54
Cloning of the human claudin-2 5’-flanking region revealed a TATA-less promoter with conserved binding sites in mouse and human for Caudal- Related Homeodomain Proteins and Hepatocyte Nuclear Factor-1 Takanori Sakaguchi*, Xiubin Gu*, Heidi M. Golden*, EunRan Suh $ , David B. Rhoads # , and Hans-Christian Reinecker* * Gastrointestinal Unit, Department of Medicine, Center for the Study of Inflammatory Bowel Disease; # Pediatric Endocrine Unit, Department of Pediatrics; Massachusetts General Hospital & Harvard Medical School, 32 Fruit Street, Boston, Massachusetts 02114; and $ Department of Internal Medicine, University of Pennsylvania School of Medicine, Philadelphia, Pennsylvania 19104 Running Title: Organ specific regulation of claudin-2 gene expression. Abbreviations: HNF, hepatocyte nuclear factor; NF-κB, nuclear factor-kappa B; SI, sucrase isomaltase; and LPH, lactase-phlorizin hydrolase Address correspondence to: Hans-Christian Reinecker Gastrointestinal Unit Massachusetts General Hospital 32 Fruit Street, Boston, MA 02114 Phone (617) 724 2172, Fax (617) 726 3673 E-mail: [email protected] Copyright 2002 by The American Society for Biochemistry and Molecular Biology, Inc. JBC Papers in Press. Published on April 4, 2002 as Manuscript M110261200 by guest on June 15, 2018 http://www.jbc.org/ Downloaded from

Cloning of the human claudin-2 5’-flanking region revealed ... · The human colon cancer derived cell line Caco-2, ... A degenerate primer approach with primers: 5 ... fragment

  • Upload
    haliem

  • View
    213

  • Download
    1

Embed Size (px)

Citation preview

Cloning of the human claudin-2 5’-flanking region revealed a TATA-less

promoter with conserved binding sites in mouse and human for Caudal-

Related Homeodomain Proteins and Hepatocyte Nuclear Factor-1

Takanori Sakaguchi*, Xiubin Gu*, Heidi M. Golden*, EunRan Suh$, David B.

Rhoads#, and Hans-Christian Reinecker*

* Gastrointestinal Unit, Department of Medicine, Center for the Study of

Inflammatory Bowel Disease; # Pediatric Endocrine Unit, Department of

Pediatrics; Massachusetts General Hospital & Harvard Medical School, 32 Fruit

Street, Boston, Massachusetts 02114; and $ Department of Internal Medicine,

University of Pennsylvania School of Medicine, Philadelphia, Pennsylvania

19104

Running Title: Organ specific regulation of claudin-2 gene expression.

Abbreviations: HNF, hepatocyte nuclear factor; NF-κB, nuclear factor-kappa B;

SI, sucrase isomaltase; and LPH, lactase-phlorizin hydrolase

Address correspondence to: Hans-Christian Reinecker

Gastrointestinal Unit

Massachusetts General Hospital

32 Fruit Street, Boston, MA 02114

Phone (617) 724 2172, Fax (617) 726 3673

E-mail: [email protected]

Copyright 2002 by The American Society for Biochemistry and Molecular Biology, Inc.

JBC Papers in Press. Published on April 4, 2002 as Manuscript M110261200 by guest on June 15, 2018

http://ww

w.jbc.org/

Dow

nloaded from

Sakaguchi et al. 2

Summary

Claudin-2 is a structural component of tight junctions in the kidneys, liver

and intestine, but the mechanisms regulating its expression have not been

defined. The 5’-flanking region of the claudin-2 gene contains binding sites for

intestinal specific Cdx homeodomain proteins and hepatocyte nuclear factor

(HNF)-1, which are conserved in human and mouse. Both, Cdx1 and Cdx2,

activated the claudin-2 promoter in the human intestinal epithelial cell line Caco-

2. HNF-1α augmented the Cdx2- but not Cdx1-induced transcriptional activation

of the human claudin-2 promoter. In mice, HNF-1α was required for claudin-2

expression in the villus epithelium of the ileum and within the liver, but not in the

kidneys, indicating an organ specific function of HNF-1α in the regulation of

claudin-2 gene expression. Tight junction structural components, which

determine epithelial polarization and intestinal barrier function, can be regulated

by homeodomain proteins that control the differentiation of the intestinal

epithelium.

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 3

Introduction

Claudin-2 is a regulatory component of tight junctions in the liver, the

kidneys, and the epithelium of the small and large intestines (1,2). Claudin-2

expression has been demonstrated to be involved in the regulation of the

intestinal barrier function by immune modulators (3). Claudins form a family of

proteins composed of at least 24 members, which are expressed in an organ

specific manner and regulate the tissue-specific physiological properties of tight

junctions (4,5). Tight junctions not only create a primary barrier to prevent

paracellular passage of solutes and pathogens but they also restrict the lateral

diffusion of membrane lipids and proteins to maintain cellular polarity (5-8).

Evidence is mounting that claudins are actively involved in the regulation of

paracellular transport of ions through tight junctions (9,10). The modulation of

selective transport through tight junction may require the coordinated expression

of distinct claudins in a particular cell type (1,11). Therefore the regulation of

claudin expression may determine the fundamental ability of the intestinal

epithelium to modulate water- or ion-transport and barrier function. However, the

transcriptional events involved in the organ specific expression of claudins have

not been determined.

Cdx1 and Cdx2 are members of the caudal-related homeobox gene family

based on their sequence homology to the caudal gene of Drosophila

melanogaster (12-14). In vitro and in vivo studies of Cdx1 and Cdx2 suggest that

these transcription factors are important in the early differentiation and

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 4

maintenance of intestinal epithelial cells. In vitro experiments show significant

functional effects of Cdx genes on intestinal differentiation (15,16), proliferation

(15,17), and intestine-specific gene transcription (18-22). Overexpression of

Cdx2 in undifferentiated IEC-6 intestinal epithelial cells leads to the development

of a differentiated phenotype (15). Cdx1 and Cdx2 have been shown to regulate

intestine-specific gene transcription by binding to several intestine-specific

promoters (18-20,23,24). In intestinal epithelial Caco-2 cells, Cdx2 expression

induces the expression of sucrase-isomaltase (SI) and lactase-phlorizin

hydrolase (LPH), two markers of intestinal differentiation (25).

In the regulation of LPH expression Cdx2 directly interacts with

Hepatocyte Nuclear Factor (HNF)-1α (21). HNF-1α and HNF-1β are related

transcription factors which bind to DNA as homo- or hetero-dimers (26). HNF-1α

and HNF-1β are known to be important for liver-specific gene transcription, but

are also expressed in other organs, such as pancreas, kidney, stomach and

intestine (27-29).

In this report, we examined the Cdx and HNF-1α mediated regulation of

the claudin-2 promoter in the human intestinal epithelial cell line Caco-2 and

determined the claudin-2 mRNA and protein expression in HNF-1α deficient

mice. These experiments identify claudin-2 as a target of Cdx homeoproteins and

HNF-1α function in human intestinal epithelial cells. HNF-1α regulated the

complex pattern of claudin-2 expression along the crypt-villus axis of the mouse

ileum and was required for claudin-2 expression in the liver.

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 5

Experimental procedures

Antibodies and expression vectors

Polyclonal antibodies recognizing claudin-1 and claudin-2 were obtained

from Zymed Laboratories, Inc. (San Francisco, CA). Anti-Cdx1 and Cdx2

polyclonal antibodies were previously described (30,31). Antibodies for human

HNF-1α and HNF-1β were from Santa Cruz Biotechnology, Inc. (Santa Cruz,

CA). Horseradish peroxidase conjugated anti-rabbit and anti-goat antibodies

were obtained from Amersham (Arlington Heights, IL) and Santa Cruz,

respectively. Mouse Cdx1- or Cdx2-expression vectors (pRc/CMV-Cdx1 or

pRc/CMV-Cdx2) were previously described (15,18). Human HNF-1α- or HNF-

1β-expression vectors (32) and mouse HNF-1α-expression vector (pBJ5mHNF-

1α) (27) were kindly provided by Dr. Marco Pontoglio (Institut Pasteur, Paris,

France) and Dr. Gerald R. Crabtree (Stanford University School of Medicine,

CA), respectively.

Cell culture

The human colon cancer derived cell line Caco-2, human hepatocellular

carcinoma derived cell line HepG2, and mouse mesenchymal cell line NIH3T3

cells were obtained from American type Culture Collection (Rockwell, MD).

These cells were grown in Dulbeco’s modified Eagle’s medium (DMEM) (Cellgro,

Mediatech Inc., Herndon, VA), supplemented with 100 IU/ml penicillin, 100 µg/ml

streptomycin and 10 % (for HepG2 and NIH3T3) or 20% (for Caco-2) heat-

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 6

inactivated fetal calf serum (Sigma, St. Louise, MO), in a humidified 5% CO2

atmosphere at 37 °C. The human colon cancer derived cell line T-84 cells were

grown in DMEM/F-12 (1:1) with antibiotics described above and 10 % heat-

inactivated fetal calf serum.

Isolation of the 5’ flanking region of the human claudin-2 gene and cloning

of human claudin-2.

A degenerate primer approach with primers: 5’-TGG ATG GA(AG) TGT

GC(ATGC) AC(ATGC) CA(CT) -3’; 5’-GA GCA (GA)GA (AG)AA GCA (ATGC)AG

(AGTC)AT(GTA)AT (AGTC)CC-3’, corresponding to the mouse claudin-2

sequence, was used to amplify 407bp of the open reading frame of human

claudin-2. Database searches with the putative human claudin-2 sequence

identified several human claudin-2 EST clones, which were used to complement

the 5' and 3’ sequence. Additional polymerase chain reaction (PCR) with the

primers: 5’-GCT TCT ACT GAG AGG TCT G -3’; 5’-TTC TTC ACA CAT ACC

CTG-3’, and DNA sequencing was utilized to confirm the expression of the full-

length human claudin-2 sequence in T-84 cells. DNA and amino acid sequence

of human claudin-2 has been submitted to Genbank, and is available under the

accession number AF250558. The GenomeWalker Kit (Clontech, Palo Alto, CA)

was used to isolate the 5’-flanking region of the human claudin-2 gene. In brief,

the first PCR was performed with the gene specific primer 1 (5’-CAA AAG CCC

CAG AAG GCC TAG GAT GTA G -3’; +30 to +57 relative to the adenosine of

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 7

the methionine start codon of the human claudin-2 cDNA; accession number

AF250558 or AF177340) and the Adaptor Primer 1. The second PCR was done

with the gene specific primer 2 (5’-GGC AGA CCT CTC AGT AGA AGC GTC

TTC -3’; -27 to -1; corresponding to 493 – 519 sequence of AF177340) and the

Adaptor Primer 2. The longest PCR fragment was purified and subcloned into

pCR2.1 vector (Invitrogen, Carlsbad, CA). The resulting plasmid was designated

as pCR-hCL2p and sequenced.

Deletion constructs, mutagenesis, and reporter gene assay

The KpnI / XhoI fragment of pCR-hCL2p was subcloned into the KpnI /

XhoI site of the pGL3B vector (Promega, Madison, WI). Various length

fragments of the 5’-flanking region of the human claudin-2 gene were amplified

by PCR and subcloned into pGL3B. To obtain –62 construct, HindIII / XbaI

fragment from –84 construct was ligated to EcoRI / HindIII-digested –84

construct with complementary 38-base oligonucleotides (designated as -62wt,

from –62 to –31: sense, 5’-AAT TCA TAT TTA ATC TGG TTT ATG GAT TTT

TTT TAG GT-3’; antisense, 5’-CTA GAC CTA AAA AAA ATC CAT AAA CCA

GAT TAA ATA TG) with 5’-EcoRI and 3’-XbaI overhangs (underlined). To make

mutant claudin-2 promoter constructs, mutated 38-base oligonucleotides were

substituted for wildtype sequence. For Mut 1, Mut 2, and Mut 1+2, 5’-AAT TCA

TAT TTA ATC TGG TGG CTG GAT TTT TTT TAG GT-3’, 5’-AAT TCA TAT TTA

ATC TGG TTT ATG GAT TTT TTG GCG GT-3’, and 5’-AAT TCA TAT TTA ATC

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 8

TGG TGG CTG GAT TTT TTG GCG GT-3’ were used, respectively. (Only sense

strands are shown, nucleotide substitutions are indicated in bold letters.)

For reporter assays, a DNA transfection mixture was prepared consisting

of 1 µg of the reporter construct and 20 ng of pRL-CMV (Promega) as an internal

control. Cells were split onto 6-well plates 18 hr before transfection. Cell

confluency at transfection was 40-60 %. The individual DNA mixtures were

transfected with LipofectAMINE Plus (Life Technologies, Gaithersburg, MD)

according to the manufacturer’s protocol. For co-transfection experiments, 0.5

µg of the expression vector was transfected along with reporter vectors.

pcDNA3.1 vector (Invitrogen) was used to equalize the amount of transfected

DNA. Cells were harvested 48 hrs after transfection, and the luciferase activity

was measured using Dual-Luciferase Reporter Assay System (Promega) and a

luminometer. Transfection efficiencies were normalized to renilla luciferase

activity of the pRL-CMV vector and results are expressed as mean relative

luciferase activity ± SD, of at least three independent experiments.

Electrophoretic Mobility Shift Assay (EMSA)

Nuclear proteins were prepared as previously described (33). Cytosolic

fractions obtained during this procedure were separated for Western blot

analysis. The double-stranded oligonucleotides, -62wt, Mut1, Mut2 and Mut1+2,

were used as probes or cold competitors to analyze the interaction between Cdx

protein and DNA. The HNF-1 wild type probe from –67 to –51 of human claudin-

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 9

2 gene sequence consisted of complementary 29-nucleotide oligonucleotides

(sense; 5’-AAT TCC TGG TCA ATA TTT AAT CTG T -3’, antisense; 5’-CTA GAC

AGA TTA AAT ATT GAC CAG G) with 5’-EcoRI and 3’-XbaI overhangs

(underlined). Mutant HNF complementary oligonucleotides were following:

sense; 5’-AAT TCC TAA TTC AGG TTT AAT CTG T-3’, antisense; 5’-CTA GAC

AGA TTA AAC CTG AAT TAG G-3’ (nucleotide substitutions are indicated by

bold).

The probes were labeled with Klenow enzyme by fill-in incorporation with

nucleotide triphosphates, including [α-32P] dATP. The binding reaction was

performed as previously described (34). For a competition assay, a 100-fold

excess of unlabeled oligonucleotide was added to the reaction. To perform

supershift assay, the binding mixtures were incubated for 10 minutes at room

temperature in the presence of 1 µL of antibodies. Samples were fractionated on

4% non-denaturing polyacrylamide gel in 0.5X TBE buffer. The resultant DNA-

protein complexes were detected by autoradiography.

Western blot analysis

The protein concentration of each sample was quantified by the Bradford

method. Samples were electrophoresed through a 4-20 % gradient SDS

polyacrylamide gel and transferred onto polyvinylidene difluoride (PVDF)

membranes (Millipore, Bedford, MA). The blots were blocked overnight at 4 °C

with 10 % dry milk in PBS containing 0.1 % Tween 20 (PBS-T), followed by the

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 10

incubation for 3 hours at room temperature (RT) with primary antibodies diluted

in blocking buffer at 1:1,000. After washing in PBS-T for 30 minutes, the blots

were incubated with secondary antibodies diluted in blocking buffer for 45

minutes at RT. The hybridized bands were detected by ECL kit (Amersham),

according to the manufacturer’s instruction.

RNA extraction and Northern Blot analysis

Total RNA was isolated from tissues using Trizol reagent (Life

Technologies). Total RNA (30 µg) was electrophoresed in a 1 % agarose

formaldehyde gel and transferred to a nylon membrane (Magna NT,

MicroSeparations Inc., Westbrough, MA) by capillary blotting. Probes were

labeled with [α-32P] dCTP using Rediprime Random Primer Labeling Kit

(Amersham). Membranes were hybridized with radio-labeled probes in Quickhyb

solution (Stratagene, La Jolla, CA) at 65 °C for 1 hour. The membranes were

washed with 0.1% SDS, 2X sodium chloride sodium citrate buffer at RT for 15

minutes and at 65 °C for 10 minutes. The blots were analyzed by

autoradiography. The probes used to detect claudin-1, claudin-2 and

glyceraldehyde-3-phosphate dehydrogenase (GAPDH) were described

previously (3). Other probes were: Cdx1, a 0.9 kb HindIII / XbaI fragment of

pRc/CMV-Cdx1; Cdx2, a 0.9 kb HindIII fragment of pRc/CMV-Cdx2; and HNF-

1α, a 0.4 kb SmaI fragment of pBJ5mHNF1α. The HNF-1α probe derives from a

unique sequence in HNF-1α cDNA and does not cross-hybridize with HNF-1β.

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 11

Northern blots were densitometrically analyzed and gene specific mRNA

expression levels were normalized to GAPDH mRNA expression levels in the

same samples, and expressed as mean density/area calculated from 3

independent experiments.

Tissue preparation and immunohistochemistry

Mice carrying the HNF-1α null allele were obtained from Dr. Frank J.

Gonzalez (NIH, MD) (35). Homozygous HNF-1α null and wildtype littermates

were obtained by mating heterozygous carriers. All animal experiments were

performed in accordance with National Institutes of Health guidelines and

protocols approved by the Subcommittee on Research Animal Care at our

institute. The liver and kidney were removed and washed with ice-cold PBS.

Segments of 2 cm from the most proximal jejunum and most distal ileum were

collected. For immunostaining, small tissue blocks were mounted in OCT

compound and frozen in dry ice-ethanol. For RNA extraction, small pieces of

tissues were snap-frozen at –80 °C.

Cryosections of frozen tissues with 4-µm thickness were prepared. The

sections were air-dried and fixed in methanol at -20 °C for 10 minutes followed by

rehydration in PBS at 4 °C for 30 minutes as previously described (2). The

sections were blocked with 0.5 % normal donkey serum in PBS (blocking

solution) for 1 hour at 20 °C, and incubated with primary antibodies or normal

rabbit serum diluted at 1:100 with blocking solution for 3 hours at RT. After three

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 12

washes with PBS, the slides were incubated at RT with FITC-labeled anti-rabbit

antibody (Vector Laboratories, Burligame, CA) diluted at 1:500 with blocking

solution for 1 hour in the dark and analyzed with an AX-70 Olympus fluorescent

microscope.

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 13

Results

Isolation of the 5’-flanking region of the human claudin-2 gene

The 5’-flanking region of the claudin-2 gene isolated from the human

genomic library and T-84 cell-derived claudin-2 cDNA sequence were confirmed

by sequence comparison with the human genomic clone AL158821. BLAST

search revealed that the gene encoding human claudin-2 is located on

chromosome X, mapping to q22.3-23. The claudin-2 mRNA expressed in T-84

cells contains an open reading frame of 693bp. Human and mouse claudin-2

have a high sequence identity of 87% on mRNA level and 93% identity on amino

acid level.

Comparison with the mouse claudin-2 promoter in genomic databases

revealed that the promoters of the human and mouse claudin-2 genes possess a

remarkable homology of 84% for the region of -1 to –400 (Fig. 1). The mouse

claudin-2 cDNA (Genbank AK004990) recovered by cap-trapping revealed the

putative transcriptional start site at 152 basepair (bp) upstream of the

translational start codon. The transcriptional initiation site is located within a

consensus initiator (Inr) element (NCANNNNN) (36,37).

The promoters of the human and mouse claudin-2 genes have no TATA

box near the putative transcriptional initiation site (Fig. 1). However, a CAAT box

is located at –60 to -63 bp and two E-boxes (CANNTG) at –198 to -195 bp and

–67 to –62 bp (Fig. 1), suggesting that regulatory elements to initiate gene

transcription are present.

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 14

The human claudin-2 promoter contains two sites for the intestine-specific

homeodomain protein family Cdx (18), designated CdxA and CdxB (Fig. 1). The

promoter has also binding sites for HNF-1 and HNF-3β (38), as well as putative

AP-1 (39), NF-κB (40) and GATA (41) binding sites. Particularly, the first Cdx

binding site CdxA, the HNF-1, HNF-3β and GATA binding sites are conserved in

human and mouse claudin-2 promoters (Fig. 1).

To identify the regions involved in regulating claudin-2 gene transcription,

sequentially deleted 5’-flanking regions (–1041, -393, -84, -62 and –31 to +148)

were cloned into the reporter plasmid pGL3B. Reporter constructs were

transfected into intestinal epithelial cell line Caco-2, hepatic cell line HepG2 or

fibroblast cell line NIH3T3. In Caco-2 cells the claudin-2 promoter fragments

containing –1040 to –62 bp of the 5’-flanking region induced a 18 - 29-fold

increase in relative luciferase activity above that observed after transfection with

the control null reporter construct (Fig. 2). In contrast, the same promoter regions

achieved only a 7-11-fold increase when transfected into HepG2 and NIH3T3

cells (Fig. 2).

Removal of the putative AP-1 and the NF-κB sites decreased the

promoter activity slightly. Disruption of the HNF-1 binding site in the -62 bp

construct did not alter the promoter activity significantly in Caco-2 cells. However,

removal of the Cdx binding sites resulted in a loss of promoter activity in Caco-2,

HepG2 and NIH3T3 cells (Fig. 2).

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 15

Claudin-2 promoter activity is regulated by Cdx homeodomain protein

overexpression in Caco-2 cells.

To examine the function of the two Cdx sites, mutations were introduced

into CdxA (Mut 1), CdxB (Mut 2), or both (Mut 1+2) (Fig. 3A). As shown in Figure

3B, mutation in CdxA (Mut 1) or CdxB (Mut 2) sites decreased the promoter

activity to 30% and 61% of that observed with -62 wildtype construct,

respectively. When both sites were mutated (Mut 1+2), promoter activity was

decreased to 15% of the wildtype construct, comparable to the –31 construct

lacking both Cdx sites.

Next we determined the ability of Cdx1 and Cdx2 to activate the claudin-2

promoter. As shown in Figure 4A, Cdx2 but not Cdx1 protein was detectable in

nuclear proteins from Caco-2 cells. Transient expression with either Cdx1 or

Cdx2 alone or in combination resulted in the strong expression of these proteins

in the nuclei of Caco-2 cells 48 hours after transfection (Fig. 4A). Ectopic

expression of Cdx1 did not alter the expression level of Cdx2, nor did Cdx2-

overexpression induce Cdx1 protein expression in the nuclei (Fig. 4A).

As shown in Figure 4B, Cdx1-overexpression resulted in 3.5-fold increase

of the promoter activity driven by -62 construct containing both intact Cdx sites in

Caco-2 cells (92-fold relative to the activity of null pGL3B vector). In contrast,

Cdx2-overexpression increased the activity of the same construct up to 6.7-fold

(177-fold of null pGL3B vector activity).

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 16

Overexpression of Cdx1 and Cdx2 together did not significantly increase

the activity above the values achieved by Cdx2 alone (Fig. 4B). Although

mutation of either CdxA (Mut 1) or CdxB (Mut 2) site retained the ability to

respond to Cdx2-overexpression, promoter activities induced by Cdx2-

overexpression were less than 25 % and 43 % of that observed in -62 construct,

respectively. Similarly, Mut 1 and Mut 2 constructs were less sensitive to Cdx1-

overexpression. In the absence of both Cdx sites neither Cdx1- nor Cdx2-

overexpression induced a significant induction of reporter gene transcription in

Caco-2 cells (Fig. 4B). The ability of Cdx2 to induce a stronger induction of

claudin-2 promoter activity in comparison to Cdx1 was specific for Caco-2 cells.

As demonstrated in Figure 4C, Cdx1 and Cdx2 enhanced claudin-2 promoter

activity in fibroblasts 2.7-fold and 2.8-fold, respectively, whereas Cdx1 induced a

2.9 fold and Cdx2 a 6.7-fold higher promoter activity in Caco-2 cells.

Cdx-2 binds to the Cdx responsive elements of the human claudin-2 promoter

To further define the interaction between Cdx2 and the two Cdx sites,

EMSA and supershifts were performed with nuclear proteins from Caco-2 cells.

These experiments were carried out in post-confluent Caco-2 cells since it was

shown that specific Cdx-DNA complexes can be obscured by unspecific binding

of unknown peptides in nuclear proteins from pre-confluent Caco-2 cells (22).

As shown in Figure 5A, three DNA-protein complexes (A, B and C) were

observed when binding reactions were carried out with radio-labeled wildtype

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 17

oligonucleotide containing both intact Cdx sites (Fig.5A, lane 1). Unlabeled Mut 2

oligonucleotide with an intact CdxA but a mutated CdxB competed with the

formation of all three complexes whereas unlabeled Mut 1 oligoncleotide with a

mutated CdxA but an intact CdxB prevented only the formation of complex B

(Fig. 5A, lane 3 and 4). In supershift assays, anti-Cdx2 antibody shifted only

complex A to reveal two distinct Cdx2-containing protein-DNA complexes (Fig.

5A, lane 7). In contrast, anti-Cdx1 antibody did not affect the mobility of the

complexes (Fig. 5A, lane 6). The Cdx2-containing complex A was also formed

with radio-labeled Mut 2 oligonucleotide used as a probe (Fig. 5A lane 10),

suggesting that this complex is preferentially formed with CdxA. Complex B did

not form on the Cdx sites since this complex was detected and consequently

competed by all three mutated oligonucleotides (Fig. 5A lane 3-5, 9, 11 and 13).

Mutation in CdxA site greatly reduced the formation of complex C, suggesting the

formation of complex C is dependent on this site (Fig. 5A, lane 8 and 12).

Within the SI gene promoter two adjacent Cdx consensus sites may be

able to direct the formation of Cdx2 homodimers (18). We therefore further

characterized the potential coordination of Cdx2 binding by the two Cdx sites. In

these experiments increasing amounts of nuclear proteins of Cdx2-transfected

Caco-2 cells were used. Complex A, which was supershifted by anti-Cdx2

antibody, was observed even in the absence of CdxA site when more nuclear

protein was used (Fig. 5B, lane 7 and 8). However, most of the Cdx2 containing

complexes formed in the presence of CdxA site and did not require the CdxB site

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 18

(Fig. 5, lane 2 - 4, and 10 -12). In addition, when 12 µg of nuclear proteins from

Cdx2-transfected Caco-2 cells were used, an additional complex D was

observed, which was shifted by anti-Cdx2 antibody (Fig. 5B, lane 3, 4, 11 and

12). Although we could not visualize an additional supershifted band derived from

complex D, it may correspond to Cdx2 homodimers, which could not be

distinguished in supershifts from monomeric complexes (18).

HNF-1 enhances Cdx2 mediated activation of human claudin-2 promoter in

Caco-2 cells.

Cdx2 has been shown to regulate intestinal specific LPH gene expression

in synergy with HNF-1α (21). We therefore determined if the HNF-1 site in the

human claudin-2 promoter could contribute to transcriptional regulation. We

compared the effect of HNF-1α- and HNF-1β-overexpression, since both proteins

share highly homologous DNA binding domains but have distinct activation

domains (42). As shown in Figure 6A, Cdx1- and Cdx2-overexpression resulted

in a 3-fold and 5-fold increase of the promoter activity driven by –84 construct

(100-fold and 170-fold relative to that of null pGL3B vector), respectively.

However, transfection of either HNF-1α or HNF-1β alone was not able to

increase promoter activity. In contrast, co-transfection of HNF-1α together with

Cdx2 but not Cdx1 resulted in a 9-fold increase of the promoter activity (293-fold

of the null pGL3B activity) (Fig. 6A). Disruption of the HNF-1 site in –84 reporter

construct prevented a synergistic co-operation of HNF-1α and Cdx2 (Fig. 6A, -62

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 19

construct). As shown in Figure 6B, transfection of Caco-2 cells with HNF-1α

expression constructs resulted in an increase of HNF-1α expression in both the

cytosolic and nuclear protein fractions. In contrast, Cdx2 was exclusively

expressed in nuclear protein fractions of Caco-2 cells even after ectopic

expression (Fig. 6B).

HNF-1 binds its recognition sequence within the human claudin-2 promoter.

To further determine the interaction between HNF-1 proteins and HNF-1

binding site in the human claudin-2 promoter, EMSA and supershifts were

performed with nuclear proteins from Caco-2 cells. As shown in Figure 7, a single

DNA-protein complex was observed when nuclear proteins from mock-

transfected Caco-2 cells was used (lane 1). Addition of 100-fold excess of

unlabeled wildtype but not mutant oligonucleotide prevented the formation of this

complex (Fig. 7, lane 2 and 3). The HNF-1 consensus sequence-protein

complex was supershifted efficiently by anti-HNF-1α antibody but only to a small

extent by anti-HNF-1β antibody (Fig. 7, lane 4 and 5). Transfection with either

HNF-1α- or Cdx2 alone did not alter the formation of this complex (Fig. 7, lane 6

and 7). In contrast, co-transfection with Cdx2 and HNF-1α together resulted in

the increased formation of the complex, which was supershifted by anti-HNF-1α

antibody (Fig. 7, lane 8 and 9).

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 20

HNF-1 is an organ specific regulator of claudin-2 expression.

The in vitro experiments identified HNF-1α as a potential regulator of

claudin-2 expression. Cdx1 and Cdx2 expression is restricted to the intestine,

whereas HNF-1α is also a regulator of gene expression in the liver and kidneys,

organs in which claudin-2 is expressed (1,5,42). In contrast to Cdx2 deficient

animals (43), HNF-1α deficient mice are viable and survive to adulthood (35,44).

We utilized these mice to determine the potential contribution of HNF-1α in the

expression of claudin-2 in different organs. Analysis of the claudin-2 mRNA and

protein expression in these animals revealed that HNF-1α was required for

expression of claudin-2 in the liver (Fig. 8A). Claudin-2 mRNA and protein

expression was absent in the liver of HNF-1α deficient animals whereas claudin-

1 mRNA expression was unaffected (Fig. 8A and B). In contrast, HNF-1α was not

required for claudin-2 mRNA and protein expression in the kidneys (Fig. 8A and

B).

We next analyzed the expression of claudin-1 and claudin-2 along the

cephalo-caudal and crypt-villus axis in wildtype and HNF-1α deficient mice, since

the in vitro experiments suggest the ability of HNF-1α to regulate claudin-2

expression in the presence of Cdx homedomain proteins in intestinal epithelial

cells. Densitometric analysis of Northern blots after normalization to GAPDH

mRNA expression demonstrated that claudin-2 mRNA was differentially

expressed along the cephalo-caudal axis. In wildtype mice, claudin-2 mRNA was

expressed at 17.2 ± 1.7–fold higher levels in the ileum than in the jejunum (Fig.

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 21

9A and B), in good agreement with the recent analysis of claudin-2 protein

expression in the rat intestine (2). In contrast, claudin-1 mRNA was expressed at

2.4 ± 0.4–fold higher levels in the jejunum than in the ileum (Fig. 9A and B).

Claudin-2 mRNA expression pattern correlated with Cdx1 mRNA expression in

the same intestinal segments, which increased 5 ± 0.5-fold from the jejunum to

the ileum (Fig.9A). However, Cdx2 mRNA expression levels were similar in the

jejunum and ileum (Fig. 9A). In the absence of HNF-1α, claudin-2 expression

decreased by 55 ± 10 % in the ileum (Fig. 9A and B). This regulation was specific

for claudin-2, since claudin-1 mRNA expression was not altered in the absence

of HNF-1α in the mouse jejunum and ileum (Fig. 9A and B).

The reduction of claudin-2 mRNA expression could be due to an overall

reduction of claudin-2 gene transcription or a reduced expression in specific

intestinal epithelial cell subsets. We therefore determined the expression and

subcellular distribution of claudin-1 and claudin-2 proteins in the ileum by

immunostaining along the crypt-villus axis in wildtype and HNF-1α deficient mice

(Fig. 9C). Claudin-2 protein was expressed in tight junctions of the crypt and

villus epithelium of the ileum in wildtype mice (Fig. 9C, panel B and G). In the

absence of HNF-1α, claudin-2 expression was restricted to the tight junctions of

the crypt epithelium (Fig. 9C, panel A and F). Claudin-1 expression was not

altered in the absence of HNF-1α, and was observed in tight junctions of the

crypt and villus epithelium of the ileum in both wildtype and HNF-1α deficient

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 22

mice (Fig. 9C, panel C, D, H and I). Incubation with rabbit control serum did not

result in detectable immunoreactivity (Fig. 9C, panel E and J).

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 23

Discussion

The diverse claudin family of tight junction associated proteins has the

potential of directing the variability of paracellular transport and barrier functions

within gastro-intestinal organs (5). Recent evidence demonstrated that claudins

are not only involved in the induction of tight junction formation, but are also able

to regulate water and ion specific paracellular transport mechanisms (9,10). Loss

of claudin-16 results in the inability to absorb magnesium in the thick ascending

limb of Henle (10). Claudin-4 expression resulted in the specific decrease in

absolute sodium permeability, whereas claudin-2 appeared to increase

paracellular conductance in kidney epithelial cells without changing the

paracellular transport of inert compounds (9). The molecular mechanisms

orchestrating the organ specific expression of claudin-2 are unknown. In this

report we provide the first insights into the transcriptional activation events, which

regulate the complex expression pattern of claudin-2.

We demonstrate that the mouse and human claudin-2 promoters contain

conserved binding sequences for Cdx homeodomain proteins and for the POU

homedomain family member HNF-1α. Cdx1 and Cdx2, intestine specific

homeobox proteins, play an important role for the transcription of the intestine

specific expression of several genes such as SI (15), LPH (21), and guanylyl

cyclase C (45). HNF-1α and HNF-1β were first identified as liver-enriched

transcription factors involved in the expression of several plasma proteins,

including albumin and clotting factors (49) and can act either as homo- or

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 24

heterodimers (26). There is increasing evidence that HNF-1α is crucial for the

transcription of the intestine-specific genes such as LPH (21) and SI (29,50).

Our experiments provide the first demonstration that Cdx homeodomain

proteins can initiate transcriptional activation of a TATA-less promoter. In

contrast HNF-family members have been demonstrated to activate tissue type

specific expression of Ksp cadherin (cadherin 16), which lacks TATA-boxes (48).

Similar to the SI and LPH genes, the claudin-2 promoter has two putative

Cdx binding sites. The Cdx binding site containing region of the claudin-2

promoter mediated basal transcriptional activation in intestinal epithelial cells, but

also had activity in fibroblasts and HepG2 cells. Similar to our results, Cdx

consensus binding site containing promoter have been demonstrated to induce

transcriptional activation in fibroblasts without Cdx proteins by undetermined

mechanisms (47). This promoter region may comprise a core promoter, which

contains transcriptional elements sensitive to activation by factors in none-

epithelial cells in the absence of Cdx and HNF-1α. In addition to tissue specific

expressed transcription factors like Cdx1 and Cdx2, silencer, binding upstream of

the investigated promoter region, may be necessary to direct tissue type specific

expression of claudin-2.

Both, Cdx1 and Cdx2 can interact with the Cdx consensus sites within the

claudin-2 promoter, although Cdx2 is the more potent activator of the claudin-2

gene transcription in Caco-2 cells. The stronger induction of claudin-2 promoter

activity by Cdx2 in comparison to Cdx1 was specific for Caco-2 cells, suggesting

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 25

that in these cells Cdx2 may co-operate with other factors enhancing its

transcriptional activity. Our experiments identified HNF-1α as a potential

candidate, since it was able to enhance Cdx2 but not Cdx1 induced claudin-2

promoter activity in Caco-2 cells.

Our results are consistent with the previous observation that Cdx2 is more

effective than Cdx1 in transcriptional activation of the clusterin gene promoter

(46). Although both Cdx binding sites were required for full transcriptional activity

of the human claudin-2 promoter in Caco-2 cells, Cdx2 binding occurred primarily

at the CdxA site. The second CdxB site may serve primarily to support Cdx2

homodimer or oligomer formation as has been proposed for the two Cdx sites in

the SI promoter (18). Alternatively, additional transcription factors may require

CdxB to bind and enhance Cdx2 mediated transcription. The involvement of

additional transcriptional activators may be particularly necessary in the

activation of the mouse claudin-2 promoter, in which the second Cdx-binding site

present in the human promoter is not completely preserved.

In Caco-2 cells, HNF-1α was able to enhanced claudin-2 promoter activity

only in the presence of overexpressed Cdx2. HNF-1α has been demonstrated to

synergize with Cdx2 to induce LPH gene transcription (21). However, whereas

the LPH promoter can be activated by the expression of HNF-1α alone,

activation of the claudin-2 promoter by HNF-1α in Caco-2 cells was dependent

on the recruitment of overexpressed Cdx2 to its binding site. This co-operation

was specific for HNF-1α since HNF-1β failed to enhance Cdx2-mediated

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 26

activation of the claudin-2 promoter. These results are similar to the previous

observations that HNF-1β was less potent than HNF-1α as a transactivator of

LPH (51), SI (29), and α1-antitrypsin (52) genes. Collectively, the promoter

analysis revealed the ability of Cdx homeodomain proteins and HNF-1α to bind

to their recognition sequences in the claudin-2 promoter and to regulate the

activation of this promoter in Caco-2 cells.

We analyzed wildtype and HNF-1α deficient mice to assess the role of

HNF-1α in the regulation of claudin-2 expression. These experiments indicate

that HNF-1α can regulate claudin-2 expression in an organ specific manner.

HNF-1α was required for claudin-2 expression in the liver. HNF-1α deficient mice

have enlarged fatty livers and dysregulated fatty acid homeostasis, which have

been traced in part to a reduced expression of liver fatty acid-binding protein

(53). It is currently not clear if the lack of claudin-2 contributes to the disturbed

liver function in HNF-1α deficient animals.

In contrast, claudin-2 mRNA and protein expression in proximal tubules of

the kidneys were not altered in the absence of HNF-1α. In the kidney HNF-3 may

compensate for the lack of HNF-1α in the activation of the claudin-2 promoter.

HNF-3 has recently been shown to mediate the kidney specific expression of

Ksp-cadherin through a motif similar to the HNF-3-CAAT box containing

sequence found in the claudin-2 promoter partially overlapping with the HNF-1

consensus sequence (48).

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 27

In the absence of HNF-1α, claudin-2 was still expressed in the small

intestine, although its expression was restricted to the crypt epithelium.

The loss of claudin-2 expression in intestinal villi epithelium may be due to the

lack of HNF-1α, which has been demonstrated to be predominantly expressed in

the intestinal epithelial cells of the small intestinal villi (28).

HNF-1α, Cdx1, and Cdx2, are differential expressed along the crypt-villus

axis of the small intestine (28,31). Cdx1 expression has been demonstrated to

localize to intestinal crypts, whereas Cdx2 expression was observed to extend

into small intestinal villi (31). However, recent experiments with antibodies

recognizing phosphorylated Cdx2 demonstrated activated Cdx2 in small

intestinal crypts (54). If the regulation of claudin-2 expression in mice

corresponds to its regulation in Caco-2 cells, HNF-1α may be required to

enhance Cdx2 mediated claudin-2 expression in the intestinal villi, whereas Cdx1

and/or Cdx2 may drive the remaining expression of claudin-2 in the crypt

epithelium of HNF-1α deficient mice.

The function of HNF-1α in the transcriptional regulation of claudin-2

expression was specific since claudin-1 expression was not regulated in the

absence of HNF-1α in the jejunum or ileum. The different transcriptional

regulation of claudin-1 was further apparent in the distinct expression pattern

along the cephalo-caudal axis and the unaltered expression along the in crypt-

villus axis in the absence of HNF-1α. The impact of HNF-1α gene disruption on

the gut has not been examined in detail. The loss of claudin-2 expression in the

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 28

ileal villi and the liver may contribute to the severe phenotype of the HNF-1α

deficient mice. HNF-1α gene disruption in mice leads to dwarfism due to a

reduced Insulin-like growth factor (IGF)-1 synthesis, and an early onset form of

type 2 diabetes mellitus due to impaired glycolytic signaling (35,44,55). However,

impaired intestinal and liver specific secretive or absorptive function may relate to

these phenotypes. Further analysis of the HNF-1α deficient mice should prove

valuable to uncover additional roles of claudin-2 in the regulation of organ

specific functions.

Our studies suggest that the expression of claudin-2 is under the

regulatory control of HNF-1α in the liver and small intestinal villi in mice. Whereas

in the liver HNF-1α is required for claudin-2 expression, in the intestine HNF-1α

may co-operate with additional factors to extend claudin-2 expression from the

crypt into the functionally distinct villus intestinal epithelial cell compartment. It

needs to be determined if the augmentation of claudin-2 gene expression by

HNF-1α in this compartment is dependent on Cdx2 as observed in Caco-2 cells.

Together our experiments support a model in which claudin-2 expression is

governed by distinct organ specific transcriptional mechanisms involving

homeodomain proteins.

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 29

Acknowledgements

This work was supported by National Institutes of Health Grants DK51003,

DK54427 and DK33506 (H.-C.R.), by U. S. Public Health Service Grant DK54399

(DBR) and by March of Dimes Grant#1-FY99-221 (DBR). The authors gratefully

acknowledge Frank J. Gonzalez for providing the HNF-1α deficient mice, Taro

Akiyama for developing the PCR genotyping protocol, and Lihua Zhang for

technical assistance.

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 30

References

1. Furuse, M., Fujita, K., Hiiragi, T., Fujimoto, K., and Tsukita, S. (1998)

J Cell Biol 141(7), 1539-50

2. Rahner, C., Mitic, L. L., and Anderson, J. M. (2001)

Gastroenterology 120(2), 411-22.

3. Kinugasa, T., Sakaguchi, T., Gu, X., and Reinecker, H. C. (2000)

Gastroenterology 118(6), 1001-11

4. Tsukita, S., and Furuse, M. (2000) J Cell Biol 149(1), 13-6.

5. Tsukita, S., Furuse, M., and Itoh, M. (2001)

Nature Reviews Molecular Cell Biology 2(4), 285-93

6. Schneeberger, E. E., and Lynch, R. D. (1992)

Am J Physiol 262(6 Pt 1), L647-61

7. Anderson, J. M., and Van Itallie, C. M. (1995)

Am J Physiol 269(4 Pt 1), G467-75

8. Madara, J. (1998) Ann. Rev. Physiol. 60, 143-159

9. Van Itallie, C., Rahner, C., and Anderson, J. M. (2001)

J Clin Invest 107(10), 1319-27.

10. Simon, D. B., Lu, Y., Choate, K. A., Velazquez, H., Al-Sabban, E.,

Praga, M., Casari, G., Bettinelli, A., Colussi, G., Rodriguez-Soriano, J.,

McCredie, D., Milford, D., Sanjad, S., and Lifton, R. P. (1999)

Science 285(5424), 103-6

11. Morita, K., Furuse, M., Fujimoto, K., and Tsukita, S. (1999)

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 31

Proc Natl Acad Sci U S A 96(2), 511-6

12. Charite, J., de Graaff, W., Consten, D., Reijnen, M. J., Korving, J., and

Deschamps, J. (1998) Development 125(22), 4349-58

13. Duprey, P., Chowdhury, K., Dressler, G. R., Balling, R., Simon, D.,

Guenet, J. L., and Gruss, P. (1988) Genes Dev 2(12A), 1647-54

14. James, R., Erler, T., and Kazenwadel, J. (1994)

J Biol Chem 269(21), 15229-37

15. Suh, E., and Traber, P. G. (1996) Mol Cell Biol 16(2), 619-25.

16. Soubeyran, P., Andre, F., Lissitzky, J. C., Mallo, G. V., Moucadel, V.,

Roccabianca, M., Rechreche, H., Marvaldi, J., Dikic, I., Dagorn, J. C., and

Iovanna, J. L. (1999) Gastroenterology 117(6), 1326-38.

17. Lynch, J., Suh, E. R., Silberg, D. G., Rulyak, S., Blanchard, N., and

Traber, P. G. (2000) J Biol Chem 275(6), 4499-506.

18. Suh, E., Chen, L., Taylor, J., and Traber, P. G. (1994)

Mol Cell Biol 14(11), 7340-51.

19. Lambert, M., Colnot, S., Suh, E., L'Horset, F., Blin, C., Calliot, M. E.,

Raymondjean, M., Thomasset, M., Traber, P. G., and Perret, C. (1996)

Eur J Biochem 236(3), 778-88

20. Troelsen, J. T., Mitchelmore, C., Spodsberg, N., Jensen, A. M., Noren, O.,

and Sjostrom, H. (1997) Biochem J 322(3), 833-8

21. Mitchelmore, C., Troelsen, J. T., Spodsberg, N., Sjostrom, H., and

Noren, O. (2000) Biochem J 346 (2), 529-35

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 32

22. Fang, R., Santiago, N. A., Olds, L. C., and Sibley, E. (2000)

Gastroenterology 118(1), 115-27.

23. Drummond, F., Sowden, J., Morrison, K., and Edwards, Y. H. (1996)

Eur J Biochem 236(2), 670-81

24. Drummond, F. J., Sowden, J., Morrison, K., and Edwards, Y. H. (1998)

FEBS Lett 423(2), 218-22

25. Lorentz, O., Duluc, I., Arcangelis, A. D., Simon-Assmann, P., Kedinger, M.,

and Freund, J. N. (1997) J Cell Biol 139(6), 1553-65.

26. Mendel, D. B., Hansen, L. P., Graves, M. K., Conley, P. B., and

Crabtree, G. R. (1991) Genes Dev 5(6), 1042-56

27. Kuo, C. J., Conley, P. B., Hsieh, C. L., Francke, U., and Crabtree, G. R.

(1990) Proc Natl Acad Sci USA 87(24), 9838-42

28. Serfas, M. S., and Tyner, A. L. (1993) Am J Physiol 265(3 Pt 1), G506-13

29. Boudreau, F., Zhu, Y., and Traber, P. G. (2001)

J Biol Chem 276(34), 32122-8

30. Silberg, D. G., Furth, E. E., Taylor, J. K., Schuck, T., Chiou, T., and

Traber, P. G. (1997) Gastroenterology 113(2), 478-86.

31. Silberg, D. G., Swain, G. P., Suh, E. R., and Traber, P. G. (2000)

Gastroenterology 119(4), 961-71.

32. Bach, I., and Yaniv, M. (1993) EMBO Journal 12(11), 4229-42

33. Awane, M., Andres, P. G., Li, D. J., and Reinecker, H. C. (1999)

J Immunol 162(9), 5337-44

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 33

34. Traber, P. G., Wu, G. D., and Wang, W. (1992) Mol Cell Biol 12(8), 3614-27

35. Lee, Y. H., Sauer, B., and Gonzalez, F. J. (1998)

Mol Cell Biol 18(5), 3059-68

36. Suzuki, Y., Tsunoda, T., Sese, J., Taira, H., Mizushima-Sugano, J., Hata,

H., Ota, T., Isogai, T., Tanaka, T., Nakamura, Y., Suyama, A., Sakaki, Y.,

Morishita, S., Okubo, K., and Sugano, S. (2001)Genome Res 11(5), 677-84.

37. Smale, S. T. (1997) Biochim Biophys Acta 1351(1-2), 73-88

38. Cereghini, S. (1996) FASEB J 10(2), 267-82

39. Wasylyk, C., Wasylyk, B., Heidecker, G., Huleihel, M., and Rapp, U. R.

(1989) Mol Cell Biol 9(5), 2247-50.

40. Gilmore, T. D. (1990) Cell 62(5), 841-3.

41. Orkin, S. H. (1992) Blood 80(3), 575-81.

42. Pontoglio, M. (2000) J Am Societ Nephrol 11 Suppl 16, S140-3

43. Chawengsaksophak, K., James, R., Hammond, V. E., Kontgen, F., and

Beck, F. (1997) Nature 386(6620), 84-7

44. Pontoglio, M., Barra, J., Hadchouel, M., Doyen, A., Kress, C., Bach, J. P.,

Babinet, C., and Yaniv, M. (1996) Cell 84(4), 575-85

45. Park, J., Schulz, S., and Waldman, S. A. (2000)

Gastroenterology 119(1), 89-96

46. Suh, E., Wang, Z., Swain, G. P., Tenniswood, M., and Traber, P. G. (2001)

Am J Physiol 280(1), G149-56

47. Xu, F., Li, H., and Jin, T. (1999) J Biol Chem 274(48), 34310-6.

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 34

48. Whyte, D. A., Li, C., Thomson, R. B., Nix, S. L., Zanjani, R., Karp, S. L.,

Aronson, P. S., and Igarashi, P. (1999) Am J Physiol 277(4 Pt 2), F587-98

49. Mendel, D. B., Khavari, P. A., Conley, P. B., Graves, M. K., Hansen, L. P.,

Admon, A., and Crabtree, G. R. (1991) Science 254(5039), 1762-7

50. Wu, G. D., Chen, L., Forslund, K., and Traber, P. G. (1994)

J Biol Chem 269(25), 17080-5

51. Krasinski, S. D., Van Wering, H. M., Tannemaat, M. R., and Grand, R. J.

(2001) Am J Physiol 281(1), G69-84

52. Hu, C., and Perlmutter, D. H. (1999) Am J Physiol 276(5 Pt 1), G1181-94

53. Akiyama, T. E., Ward, J. M., and Gonzalez, F. J. (2000)

J Biol Chem 275(35), 27117-22

54. Rings, E. H., Boudreau, F., Taylor, J. K., Moffett, J., Suh, E. R., and

Traber, P. G. (2001) Gastroenterology 121(6), 1437-50.

55. Dukes, I. D., Sreenan, S., Roe, M. W., Levisetti, M., Zhou, Y. P., Ostrega, D.,

Bell, G. I., Pontoglio, M., Yaniv, M., Philipson, L., and Polonsky, K. S. (1998)

J Biol Chem 273(38), 24457-64

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 35

Legends

Figure 1. Sequence analysis of the 5’-flanking region of the human and mouse

claudin-2 gene. The nucleotides differing between two species are highlighted in

gray. Potential binding motifs are underlined. The regions used for the reporter

constructs are indicated by the arrows and the corresponding basepair (bp)

numbers.

Figure 2. Analysis of human claudin-2 promoter deletion constructs.

The reporter constructs containing sequentially deleted 5’-flanking fragments

were prepared and transfected into Caco-2, HepG2 and NIHT3T cells as

described in Materials and Methods. Results are expressed as relative luciferase

activity of three different experiments carried out in triplicate (mean ± SD). The

mean value of cells transfected with null pGL3B vector was set to 1.

Figure 3. Mutational analysis of the Cdx binding sites within the human claudin-

2 promoter. A, Sequences of wildtype and mutated Cdx consensus motifs. Cdx

consensus sequences are underlined and mutations are in bold. B, Reporter

gene assay. Transfection into Caco-2 cells and luciferase assay were done.

Results are expressed as relative luciferase activity of three different experiments

carried out in triplicate (mean ± SD). The mean value of cells transfected with

null pGL3B vector was set to 1.

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 36

Figure 4. Cdx1- and Cdx2-overexpression activates the human claudin-2

promoter. A, Western blot analysis of nuclear Cdx1 and Cdx2 protein expression

in Caco-2 cells. Caco-2 cells were transfected with Cdx1- and/or Cdx2-

expression vectors and 10µg of nuclear proteins were analyzed. B and C,

Reporter gene analysis. Caco-2 and NIH3T3 cells were transfected with Cdx1-

and/or Cdx2-expression vectors in the presence of the indicated reporter

constructs and luciferase assays were performed. Results are expressed as

relative luciferase activity of three different experiments carried out in triplicate

(mean ± SD). The mean value of cells transfected with pGL3B vector in the

absence of expression vectors was set to 1.

Figure 5. Interaction of Cdx2 with the Cdx binding sites of the human claudin-2

promoter. A, Preferential binding of Cdx2 to the upstream Cdx site. EMSA was

performed using 4 µg of nuclear proteins from post-confluent Caco-2 cells.

Competitions were done with 100-fold excess of indicated oligonucleotides.

Supershift assays were done by addition of 1µl of either anti-Cdx1 or anti-Cdx2

antibody. The sequences of oligonucleotides (Wt, Mut 1, Mut 2, and Mut 1+2)

are given in Figure 3A. The DNA-protein complexes (A, B, and C) and the

supershifted bands are indicated by arrows and white arrowheads, respectively.

B, Concentration-dependent interaction between Cdx consensus sites and Cdx2.

Nuclear proteins from Cdx2-transfected Caco-2 cells were incubated with the

indicated labeled probes. DNA-protein complexes are indicated by black arrows.

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 37

The white arrow in lane 7 indicates the complex A. The supershifted bands in the

presence of anti-Cdx2 antibody are indicated by white arrowheads.

Figure 6. The effects of HNF-1α- and HNF-1β-overexpression on the human

claudin-2 promoter. A, Reporter gene analysis. Caco-2 cells were transfected

with 0.5 µg of indicated expression vectors in the presence of reporter constructs,

and luciferase assay was performed. Results are expressed as relative

luciferase activity of three different experiments carried out in triplicate (mean ±

SD). The mean value of cells transfected with pGL3B in the absence of

expression vectors was set at 1. B, Western blot analysis of Cdx2 and HNF-1α

proteins in the cytosol and the nuclear protein fractions. Caco-2 cells were

transfected with 0.5 µg of HNF-1α- and/or Cdx2-expression vectors. Two days

after transfection, cytosol and nuclear protein (NE) fractions were prepared.

Equal amounts of proteins (25 µg per lane) were analyzed.

Figure 7. Interactions between HNF-1α and HNF-1β with the HNF-1 binding site

in the human claudin-2 promoter. EMSA was performed using 4 µg of nuclear

proteins of Caco-2 cells transfected with the indicated vectors. Competitions

were done with a 100-fold excess of wild type (Wt) or mutant (Mut)

oligonucleotide. Supershift assays were done by addition of 1µl of anti-HNF-1α

or anti-HNF-1β antibody. The oligonucleotide sequences are given in Materials

and Methods. The specific DNA-protein complex is indicated by an arrow. The

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 38

supershifted bands by anti-HNF-1α and anti-HNF-1β antibodies are indicated by

black and white arrowheads, respectively.

Figure 8. Claudin-1 and claudin-2 expression in the liver and kidneys of HNF-1 α

deficient and wildtype mice. A, Northern blot analysis of claudin-1 and claudin-2

mRNAs in the liver and kidneys of HNF-1α deficient (-/-) and wild type littermates

(+/+). Total RNA (30 µg) was electrophoresed, transferred to a nylon membrane

and hybridized with the indicated probes. B, Immunostaining for claudin-2 protein

in the liver and kidneys of HNF-1α deficient (-/-) and wild type littermates (+/+).

Arrows indicate claudin-2 expression in tight junctions. CV, central vein of the

liver.

Figure 9. Claudin-1 and claudin-2 expression in the small intestine of HNF-1α

deficient and wildtype mice. A, Northern blot analysis of claudin-1, claudin-2,

Cdx1, Cdx2, and HNF-1α mRNA expression in the jejunum and ileum of HNF-1α

deficient (-/-) and wildtype (+/+) mice. B, Densitometric analysis of claudin-1

(open bars) and claudin-2 (black bars) mRNA expression in the presence or

absence of HNF-1α gene. Expression levels of claudin-1 and claudin-2 mRNAs

were normalized for GAPDH mRNA levels in the same RNA isolations and

expressed as relative density per area (mean ± SD, n = 3). C, Immunostaining of

claudin-2 and claudin-1 protein in the ileum of HNF-1α deficient (-/-) and wildtype

(+/+) mice. Frozen sections were stained with either anti-claudin-2 (panel A, B, F

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Sakaguchi et al. 39

and G) or anti-claudin-1 (panel C, D, H and I) antibody and FITC-labeled anti-

rabbit secondary antibody. Panel E and J, control staining with rabbit serum and

secondary antibody. Arrows indicate stainings of claudins in tight junctions

(Original magnification in panel A-E 40x, in panel F-J 100 x).

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from

Hans-Christian ReineckerTakanori Sakaguchi, Xiubin Gu, Heidi M. Golden, EunRan Suh, David B. Rhoads and

αproteins and hepatocyte nuclear factor-1with conserved binding sites in mouse and human for caudal-related homeodomain Cloning of the human claudin-2 5'-flanking region revealed a TATA-less promoter

published online April 4, 2002J. Biol. Chem. 

  10.1074/jbc.M110261200Access the most updated version of this article at doi:

 Alerts:

  When a correction for this article is posted• 

When this article is cited• 

to choose from all of JBC's e-mail alertsClick here

by guest on June 15, 2018http://w

ww

.jbc.org/D

ownloaded from