/jimmunol.164.6.3157
S. Roy Himes
*Hanson Center for Cancer Research, Adelaide, South Australia;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Raymond Reeves
†Department of Biochemistry and Biophysics, Washington State University, Pullman, WA 99164; and
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Joanne Attema
‡Division of Biochemistry and Molecular Biology, John Curtin School of Medical Research, Australian National University, Canberra, Australia
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Mark Nissen
†Department of Biochemistry and Biophysics, Washington State University, Pullman, WA 99164; and
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Ying Li
†Department of Biochemistry and Biophysics, Washington State University, Pullman, WA 99164; and
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
M. Frances Shannon
‡Division of Biochemistry and Molecular Biology, John Curtin School of Medical Research, Australian National University, Canberra, Australia
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • Article
  • Figures & Data
  • Info & Metrics
  • PDF
Loading

Abstract

The high-mobility group I(Y) (HMGI(Y)) family of proteins plays an important architectural role in chromatin and have been implicated in the control of inducible gene expression. We have previously shown that expression of HMGI antisense RNA in Jurkat T cells inhibits the activity of the IL-2 promoter. Here we have investigated the role of HMGI(Y) in controlling IL-2 promoter-reporter constructs as well as the endogenous IL-2 gene in both Jurkat T cells and human PBL. We found that the IL-2 promoter has numerous binding sites for HMGI(Y), which overlap or are adjacent to the known transcription factor binding sites. HMGI(Y) modulates binding to the IL-2 promoter of at least three transcription factor families, AP-1, NF-AT and NF-κB. By using a mutant HMGI that cannot bind to DNA but can still interact with the transcription factors, we found that DNA binding by HMGI was not essential for the promotion of transcription factor binding. However, the non-DNA binding mutant acts as a dominant negative protein in transfection assays, suggesting that the formation of functional HMGI(Y)-containing complexes requires DNA binding as well as protein:protein interactions. The alteration of HMGI(Y) levels affects IL-2 promoter activity not only in Jurkat T cells but also in PBL. Importantly, we also show here that expression of the endogenous IL-2 gene as well as proliferation of PBL are affected by changes in HMGI(Y) levels. These results demonstrate a major role for HMGI(Y) in IL-2 expression and hence T cell proliferation.

Activation of Th cells requires the specific binding of the TCR to an Ag-MHC class II complex on the surface of APCs as well as costimulation through the interaction of cell-surface molecule pairs such as CD28 on T cells and B7 on APCs (1, 2). Signal transduction through the TCR and CD28 induces the activation and nuclear expression of a number of transcription factors, including NF-κB/Rel, NF-AT and AP-1 family proteins (reviewed in Refs. 3 and 4). This cascade of events results in the coordinate expression of a variety of genes including the cytokine IL-2, which functions as a major growth factor for T cells, and expression of IL-2Rα, which allows the formation of a high-affinity receptor for IL-2. The resulting autocrine loop can act as a strong growth stimulus allowing a specific clonal expansion of the T cell.

The requirements for activation of IL-2 gene expression in response to T cell activation have been studied in detail. Initially, two major TCR responsive regions were identified in the IL-2 promoter (reviewed in Refs. 3 and 5). These were termed Ag response elements (ARRE),4 i.e., ARRE-1 (or NF-IL-2A) and ARRE-2 (or NF-IL-2E). The ARRE-2 region was the first to be identified as a composite binding site for NF-AT and AP-1 proteins, whereas ARRE-1 was thought to function via the binding of AP-1 and Oct transcription factors (6, 7, 8, 9, 10). Recently, the ARRE-1 region has been found to bind a complex of NF-AT and AP-1 as well as Oct (11). Several other regions of the promoter have also recently been identified as binding sites for NF-AT alone or as complexes with AP-1 (11). The mechanism of costimulation through the CD28 receptor involves enhanced nuclear expression of NF-κB and c-Rel transcription factors (12, 13, 14), as well as phosphorylation of c-Jun by JNK kinase, resulting in increased transactivation competence for AP-1 (15, 16). The region of the IL-2 gene that responds to CD28 (the CD28RR) is composed of a variant NF-κB site (referred to as the CD28RE), which has a high affinity for c-Rel-containing complexes, as well as an adjacent AP-1 site. These elements cooperate to lead to increased IL-2 promoter activity (17, 18, 19, 20).

The highly inducible nature of the IL-2 promoter seems to be the result of coordinate binding of many transcription factors to their recognition sequences on the promoter leading to the assembly of a functional unit (3, 4, 11, 14, 21, 22, 23, 24). Such a unit has been termed an enhanceosome on the IFN-β promoter (25). The family of nuclear proteins known as high-mobility group I(Y) (HMGI(Y)) are known to play a major role in the assembly of the IFN-β enhanceosome (26, 27, 28). We have previously shown that HMGI(Y) is essential for the selective binding of c-Rel to the IL-2 CD28RE and that modulating HMGI(Y) levels affects not only the activity of the CD28RR but of the entire promoter (14).

HMGI(Y) proteins are small nonhistone nuclear proteins that bind the narrow minor groove of A:T sequence-rich B form DNA and have numerous modes of action by which they modify gene transcription (reviewed in Ref. 29). The HMGI(Y) family of proteins consist of three members, HMGI and HMGY, which are produced by alternative splicing of mRNA from the same gene locus, and HMGIC, which is coded for by a related gene at a separate locus (29). These proteins have been classed as architectural transcription factors because they do not act as transactivators in their own right but modify the function of other proteins (30). HMGI(Y) proteins can interact directly with several families of transcription factors including Rel/NF-κB, bZip, Ets, homeodomain, and Pou domain proteins and lead to an alteration in their DNA binding to sites that either overlap or are adjacent to A:T-rich HMGI(Y) binding sites (29). The binding of HMGI(Y) also induces structural changes in DNA substrates (31, 32) that, in turn, often leads to alterations in the assembly of transcription factors into higher order functional complexes (29). In addition, HMGI(Y) appears to play a critical role in chromatin architecture (33) and has been shown to interact specifically with isolated nucleosome core particles (34), to alter the rotational setting of DNA on the surface of nucleosomes (35), and to antagonize H1-mediated transcriptional repression (36, 37).

Here we show that HMGI(Y) can bind to numerous sites across the IL-2 proximal promoter region and modulate the binding of the major transcription factor families that are thought to control IL-2 gene transcription. Functional studies in both Jurkat T cells and primary T cells show that HMGI(Y) plays a major role in the regulation of the IL-2 gene and hence T cell proliferation.

Materials and Methods

Cell culture and transfection

The basic medium for Jurkat cell culture was RPMI 1640 medium containing 10% FCS, supplemented with l-glutamine and penicillin-gentamicin antibiotics (RPMIJ). The basic medium for PBL culture was as above but contained 20% FCS, 100 μM 2-ME, and 5% conditioned media (RPMIL). Mononuclear cells were isolated from peripheral blood using lymphoprep (Nycomed Pharma AS, Oslo, Norway) and cultured for 4 days in RPMIL and PHA (5 μg/ml) (Boehringer Mannheim, Mannheim, Germany). Cells were then stimulated with 5 ng/ml of IL-12 (PharMingen, San Diego, CA) for 4 h to promote transition to G1 and panned to remove macrophages. Nonadherent PBLs were pelleted and resuspended in RPMIL at 1 × 107 cells in 400 μl media. PBLs were assayed for CD3, CD4, and CD8 expression using the Cyto-Stat assay kit from Coulter (Palo Alto, CA). Ninety-four percent of the cells were CD3 positive, 74% were CD4 positive, and 20% were CD8 positive. Cells were transfected by electroporation using a Bio-Rad Gene Pulser II (Richmond, CA) at 290 V with a capacitance of 975 μF. Jurkat cells were resuspended at 5 × 106 cells in 400 μl media and electroporated at 270 V, 975 μF capacitance. The efficiency of transfection was determined using the pCMVGFP expression plasmid and flow cytometry (Epics XL-MCL; Coulter) and ranged from 2 to 4% green fluorescence protein (GFP) positive in PBLs and 15 to 25% GFP positive in Jurkat cells. Transfected cells were incubated in RPMIL with 10% conditioned media (PBLs) or RPMIJ media (Jurkats) for 24 h. Cells were then sorted for high fluorescence using a FACStarPlus cell sorter (Becton Dickinson, Mountain View, CA), and the top 1–2% for PBLs and 10–15% for Jurkats were collected and used in IL-2 ELISA and proliferation assays.

Reporter assay, ELISA, and proliferation assay

pRcCMV and pcDNA3.1/Zeo were obtained from Invitrogen (San Diego, CA), and pRcCMVIGMH, pIL-2luc, and pHIVluc have been previously described (14). Standard site-specific mutagenesis procedures, the details of which will be reported elsewhere (Li et al., unpublished observations), were used to create a non-DNA binding mutant form of the HMGI protein designated HMGI(mII,mIII) starting with the wild-type human HMGI cDNA, clone 7C (38). Briefly, HMGI(mII,mIII) had four proline to alanine substitutions introduced at amino acid residues 57, 61, 83, and 87 located in its second and third DNA-binding domains, the primary regions of the protein that interact with the minor groove of A:T-rich substrates (39). As a consequence, the recombinant HMGI(mII,mIII) protein lacks the ability to specifically bind with high affinity to A:T-rich DNA sequences in vitro but retains its ability for specific protein-protein interactions with other transcription factors (Li et al., unpublished observations). For expression in transfected mammalian cells, the mutant HMGI(mII,mIII) cDNA was subcloned into the pcDNA3.1/Zeo plasmid vector (Invitrogen) and used as described above. The HMGY cDNA was also subcloned into pcDNA3.1 to generate pcDNAHMGY for expression of the protein in cells. Jurkat cells used in the reporter assay were cotransfected with 5 μg of pIL-2luc or pHIVluc together with pRcCMVIGMH, pcDNAHMGI(mII,mIII), or pcDNAHMGY, in combination with the amount of parent plasmid needed to normalize DNA at 10 μg. At 24 h posttransfection, cells were pelleted and resuspended in phenol red free RPMIJ media at 1 × 106 cells/ml. Then, 100 μl of cells/well were plated in 96-well culture plates and stimulated with 50 ng/ml PMA and 1 μM Ca2+ ionophore (P/I) (A23187; Boehringer Mannheim) or with the above and a 1/10,000 dilution of ascites fluid containing activating Ab to the CD28 receptor (α-CD28) (Bristol-Myers Squibb, New York, NY) for 9 h. Cells were then harvested and assayed for luciferase activity using a 96-well plate Luminometer (TopCount; Packard, Meriden, CT) as previously described (14). PBLs were cotransfected with 5 μg of pIL-2luc and either 10 μg of pRcCMVIGMH, pcDNAHMGI (mII,mIII), or pcDNAHMGY and the appropriate parent plasmid. At 24 h posttransfection, cells were stimulated with PHA (5 μg/ml) for 16 h. Cells were then harvested and cell extracts assayed for luciferase activity as previously described (40).

ELISA and proliferation assays were performed with PBL or Jurkat cells transfected with 10 μg of the appropriate expression plasmid together with 5 μg of pCMVGFP for cell sorting. Then, 1 × 104 sorted cells in 100 μl were plated in 96-well tissue culture plates and stimulated with either PHA for 24 h (PBL) or P/I+α-CD28 for 8 h (Jurkat), and supernatants were assayed for IL-2 using a Quantikine ELISA kit (R&D Systems, Minneapolis, MN). Proliferation was assayed in similarly treated cells using the chemiluminescent bromodeoxyuridine incorporation kit from Boehringer Mannheim.

Production and purification of recombinant proteins

Recombinant hexahistidine-tagged full-length Fos and Jun proteins, as well as truncated Fos116–211 and Jun224–334 proteins, were prepared and purified by affinity chromatography using a Ni-NTA-agarose column (Qiagen, Chatsworth, CA) as described (41). Fos and Jun proteins were corenatured in vitro into the heterodimer transcription factor AP-1 by step-wise dialysis from 6 M urea with the final buffer containing 25 mM sodium phosphate, pH 7.6, 5% glycerol, and 5 mM DTT (42). Full-length recombinant human HMGI protein (i.e., the unspliced member of the HMGI(Y) protein family; Ref. 29) was produced using the expression vector pET7C carrying the wild-type human HMGI cDNA (38) as described (43, 44). NF-ATp and c-Rel were prepared as described (45, 46). The purity of each recombinant preparation was assessed by SDS-PAGE (47) Protein concentrations were determined spectrophotometrically employing either a Bio-Rad protein assay kit or using the extinction coefficient ε220 = 74,000 L/mol · cm for wild-type HMGI protein (48).

Preparation of nuclear extracts

Nuclear extracts were prepared from Jurkat T cells stimulated with PMA (20 ng/ml), Ca2+ ionophore (1 μm), and CD28 Ab (1:10,000) for 1 or 6 h. Nuclear extract preparation was as previously described (14).

In vitro DNase I footprinting

A cloned 410 bp XhoI/HindIII restriction fragment encompassing nucleotides (nt) −360 to +50 of the human IL-2 gene proximal promoter (49) was the starting fragment for use in in vitro protein footprinting. Promoter subfragments were isolated by either selective restriction enzyme digestion or PCR amplification techniques (47). A fragment from −180 to −60 was used for some of the footprinting experiments. Standard gel electrophoretic procedures (47) were used to isolate all DNA fragments followed by purification on a Qiagen column as described by the manufacturer. Restriction enzyme fragments were 5′-end radiolabeled with T4 polynucleotide kinase and [γ-32P]ATP. The ends of PCR fragments were selectively radiolabeled by incorporating, during the final few cycles of the amplification reaction, either one or the other of the two PCR primers that had been 5′ radiolabeled with T4 polynucleotide kinase and [γ-32P]ATP.

Footprinting of both the HMGI and AP-1 recombinant proteins on promoter DNA fragments radiolabeled on one 5′ end, employing the nuclease DNase I, followed published protocols (35, 44). For each protein and DNA substrate, optimal conditions for footprinting were empirically determined. Single-stranded DNA cleavage products were then separated by electrophoresis on a 6% sequencing gel with Maxam-Gilbert “G-lane” chemical cleavage products of control DNA fragments serving as reference standards (47). Band intensities were quantitatively analyzed using a PhosphorImager machine and ImageQuant software (Molecular Dynamics, Sunnyvale, CA).

EMSA

Oligonucleotides for gel shift analysis had the following sequences: IL-2 CD28RR, 5′-TGGGGGTTTAAAGAAATTCCAGAGAGTCATCAG-3′; CD28APm, 5′-TGGGGGTTTAAAGAAATTCCAGAGAGagATCAG-3′; CD28REm, 5′-TGGGGGTTTAAAGccATTCCAGAGAGTCATCAG-3′; IL-2 NF-IL-2A, 5′-AAGTCTTTGAAAATATGTGTAATATGTAAAACATTTTGA-3′; GM170, 5′-GATCCTGTAGGAAACAGGGGCTTGAGTCACTCCAG-3′; TRE, 5′-GTGGAATTCGGGGAAGTGACTCAGCGCTCGGACG-3′. The lower case letters represent the mutations introduced into the CD28RR to generate the AP mutant or the RE mutant. Double-stranded oligonucleotides were end labeled using T4 polynucleotide kinase (40). Affinity-purified recombinant NF-ATp and truncated Fos and Jun proteins were used in binding reactions with the NF-IL-2A or GM170 radiolabeled oligonucleotides (0.1 ng) in a buffer containing 20 mM HEPES, 5 mM KCl, 20 mM NaCl, 2 mm MgCl2, 2 mM DTT, and 5 μg purified BSA (New England Biolabs, Beverly, MA) (50). Protein:DNA complexes was resolved on a nondenaturing 5% polyacrylamide gel containing 0.5× TBE buffer (0.5 mM Tris, 42 mM boric acid, 1 mM EDTA, pH 8.3). Binding of recombinant proteins to the CD28RR was conducted in 10 mM Tris, pH 7.5, 10 mM MgCl2, 5 mM EDTA, pH 7.5, 10 mM DTT, 0.2% Nonidet P-40, 1% glycerol, 0.4% sucrose, 0.5 mg/ml BSA, and 100 ng poly(dG:dC) in 20 μl reactions. A total of 0.2 ng of radiolabeled probe was generally used with the amounts of recombinant proteins indicated for individual experiments. Reactions were separated on 5% 0.5× TBE polyacrylamide gels. EMSAs for nuclear extract were conducted as previously described except that 5% glycerol was used instead of Ficoll and 0.5 mM PMSF was included. Competitor double-stranded oligonucleotides or Abs was added to the binding reactions at the concentrations indicated in individual experiments and incubated for 10 min before the addition of the radiolabeled probe. All gels were dried and exposed to x-ray film. Digital images were produced on a Fujifilm LAS1000 Plus CCD camera (Tokyo, Japan) and Image Gauge software (Fujifilm).

Western blot analysis

For Western blot, Jurkat cells were transfected, sorted, and stimulated as above, except 1 × 106 cells were recovered from the GFP-positive population. A similar analysis using PBLs was not feasible due to the low transfection efficiency of these cells and the large numbers of cells required for the assay. Nuclei were isolated from cells and HMG I(Y) proteins were extracted with 5% perchloric acid and precipitated with TCA (44). Proteins were separated on 18% denaturing polyacrylamide gels, and HMGI(Y) was detected by Western blot using a 1:250 dilution of a rabbit polyclonal anti-HMGI(Y) Ab (51). The enhanced chemiluminescence detection system (Amersham, Arlington Heights, IL) was used, and the resultant bands were detected by x-ray film and scanned by densitometry (Bio-Rad).

Results

HMGI(Y) binds to multiple elements within the IL-2 promoter

We have previously shown that transcription from a transfected IL-2 promoter reporter construct was inhibited by coexpression of antisense RNA for HMGI(Y) proteins (14). A comparison of the sensitivity of the promoters for a number of T cell-expressed genes (IL-2, IL-3, GM-CSF, and the HIV long terminal repeat (LTR)) to HMGI(Y) depletion found that IL-2 was the most significantly affected by changes in HMGI(Y) levels (data not shown). The IL-2 proximal promoter contains ∼65% A:T base pairs and has many potential binding sites for HMGI(Y) with A:T sequences present within or adjacent to several known transcription factor binding sites. To identify the binding sites for HMGI(Y) on the IL-2 promoter, DNase I footprinting was performed on a fragment of the IL-2 promoter from nt −360 to +50 (Fig. 1⇓). DNase I footprinting was conducted on both DNA strands by labeling each end of the DNA fragment independently. Fig. 1⇓A represents footprints obtained with the DNA labeled at −360, and in Fig. 1⇓B the DNA was labeled at +50. As shown in Fig. 1⇓, A and B, footprints for HMGI(Y) were observed at a number of locations across the proximal promoter (indicated by vertical bars and filled arrowheads). Fig. 1⇓C shows the location of these HMGI(Y) binding sites in relation to the sequence and the known transcription factor binding sites on the promoter. Because different binding sites within the promoter have different affinities for HMGI(Y) (data not shown), only those sites that are consistently seen in independent replicate experiments are indicated. The HMGI(Y) footprints are within or adjacent to the −45 NF-AT/TATA-2 site (11), the NF-IL-2A region, the NF-IL-2B region, the CD28 response region (CD28RR), and the −285 NF-AT site in ARRE-2. All of these elements have been shown to be important in IL-2 promoter activity (reviewed in Refs. 3 and 5). Interestingly, at the concentrations used in this assay, HMGI(Y) was also observed to specifically footprint at the −30 TATA-1 sequence (Fig. 1⇓), a region of the IL-2 promoter that is known to be protected in vivo by proteins in both unstimulated and stimulated T cells (21). As a control for these experiments, a Fos/Jun (AP-1) heterodimer was also footprinted to the promoter DNA, both alone and in combination with the HMGI(Y) protein (Fig. 1⇓, A and B). AP-1 was used at two different concentrations, which were empirically determined, in Fig. 1⇓, A and B. As expected, at the high concentration of input protein used in Fig. 1⇓A, the Fos/Jun complex altered the footprinting pattern across the entire promoter but specific footprints were seen across known AP-1 sites in the promoter region (indicated by vertical dashed lines), including the low-affinity NF-IL-2A (ARRE-1) site located between nt −80 and −90. These specific footprints were confirmed in Fig. 1⇓B at lower concentrations of AP-1. However, when both HMGI(Y) and Fos/Jun were combined together in the reaction mixture, binding of Fos/Jun to the NF-IL-2A site and the CD28RR was considerably reduced, and this is most clearly seen at the higher concentrations of protein used in Fig. 1⇓A (see below).

           FIGURE 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
FIGURE 1.

A and B, DNase I footprinting of HMGI(Y) and AP-1 on IL-2 promoter DNA. The DNA fragment was labeled at either −360 (A) or +50 (B). The proteins added are indicated at the top of the gels with a Maxam-Gilbert “G” chemical cleavage of the naked DNA serving as a sequence reference marker. A+H represents the simultaneous addition of HMGI(Y) and AP-1. Lanes 1–5 and 6–10 are two different loadings of the same samples on the sequencing gel that have been electrophoresed for different lengths of time. The nucleotide numbers relative to the major transcriptional start site are indicated next to the G lanes. The recognition sequences for known protein transcription factors and vertical lines next to the nucleotide numbers indicate functional elements. A vertical thick solid line and a filled arrowhead indicate the footprinted areas protected by HMGI, and a vertical broken line indicates those protected by AP-1. C, Sequence of the IL-2 proximal promoter showing HMGI(Y) footprints relative to known transcription factor binding sites. The HMGI(Y) footprints that are consistently observed are marked by lower case letters in the sequence and filled dots underneath. The sequence is numbered relative to the transcription start site of +1. Horizontal lines above the sequence indicate transcription factor binding sites and control elements.

HMGI(Y) modulates AP-1 and NF-AT binding to the IL-2 promoter

HMGI(Y) has previously been shown to promote or inhibit transcription factor binding to their regulatory elements overlapping or adjacent to the HMGI(Y) binding sites (reviewed in Ref. 29). We have previously shown that HMGI(Y) can modulate the binding of c-Rel to the CD28RR on the IL-2 promoter. We wished to determine whether HMGI(Y) affected the binding if other transcription factors to the IL-2 promoter. Several of the regions on the IL-2 promoter to which HMGI(Y) binds have been described as NF-AT and AP-1 binding sites (reviewed in Refs. 3 and 4). We chose to examine the effect of HMGI(Y) on NF-AT and AP-1 binding to two of these sites because of their distinct properties. The NF-IL-2A (ARRE-1) region has previously been reported to cooperatively bind NF-AT and AP-1 to form a higher order NF-AT complex on the NF-IL-2A region (11). The second site we examined, the CD28RR, has also been reported to bind NF-ATp (11, 46) and AP-1, but the binding sites are in a distinct configuration and cooperative binding is not observed (46).

At low protein concentrations, recombinant truncated Fos/Jun alone does not stably bind to the low-affinity AP-1 site in the NF-IL-2A site in EMSA (Fig. 2⇓A, lane 1). However, the addition of recombinant HMGI(Y) resulted in the appearance of an AP-1 band (Fig. 2⇓A, lanes 2–5). The intensity of the AP-1 band initially increased (up to 2 ng HMGI(Y)) and then decreased with the addition of increasing amounts of HMGI(Y) (Fig. 2⇓A). These results demonstrating that high concentrations of HMGI(Y) inhibit Fos/Jun binding are consistent with the previously described footprinting results obtained with varying concentrations of these proteins (Fig. 1⇑A). Recombinant NF-ATp protein was added to determine the effect of HMGI(Y) proteins on formation of the NF-AT/AP-1 complex. Under the binding conditions used here, NF-ATp did not bind alone to the probe and the formation of the higher-order NF-AT complex did not occur (Fig. 2⇓A, lane 6). Addition of HMGI(Y) protein resulted in a dose-dependent increase in formation of the NF-AT complex as well as the previously observed increase in AP-1 binding (Fig. 2⇓A, lanes 7–10). At higher levels of HMGI(Y), the intensity of the NF-AT complex did not significantly decline as was seen for AP-1 alone (Fig. 2⇓A, lanes 7–10). The amount of AP-1 and NF-ATp used in these experiments was sufficient for strong binding to the consensus sites contained in an oligonucleotide from the GM-CSF enhancer (GM170) that can bind NF-ATp or AP-1 alone and also form a higher-order complex (Fig. 2⇓A, lanes 11–15). This oligonucleotide showed no binding of HMGI(Y), and neither the binding of AP-1 or NF-ATp alone nor the formation of the NF-AT higher-order complex were affected by addition of HMGI(Y) at any concentration (Fig. 2⇓A, lanes 11–15).

           FIGURE 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
FIGURE 2.

A, HMGI(Y) promotes AP-1 and NF-ATp binding to the NF-IL-2A region of the IL-2 promoter. EMSAs were performed using radiolabeled oligonucleotides spanning the NF-IL-2A region of IL-2 (lanes 1–10) or the GM170 region of the GM-CSF enhancer (lanes 11–15) and recombinant truncated NF-ATp (0.4 ng), truncated c-Fos/c-Jun (tAP-1) (2 ng), and HMGI(Y) (1, 2, 4, and 8 ng) proteins. The results were visualized using a Molecular Dynamics PhosphorImager. The proteins and probes used in the individual lanes are indicated above the lanes, and the positions of the free DNA and the individual protein:DNA complexes are indicated. B, HMGI(Y) promotes AP-1 but inhibits NF-ATp binding to the IL-2 CD28RR. EMSAs were performed as above using full-length recombinant c-Fos and c-Jun (AP-1) at 12 ng (lanes 1, 4, and 7), 6 ng (lanes 2, 5, and 8), or 3 ng (lanes 3, 6, and 9) and the CD28RR radiolabeled oligonucleotide. In lanes 10–13, 0.1 ng NF-ATp was added with 3 ng AP-1. HMGI protein was added to the binding reactions at 2 ng (lanes 4–6), 20 ng (lanes 7–9), or 5 (lane 11), 10 (lane 12), or 20 ng (lane 13). EMSAs were processed, and the diagram was labeled as in A.

Recombinant full-length c-Fos/c-Jun (AP-1) bound to the CD28RR in a dose-dependent manner, without the addition of HMGI(Y) (Fig. 2⇑B, lanes 1–3). An increase in AP-1 binding was observed with the addition of HMGI(Y) (Fig. 2⇑B, compare lanes 2, 5, and 8 or lanes 3, 6, and 9), but relatively high levels of HMGI(Y) were required to observe this effect (20 ng compared with 2 ng for the promotion of AP-1 binding to the ARRE-1 region). We have previously shown that HMGI(Y) promotes the binding of c-Rel to an adjacent site (CD28RE) in the CD28RR (14). NF-AT proteins can also bind to the CD28RR, and we have previously shown that HMGI(Y) can promote truncated recombinant NF-ATp binding to the CD28RR (46). Therefore, we examined whether HMGI(Y) could promote the formation of an NF-ATp/AP-1 complex on this site. At the concentrations used to promote AP-1 binding here (20 ng), the binding of NF-ATp was in fact inhibited (Fig. 2⇑B, lanes 10–13). Similar results were observed whether full-length AP-1 (Fig. 2⇑C) or truncated AP-1 (data not shown) was used in the binding reactions.

These results show that HMGI(Y) can modulate the binding of transcription factors that are required for IL-2 promoter function to at least two major control regions of the IL-2 proximal promoter. However, it appears that the ratio of HMGI(Y) to AP-1 or NF-ATp may be important in determining the functional outcome of the interaction.

HMGI(Y) modulates nuclear AP-1 binding to the CD28RR

The CD28RR consists of an AP-1 binding site as well as a c-Rel binding site (CD28RE). We have previously shown that the c-Rel-containing complexes from activated T cell nuclear extracts require HMGI(Y) to bind to the CD28RE (14). To determine whether nuclear AP-1 was also dependent on HMGI(Y), we examined binding of AP-1 from nuclear extracts of P/I/CD28-activated Jurkat T cells. Because the Rel and AP-1 complexes migrate at the same position on EMSA gels using the intact CD28RR (17), it was necessary to generate EMSA probes with mutations in either the CD28RE (RE mutant) or the AP-1 site (AP mutant). Binding of nuclear extracts to these probes showed that an inducible complex could bind to the RE mutant in extracts made both 1 and 6 h following stimulation with P/I/CD28 (Fig. 3⇓A, lanes 4–6). These complexes migrated at the same position as the complexes binding to the wild-type probe or the Rel-containing complexes binding to the AP mutant (Fig. 3⇓A, lanes 1–3 and lanes 7–9). The identity of the complexes binding to the RE mutant was confirmed by competition experiments where the RE mutant, the wild-type CD28RR, and a consensus AP-1 site (TRE) were able to compete for complex formation but the AP mutant was not (Fig. 3⇓B). Thus, these complexes contain AP-1-like proteins. To determine whether the formation of the AP-1-like complex was dependent on HMGI(Y), an anti-HMGI(Y) Ab was added to the binding reactions. The addition of the HMGI(Y) Ab to the binding reactions reduced AP-1 binding to the RE mutant (Fig. 3⇓C, lanes 1–3), as it did the c-Rel complexes binding to the AP mutant (Fig. 3⇓C, lanes 5–7). A control Ab (C) did not affect complex formation (Fig. 3⇓C, lanes 4, 8, and 11). We have previously shown that HMGI(Y) Ab removes the inducible complexes containing c-Rel from the CD28RR but does not affect NF-κB complexes binding to a distinct NF-κB site (14). AP-1 binding to a TRE was not significantly reduced by the addition of the HMGI(Y) Ab (Fig. 3⇓C, lanes 9 and 10).

           FIGURE 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
FIGURE 3.

Binding of AP-1 to the CD28RR from Jurkat T cell extracts requires HMGI(Y). A, EMSAs were performed with nuclear extracts prepared from Jurkat T cells either unstimulated (lanes 1, 4, and 7) or activated with P/I/CD28 for 1 h (lanes 2, 5, and 8) or 6 h (lanes 3, 6, and 9) using radiolabeled CD28RR (lanes 1–3), RE mutant (lanes 4–6), or AP mutant lanes 7–9) oligonucleotides. The inducible protein DNA complexes are indicated by an arrow. B, Competition experiments using the RE mutant as the radiolabeled probe and extracts from cells activated for 6 h. The CD28RR wild-type sequence (lanes 2–4), AP mutant (lanes 5–7), RE mutant (lanes 8–10), or a TRE (AP-1 consensus binding site) (lanes 11–13) were used as competitors at 1, 10, and 20 ng per reaction as indicated. C, The effect of addition of HMGI(Y) Ab on complexes binding to the RE mutant (lanes 1–4), the AP mutant (lanes 5–8), and the TRE (lanes 9–11). The HMGI(Y) Ab (α-H; 1 μl (lanes 2 and 6) or 2 μl (lanes 3, 7, and 10)) or a nonspecific serum (C; 2 μl) was added to the binding reactions before the addition of probe. The AP-1 complexes are indicated. In A–C, only those parts of the gel containing the inducible bands are shown. D, HMGI promotes recombinant AP-1 binding to the RE mutant. EMSAs containing 3 ng recombinant AP-1 and increasing concentrations of HMGI (0.5, 1, 2.5, 5, 10, and 20 ng) were performed with either the wild-type CD28RR probe (lanes 1–8) or the RE mutant (lanes 9–16). Lanes 1 and 9 had no protein added, and lanes 2 and 10 had only AP-1. The positions of the HMGI and AP-1 bands are indicated.

To confirm that HMGI(Y) could promote AP-1 binding to the RE mutant, recombinant HMGI was titrated into binding reactions with a fixed amount of AP-1 using either the wild-type CD28RR or the RE mutant. As described above, AP-1 binding to the CD28RR was increased only at high concentrations of HMGI (Fig. 3⇑D, lanes 7 and 8). In contrast, a dose-dependent increase in AP-1 binding was observed with increasing levels of HMGI starting at 1 ng on the RE mutant (Fig. 3⇑D, lanes 9–16). We also observed that HMGI(Y) no longer bound to the RE mutant (Fig. 3⇑D), probably because the RE mutant affects the A:T stretch within the CD28RE to which HMGI(Y) most likely binds. These results 1) show that AP-1 binding in nuclear extracts is influenced by HMGI(Y) and 2) imply that the ability of HMGI(Y) to promote AP-1 binding is not dependent on DNA binding by HMGI(Y).

HMGI(Y) binding to DNA is not required for promotion of transcription factor binding to the CD28RR

The results above imply that the promotion of transcription factor binding may not always require DNA:HMGI(Y) interactions. To test this further, site-specific proline to alanine mutations were introduced into the second and third DNA binding domains of HMGI (two substitutions in each binding domain) to create a form of the protein (designated HMGI(mII,mIII)) that could not specifically bind to A:T-DNA under the assay conditions employed. We tested the ability of HMGI(mII,mIII) to bind to the IL-2 promoter both in footprinting and in gel shift assays. At concentrations of protein where recombinant HMGI or HMGY bound specifically to the CD28RR, HMGI(mII,mIII) protein did not bind (Fig. 4⇓A). The same result was obtained in footprinting assays across the IL-2 promoter from −180 to −60 (Fig. 4⇓B). Four footprints for HMGI were observed across this region of the promoter (Fig. 4⇓B, lanes 2–5). In contrast, HMGI(mII,mIII) did not form specific footprints across any of these regions (Fig. 4⇓B, lanes 6–9).

           FIGURE 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
FIGURE 4.

A and B, HMGI(mII,mIII) does not bind specifically to the IL-2 promoter. A, EMSAs were performed using the IL-2CD28RR oligonucleotide as a probe and 1, 10, and 20 ng of HMGI (lanes 1–3), HMGY (lanes 4–6), and HMGI(mII,mIII) (lanes 7–9). The HMGI/Y bands and the free DNA are indicated. B, DNase I footprinting assay using the −180 to −60 region of the IL-2 promoter. HMGI (lanes 2–5) or HMGI(mII,mIII) (lanes 6–9) at 10, 20, 50, or 100 ng were used in the footprinting reactions. Solid lines indicate the HMGI footprints, and the names of the regions are shown. C and D, HMGI(mII,mIII) can promote but not inhibit transcription factor binding to the IL-2 CD28RR. Increasing amounts (0.5, 1, 2.5, 5, 10, and 20 ng) of HMGI (lanes 4–9) or HMGI(mII,mIII) (lanes 10–15) were added to binding reactions containing the IL-2CD28RR probe and 3 ng AP-1 and 0.1 ng NF-ATp. Reactions containing NF-ATp alone (lane 1), AP-1 alone (lane 2), or AP-1 and NF-ATp together (lane 3) were also analyzed. The positions of the transcription factor:DNA complexes are indicated. D, Similar binding reactions as described in C were conducted using bacterial extract containing the Rel homology domain of c-Rel and increasing amounts (1, 5, 10, and 20 ng) of HMGI (lanes 2–5) or HMGI(mII,mIII) (lanes 6–9). Lane 1 contains c-Rel alone.

We also tested whether HMGI(mII,mIII) could interact with transcription factors by generating an affinity matrix for both HMGI and HMGI(mII,mIII). The results of binding experiments using these affinity matrices in binding experiments demonstrated that both the wild-type HMGI and mutant HMGI(mII,mII) proteins could bind specifically to the AP-1 and NF-AT proteins but could not interact with themselves or with each other (data not shown).

We then investigated whether HMGI(mII,mIII) could alter transcription factor binding when added to binding assays using the CD28RR together with AP-1 and NF-ATp. The addition of HMGI(mII,mIII) lead to a dose-dependent increase in AP-1 binding but did not inhibit NF-ATp binding as did wild-type HMGI (Fig. 4⇑C, lanes 10–15). We observed that HMGI(mII,mIII) generated a more consistent dose dependence compared with the wild-type HMGI (Fig. 4⇑C, compare lanes 4–9 and 10–15). c-Rel binding to the CD28RR was also increased by HMGI(mII,mIII) (Fig. 4⇑D, lanes 1 and 6–9). In addition, inhibition of transcription factor binding sometimes seen at high concentrations of HMGI was never observed with HMGI(mII,mIII). In summary, these results show that HMGI binding to DNA is involved in the inhibition of binding but is not required for the promotion of transcription factor binding.

Alteration of HMGI(Y) expression results in the modulation of IL-2 promoter activity in both Jurkat and primary T cells

To determine whether HMGI(Y) had a functional role in IL-2 gene transcription, HMGI(Y) levels were altered in the cell by antisense or overexpression studies. The pRcCMVIGMH plasmid expressing antisense HMGI was cotransfected into Jurkat T cells with pIL2luc or pHIVluc. The HIV LTR responds to P/I and CD28 stimulation through NF-κB sites that closely match consensus sites and contain no potential binding sites for HMGI(Y) (our unpublished observations).

As previously reported (14), the expression of antisense HMGI RNA resulted in a 72% decrease in reporter activity in P/I-treated Jurkat cells and a 75% inhibition in P/I+α-CD28-treated cells at the maximum dose of antisense expression plasmid (10 μg; Fig. 5⇓A). The pHIVluc reporter transfections showed only a small reduction in promoter activity at the maximum dose of antisense plasmid (Fig. 5⇓A). To confirm that the antisense HMGI RNA was reducing the level of HMGI(Y) protein in the cells, a Western blot, using an anti-HMGI polyclonal Ab, was performed on nuclear extracts from transfected cells either unstimulated or P/I-stimulated following transfection. Densitometry scanning of the blot showed that there was a significant reduction in the levels of HMGI(Y) protein in the antisense-expressing cells (Fig. 5⇓B). This experiment also showed that the amount of HMGI(Y) in the nuclei of Jurkat T cells was increased almost 3-fold by P/I activation (Fig. 5⇓B).

           FIGURE 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
FIGURE 5.

Inhibition of HMGI(Y) production reduces IL-2 promoter activity in Jurkat T cells and PBLs. A, Jurkat T cells were transfected with increasing amounts of pRcCMVIGMH and the pIL-2luc or pHIVluc reporter plasmids. Transfected cells were treated with P/I (□) or P/I+α-CD28 (▪) before luciferase assays were performed. Luciferase activity is expressed as counts per second (CPS) readings from the Packard TopCount luminometer. B, Western blot showing the decrease in HMGI(Y) levels in cells expressing antisense HMGI. Cells were transfected with either pRcCMV (lanes 1 and 3) or pRcCMVIGMH antisense expression plasmid (lanes 2 and 4). Protein was extracted from nuclei of either unstimulated (lanes 1 and 2) or P/I-stimulated (lanes 3 and 4) cells, resolved on SDS polyacrylamide gels, and HMGI(Y) was detected by Western blot. Recombinant HMGI protein (lane 5) was also loaded as a reference. Numbers below represent the densitometry quantitation of HMGI(Y) protein bands. C, PBLs, pretreated to generate blasts, were transfected with the pRcCMV or pRcCMVIGMH plasmids, and the pIL-2luc reporter plasmid and luciferase activity was measured either before (□) or following PHA treatment (▪). Columns represent the means of five replicate assays, and the bars show the SEMs.

HMGI(Y) expression in transformed cell lines is generally high (29), hence the Jurkat T cell leukemia may display an aberrant role for HMGI(Y) proteins in IL-2 promoter activity. In contrast, HMGI(Y) levels are quite low or not detectable in most normal cell types. Therefore, the dependence of the IL-2 promoter on HMGI(Y) proteins for activation of transcription was also examined in normal PBL. PBLs isolated from blood were cultured for 4 days as described to generate blasts, transfected by electroporation, rested for 24 h, and restimulated with PHA. The IL-2 promoter transfected into PBLs had activity in cells that were not stimulated posttransfection most likely because of the primary stimulation with PHA before transfection to generate T cell blasts. However, these cells showed a 2- to 3-fold increase in promoter activity when stimulated with PHA posttransfection (Fig. 5⇑C). Transfection with pRcCMVIGMH resulted in an 84% inhibition of the reporter activity in unstimulated PBLs and an 81% inhibition in PHA-stimulated cells (Fig. 5⇑C). These results show that the activity of a transfected IL-2 promoter is dependent on HMGI(Y) in either a T cell line or in primary T cells.

We also tested whether increased levels of HMGI(Y) protein affected IL-2 promoter activity. Transfection of the pcDNAHMGY expression plasmid into Jurkat T cells together with the pIL-2luc or the pHIVluc increased luciferase activity ∼2-fold for IL-2 but had no effect on HIV LTR activity (Fig. 6⇓A). Because the mutant HMGI(mII,mIII) protein also promoted transcription factor binding, its effect in transfection assays was also tested. Cotransfection of an expression plasmid for the HMGI(mII,mIII) protein resulted in significant inhibition of the IL-2 promoter (60% in P/I and 73% in P/I+α-CD28-treated cells; Fig. 6⇓B) at the highest amounts of transfected plasmid. The pHIVluc reporter plasmid showed no loss of reporter activity with expression of the HMGI(mII,mIII) protein (Fig. 6⇓B). Thus, this non-DNA binding mutant appears to act as a dominant negative protein for IL-2 gene transcription.

           FIGURE 6.
  • Download figure
  • Open in new tab
  • Download powerpoint
FIGURE 6.

A, Expression of HMGY in Jurkat T cells increases IL-2 promoter activity. Jurkat cells were transfected with either the pIL-2luc or the pHIVluc reporter plasmids together with the indicated amounts of the pcDNAHMGY expression plasmid or the control pcDNA3.1 plasmid. Cells were stimulated with P/I (□) or P/I+α-CD28 (▪) for 8 h before harvesting for luciferase assays. pHIVluc transfected cells were only stimulated with P/I+α-CD28. Luciferase activity is expressed as counts per second (CPS) as measured by a Packard TopCount scintillation counter. B, Expression of HMGI(mII,mIII) in Jurkat cells inhibits IL-2 promoter activity. The pIL-2luc or pHIVluc reporter plasmids were transfected into Jurkat T cells together with the indicated amounts of the pcDNAHMGI(mII.mIII) expression plasmid or the control pcDNA plasmid. Cells were activated and analyzed, and results are presented as described in A.

Taken together these results show that modulating HMGI(Y) levels or function in either Jurkat T cells or PBLs affects IL-2 promoter activity without a general effect on transcription as measured by HIV LTR function.

IL-2 production and primary T cell proliferation are dependent on HMGI(Y)

It is possible that the transfected IL-2 promoter may be more highly dependent on HMGI(Y) than its endogenous counterpart because of the likely differences in chromatin configuration on chromosomal and plasmid DNA. To monitor the effect of HMGI(Y) on expression from the endogenous IL-2 gene, the amount of IL-2 protein secreted into the supernatant of cells transfected with pRcCMVIGMH and pcDNAHMGI(mII,mIII) was assayed by ELISA. To enrich for transfected cells, 5 μg of pCMVGFP was cotransfected with the HMGI plasmids above, and the cells were subsequently sorted for high level GFP expression by FACS. The sorted cells were then stimulated with either PHA for PBLs or P/I+α-CD28 for Jurkat cells to induce IL-2 expression. In PBLs and Jurkats, significant inhibition of IL-2 production was detected in cells transfected with RcCMVIMGH (78 and 88%, respectively) or pcDNAHMGI(mII,mIII) (47 and 55%, respectively) (Table I⇓). These results imply that the role of HMGI(Y) may be extended to endogenous IL-2 gene transcription in both Jurkat T cells and primary T cells.

View this table:
  • View inline
  • View popup
Table I.

Decrease in IL-2 production with inhibition of HMGI(Y)

If IL-2 production in PBLs is dependent on HMGI(Y), then PBL proliferation may also be affected by changes in HMGI(Y) levels in the cells. To test the effect of inhibition of HMGI(Y) on cell growth, the IL-2-dependent PBLs and the IL-2-independent Jurkat cell line were transfected with either the parent plasmids or the expression plasmids RcCMVIGMH and pcDNAHMGI(mII,mIII) together with the pCMVGFP plasmid to allow sorting of transfected cells by FACS. Proliferation of PBLs was inhibited by RcCMVIGMH (62%) or pcDNAHMGI(mII,mIII) (29%), whereas the effect of these plasmids on growth of the Jurkat cell line was not significant (Fig. 7⇓, A and B). To test whether increased HMGI(Y) levels also altered proliferation, PBLs were transfected with the pcDNAHMGY plasmid and pCMVGFP to allow for sorting of transfected cells. Overexpression of HMGY in PBLs increased proliferation of these cells by ∼2-fold (Fig. 7⇓C).

           FIGURE 7.
  • Download figure
  • Open in new tab
  • Download powerpoint
FIGURE 7.

Altering HMGI(Y) levels or function affects T cell proliferation. PBLs and Jurkat cells were transfected with either 10 μg of the antisense expression plasmid pRcCMVIGMH or the parent plasmid pRcCMV (A), and pcDNAHMGI(mII,mIII) or the parent plasmid, pcDNA3.1 (B) PBLs were also transfected with the pcDNAHMGY plasmid or the parent plasmid (C). All cells were cotransfected with 5 μg of pCMVGFP and subsequently sorted by FACS for GFP expression. Cells were then stimulated with PHA in the presence of bromodeoxyuridine and assayed for proliferation using enzyme-conjugated Ab to bromodeoxyuridine and a chemiluminescent substrate. Values are given as counts per second measured by a 96-well plate luminometer, and the columns represent the mean, and error bars the SEM, of four replicate assays.

These results show that IL-2 production in both PBLs and Jurkats is dependent on the correct level of HMGI(Y). This dependence can also be seen at the level of proliferation in PBLs that are IL-2 dependent but not in Jurkats that are IL-2 independent.

Discussion

The expression of the IL-2 and IL-2Rα-chain genes is an essential component of the formation of the autocrine loop that drives T cell proliferation and clonal expansion following an immune stimulus. We have shown here that HMGI(Y), a protein involved in regulating promoter architecture, is required for IL-2 promoter activity and T cell proliferation. This requirement for proliferation appears to stem from the involvement of HMGI(Y) in controlling transcription from both the IL-2 and IL-2Rα-chain genes in response to mitogenic stimuli. It has previously been shown that HMGI(Y) participates in the inducible expression of the human IL-2Rα gene by facilitating the assembly of multiprotein, enhanceosome-like complexes on both an upstream enhancer element that has binding sites for HMGI(Y), Elf-1 (an Ets family protein), Stat5, and a GATA family protein (52) and on the proximal promoter element that contains binding sites for HMGI(Y), Elf-1, and NF-κB proteins (53). Likewise, the present results obtained from both in vitro footprinting, EMSA, and transfection experiments strongly suggest that HMGI(Y) facilitates the formation of a functional multiprotein complex consisting of a number of different transcription factors on the proximal promoter of the IL-2 gene during transcriptional activation in vivo.

Here, and elsewhere (14), we have demonstrated that HMGI(Y) modulates the binding to the IL-2 promoter of NF-AT, AP-1, and c-Rel; three transcription factor families that play important roles in IL-2 promoter activity. This has been shown both for recombinant proteins and proteins present in nuclear extracts. In each case, the transcription factor binding site constitutes a nonconsensus element that differs markedly in sequence from the high-affinity consensus sites. This is generally, but not always, because of the presence of the A:T sequences required for HMGI(Y) binding; consequently, these sites are weakly recognized by their cognate transcription factors. Therefore, HMGI(Y) serves to lower the threshold at which transcription factors can bind to and activate these weak sites. The presence of these nonconsensus sites in the IL-2 promoter appears to be important for the T cell-restricted expression of the promoter. It has been shown that mutation of some of these sites to consensus high-affinity sites weakens the induction dependence or T cell specificity of the promoter (54, 55, 56). There is no evidence that there are T cell-restricted members of the transcription factor families that may have selectivity for the IL-2 promoter sites. Instead, it is likely that the weak interactions together with the need for proteins such as HMGI(Y) and cooperative binding of many of the protein complexes is a requirement to generate these characteristics. Thus, HMGI(Y) may play an indirect role in the T cell specificity and induction dependence of the IL-2 promoter.

The role of HMGI(Y) in the assembly of a functional enhanceosome has been best studied for the IFN-β promoter (57). It has been shown that HMGI(Y) promotes the coordinate binding of members of the NF-κB, activating transcription factor, and IFN-γ regulatory factor families of transcription factors to the IFN-β promoter leading to the assembly of a functional complex known as an enhanceosome (57). This enhanceosome is then thought to recruit coactivator complexes such as CREB binding protein to lead to chromatin reorganization and transcription activation (58). It has recently been shown that the recruitment of transcription factors to the IFN-β promoter requires the binding of HMGI(Y) to DNA and that HMGI(Y):transcription factor interaction is not crucial for this recruitment (59). The IFN-β promoter DNA contains an intrinsic bend that is straightened by HMGI(Y) binding, and this is important to allow transcription factor binding (27). However, HMGI(Y):transcription factor interactions are required for the completion of the enhanceosome assembly process on the IFN-β gene (59). We have found here that binding of either c-Rel or c-Fos/c-Jun to the CD28RR of the IL-2 promoter does not require HMGI(Y) binding to DNA. We have shown this by using either a mutant of the CD28RR that can no longer bind HMGI(Y) or a mutant HMGI protein that does not bind DNA specifically. However, the non-DNA binding mutant of HMGI can still interact with the NF-ATp and AP-1 transcription factors and may promote transcription factor binding by protein:protein interactions. The intrinsic structure of the IL-2 promoter region is not known but it is possible that the promoter DNA does not require structural alteration to allow transcription factor binding. However, the HMGI(mII,mIII) non-DNA binding mutant acted as a dominant negative protein in transfection assays. This result implies that correct DNA binding as well as protein recruitment is important for the assembly of a functional complex. When expressed at sufficiently high levels, the dominant negative mutant may compete with wild-type protein in the cells to form nonfunctional complexes. Thus, it is likely that for both the IL-2 and IFN-β promoters, HMGI(Y) needs to both bind to DNA and interact with transcription factors to generate a functional complex, although the order of events may differ.

High concentrations of HMGI appear to inhibit the binding of certain transcription factors at specific sites, e.g., AP-1 to the NF-IL-2A site and NF-ATp to the CD28RR. This inhibition was dependent on DNA binding by HMGI and would appear to be a direct competitive effect. It has previously been shown that HMGI(Y) can inhibit NF-AT binding to the IL-4 promoter (60, 61). In the case of IL-4, this inhibition of DNA binding translates into the ability of increased HMGI(Y) to inhibit IL-4 promoter function (61). In contrast, our results strongly suggest that HMGI(Y) is an activator of the IL-2 promoter. The data that support this conclusion are 1) antisense HMGI RNA expression strongly inhibits IL-2 promoter activity in both Jurkat T cells and PBLs, 2) increased expression of HMGI, either in response to P/I activation or by cotransfection of an expression plasmid, correlates with increased IL-2 promoter activity, and 3) production of IL-2 from the endogenous IL-2 gene is inhibited by antisense HMGI RNA in both Jurkat T cells and PBLs. These experiments not only show that a promoter in the context of a reporter plasmid, but also the endogenous gene, in both a T cell line and normal T cells is dependent on HMGI(Y) for activity. The ability of HMGI(Y) to promote transcription factor binding may be an important aspect of the positive function of HMGI(Y) on the IL-2 promoter. Whether inhibition of transcription factor binding is an artifact of high concentrations of HMGI or plays a role in the removal of inhibitors or down-regulation of the promoter following activation remains to be determined. One possible explanation for the results observed here is that at high concentrations of HMGI(Y) following T cell activation NF-ATp is displaced by HMGI(Y), which then promotes the binding of c-Rel. The appearance of NF-ATp in the nucleus (4) at early times following activation is likely to proceed the increased levels of c-Rel that requires new protein synthesis (12). A time-dependent change in the proteins that bind to the CD28RR may play a role in correct activation.

We have previously shown that the GM-CSF promoter (14) and the IL-3 promoter (our unpublished observations) are dependent on HMGI(Y) for activity. All of these cytokines are expressed following a primary T cell activation, before the development of effector T cell function (62, 63). In contrast, IL-4 requires at least three cell divisions before expression is detected (64, 65) and is the hallmark of the Th2 effector phenotype (62, 63). Signaling from the IL-4 receptor has also been shown to lead to phosphorylation of HMGI(Y) and a consequent reduction in DNA binding (66). Because IL-4 activation of T cells leads to increased IL-4 synthesis, it is intriguing to speculate that the phosphorylation of HMGI(Y) and its removal from the DNA may be one mechanism by which this increased IL-4 production is achieved (61). The same signaling pathway may consequently decrease IL-2 production. It may be of interest to examine the role of HMGI(Y) in expression of other Th1- or Th2-specific cytokines and to test the consequences of different levels of HMGI(Y) on the expression of these cytokines and the development of the T cell subtypes.

Alteration of the level of HMGI(Y) in PBLs leads to an effect on proliferation of these cells. It would appear that the expression of antisense HMGI RNA does not have a general effect on the cell cycle machinery because antisense HMGI RNA expression had little effect on the proliferation of Jurkat T cells. This change in proliferation in PBLs is likely to be the result of changes in IL-2 expression in these cells, although we have not definitively proven this link. Attempts to rescue the proliferative defect in HMGI antisense RNA-expressing cells lead to only a small reversal of this inhibition. This may be due to the fact that the expression of the IL-2Rα-chain gene is also controlled by HMGI(Y) as discussed above. The IL-2Rα-chain is required for high-affinity IL-2 receptor complex formation (reviewed in Ref. 67). Indeed, the PBLs expressing antisense HMGI RNA had a reduced level of IL-2Rα-chain on the cell surface as measured by FACS analysis (data not shown).

We have found here that HMGI(Y) levels increase in Jurkat T cells in response to P/I treatment. PMA treatment has previously been shown to increase HMGI(Y) levels in other cell types (68, 69), as have other signals such as exposure to growth factors such as serum (70, 71) and epidermal growth factor (72) or proinflammatory agent such as endotoxin and IL-1β (73). The increase observed in HMGI(Y) levels may directly translate into increased promoter activity mediated by HMGI(Y) interactions with the other inducible transcription factors such as c-Rel or AP-1.

The results presented here show that the architectural transcription factor HMGI(Y) is critical for the correct regulation of IL-2. The finding that HMGI(Y) not only affects the function of a transfected IL-2 promoter but also the endogenous IL-2 gene and T cell proliferation implies that the level of HMGI(Y) in cells is a critical determinant of IL-2 gene activation and that modulating such levels either by physiological or pharmacological means may provide a means of modulating the immune response.

Acknowledgments

We thank Dr. D. Tremethick for supplying us with recombinant Fos and Jun proteins and also for many helpful discussions during the course of these experiments. We also thank Dr. Peter Cockerill for recombinant NF-ATp and truncated fos and Jun proteins and Dr. Binks Wattenberg for the pCMVGFP plasmid.

Footnotes

  • ↵1 This work was supported, in part, by National Institutes of Health Grant RO1-GM46352 (to R.R) and a National Health and Medical Research Council (Australia) Program Grant (to M.F.S.).

  • ↵2 Current address: Department of Microbiology, University of Queensland, Brisbane, Australia.

  • ↵3 Address correspondence and reprint requests to Dr. M. Frances Shannon, Division of Biochemistry and Molecular Biology, John Curtin School of Medical Research, Australian National University, P.O. Box 334, Canberra ACT 2601, Australia. E-mail address: frances.shannon{at}anu.edu.au

  • ↵4 Abbreviations used in this paper: ARRE, Ag response element; HMGI(Y), high-mobility group protein I or Y; CD28RR, CD28 response region; CD28RE, CD28 response element; P/I, PMA and calcium ionophore; LTR, long terminal repeat; GFP, green fluorescence protein.

  • Received September 22, 1999.
  • Accepted January 4, 2000.
  • Copyright © 2000 by The American Association of Immunologists

References

  1. ↵
    Janeway, C. A., K. Bottomly. 1994. Signals and signs for lymphocyte responses. Cell 76: 275
    OpenUrlCrossRefPubMed
  2. ↵
    Lanzavecchia, A., G. Lezzi, A. Viola. 1999. From TCR engagement to T cell activation: a kinetic view of T cell behavior. Cell 96: 1
    OpenUrlCrossRefPubMed
  3. ↵
    Serfling, E., A. Avots, M. Neumann. 1995. The architecture of the interleukin-2 promoter: a reflection of T lymphocyte activation. Biochim. Biophys. Acta 1263: 181
    OpenUrlPubMed
  4. ↵
    Rao, A., C. Luo, P. Hogan. 1997. Transcription factors of the NFAT family: regulation and function. Annu. Rev. Immunol. 15: 707
    OpenUrlCrossRefPubMed
  5. ↵
    Jain, J., C. Loh, A. Rao. 1995. Transcriptional regulation of the IL-2 gene. Curr. Opin. Immunol. 7: 333
    OpenUrlCrossRefPubMed
  6. ↵
    Boise, L. H., B. Petryniak, X. Mao, C. H. June, C-Y. Wang, T. Lindsten, R. Bravo, K. Kovary, J. M. Leiden, C. B. Thompson. 1993. The NFAT-1 DNA binding complex in activated T cells contains Fra-1 and JunB. Mol. Cell Biol. 13: 1911
    OpenUrlAbstract/FREE Full Text
  7. ↵
    Brabietz, T., I. Pietrowski, E. Serfling. 1991. The immunosupressive FK506 and cyclosporin A inhibit the generation of protein factors binding to two purine boxes of the interleukin-2 enhancer. Nucleic Acids Res. 19: 61
    OpenUrlAbstract/FREE Full Text
  8. ↵
    Jain, J., P. J. McCaffrey, V. E. Valge-Archer, A. Rao. 1992. Nuclear factor of activated T cells contains Fos and Jun. Nature 356: 801
    OpenUrlCrossRefPubMed
  9. ↵
    Ullman, K. S., W. M. Flanagen, C. A. Edwards, G. R. Crabtree. 1991. Activation of early gene expression in T lymphocytes by Oct-1 and an inducible protein, OAP40. Science 254: 558
    OpenUrlAbstract/FREE Full Text
  10. ↵
    Ullman, K. S., J. P. Northorp, A. Admon, G. R. Crabtree. 1993. Jun family members are controlled by a calcium-regulated, cyclosporin A-sensitive signaling pathway in activated T lymphocytes. Genes Dev. 7: 188
    OpenUrlAbstract/FREE Full Text
  11. ↵
    Rooney, J. W., Y-L. Sun, L. H. Glimcher, T. Hoey. 1995. Novel NFAT sites that mediate activation of the interleukin-2 promoter in response to T-cell receptor stimulation. Mol. Cell Biol. 15: 6299
    OpenUrlAbstract/FREE Full Text
  12. ↵
    Bryan, R. G., Y. Li, J-H. Lai, M. Van, N. C. Rice, R. R. Rich, T-H. Tan. 1994. Effect of CD28 signal transduction on c-Rel in human peripheral blood T cells. Mol. Cell Biol. 14: 7933
    OpenUrlAbstract/FREE Full Text
  13. ↵
    Kontgen, F., R. J. Grumont, A. Strasser, D. Metcalf, M. Li, D. Tarlinton, S. Gerondakis. 1995. Mice lacking the c-rel protooncogene exhibit defects in lymphocyte proliferation, humoral immunity and interleukin-2 expression. Genes Dev. 9: 1965
    OpenUrlAbstract/FREE Full Text
  14. ↵
    Himes, S. R., L. S. Coles, R. Reeves, M. F. Shannon. 1996. High-mobility-group protein I(Y) is required for function and for c-Rel but not Rel A binding to CD28 response elements within the GM-CSF and IL-2 promoters. Immunity 5: 479
    OpenUrlCrossRefPubMed
  15. ↵
    Su, B., E. Jacinto, M. Hibi, T. Kallunki, M. Karin, Y. Ben-Neriah. 1994. JNK is involved in signal integration during costimulation of T lymphocytes. Cell 77: 727
    OpenUrlCrossRefPubMed
  16. ↵
    Rincon, M., R. A. Flavell. 1994. AP-1 transcriptional activity requires both T-cell receptor-mediated and costimulatory signals in primary T lymphocytes. EMBO J. 13: 4370
    OpenUrlPubMed
  17. ↵
    McGuire, K. L., M. Iacobelli. 1997. Involvement of Rel, Fos and Jun proteins in binding activity to the IL-2 promoter CD28 response element/AP-1 sequence in human T cells. J. Immunol. 159: 1319
    OpenUrlAbstract
  18. ↵
    Butscher, W. G., C. Powers, M. Olive, C. Vinson, K. Gardner. 1998. Coordinate transactivation of the interleukin-2 CD28 response element by c-Rel and ATF-1/CREB2. J. Biol. Chem. 273: 552
    OpenUrlAbstract/FREE Full Text
  19. ↵
    Parra, E., K. McGuire, G. Hedlund, M. Dohlsten. 1998. Overexpression of p65 and c-Jun substitute for B7-1 costimulation by targeting the CD28RE within the IL-2 promoter. J. Immunol. 160: 5374
    OpenUrlAbstract/FREE Full Text
  20. ↵
    Shapiro, V. S., K. E. Truitt, J. B. Imboden, A. Weiss. 1997. CD28 mediates transcriptional upregulation of the IL-2 promoter through a composite CD28RE and NF-IL-2B AP-1 sites. Mol. Cell Biol. 17: 4051
    OpenUrlAbstract/FREE Full Text
  21. ↵
    Brunvand, M. W., A. Krumm, M. Groudine. 1993. In vivo footprinting of the human IL-2 gene reveals a nuclear factor bound to the transcription start site in T cells. Nucleic Acids Res. 21: 4824
    OpenUrlAbstract/FREE Full Text
  22. ↵
    Garrity, P. A., D. Chen, E. V. Rothenberg, B. J. Wold. 1994. Interleukin-2 transcription is regulated in vivo at the level of coordinate binding of both constitutive and regulated factors. Mol. Cell Biol. 14: 2159
    OpenUrlAbstract/FREE Full Text
  23. ↵
    Rothenberg, E. V., S. B. Ward. 1996. A dynamic assembly of diverse transcription factors integrates activation and cell-type information for interleukin 2 gene regulation. Proc. Natl. Acad. Sci. USA 93: 9358
    OpenUrlAbstract/FREE Full Text
  24. ↵
    Ward, S. B., G. Hernandez-Hoyos, F. Chen, M. Waterman, R. Reeves, E. V. Rothenberg. 1998. Chromatin remodeling of the interleukin-2 gene: distinct alterations in the proximal versus distal enhancer regions. Nucleic Acids Res. 12: 2923
    OpenUrl
  25. ↵
    Thanos, D., T. Maniatis. 1995. Virus inducton of human IFNβ gene expression requires the assembly of an enhanceosome. Cell 83: 1091
    OpenUrlCrossRefPubMed
  26. ↵
    Thanos, D., T. Maniatis. 1992. The high mobility group protein HMGI(Y) is required for NF-κB-dependent virus induction of the human IFN-β gene. Cell 71: 777
    OpenUrlCrossRefPubMed
  27. ↵
    Flavo, J. V., D. Thanos, T. Maniatis. 1995. Reversal of intrinsic DNA bends in the IFNβ gene enhancer by transcription factors and the architectural protein HMGI(Y). Cell 83: 1101
    OpenUrlCrossRefPubMed
  28. ↵
    Yie, J., S. Liang, M. Merika, D. Thanos. 1997. Intra- and intermolecular cooperative binding of high mobility group protein I(Y) to the β-interferon promoter. Mol. Cell Biol. 17: 2649
    OpenUrlAbstract/FREE Full Text
  29. ↵
    Bustin, M., R. Reeves. 1996. High mobility-group chromosomal proteins: architectural components that facilitate chromatin function. Prog. Nucl. Acid Res. Mol. Biol. 54: 35
    OpenUrlCrossRefPubMed
  30. ↵
    Grosschedl, R., K. Giese, J. Page. 1994. HMG domain proteins: architectural elements in the assembly of nucleoprotein structures. Trends Genet. 10: 94
    OpenUrlCrossRefPubMed
  31. ↵
    Lehn, D., T. Elton, K. Johnson, R. Reeves. 1988. A conformational study of the sequence specific binding of HMG-I(Y) with the bovine interleukin-2 cDNA. Biochem. Int. 16: 963
    OpenUrlPubMed
  32. ↵
    Nissen, M. S., R. Reeves. 1995. Changes in superhelicity are introduced into closed circular DNA by binding of high mobility group protein I/Y. J. Biol. Chem. 270: 4355
    OpenUrlAbstract/FREE Full Text
  33. ↵
    Reeves, R., M. S. Nissen. 1995. Cell cycle regulation and functions of HMG-I(Y). Prog. Cell Cycle Res. 1: 337
    OpenUrl
  34. ↵
    Reeves, R., M. Nissen. 1993. Interaction of nohistone proteins HMG-I(Y) with nucleosome core particles. J. Biol. Chem. 268: 31137
    OpenUrl
  35. ↵
    Reeves, R., A. P. Wolffe. 1996. Substrate structure influences binding of the non-histone protein HMG-I(Y) to free and nucleosomal DNA. Biochemistry 35: 5063
    OpenUrlCrossRefPubMed
  36. ↵
    Kas, E., L. Poljak, Y. Adachi, U. K. Laemmli. 1993. A model for chromatin opening: stimulation of topoisomerase II and restriction enzyme cleavage of chromatin by distamycin. EMBO J. 12: 115
    OpenUrlPubMed
  37. ↵
    Zhoa, K., E. Kas, E. Gonzalez, U. K. Laemmli. 1993. SAR-dependent mobilization of histone H1 by HMG-I/Y in vitro: HMG-I/Y is enriched in H1 depleted chromatin. EMBO J. 12: 3237
    OpenUrlPubMed
  38. ↵
    Johnson, K., D. Lehn, R. Reeves. 1989. Alternative processing of mRNAs encoding mammalian “High Mobility Group” proteins HMG-I and HMG-Y. Mol. Cell Biol. 9: 2114
    OpenUrlAbstract/FREE Full Text
  39. ↵
    Huth, J. R., C. A. Bewley, M. S. Nissen, J. N. S. Evans, R. Reeves, A. M. Gronenborn, G. M. Clore. 1997. The solution structure of an HMG-I(Y)-DNA complex defines a new architectural minor groove binding motif. Nat. Struct. Biol. 4: 657
    OpenUrlCrossRefPubMed
  40. ↵
    Himes, S. R., L. S. Coles, R. Katsikeros, R. K. Lang, M. F. Shannon. 1993. HTLV-1 Tax activation of the GM-CSF and G-CSF promoters requires the interaction of NF-κB with other transcription factor families. Oncogene 8: 3189
    OpenUrlPubMed
  41. ↵
    Abate, C., D. Luk, E. Gagne, R. G. Roeder, T. Curran. 1990. Fos and Jun cooperate in transcriptional regulation via heterologous activation domains. Mol. Cell Biol. 10: 5532
    OpenUrlAbstract/FREE Full Text
  42. ↵
    Ng, K. W., P. Ridgway, D. R. Cohen, D. J. Tremethick. 1997. The binding of a Fos/Jun heterodimer can completely disrupt the structure of a nucleosome. EMBO J. 16: 2072
    OpenUrlAbstract
  43. ↵
    Nissen, M., T. A. Langan, R. Reeves. 1991. Phosphorylation by cdc2 kinase modulates DNA binding activity of HMG-I nonhistone chromatin protein. J. Biol. Chem. 266: 19945
    OpenUrlAbstract/FREE Full Text
  44. ↵
    Reeves, R., M. S. Nissen. 1999. Purification and assays for high mobility group HMG-I(Y) protein function. Methods Enzymol. 304: 155
    OpenUrlCrossRefPubMed
  45. ↵
    Cockerill, P. N., A. G. Bert, F. Jenkins, G. R. Ryan, M. F. Shannon, M. A. Vadas. 1993. Human GM-CSF enhancer function is associated with cooperative interactions between AP-1 and NFATp/c. Mol. Cell Biol. 15: 2071
    OpenUrlAbstract/FREE Full Text
  46. ↵
    Shang, C., J. Attena, D. Cakouros, P. N. Cockerill, and M. F. Shannon. 2000. Nuclear factor of activated T cells contributes to the function of the CD28 response region of the GM-CSF promoter. Int. Immunol. 11:1945.
  47. ↵
    Ausubel, F., R. Brent, R. Kingston, D. Moore, J. Seidman, J. Smith, J. Struhl. 1988. Current Protocols in Molecular Biology John Wiley & Sons, New York.
  48. ↵
    Reeves, R., M. Nissen. 1990. The A-T-rich DNA-binding domain of high mobility group protein HMG-I: a novel peptide motif for recognizing DNA structure. J. Biol. Chem. 265: 8573
    OpenUrlAbstract/FREE Full Text
  49. ↵
    Holbrook, N. J., K. A. Smith, A. J. Fornace, C. M. Comeau, R. L. Wiskocil, G. R. Crabtree. 1984. T-cell growth factor: complete nucleotide sequence and organization of the gene in normal and malignant cells. Proc. Natl. Acad. Sci. USA 81: 1634
    OpenUrlAbstract/FREE Full Text
  50. ↵
    Jenkins, F., P. N. Cockerill, D. Bohmann, M. F. Shannon. 1995. Multiple signals are required for function of the human GM-CSF gene promoter in T cells. J. Immunol. 155: 1240
    OpenUrlAbstract/FREE Full Text
  51. ↵
    Disney, J. E., K. R. Johnson, N. S. Magnuson, S. R. Sylvester, R. Reeves. 1989. High-mobility group protein HMG-I localizes to G/Q and C-bands of human and mouse chromosomes. J. Cell Biol. 109: 1975
    OpenUrlAbstract/FREE Full Text
  52. ↵
    John, S., C. M. Robbins, W. J. Leonard. 1996. An IL-2 response element in the human IL-2 receptor-α chain promoter is a composite element that binds Stat5, Elf-1, HMG-I(Y) and a GATA family protein. EMBO J. 15: 5627
    OpenUrlPubMed
  53. ↵
    John, S., R. Reeves, J. X. Lin, R. Child, J. M. Leiden, C. B. Thompson, W. J. Leonard. 1995. Regulation of cell-type-specific interleukin-2 receptor α chain gene expression: potential role of physical interactions between Elf-1, HMG-I(Y), and NF-κB family proteins. Mol. Cell Biol. 15: 1786
    OpenUrlAbstract/FREE Full Text
  54. ↵
    Briegel, K., B. Hentsch, I. Pfeuffer, E. Serfling. 1991. One base pair change abolishes T cell restricted activity of a κB-like proto-enhancer element from the interleukin 2 promoter. Nucleic Acids Res. 19: 5929
    OpenUrlAbstract/FREE Full Text
  55. ↵
    Hentsch, B., A. Mouzaki, I. Pfeuffer, D. Rungger, E. Serfling. 1992. The weak, fine-tuned binding of ubiquitous transcription factors to the IL-2 enhancer contributes to its T cell-restricted activity. Nucleic Acids Res. 20: 2657
    OpenUrlAbstract/FREE Full Text
  56. ↵
    Novak, T. J., P. M. White, E. V. Rothenberg. 1990. Regulatory anatomy of the murine interleukin-2 gene. Nucleic Acids Res. 18: 4523
    OpenUrlAbstract/FREE Full Text
  57. ↵
    Carey, M.. 1998. The enhanceosome and transcriptional synergy. Cell 92: 5
    OpenUrlCrossRefPubMed
  58. ↵
    Merika, M., A. J. Williams, G. Chen, T. Collins, D. Thanos. 1998. Recruitmant of CBP/p300 by the IFN-β enhanceosome is required for synergistic activation of transcription. Mol. Cell 1: 277
    OpenUrlCrossRefPubMed
  59. ↵
    Yie, J., M. Merika, N. Munshi, G. Chen, D. Thanos. 1999. The role of HMGI(Y) in the assembly and function of the IFN-β enhanceosome. EMBO J. 18: 3074
    OpenUrlAbstract
  60. ↵
    Chuvpilo, S., C. Schomberg, R. Gerwig, A. Heinfling, R. Reeves, F. Grummt, E. Serfling. 1993. Multiple closely linked NFAT/octamer and HMG-I(Y) binding sites are part of the interleukin-4 promoter. Nucleic Acids Res. 21: 5694
    OpenUrlAbstract/FREE Full Text
  61. ↵
    Klein-Hessling, S., G. Schneider, A. Heinfling, S. Chuvpilo, E. Serfling. 1996. HMG-I(Y) interferes with the DNA binding of NF-AT factors and the induction of the interleukin 4 promoter. Proc. Natl. Acad. Sci. USA 93: 15311
    OpenUrlAbstract/FREE Full Text
  62. ↵
    Abbas, A. K., K. M. Murphy, A. Sherr. 1996. Functional diversity of helper T lymphocytes. Nature 383: 787
    OpenUrlCrossRefPubMed
  63. ↵
    O’Garra, A.. 1998. Cytokines induce the development of functionally heterogeneous T helper subsets. Immunity 8: 275
    OpenUrlCrossRefPubMed
  64. ↵
    Gett, A. V., P. D. Hodgkin. 1998. Cell division regulates the T cell cytokine repetoire, revealing a mechanism underlying immune class regulation. Proc. Natl. Acad. Sci. USA 95: 9488
    OpenUrlAbstract/FREE Full Text
  65. ↵
    Bird, J. J., D. R. Brown, A. C. Mullen, N. H. Moskowitz, M. A. Mahowald, J. R. Sider, T. F. Gajewski, C.-R. Wang, S. L. Reiner. 1998. Helper T cell differentiation is controlled by the cell cycle. Immunity 9: 229
    OpenUrlCrossRefPubMed
  66. ↵
    Wang, D.-Z., P. Ray, M. Boothby. 1995. Interleukin 4-inducible phosphorylation of HMG-I(Y) is inhibited by rapamycin. J. Biol. Chem. 207: 22924
    OpenUrl
  67. ↵
    Nelson, B. H., D. M. Willerford. 1998. Biology of the interleukin-2 receptor. Adv. Immunol. 70: 1
    OpenUrlCrossRefPubMed
  68. ↵
    Friedmann, M., L. T. Holth, H. Zoghbi, R. Reeves. 1993. Organization, inducible-expression and chromosome localization of the human HMG-I(Y) nonhistone protein gene. Nucleic Acids Res. 21: 4259
    OpenUrlAbstract/FREE Full Text
  69. ↵
    Ogram, S. A., R. Reeves. 1995. Differential regulation of a multipromoter gene: selective 12-O-tetradecanoylphorbol-13-acetate induction of a single transcription start site in the HMG-I/Y gene. J. Biol. Chem. 270: 14235
    OpenUrlAbstract/FREE Full Text
  70. ↵
    Johnson, K. R., J. E. Disney, C. R. Wyatt, R. Reeves. 1990. Expression of mRNAs encoding mammalian chromosomal proteins HMG-I and HMG-Y during cellular proliferation. Exp. Cell Res. 187: 69
    OpenUrlCrossRefPubMed
  71. ↵
    Lanahan, A., J. B. Williams, L. K. Sanders, D. Nathans. 1992. Growth-factor-induced delayed early response genes. Mol. Cell Biol. 12: 3919
    OpenUrlAbstract/FREE Full Text
  72. ↵
    Holth, L. T., A. E. Thorlacius, R. Reeves. 1997. Effects of epidermal growth factor and estrogen on the regulation of the HMG-I/Y gene in human mammary epithelial cell lines. DNA Cell Biol. 16: 1299
    OpenUrlCrossRefPubMed
  73. ↵
    Pellacani, A., M. T. Chin, P. Wiesel, M. Ibanez, A. Patel, S.-F. Yet, C.-M. Hsieh, J. D. Paulauskis, R. Reeves, M.-E. Lee, M. A. Parrella. 1999. Induction of high mobility group-I(Y) protein by endotoxin and interleukin-1β in vascular smooth muscle cells. J. Biol. Chem. 274: 1525
    OpenUrlAbstract/FREE Full Text
PreviousNext
Back to top

In this issue

The Journal of Immunology: 164 (6)
The Journal of Immunology
Vol. 164, Issue 6
15 Mar 2000
  • Table of Contents
Print
Download PDF
Article Alerts
Sign In to Email Alerts with your Email Address
Email Article

Thank you for your interest in spreading the word about The Journal of Immunology.

NOTE: We only request your email address so that the person you are recommending the page to knows that you wanted them to see it, and that it is not junk mail. We do not capture any email address.

Enter multiple addresses on separate lines or separate them with commas.
The Role of High-Mobility Group I(Y) Proteins in Expression of IL-2 and T Cell Proliferation
(Your Name) has forwarded a page to you from The Journal of Immunology
(Your Name) thought you would like to see this page from the The Journal of Immunology web site.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Citation Tools
The Role of High-Mobility Group I(Y) Proteins in Expression of IL-2 and T Cell Proliferation
S. Roy Himes, Raymond Reeves, Joanne Attema, Mark Nissen, Ying Li, M. Frances Shannon
The Journal of Immunology March 15, 2000, 164 (6) 3157-3168; DOI: 10.4049/jimmunol.164.6.3157

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Share
The Role of High-Mobility Group I(Y) Proteins in Expression of IL-2 and T Cell Proliferation
S. Roy Himes, Raymond Reeves, Joanne Attema, Mark Nissen, Ying Li, M. Frances Shannon
The Journal of Immunology March 15, 2000, 164 (6) 3157-3168; DOI: 10.4049/jimmunol.164.6.3157
del.icio.us logo Digg logo Reddit logo Twitter logo Facebook logo Google logo Mendeley logo
  • Tweet Widget
  • Facebook Like

Jump to section

  • Article
    • Abstract
    • Materials and Methods
    • Results
    • Discussion
    • Acknowledgments
    • Footnotes
    • References
  • Figures & Data
  • Info & Metrics
  • PDF

Related Articles

Cited By...

More in this TOC Section

  • Efficient Identification of Tembusu Virus CTL Epitopes in Inbred HBW/B4 Ducks Using a Novel MHC Class I–Restricted Epitope Screening Scheme
  • Poor-Quality Vβ Recombination Signal Sequences and the DNA Damage Response ATM Kinase Collaborate to Establish TCRβ Gene Repertoire and Allelic Exclusion
  • Multifaceted Activities of Seven Nanobodies against Complement C4b
Show more MOLECULAR AND STRUCTURAL IMMUNOLOGY

Similar Articles

Navigate

  • Home
  • Current Issue
  • Next in The JI
  • Archive
  • Brief Reviews
  • Pillars of Immunology
  • Translating Immunology

For Authors

  • Submit a Manuscript
  • Instructions for Authors
  • About the Journal
  • Journal Policies
  • Editors

General Information

  • Advertisers
  • Subscribers
  • Rights and Permissions
  • Accessibility Statement
  • FAR 889
  • Privacy Policy
  • Disclaimer

Journal Services

  • Email Alerts
  • RSS Feeds
  • ImmunoCasts
  • Twitter

Copyright © 2022 by The American Association of Immunologists, Inc.

Print ISSN 0022-1767        Online ISSN 1550-6606