(Translated by https://www.hiragana.jp/)
Human Chorionic Somatomammotropin Enhancer Function Is Mediated by Cooperative Binding of TEF-1 and CSEF-1 to Multiple, Low-Affinity Binding Sites -- Jiang et al. 11 (9): 1223 -- Molecular Endocrinology
The Wayback Machine - https://web.archive.org/web/20080705171010/http://mend.endojournals.org/cgi/content/full/11/9/1223
help button home button Endocrine Society Molecular Endocrinology
HOME HELP FEEDBACK SUBSCRIPTIONS ARCHIVE SEARCH TABLE OF CONTENTS

This Article
Right arrow Abstract Freely available
Right arrow Full Text (PDF)
Right arrow Purchase Article
Right arrow View Shopping Cart
Right arrow Alert me when this article is cited
Right arrow Alert me if a correction is posted
Right arrow Citation Map
Services
Right arrow Email this article to a friend
Right arrow Similar articles in this journal
Right arrow Similar articles in PubMed
Right arrow Alert me to new issues of the journal
Right arrow Download to citation manager
Citing Articles
Right arrow Citing Articles via HighWire
Right arrow Citing Articles via Google Scholar
Google Scholar
Right arrow Articles by Jiang, S.-W.
Right arrow Articles by Eberhardt, N. L.
Right arrow Search for Related Content
PubMed
Right arrow PubMed Citation
Right arrow Articles by Jiang, S.-W.
Right arrow Articles by Eberhardt, N. L.
Molecular Endocrinology 11 (9): 1223-1232
Copyright © 1997 by The Endocrine Society

Human Chorionic Somatomammotropin Enhancer Function Is Mediated by Cooperative Binding of TEF-1 and CSEF-1 to Multiple, Low-Affinity Binding Sites

Shi-Wen Jiang, Miguel A. Trujillo and Norman L. Eberhardt

Endocrine Research Unit (S.-W.J., M.A.T., N.L.E.) Departments of Medicine and Biochemistry/Molecular Biology (N.L.E.) Mayo Clinic Rochester, Minnesota 55905


    ABSTRACT
 TOP
 ABSTRACT
 INTRODUCTION
 RESULTS
 DISCUSSION
 MATERIALS AND METHODS
 REFERENCES
 
The human chorionic somatomammotropin gene enhancer (CSEn) is composed of multiple enhansons (Enh) that share sequence similarities with those of the simian virus, SV40 enhancer (SVEn). The sequence homology includes two GT-IIC-like (Enh1 and Enh4) and three SphI/II-like enhansons (Enh2, Enh3, and Enh5). We previously showed that transcription enhancer factor 1 (TEF-1) and a 30-kDa placental-specific factor, chorionic somatomammotropin enhancer factor 1 (CSEF-1), bind to Enh4, which plays an essential role in enhancer function. In this study, we demonstrate that TEF-1 and CSEF-1 bind specifically to all the other GT-IIC- and SphI/II-like elements within CSEn with a broad range of binding affinities that vary between 0.005 and 0.15 that of Enh4. Each individual concatenated enhanson was able to stimulate hCS promoter activity in an orientation-independent manner in choriocarcinoma cells (BeWo) with an observed stimulation that was directly proportional to its relative binding affinity for TEF-1 and CSEF-1. These results indicate that CSEn function results from the cooperative interaction of TEF-1 and/or CSEF-1 binding to multiple, low-affinity GT-IIC- and SphI/II-like enhansons within the enhancer.


    INTRODUCTION
 TOP
 ABSTRACT
 INTRODUCTION
 RESULTS
 DISCUSSION
 MATERIALS AND METHODS
 REFERENCES
 
The human chorionic somatomammotropin (hCS) genes are members of the GH and PRL gene families. The multiple hGH and hCS genes (Fig. 1AGo) arose by relatively recent duplication of the hGH gene (1). These homologous, nearly identical genes are located within a short 50-kbp span of DNA on chromosome 17q22-q24 and yet are expressed in a strict cell-specific pattern whereby the hCS-1 and hCS-2 genes are expressed exclusively in placental syncytiotrophoblasts (1, 2). A chorionic somatomammotropin enhancer (CSEn) found 2 kb downstream of the hCS-2 gene (Fig. 1AGo) participates in the cell-specific control of hCS gene expression (3, 4, 5, 6). A minimal enhancer (Fig. 2BGo) is contained within a 240-bp element that stimulates hCS promoter activity in human placental BeWo and JEG-3 and monkey kidney COS-1 cells, but not in HeLa or pituitary GC cells (6, 7). The choriocarcinoma cell lines BeWo and JEG-3 express low levels of hCS-1, hCS-2, and hGH-2 mRNAs, but not hGH-1 mRNA (8, 9) and have served as the dominant model for studying cell-specific expression of the hCS genes (3, 4, 5, 6, 7, 8, 9). The minimal enhancer has been shown to contain several enhansons (individual DNA response elements comprising a modular enhancer) that are homologous to the GT-IIC and SphI/SphII enhansons in the SV40 enhancer, which are binding sites for transcription enhancer factor-1 (TEF-1) (3, 5, 6, 7, 10). We and others have demonstrated that the 53-kDa TEF-1 and a 30-kDa factor, chorionic somatomammotropin enhancer factor 1 (CSEF-1) that is present in placental and COS-1 cells, binds to the major GT-IIC-like enhansons in CSEn (5, 7, 10). Positive enhancer activity is correlated with the binding of CSEF-1 (7), whereas the binding of TEF-1 appears to be associated with inhibition of enhancer activity as well as basal promoter activity (11). The inhibition that occurs through TEF-1 is correlated with its ability to interact with the TATA-binding protein, TBP, and the resultant inability of the TEF-1-TBP complex to bind to the TATA element (11).



View larger version (29K):
[in this window]
[in a new window]
 
Figure 1. Schematic Diagrams of the hGH/hCS Chromosomal Locus (A), the 240-bp Minimal Human Chorionic Somatomammotropin Enhancer, CSEn2 (B), and the Individual GT-IIC- and SphI/SphII-Like Enhansons (Enh1-Enh5) That Comprise the Enhancer (C)

The hGH/hCS locus contains five genes, including the pituitary-specific hGH-1 and placental-specific hCS-5 or hCS-like (a putative pseudogene), hCS-1, hCS-2, and hGH-2 or hGH-variant. The location of the two other enhancers, CSEn5 and CSEn1, that are related (~97.5% nucleotide sequence identity) to the CSEn2 enhancer are shown, and the more detailed structure of the minimal CSEn2 enhancer is depicted (B). Arrows indicate the orientation of the GT-IIC-related (black squares) and SphI/II-related (shaded circles) enhansons. Open rectangles (FP-1 to FP-5) indicate the extent of DNaseI-protected regions in the presence of BeWo, HeLa, and GC cell nuclear extracts and the location of the 6- to 8-bp block mutations (EM1-EM8) that were used in the initial analysis of the enhancer structure (6 ). Sequences of the individual enhansons were aligned using the GCG PILEUP program (Genetics Computer Group, Madison, WI) (C). Enh5 is included in both the GT-IIC and SphI/SphII comparisons because a reasonable fit was found for both consensus types. M-CAT is the designation for the muscle-specific regulatory element that is related to the GT-IIC enhanson.

 


View larger version (109K):
[in this window]
[in a new window]
 
Figure 2. Gel Shift Analysis of the Binding of in vitro-Generated TEF-1 to the GT-IICSV and Enh1-Enh5 Enhansons

In vitro control reactions (TNT) and TEF-1-programmed reactions (TEF) were carried out as described (Materials and Methods). A mutated form (MUT) of the GT-IICSV enhanson was included as an additional nonspecific DNA binding control.

 
An essential GT-IIC enhanson, Enh4, is located within a region (FP-3, Fig. 1BGo) that is protected from deoxyribonuclease I (DNaseI) digestion in the presence of BeWo cell nuclear proteins (3, 5, 6). However, several lines of evidence indicate that additional sequences are also involved in the regulation of CSEn activity. First, footprinting analyses with nuclear extracts from placental cells detected four additional DNaseI-protected regions, FP-1, FP-2, FP-4, and FP-5 (Fig. 1BGo), within the 240-bp fragment (5, 6). Second, mutations of sequences within these footprinted regions resulted in significant reduction of CSEn activity (EM1-EM8, Fig. 1BGo), indicating that these sites are functional (5, 6). Finally, although a single copy of a CSEn enhancer harboring a mutation in the central Enh4 enhanson (EM5, Fig. 1BGo) is virtually devoid of activity, a construct containing two copies of the enhancer carrying the Enh4 mutation restored enhancer activity in placental cells (S.-W. Jiang and N. L. Eberhardt, unpublished results). These findings support the concept that functional DNA elements other than Enh4 are required for CSEn activity.

In the current studies we have performed further analysis of the additional GT-IIC- and SphI/II-like enhansons (Enh1-Enh5) within CSEn. Utilizing gel shift assays, gel supershift experiments with TEF-1 antibodies, and UV cross-linking techniques, we demonstrate that TEF-1 and CSEF-1, generated in vitro or from nuclear cell extracts, specifically bind to each of these enhansons. Interestingly, competition experiments revealed differences of as much as 85-fold in the relative binding affinities of the different enhansons. Nevertheless, each of the individual concatenated enhansons was capable of stimulating hCS promoter activity in an orientation-independent manner in transfected placental choriocarcinoma cells. These results suggest that full CSEn function depends on the cooperative binding of TEF-1 and CSEF-1 to multiple low-affinity binding sites.


    RESULTS
 TOP
 ABSTRACT
 INTRODUCTION
 RESULTS
 DISCUSSION
 MATERIALS AND METHODS
 REFERENCES
 
We have demonstrated that five regions of CSEn, designated FP-1 to FP-5 (Fig. 1BGo) (6), are protected by nuclear extracts from GC, HeLa, and placental BeWo cells. Mutations of individual GT-IIC- and SphI/SphII-like sequences in an otherwise intact enhancer within FP-2 (EM3 and EM4, Fig. 1BGo) and FP-3 (EM5, Fig. 1BGo) reduced enhancer activity dramatically (70–100%), whereas mutations within FP-1 (EM1), FP-4 (EM6), and FP-5 (EM7 and EM8) resulted in less dramatic, but significant (ca. 30–50%) reductions in CSEn function (6). Thus, factors binding to FP-1, FP-4, and FP-5 may contribute to the overall enhancer activity. With the exception of sequences within FP-5, the other four regions share extensive similarities with GT-IIC and SphI/II enhansons (Fig. 1CGo), suggesting that the binding of a common factor to these regions might account for a majority of CSEn functional activity. We therefore sought to determine whether these sequences are recognized by TEF-1 and/or CSEF-1. To test this possibility, we performed gel shift assays using the 5'-end labeled oligonucleotides (Table 1Go) with in vitro-generated TEF-1 and BeWo nuclear cell extracts containing both TEF-1 and CSEF-1.


View this table:
[in this window]
[in a new window]
 
Table 1. Oligonucleotides

 
TEF-1 and CSEF-1 Binding to Enh1-Enh5
In vitro-generated TEF-1 binding to Enh1-Enh5 oligonucleotides is shown in Fig. 2Go. All of the double-stranded Enh oligonucleotides showed a pattern of migration similar to that of the SV40 GT-IIC (GT-IICSV) oligonucleotide, whereas no retarded band is observed in unprogrammed extracts (TNT, Fig. 2Go). Thus, the shift in mobility appears to be specifically produced by TEF-1. Although the same amount of in vitro-generated TEF-1 and radioactivity were applied in each case, the intensity of the TEF-1-DNA complexes was quite different. Since the specific activity of all the probes was similar, we infer that the differences in intensity indicate differences in the relative binding affinities of TEF-1 for each of the enhansons, Enh1-Enh5. The relative binding affinity appears to be GT-IICSV ~ Enh4 {succeeds} Enh5 {succeeds} Enh1 ~ Enh3 {succeeds} Enh2.

We previously showed that nuclear extracts from BeWo cells contain an additional protein that recognizes the GT-IIC- and SphI/SphII-like motifs (7). This protein was designated CSEF-1 and was shown to have a molecular mass of approximately 30 kDa, which produces a much more rapidly migrating complex with the GT-IIC oligonucleotide in gel shift analyses. As shown in Fig. 3Go, this more rapidly migrating complex was observed in gel shift analyses with all of the Enh enhansons in the presence of BeWo nuclear cell proteins. Moreover, the general pattern of intensities with the various Enh1-Enh5 oligonucleotides was similar to that observed with the TEF-1 and CSEF-1 complexes in BeWo nuclear extracts (Fig. 3Go) as with in vitro-generated TEF-1 (Fig. 2Go).



View larger version (64K):
[in this window]
[in a new window]
 
Figure 3. Gel Shift Analysis of the Binding of BeWo Cell Nuclear Extracts to the GT-IICSV and Enh1-Enh5 Enhansons (A and B)

Experiments were performed as described in Materials and Methods. Nonspecific DNA binding activity was assessed by inclusion of a mutated GT-IICSV enhanson (MUT). The identities of TEF-1- and CSEF-1-containing complexes have been described in detail in Jiang and Eberhardt (7 ). The gel in panel A was exposed for 3 days to visualize the weaker interactions, resulting in overexposure of lanes containing the high-affinity enhansons. The lanes shown in panel B are from the same gel exposed for 6 h; however, in this case only the lanes including the GT-IICSV- and Enh4-protein complexes are shown.

 
Characterization of the Factors That Bind to the CSEn Enhanson
To verify that the complexes formed with each of the enhansons corresponded to TEF-1 and CSEF-1 binding, we characterized the complexes further. We previously demonstrated that TEF-1 is heat sensitive whereas CSEF-1 is heat resistant. Gel shift experiments with heat-treated BeWo cell extracts revealed a single, more rapidly migrating complex typical of that formed with the CSEF-1 complex (data not shown), indicating that the more slowly migrating band corresponded to TEF-1. To corroborate that TEF-1 was the factor present in the slow migrating band, a TEF-1-specific antibody was used in a supershift assay. In this experiment, the CSEF-I band cannot be clearly observed because minimal amounts of nuclear extracts were mixed with large amounts of DNA. These conditions were chosen to achieve maximal TEF-1 occupancy of its cognate site to optimize the observation of the supershifted complex. The TEF-1 antisera, but not the preimmune antisera, clearly produced a supershifted band that was observed with all of the Enh probes (Fig. 4Go). In addition, the relative amount of supershifted TEF-1-DNA complex, as judged by the relative intensity of the band, was in the same relative order as observed in simple gel shift experiments (Enh4 {succeeds} Enh5 {succeeds} Enh1 ~ Enh3 {succeeds} Enh2). Taken together, these results indicate that the more highly retarded band corresponds to TEF-I.



View larger version (97K):
[in this window]
[in a new window]
 
Figure 4. Gel Supershift Experiment with a Chicken TEF-1 Antibody (AB) or Nonimmune Serum (NI) in Reactions Containing BeWo Cell Nuclear Proteins and Enhansons Enh1-Enh5

Reaction conditions were modified (Materials and Methods) to maximize TEF-1 binding to the various oligonucleotides.

 
CSEF-1 has not been cloned, and specific antibodies are not yet available. However, UV cross-linking characterized CSEF-I as a factor migrating with an apparent molecular mass of 30 kDa (7). Using the GT-IICSV oligonucleotide, the cross-linked factor from the more rapidly migrating complex has an apparent molecular mass of 30 kDa after electrophoresis in SDS gels (Fig. 5Go). A factor with the same molecular mass was cross-linked to each of the Enh1-Enh5 oligonucleotides (Fig. 5Go), and the relative binding affinity, as judged by the intensity of the band, was similar to that observed in the previous experiments ( Figs. 2–4GoGoGo). These data are consistent with the concept that CSEF-1 binds to each of the Enh1-Enh5 enhansons with an affinity similar to that of TEF-1 binding.



View larger version (45K):
[in this window]
[in a new window]
 
Figure 5. UV Cross-Linking of CSEF-1 to the GT-IICSV and Enh1-Enh5 Enhansons

The enhanson-CSEF-1 complexes migrate with an apparent molecular mass of ~30 kDa. A negative GT-IICSVMUT control oligonucleotide was included in the gel shift experiment that was generated by complexation with BeWo cell nuclear proteins and was subjected to UV cross-linking; however, because of the absence of a shifted band (see Fig. 3Go), it was not eluted for SDS gel electrophoresis.

 
Relative Binding Affinities of TEF-1 and CSEF-1 for the GT-IIC and SphI-Like Elements
To measure more precisely the relative binding affinities of TEF-1 and CSEF-1 for Enh1-Enh5, we performed competition experiments. In this case we used a labeled GT-IICSV oligonucleotide, and competition was effected by including increasing amounts of cold Enh oligonucleotides (Enh1-Enh5, 0.06–600 nM) in gel shift assays. The intensity of bands resulting from complexes with both TEF-I and CSEF-I binding were scanned by densitometry, and competition curves were generated (Fig. 6Go). The relative affinities of each enhanson was estimated from the data in Fig. 6Go at the point at which 50% of the binding was inhibited, and the relative binding affinities are shown in Table 2Go. TEF-1 has an almost 2-fold higher binding affinity for each of the enhansons than CSEF-1, and the order of binding affinities is identical to that observed in the earlier experiments ( Figs. 2–5GoGoGoGo).



View larger version (24K):
[in this window]
[in a new window]
 
Figure 6. Competition Analysis to Establish the Relative Binding Affinities of TEF-1 (A) and CSEF-1 (B) to the Individual Enhansons Enh1-Enh5

BeWo cell nuclear extracts were incubated with the labeled GT-IICSV probe, and unlabeled competitor DNA was included at concentrations ranging from 0.06–600 nM. Enh Mut represents GT-IICSV MUT. Gel shift assays are described in Materials and Methods. The intensities of the TEF-1- and CSEF-1-DNA complexes were measured by densitometry analysis of the autoradiograms using NIH IMAGE software.

 

View this table:
[in this window]
[in a new window]
 
Table 2. TEF-1 and CSEF-1 Relative Binding Affinities

 
The Concatenated Individual Enhansons (Enh1-Enh5) Stimulate hCS Promoter Transcription
To evaluate the ability of individual enhansons to stimulate transcription, a series of constructs containing concatenated individual enhansons were cloned upstream of the hCSp.LUC gene. Tandem repeats (three to nine copies) in both orientations were placed 800 bp upstream of the hCS promoter, and their activity was assessed in transfected BeWo cells (Fig. 7Go). Almost all of these constructs exhibited significant stimulation of hCS promoter activity (1.5- to 4-fold). To evaluate the relative activities of each of the enhansons, the activities of constructs containing five copies of the individual enhansons in both orientations were averaged and plotted against the relative binding affinity as shown in Fig. 8Go. There was a linear relationship between the relative activity and the relative binding affinity for TEF-1 and CSEF-1. These results, along with our previous mutational analyses (6), strongly support the concept that, in addition to Enh4, Enh1, Enh3, Enh5, and possibly Enh2 contribute to CSEn activity by acting as relatively weak binding sites for TEF-1 and CSEF-1. The data suggest that the cooperative interaction of multiple TEF-1 and/or CSEF-1 molecules accounts for CSEn activity.



View larger version (37K):
[in this window]
[in a new window]
 
Figure 7. Enhancer Activity Associated with Concatamers of the Individual Enhansons When Cloned Upstream of the hCSp.LUC Gene

Concatamers containing three to nine repeats were cloned into the vector in both orientations (+ [syn] and - [anti] relative to the hCS promoter), and the constructs were transfected into BeWo cells. Luciferase activity was measured as described (Materials and Methods), and the fold activation was plotted. Data were analyzed by multivariate ANOVA (P{succeeds} 0.0001) and by post hoc Bonferroni t tests (asterisks indicate P < 0.05 compared with the control hCSp.LUC activity).

 


View larger version (15K):
[in this window]
[in a new window]
 
Figure 8. Relationship of Relative Stimulatory Activity and Relative Binding Affinity of the Enh1-Enh5 Enhansons

The relative binding affinities (Table 2Go) were plotted against the combined functional data (Fig. 7Go) for the constructs containing five copies of each of the enhansons. For this analysis the data for both orientations were included. Linear regression equations and r2 values are indicated for both CSEF-1 binding and TEF-1 binding.

 

    DISCUSSION
 TOP
 ABSTRACT
 INTRODUCTION
 RESULTS
 DISCUSSION
 MATERIALS AND METHODS
 REFERENCES
 
The CSEn is a typical enhancer with a modular structure that is related to the SV40 enhancer (3, 4, 5, 6, 7, 10). In previous studies we found that two factors, TEF-1 and CSEF-1, bind to the central, high affinity GT-IIC-like enhanson (Enh4) of CSEn and demonstrated that this enhanson is essential for enhancer function (6, 7). The two proteins compete with each other for this site in a mutually exclusive manner. Several lines of evidence suggest that TEF-1 represses, whereas CSEF-1 induces, CSEn activity. First, CSEn is active in BeWo and COS-1 cells that express relatively large amounts of CSEF-1 compared with TEF-1, whereas in GC and HeLa cells that only express TEF-1, CSEn lacks enhancer activity (7). Furthermore, cotransfection of TEF-1 expression constructs in BeWo cells only results in CSEn inhibition. Cotransfection with a TEF-1 antisense oligonucleotide, which inhibits TEF-1 expression, up-regulates enhancer activity in CSEF-1-expressing cells (7). Finally, we have shown that TEF-1 binds to the TATA binding protein, TBP, and inhibits its ability to bind to the TATA element (11), providing a possible mechanism by which to understand TEF-mediated inhibition of the hCS promoter activity.

DNaseI footprinting studies of the 240-bp minimal enhancer with nuclear extracts from a variety of cells has revealed five protected regions (5, 6). Four of these DNaseI-protected regions contain GT-IIC- and SphI/SphII-like sequences (Enh1-Enh5, Fig. 1AGo and B). Enh2 does not reside within a previously recognized DNaseI-protected region. Mutagenesis of sequences within each of these DNaseI-protected regions diminished CSEn activity (6), indicating that CSEn activity is governed by the interactions of multiple elements. However, the identity of factors binding to these elements has not been established, nor has the function of isolated GT-IIC- or SphI/SphII-related enhansons been tested.

In the present study, TEF-1 and CSEF-1 binding and functional activities were examined for each of the GT-IIC- and SphI/SphII-like regions. Each of the enhansons were shown to bind to TEF-1 and CSEF-1 with comparable relative binding affinities that varied considerably (~2 orders of magnitude) ( Figs. 2–6GoGoGoGoGo and Table 2Go). When the individual enhansons were concatenated (3- to 9-mers), virtually all of the enhansons stimulated hCS promoter activity in an orientation-independent manner when cloned upstream of the hCSp.LUC gene and transfected into BeWo cells (Fig. 7Go), indicating that each of these structures can contribute to CSEn function. Moreover, there was a linear relationship between the relative binding affinity of each enhanson for TEF-1 and CSEF-1 and the relative enhancer activity (Fig. 8Go). This is consistent with the concept that CSEn function is governed by the binding of multiple copies of TEF-1 and CSEF-1 to the enhansons, Enh1-Enh5. Although Enh2 displays the weakest binding and functional activity and exists within a CSEn domain that is not protected by DNaseI footprinting (6), it may contribute, nevertheless, toward mediating enhancer function. The fact that it resides within a region not protected by DNaseI may reflect a lower bound of affinity for which DNaseI footprints may be observed. Interestingly, mutation of either Enh3 or Enh4 results in 70–100% loss of enhancer function in the context of the intact enhancer (6), although there is a ~20-fold difference in the TEF-1 or CSEF-1 binding affinity for the individual enhansons (Fig. 6Go), and Enh4 is a more potent enhancer when multimerized (Fig. 7Go). This result may reflect the importance of the spatial organization of the modular enhancer and illustrates that a relatively weaker enhanson may possess inordinate functional significance in the context of the intact enhancer, which is likely the result of cooperative interactions. In this regard it is of interest that the Enh3 and Enh4 enhansons are centrally located, suggesting that they may comprise part of a core enhancer structure. Taken together these data support the concept that the cooperative interaction of TEF-1 and CSEF-1 to these multiple, low-affinity binding sites plays an essential role in mediating CSEn functional activity; however, they do not exclude the possibility that other factors may be involved in enhancer function. For example, recent evidence indicates that M-CAT (GT-IIC-like) elements involved in both muscle-specific and non-muscle-specific transcription may be modulated by additional factors that bind to flanking sequences adjacent to the M-CAT motifs (12).

Although many enhancers function through the interactions of multiple, unique proteins, several enhancers and regulatory elements appear to operate by the binding of single transcription factors to multiple, low-affinity binding sites. For example, ovalbumin gene regulation by estrogen occurs by estrogen receptor binding to several half-palindromic TGACC motifs instead of the typical palindromic estrogen response element (ERE) (13). Cooperative binding of the estrogen receptor to these weak, relatively widely spaced half-sites provides for synergistic activation of the ovalbumin gene by estrogen. In a very similar manner, estrogen regulation of the progesterone receptor gene is governed by weak, nearly palindromic EREs that are coupled to estrogen receptor half-sites (14). Regulation of the immunoglobulin heavy chain µ-enhancer is modulated by the binding of the enhancer-binding regulatory protein, NF-µNR, to four adjacent binding sites that flank the enhancer core. Like the GT-IIC- and SphI/II-like elements of CSEn, the individual NF-µNR binding sites display a range of binding affinities spanning 2 orders of magnitude (15). Nevertheless, when low- and high-affinity NF-µNR binding sites are present on the same molecule, both sites are occupied at concentrations of NF-µNR that would only be expected to occupy the high-affinity site by itself. Accordingly, the juxtaposition of low- and high-affinity binding sites within the µ-enhancer results in cooperative binding of NF-µNR to the enhancer (15). Finally, the SV40 enhancer functions through the binding of TEF-1 to multiple GT-IIC and SphI/SphII enhansons (16, 17, 18) that differ by 4- to 10-fold in their ability to bind TEF-1 (GT-IIC {succeeds} SphI {succeeds}{succeeds} SphII) (16). These differences in binding affinity closely mirror those observed with the GT-IIC- and SphI/SphII-like CSEn enhansons ( Figs. 2–6GoGoGoGoGo and Table 2Go). Because mutation of the low affinity TEF-1/CSEF-1 binding sites results in an enhancer with lower functional activity (6), these sites are important for enhancer function. Consequently, it is likely that, as is the case with the SV40 enhancer, cooperative binding of TEF-1/CSEF-1 to the multiple low-affinity sites is important for full CSEn functional activity.

TEF-1 is a member of a highly conserved family of regulatory proteins that includes the yeast factor TEC1 (19), the Aspergillus nidulans factor AbaA (20), and the Drosophila scalloped (sd) gene product (21). In addition to TEF-1, a number of distinct TEF-1 homologs from humans (22), mice (22), and chicken (23, 24) have been cloned. TEFs are involved in the regulation of a diverse set of processes, including skeletal and cardiac muscle gene expression (12, 23, 24, 25, 26), SV40 (16, 17, 18), CSEn (3, 4, 5, 6), and human papillomavirus type 16 E6 and E7 (27) enhancer control. The mouse TEF-1, TEF-3, and TEF-4 homologs display a complex pattern of expression during development that implicate these homologs in myogenesis and cardiogenesis, as well as central nervous system development and organogenesis (22). Despite the diversity of these different products, all the members of this family contain a very highly conserved TEA/ATTS DNA-binding domain that recognizes the DNA sequences related to the prototypic GT-IIC and SphI/SphII enhansons (28).

Although the exact structure of the TEA/ATTS domain is not yet known, it has been proposed that this 80-amino acid-containing domain contains either three {alpha}-helices or one {alpha}-helix and two ß-sheet structures (20, 21, 28). Mutational analysis of the putative {alpha}-helical and/or {alpha}-helical/ß-sheet structures demonstrates that the first {alpha}-helical and third {alpha}-helical/ß-sheet are critical for DNA binding; however, the carboxyl terminus of TEF-1 can also modulate DNA-binding affinity (18). Nevertheless, expression of a synthetic DNA-binding domain containing all three of the {alpha}-helical and/or {alpha}-helical/ß-sheet motifs established that these structures are sufficient to determine the binding specificity to the unrelated GT-IIC and SphI/SphII enhansons (18). Thus DNA-binding specificity of these family members resides within the TEA/ATTS domain. Given the striking similarity in the relative binding affinities of CSEF-1 and TEF-1 for the unrelated GT-IIC and SphI/SphII enhansons ( Figs. 2–6GoGoGoGoGo and Table 2Go), it seems likely that CSEF-1 is an as yet unidentified member of this family. It is noteworthy that the TEA/ATTS domain is sufficient for cooperative binding to tandemly repeated GT-IIC and Sph enhansons and that the cooperativity is required for binding to low-affinity enhansons (22). This lends further support to the concept that the multiple, low-affinity GT-IIC- and SphI/SphII-like sites within CSEn may act via cooperative binding of TEF-1 and/or CSEF-1.

It is possible that multiple low-affinity binding sites with or without interspersed high-affinity sites provide a mechanism for generating a transcriptional rheostat that senses different levels of physiological signals and generates fine-tuned control of gene expression. Such a mechanism may help to explain hCS gene expression during pregnancy, which is gradually up-regulated from nearly silent expression during early pregnancy to maximal expression in the third trimester. It should be emphasized that the mechanism by which TEF-1 and/or CSEF-1 mediate enhancer function is unknown. Overexpression of intact TEF-1 in several cell lines (11, 17) results in a dominant negative inhibition of reporter gene activity, suggesting that limiting cofactors are required for its transactivating functions (17). Using GAL4-TEF-1 chimeras, Hwang et al. (18) have been able to demonstrate that three distinct, but interdependent, domains were required for both transactivation and squelching functions, providing additional support for the limiting cofactor model. In contrast, we previously demonstrated that TEF-1-mediated transrepression may be accounted for by interactions with TBP that inhibit TBP from binding the TATA element. Interestingly, the same three TEF-1 activation/squelching domains identified earlier (18) were required for TBP binding (11). These latter results suggest an alternate model in which TEF-1 is a repressor, whose binding to the enhancer may allow it to interact with TBP and negatively regulate transcription initiation. Further evidence for such a repressor model is presented in the accompanying article in which it is shown that TEF-1 binding to multiple enhancers (CSEn2 and CSEn1 or CSEn2 and CSEn5) is associated with a composite silencer activity in pituitary GC cells (30). Accordingly, a cofactor might be required as part of a switch mechanism in cell types in which TEF-1 acts as a transactivator, but not presumably in BeWo cells in which CSEF-1 appears to mediate transactivation. Further studies will be required to elucidate these mechanisms.


    MATERIALS AND METHODS
 TOP
 ABSTRACT
 INTRODUCTION
 RESULTS
 DISCUSSION
 MATERIALS AND METHODS
 REFERENCES
 
Cell Transfection
Placental trophoblast BeWo cells (ATCC, Rockville, MD) were grown in RPMI 1640 (Life Technologies, Inc., Gaithersburg, MD) supplemented with 10% FBS (BioWhittaker, Walkersville, MD), 100 U/ml penicillin (Life Technologies, Inc.), 100 µg/ml streptomycin (Life Technologies, Inc.) and 2 mM glutamine (Life Technologies, Inc.). Cells were maintained at 37 C in an atmosphere containing 5% CO2 and 100% humidity. Transfections were performed with double CsCl2-purified plasmids (15 µg) as described in detail (6). Because of the large number of constructs that had to be evaluated, multiple transfections with different batches of cells grown at different times were required. Consequently, reporter gene activity was normalized to ß-galactosidase activity instead of protein per our normal protocol (6, 7, 11, 30, 32). This was accomplished by cotransfection of a CMV ßGal plasmid (5 µg) and subsequent analysis of ß-galactosidase activity.

Plasmid Construction
All reporter plasmids were cloned into pA3LUC (29). Construction of CSp.LUC, which contains the hCS 496-bp promoter, was described previously (6). Complementary oligonucleotides harboring CSEn sequences protected in footprinting experiments (6) were synthesized at the Molecular Biology Core Facility, Mayo Clinic (Table 1Go). We devised a simple and efficient PCR-based method to subclone all the synthetic oligonucleotides in tandem repeats with defined spacing, orientation, and repeat number (31). Briefly, phosphorylated individual enhanson monomers that contained GG and CC overhangs at the 3'-ends (Table 1Go) were mixed with a synthetic linker containing a BglII site and similarly affixed 3'-GG and 3'-CC overhangs. Enhanson monomers and synthetic linker (10:1) were mixed and ligated with T4 DNA ligase, and 200- to 300-bp DNA fragments containing tandemly repeated monomers and linkers were selected on a 2% agarose gel. Oligomers recovered after the gel had been subjected to several freeze/thaw cycles were amplified by PCR using the synthetic linker oligonucleotides as primers. Self-priming, due to the repetitive nature of the oligomers, generated large DNA PCR products. After BglII restriction digestion, clear bands corresponding to DNA fragments containing different numbers of repeated monomers bounded by the linker oligonucleotide were visible on the gel. These oligomers were ligated to the BglII-treated pA3LUC to generate constructs containing Enh1-Enh5 sequences in different numbers and orientations. Positive clones were screened by BglII digestion, and the sequence of the insert was confirmed by dideoxy-nucleotide sequencing (Molecular Biology Core Facility, Mayo Clinic).

Data Analysis
Data were subjected to multivariate ANOVA using post hoc Bonferroni t tests to assess individual differences among the multiple comparisons.

Gel Shift Assays
The pXJ40-TEF-1A plasmid (17), generously provided by Dr. Pierre Chambon and Irwin Davidson (University of Strasbourg, Strasbourg, France), was used to generate TEF-1 protein by in vitro translation (TNT, Promega, Madison, WI) using pBluescript in a mock translation reaction for negative control as described previously (7). Large-scale nuclear extracts were isolated from cultured BeWo cells according to the method of Dignam et al. (32).

Gel shift probes, which contain the same sequences as the oligonucleotides listed in Table 1Go, except for the absence of protruding 5'-GG and 3'-CC ends, were 5'-end labeled with [{gamma}-32P]ATP (Amersham Corp., Arlington Heights, IL) and with polynucleotide kinase to a specific activity 2 x 106 cpm/pmol. The labeled probe was purified through a Bio-Gel P-60 (Bio-Rad, Richmond, CA) column. Probe (30,000 cpm) and 2.5 µl of in vitro-generated TEF or 20 µg of BeWo nuclear extract were used for gel shift analyses under conditions described in detail previously (33).

The TEF-1 and CSEF-1 affinities to Enh1-Enh5 were measured by gel shift competition experiments. Increasing concentrations (0.05 to 500 nM) of unlabeled double-stranded oligonucleotides were mixed with labeled GT-IICSV probe (30,000 cpm) and BeWo cell nuclear extracts (20 µg). After autoradiography, the TEF-1-DNA and CSEF-1-DNA complexes were analyzed by densitometry (NIH Image).

Rabbit anti-chicken TEF-1 antibody was generously provided by Drs. Charles Ordahl and Iain Farrance (University of California San Francisco) and used for gel supershift experiments. Compared with normal gel shift analyses, less nuclear extract (10 µg) and more DNA probe (80,000 cpm) was used to enhance the sensitivity. After a 30-min incubation, 3 µl TEF-1 antibody were added to the binding reaction and the incubation was continued for an additional 15 min. After electrophoresis, the gel was dried and exposed to Kodak x-ray film for 2 days.

UV Cross-Linking
For cross-linking studies, 120,000 cpm of DNA probe and 80 µg BeWo nuclear extracts were incubated and subsequently resolved by electrophoresis. The wet nondenatured gel was placed on ice and irradiated with UV light (Stratalinker, Stratagene, La Jolla, CA) for 1 h. Autoradiography was performed overnight at 4 C with an intensifying screen on top of the gel. Gel slices containing the CSEF-1 complexes were excised and soaked in 200 µl 2x SDS-PAGE loading buffer (100 mM Tris·HCl (pH 7.6), 300 mM KCl, 1 mM EDTA, 10 mM dithiothreitol) at 4 C for 30 min. The cross-linked CSEF-1-DNA complex was resolved by 10% SDS-PAGE. The gel was dried and exposed to Kodak x-ray film with intensifying screens at -20 C for 2 days.


    ACKNOWLEDGMENTS
 
The authors wish to express their appreciation to Drs. Pierre Chambon and Irwin Davidson for the pXJ140 TEF-1 expression plasmid and to Drs. Charles Ordahl and Iain Farrance for the generous gift of the chicken TEF-1 antibody. We thank Ruth Kiefer for preparing the manuscript.


    FOOTNOTES
 
Address requests for reprints to: Dr. Norman L. Eberhardt, Endocrine Research Unit, 4–407 Alfred, Mayo Clinic, Rochester, Minnesota 55905.

This work was supported by NIH Grants DK-41206 and DK-51492 (to N.L.E.).

Received for publication November 12, 1996. Revision received March 18, 1997. Accepted for publication May 22, 1997.


    REFERENCES
 TOP
 ABSTRACT
 INTRODUCTION
 RESULTS
 DISCUSSION
 MATERIALS AND METHODS
 REFERENCES
 

  1. Miller WL, Eberhardt NL 1983 Structure and evolution of the growth hormone gene family. Endocr Rev 4:97–130[Medline]
  2. Walker WH, Fitzpatrick SL, Barrera-Saldana HA, Resendez-Perez D 1991 The human placental lactogen genes: structure, function, evolution and transcriptional regulation. Endocr Rev 12:316–328[Abstract]
  3. Walker WH, Fitzpatrick SL, Saunders GF 1990 Human placental lactogen transcriptional enhancer. Tissue specificity and binding with specific proteins. J Biol Chem 265:12940–12948[Abstract/Free Full Text]
  4. Fitzpatrick SL, Walker WH, Saunders GF 1990 DNA sequences involved in the transcriptional activation of a human placental lactogen gene. Mol Endocrinol 4:1815–1826[Abstract]
  5. Jacquemin P, Oury C, Peers B, Morin A, Belayew A, Martial JA 1994 Characterization of a single strong tissue-specific enhancer downstream from the three human genes encoding placental lactogen. Mol Cell Biol 14:93–103[Abstract/Free Full Text]
  6. Jiang SW, Eberhardt NL 1994 The human chorionic somatomammotropin gene enhancer is composed of multiple DNA elements that are homologous to several SV40 enhansons. J Biol Chem 269:10384–10392[Abstract/Free Full Text]
  7. Jiang SW, Eberhardt NL 1995 Involvement of a protein distinct from transcription enhancer factor-1 (TEF-1) in mediating human chorionic somatomammotropin gene enhancer function through the GT-IIC enhanson in choriocarcinoma and COS cells. J Biol Chem 270:13906–13915[Abstract/Free Full Text]
  8. Nickel BE, Bock ME, Nachtigal MW, Cattini PA 1993 Differential expression of human placental growth hormone variant and chorionic somatomammotropin genes in choriocarcinoma cells treated with methotrexate. Mol Cell Endocrinol 91:159–166[CrossRef][Medline]
  9. Nickel BE, Cattini PA 1991 Tissue-specific expression and thyroid hormone regulation of the endogenous placental growth hormone variant and chorionic somatomammotropin genes in a human choriocarcinoma cell line. Endocrinology 128:2353–2359[Abstract]
  10. Lytras A, Cattini PA 1994 Human chorionic somatomammotropin gene enhancer activity is dependent on the blockade of a repressor mechanism. Mol Endocrinol 8:478–489[Abstract]
  11. Jiang SW, Eberhardt NL 1996 TEF-1 transrepression in BeWo cells is mediated through interactions with the TATA-binding protein, TBP. J Biol Chem 271:9510–9518[Abstract/Free Full Text]
  12. Larkin SB, Farrance IK, Ordahl CP 1996 Flanking sequences modulate the cell specificity of M-CAT elements. Mol Cell Biol 16:3742–3755[Abstract]
  13. Kato S, Tora L, Yamauchi J, Masushige S, Bellard M, Chambon P 1992 A far upstream estrogen response element of the ovalbumin gene contains several half-palindromic 5'-TGACC-3' motifs acting synergistically. Cell 68:731–742[CrossRef][Medline]
  14. Kraus WL, Montano MM, Katzenellenbogen BS 1994 Identification of multiple, widely spaced estrogen-responsive regions in the rat progesterone receptor gene. Mol Endocrinol 8:952–969[Abstract]
  15. Scheuermann RH 1992 The tetrameric structure of NF-mu NR provides a mechanism for cooperative binding to the immunoglobulin heavy chain mu enhancer. J Biol Chem 267:624–634[Abstract/Free Full Text]
  16. Davidson I, Xiao JH, Rosales R, Staub A, Chambon P 1988 The HeLa cell protein TEF-1 binds specifically and cooperatively to two SV40 enhancer motifs of unrelated sequence. Cell 54:931–942[CrossRef][Medline]
  17. Xiao JH, Davidson I, Matthes H, Garnier JM, Chambon P 1991 Cloning, expression, and transcriptional properties of the human enhancer factor TEF-1. Cell 65:551–568[CrossRef][Medline]
  18. Hwang JJ, Chambon P, Davidson I 1993 Characterization of the transcription activation function and the DNA binding domain of transcriptional enhancer factor-1. EMBO J 12:2337–2348[Medline]
  19. Laloux I, Jacobs E, Dubois E 1994 Involvement of SRE element of Ty1 transposon in TEC1-dependent transcriptional activation. Nucleic Acids Res 22:999–1005[Abstract/Free Full Text]
  20. Andrianopoulos A, Timberlake WE 1994 The Aspergillus nidulans abaA gene encodes a transcriptional activator that acts as a genetic switch to control development. Mol Cell Biol 14:2503–2515[Abstract/Free Full Text]
  21. Campbell S, Inamdar M, Rodrigues V, Raghavan V, Palazzolo M, Chovnick A 1992 The scalloped gene encodes a novel, evolutionarily conserved transcription factor required for sensory organ differentiation in Drosophila. Genes Dev 6:367–379[Abstract/Free Full Text]
  22. Jacquemin P, Hwang JJ, Martial JA, Dolle P, Davidson I 1996 A novel family of developmentally regulated mammalian transcription factors containing the TEA/ATTS DNA binding domain. J Biol Chem 271:21775–21785[Abstract/Free Full Text]
  23. Stewart AF, Larkin SB, Farrance IK, Mar JH, Hall DE, Ordahl CP 1994 Muscle-enriched TEF-1 isoforms bind M-CAT elements from muscle-specific promoters and differentially activate transcription. J Biol Chem 269:3147–3150[Abstract/Free Full Text]
  24. Azakie A, Larkin SB, Farrance IK, Grenningloh G, Ordahl CP 1996 DTEF-1, a novel member of the transcription enhancer factor-1 (TEF-1) multigene family. J Biol Chem 271:8260–8265[Abstract/Free Full Text]
  25. Farrance IK, Mar JH, Ordahl CP 1992 M-CAT binding factor is related to the SV40 enhancer binding factor, TEF-1. J Biol Chem 267:17234–17240[Abstract/Free Full Text]
  26. Farrance IK, Ordahl CP 1996 The role of transcription enhancer factor-1 (TEF-1) related proteins in the formation of M-CAT binding complexes in muscle and non-muscle tissues. J Biol Chem 271:8266–8274[Abstract/Free Full Text]
  27. Ishiji T, Lace MJ, Parkkinen S, Anderson RD, Haugen TH, Cripe TP, Davidson I, Chambon P, Turek LP 1992 Transcriptional enhancer factor (TEF)-1 and its cell-specific co-activator activate human papillomavirus-16 E6 and E7 oncogene transcription in keratinocytes and cervical carcinoma cells. EMBO J 11:2271–2281[Medline]
  28. Burglin TR 1991 The TEA domain: a novel, highly conserved DNA-binding motif [letter]. Cell 66:11–12[CrossRef][Medline]
  29. Maxwell IH, Harrison GS, Wood WM, Maxwell F 1989 A DNA cassette containing a trimerized SV40 polyadenylation signal which efficiently blocks spurious plasmid-initiated transcription. Biotechniques 7:276–280[Medline]
  30. Jiang SW, Trujillo MA, Eberhardt NL 1997 The placental human chorionic somatomammotropin enhancers form a composite silencer in pituitary cells in vitro. Mol Endocrinol 11:1233–1244[Abstract/Free Full Text]
  31. Jiang SW, Trujillo MA, Eberhardt NL 1996 An efficient method for generation and subcloning of tandemly repeated DNA sequences with defined length, orientation and spacing. Nucleic Acids Res 24:3278–3279[Abstract/Free Full Text]
  32. Dignam JD, Lebovitz RM, Roeder RG 1983 Accurate transcription initiation by RNA polymerase II in a soluble extract from isolated mammalian nuclei. Nucleic Acids Res 11:1475–1489[Abstract/Free Full Text]
  33. Jiang SW, Shepard AR, Eberhardt NL 1995 An initiator element is required for maximal human chorionic somatomammotropin gene promoter and enhancer function. J Biol Chem 270:3683–3692[Abstract/Free Full Text]



This article has been cited by other articles:


Home page
Cancer Res.Home page
T. Hucl, J. R. Brody, E. Gallmeier, C. A. Iacobuzio-Donahue, I. K. Farrance, and S. E. Kern
High Cancer-Specific Expression of Mesothelin (MSLN) Is Attributable to an Upstream Enhancer Containing a Transcription Enhancer Factor Dependent MCAT Motif
Cancer Res., October 1, 2007; 67(19): 9055 - 9065.
[Abstract] [Full Text] [PDF]


Home page
Mol. Endocrinol.Home page
Y. Jin, L. D. Norquay, X. Yang, S. Gregoire, and P. A. Cattini
Binding of AP-2 and ETS-Domain Family Members Is Associated with Enhancer Activity in the Hypersensitive Site III Region of the Human Growth Hormone/Chorionic Somatomammotropin Locus
Mol. Endocrinol., March 1, 2004; 18(3): 574 - 587.
[Abstract] [Full Text] [PDF]


Home page
J. Biol. Chem.Home page
K. Yamada, H. Ogawa, S.-i. Honda, N. Harada, and T. Okazaki
A GCM Motif Protein Is Involved in Placenta-specific Expression of Human Aromatase Gene
J. Biol. Chem., November 5, 1999; 274(45): 32279 - 32286.
[Abstract] [Full Text] [PDF]


Home page
J. Biol. Chem.Home page
R. A. Rachubinski, S. L. Marcus, and J. P. Capone
The p56lck-interacting Protein p62 Stimulates Transcription via the SV40 Enhancer
J. Biol. Chem., June 25, 1999; 274(26): 18278 - 18284.
[Abstract] [Full Text] [PDF]


Home page
Mol. Endocrinol.Home page
S.-W. Jiang, K. Wu, and N. L. Eberhardt
Human Placental TEF-5 Transactivates the Human Chorionic Somatomammotropin Gene Enhancer
Mol. Endocrinol., June 1, 1999; 13(6): 879 - 889.
[Abstract] [Full Text]


Home page
Mol. Endocrinol.Home page
Y. Sun and M. L. Duckworth
Identification of a Placental-Specific Enhancer in the Rat Placental Lactogen II Gene That Contains Binding Sites for Members of the Ets and AP-1 (Activator Protein 1) Families of Transcription Factors
Mol. Endocrinol., March 1, 1999; 13(3): 385 - 399.
[Abstract] [Full Text]


Home page
J. Biol. Chem.Home page
Z. Wang and S. Melmed
Functional Map of a Placenta-specific Enhancer of the Human Leukemia Inhibitory Factor Receptor Gene
J. Biol. Chem., October 2, 1998; 273(40): 26069 - 26077.
[Abstract] [Full Text] [PDF]


Home page
Mol. Endocrinol.Home page
S.-W. Jiang and N. L. Eberhardt
The Human Chorionic Somatomammotropin Enhancers Form a Composite Silencer in Pituitary Cells in Vitro
Mol. Endocrinol., August 1, 1997; 11(9): 1233 - 1244.
[Abstract] [Full Text]


This Article
Right arrow Abstract Freely available
Right arrow Full Text (PDF)
Right arrow Purchase Article
Right arrow View Shopping Cart
Right arrow Alert me when this article is cited
Right arrow Alert me if a correction is posted
Right arrow Citation Map
Services
Right arrow Email this article to a friend
Right arrow Similar articles in this journal
Right arrow Similar articles in PubMed
Right arrow Alert me to new issues of the journal
Right arrow Download to citation manager
Citing Articles
Right arrow Citing Articles via HighWire
Right arrow Citing Articles via Google Scholar
Google Scholar
Right arrow Articles by Jiang, S.-W.
Right arrow Articles by Eberhardt, N. L.
Right arrow Search for Related Content
PubMed
Right arrow PubMed Citation
Right arrow Articles by Jiang, S.-W.
Right arrow Articles by Eberhardt, N. L.


HOME HELP FEEDBACK SUBSCRIPTIONS ARCHIVE SEARCH TABLE OF CONTENTS
Endocrinology Endocrine Reviews J. Clin. End. & Metab.
Molecular Endocrinology Recent Prog. Horm. Res. All Endocrine Journals