You are on page 1of 16

INFECTION AND IMMUNITY, Feb. 2008, p. 796811 Vol. 76, No.

2
0019-9567/08/$08.000 doi:10.1128/IAI.00093-07
Copyright 2008, American Society for Microbiology. All Rights Reserved.

Modulation of Intestinal Goblet Cell Function during Infection by an


Attaching and Effacing Bacterial Pathogen
Kirk S. B. Bergstrom,1 Julian A. Guttman,2 Mohammad Rumi,1 Caixia Ma,1 Saied Bouzari,1
Mohammed A. Khan,1 Deanna L. Gibson,1 A. Wayne Vogl,3 and Bruce A. Vallance1*
Division of Gastroenterology, BCs Childrens Hospital,1 Michael Smith Laboratories,2 and Department of
Cellular and Physiological Sciences, Division of Anatomy and Cell Biology,3 University of
British Columbia, Vancouver, British Columbia, Canada
Received 17 January 2007/Returned for modification 8 March 2007/Accepted 27 October 2007

The attaching and effacing (A/E) bacterial pathogens enteropathogenic Escherichia coli and enterohemor-
rhagic E. coli and the related mouse pathogen Citrobacter rodentium colonize their hosts intestines by infecting
the apical surfaces of enterocytes, subverting their function, and they ultimately cause diarrhea. Surprisingly,
little is known about the interactions of these organisms with goblet cells, which are specialized epithelial cells
that secrete the protective molecules Muc2 and trefoil factor 3 (Tff3) into the intestinal lumen. C. rodentium
infection leads to dramatic goblet cell depletion within the infected colon, yet it is not clear whether C.
rodentium infects goblet cells or if this pathology is pathogen or host mediated. As determined by immuno-
staining and PCR, both the number of goblet cells and the expression of genes encoding Muc2 and Tff3 were
significantly reduced by day 10 postinfection. While electron microscopy and immunostaining revealed that C.
rodentium directly infected a fraction of colonic goblet cells, C. rodentium localization did not correlate with
goblet cell depletion. To assess the role of the host immune system in these changes, Rag1 knockout (KO) (T-
and B-cell-deficient) mice were infected with C. rodentium. Rag1 KO mice did not exhibit the reduction in the
number of goblet cells or in mediator (Muc2 and Tff3) expression observed in infected immunocompetent mice.
However, reconstitution of Rag1 KO mice with T and B lymphocytes from C57BL/6 mice restored the goblet cell
depletion phenotype during C. rodentium infection. In conclusion, these studies demonstrated that while colonic
goblet cells can be subject to direct infection and potential subversion by A/E pathogens in vivo, it is the host
immune system that primarily modulates the function of these cells during infection.

Enteropathogenic Escherichia coli (EPEC) and enterohem- cluding A/E pathogens, intestinal epithelial cells can be subject
orrhagic E. coli (EHEC) are food- and waterborne pathogens to direct modulation by the pathogens (4, 22, 50, 67). A/E
that are attaching and effacing (A/E) bacteria, a class of non- pathogens utilize a type III secretion system (T3SS) to secrete
invasive enteric bacterial pathogens (49). These bacteria infect various bacterial effectors encoded in their genomes (e.g., the
the apical surfaces of intestinal epithelial cells, causing a sig- transmembrane intimin receptor Tir) (18, 31) directly into host
nature histopathology characterized by intimate attachment to cells to cause disease. These virulence factors act in an orches-
the host plasma membrane and localized effacement of the trated manner to subvert intracellular signaling pathways
brush border immediately surrounding the attached bacteria within host cells, altering various cellular processes, including
(16, 74). As a result of this infection and the resulting host cytoskeletal (61), organelle (39, 47), and barrier (10, 19, 39)
inflammatory response, EPEC and EHEC cause severe diar- functions. This strategy allows the bacteria not only to inti-
rhea and other complications, leading to the death of hundreds mately attach to and form A/E lesions on epithelial surfaces (6,
of thousands of children worldwide each year (7, 28, 49). Al- 32) but also to suppress inflammatory responses and host de-
though these pathogens are human specific and thus difficult to fenses (44, 65). Through the release of various cytokines, host
model in vivo, Citrobacter rodentium is a natural A/E bacterial immune cells can also modulate intestinal epithelial function
pathogen of mice that is closely related to EPEC and EHEC by altering epithelial cell proliferation (5, 54), migration (8),
(11). C. rodentium infection causes colonic epithelial cell pro- and permeability (24, 75). Such immunomodulation of epithe-
liferation and crypt elongation, as well as inflammation and lial function is thought to represent a critical effector mecha-
diarrhea (20, 38). Because C. rodentium produces A/E lesions nism by which the host is able to mediate clearance of invading
that are virtually indistinguishable from those produced by enteric pathogens, as demonstrated with diverse classes of
EPEC and EHEC (38), it has been widely used as a model to pathogens, such as viruses (5) and helminths (8, 35, 48). How-
study A/E bacterial pathogenesis in vivo (11, 12, 21, 25). ever, while this mechanism has been characterized best for
During infection by several enteric bacterial pathogens, in- parasitic infections, the role of immunomodulation of intesti-
nal epithelial cells during enteric bacterial infections, including
A/E pathogen infections, remains largely undefined.
* Corresponding author. Mailing address: ACB, Rm. K4-188, 4480 Infection by several enteric pathogens, including C. roden-
Oak Street, BCs Childrens Hospital, Vancouver, British Columbia, tium, leads to a dramatic reduction in the number of pheno-
Canada V6H 3V4. Phone: (604) 875-2345, ext. 5118. Fax: (604) 875-
3244. E-mail: bvallance@cw.bc.ca.
typically distinct goblet cells, which is termed goblet cell de-
K.S.B.B. and J.A.G. contributed equally to this work. pletion (38). Intestinal goblet cells are highly polarized

Published ahead of print on 5 November 2007. secretory cells that are present throughout the intestinal tract

796
VOL. 76, 2008 GOBLET CELLS AND C. RODENTIUM INFECTION 797

but are most abundant in the distal colon and rectum (59), protocols employed were approved by the University of British Columbias
where they make up 16% of the total epithelial cell population Animal Care Committee and were in direct accordance with guidelines drafted
the Canadian Council on the Use of Laboratory Animals.
in mice (29). These specialized epithelial cells are thought to Bacterial strains and infection of mice. Mice were infected by oral gavage with
play an important protective role in the intestine by synthesiz- 0.1 ml of an overnight culture in Luria broth containing approximately 2.5 108
ing and secreting several mediators, including the mucin CFU of wild-type C. rodentium (formerly Citrobacter freundii biotype 4280 strain
MUC2 (46) and the small peptide trefoil factor 3 (TFF3) (63). DBS100) (33). For transmission electron microscope (TEM) studies, mice were
also infected with a mutant escN C. rodentium strain lacking a functional T3SS
MUC2 (in mice, Muc2) is a high-molecular-weight glycopro-
(12).
tein that is stored within granules in the apical compartment of Tissue collection. Uninfected control mice or mice at days 6 and 10 postin-
the cell. Under basal conditions or under the influence of host fection (p.i.) were anesthetized with Halothane and killed by cervical dislocation,
or bacterial stimuli (14), goblet cells release MUC2-containing and colons resected for further analysis; the colons were divided in half to
granules into the lumen, where they hydrate and form the separate the proximal and distal portions. Tissues were immediately placed in
10% neutral buffered formalin (Fisher) for histological studies, or they were
structural basis for the mucus gel layer overlying the intestinal placed in RNA later (Qiagen) and stored at 86C for subsequent RNA extrac-
epithelium (69). This mucus layer plays important physiologi- tion or in 4% paraformaldehyde for subsequent freezing and cryosectioning. The
cal roles in the gut; it simultaneously lubricates the intestinal paraformaldehyde-fixed tissues were washed in phosphate-buffered saline (PBS),
surface, limits passage of luminal molecules into the mucosa, incubated in 20% sucrose in PBS overnight at 4C, and then embedded in
functions as a dynamic defensive barrier against enteric patho- Shandon Cryomatrix embedding medium (Thermoelectron Corporation), and
the mold was frozen by partial immersion in liquid N2-precooled 2-methylbutane
gens (14, 59), and acts as a substrate and niche which the and stored at 20C until it was used.
commensal flora can colonize and from which this flora can Bacterial counting. Whole mouse colons, including stools, were washed thor-
derive nutrients (57). TFF3 (in mice, Tff3) is another goblet oughly in PBS (pH 7.4), placed in 1.5 ml of PBS, and homogenized at 19,000 rpm
cell-derived molecule belonging to a family of small cysteine- for 45 s using a Polytron homogenizer (Kinematica). Tissue homogenates were
serially diluted in PBS, plated onto Luria broth agar plates, and incubated
rich secretory peptides that are expressed in a region-specific
overnight at 37C, and bacterial colonies were counted the following day.
manner throughout the gastrointestinal tract (63). A potent RNA extraction and quantitative RT-PCR. Colon tissues stored in RNAlater
inducer of cell migration and an inhibitor of apoptosis (64), (Qiagen) at 86C were thawed on ice and weighed, and the total RNA was
TFF3 plays a critical role in wound healing by promoting ep- extracted using a Qiagen RNeasy kit by following the manufacturers instruc-
ithelial restitution following mucosal injury (45). In addition, tions. Tissues were homogenized in 1 ml of buffer RLT (supplied with the Qiagen
RNeasy kit) using a Polytron homogenizer for 1 min at 26,000 rpm. Total RNA
TFF3 is thought to synergize with colonic mucins to enhance was quantified using a Bio-Rad SmartSpec (Bio-Rad), and 1 to 2 g of RNA was
the protective barrier properties of the mucus layer against reverse transcribed using a Qiagen Omniscript reverse transcription (RT) kit
bacterial toxins (36). Evidence of the importance of goblet cells (Qiagen) according to the manufacturers instructions. cDNA was diluted 1:25 in
in maintaining overall health has come from studies of mice RNase- and DNase-free H2O, and 5 l was added to a 15-l PCR mixture.
Conventional semiquantitative PCR was carried out with an Eppendorf Master-
lacking either Muc2 or Tff3. These mice are highly susceptible
cycler, using the primers for murine Muc2, Tff3, or -actin as a housekeeping
to experimental colitis (45) or have profound defects in intes- control. The sequences of all primer sets used, the PCR conditions, and the cycle
tinal homeostasis under basal conditions (70). numbers are shown in Table 1. Agarose gels were stained with SYBR Safe DNA
Considering the critical role that goblet cells appear to play gel stain (Molecular Probes) and visualized with a Chemi Doc XRS system
in host defense against enteric pathogens, the observation that (Bio-Rad). Densitometric analysis was carried out using ImageJ software 1.38x
(downloaded from the National Institutes of Health website [http://rsb.info.nih
these cells are depleted during C. rodentium infection may .gov/ij/download.html]). For quantitative PCR, Bio-Rad Supermix was used at a
have important implications regarding the pathogenesis of this 1:2 dilution, and real-time PCR was carried out using a Bio-Rad MJ Mini-
infection, as well as infection by clinically important pathogens, Opticon. Quantitation was carried out using GeneEx Macro OM 3.0 software,
including Shigella (53, 60) and Campylobacter (37), where the which employs the 2Ct method for real-time quantification of gene expression.
Melting point analysis confirmed the specificity for each of the PCRs, and the
goblet cell depletion phenotype is also observed. At present,
PCR efficiency for each of primer set was incorporated into the final calculations.
whether the goblet cell depletion seen during C. rodentium Histological staining. Briefly, 5-m paraffin sections were deparaffinized by
infection reflects the death or functional alteration of goblet heating them at 55 to 65C for 10 min, cleared with xylene, and rehydrated
cells is not clear, nor has the expression of goblet cell media- through an ethanol gradient to water. For periodic acid-Schiff (PAS) staining,
tors been assessed in this model. Similarly, it is not clear if this standard histological techniques were used. For immunostaining, either rabbit
polyclonal antisera that recognized the murine colonic mucin Muc2 (1:50; a gift
pathology reflects direct infection and subversion of goblet cell
from Jan Dekker), rabbit polyclonal antisera generated against rat Tff3 (1:200; a
function by C. rodentium, perhaps in an attempt to bypass gift from D. Podolsky), or rat antisera against C. rodentium Tir (1:500; a gift from
mucosal defenses, or alternatively, if the goblet cell depletion W. Deng) were used as the primary antibody. The primary antibodies were
is mediated by the host as a currently cryptic form of host diluted in PBS containing 1% bovine serum albumin. For immunoperoxidase
defense. We hypothesized that goblet cell function during C. staining, antigen retrieval was performed by incubating deparaffinized, rehy-
drated slides in 10 mM citric acid (pH 6.0) at 90 to 100C for 20 min, followed
rodentium infection is subject to modulation by both the patho- by cooling to room temperature. Horseradish peroxidase-linked goat anti-rabbit
gen and components of the hosts immune system. With this immunoglobulin G (IgG) (1:200; Genetex) was used as the secondary antibody.
hypothesis in mind, the current study addressed the mecha- All antibody incubations were carried out in the presence of 0.2% Triton X-100
nisms underlying the intestinal goblet cell depletion that occurs (Sigma) to facilitate cell permeabilization. The immunoreaction was developed
using SigmaFast DAB substrate (Sigma). Stained sections were washed in water,
during C. rodentium infection.
counterstained with Gills hematoxylin, dehydrated using ethanol, cleared using
xylene, and mounted using Entellan (EM Biosciences). For double-immunofluo-
rescence studies with anti-Muc2 and anti-Tir, no permeabilization or antigen
MATERIALS AND METHODS
retrieval methods were used to minimize staining for untranslocated Tir still
Mice. Six- to 8-week-old female C57BL/6 mice and Rag1 knockout (KO) mice present within the bacteria. To obtain frozen sections, 6-m sections were cut,
(with a C57BL/6 background) were purchased from Jackson Laboratories (Bar placed onto Superfrost/Plus slides (Fisher), and subsequently stained with rat
Harbor, ME). Mice were kept in sterilized, filter-topped cages, handled in tissue anti-mouse CD3 (clone 17A2; 1:300; BioLegend) overnight at 4C or with rat
culture hoods, and fed autoclaved food and water under specific-pathogen-free anti-mouse CD45R/B220 (clone RA3-6B2; 1:100; Becton Dickinson) for 1 h at
conditions. Sentinel animals were routinely tested for common pathogens. The room temperature. Immunofluorescent labeling for all stains was carried out
798 BERGSTROM ET AL. INFECT. IMMUN.

TABLE 1. Primer sets and PCR conditions used in this study


Primera No. of cycles
PCR cycle conditions
Target mRNA (endpoint
(denaturation/annealing/extension)b
Direction Sequence PCR only)

Muc2 Forward 5-CTGACCAAGAGCGAACACAA-3 94C for 30 s/55C for 30 s/72C for 45 s 23


Reverse 5-CATGACTGGAAGCAACTGGA-3

Tff3c Forward 5-CAGATTACGTTGGCCTGTCTCC-3 94C for 30 s/60C for 30 s/72C for 30 s 30


Reverse 5-ATGCTTGCTACCCTTGGACCAC-3

TNF-d Forward 5-ATGAGCACAGAAAGCATGATC-3 94C for 30 s/59C for 30 s/72C for 45 s NAe
Reverse 5-TACAGGCTTGTCACTCGAATT-3

IL-17A (IL-17)f Forward 5-GCTCCAGAAGGCCCTCAGA-3 94C for 30 s/60C for 30 s/72C for 30 s NA
Reverse 5-CTTTCCCTCCGCATTGACA-3

-Actin Forward 5-CAGCTTCTTTGCAGCTCCTT-3 94C for 30 s/55 to 60C for 30 s/72C for 30 s 23 or 30
Reverse 5-CTTCTCCATGTCGTCCCAGT-3
a
The same primer sets were used for end point and quantitative PCR experiments described in text.
b
In all PCR experiments there was an initial denaturation step of 95C for 5 min before PCR cycling, and in end point PCR experiments there was an extension
step of 72C for 10 min after the final cycle.
c
Tff3 primer sequences were obtained from reference 27.
d
TNF- primer sequences were obtained from reference 62.
e
NA, not applicable.
f
IL-17 primer sequences were obtained from reference 23.

with the appropriate secondary antibody using AlexaFluor 488-conjugated goat the tissue was washed with sterile PBS. Tissues were then placed in a light-tight
anti-rat IgG, AlexaFluor 568-conjugated goat anti-rabbit IgG, or AlexaFluor specimen chamber that is part of an in vivo imaging system (IVIS; Xenogen,
568-conjugated goat anti-rat IgG (Invitrogen). Tissues were mounted using Pro- Alameda, CA). The bacterial signals were quantified using the software program
Long Gold Antifade (Invitrogen) that contained 4,6-diamidino-2-phenylindole LIVING IMAGE (Xenogen) as an overlay on Igor (Wavemetrics, Seattle, WA).
(DAPI) for DNA staining. Sections were viewed at 350, 488, and 594 nm with a To determine the anatomical location, a pseudocolor image showing the light
Zeiss AxioImager microscope. Images were obtained using a Zeiss AxioImager intensity (blue [least intense] to red [most intense]) was generated using LIVING
microscope equipped with an AxioCam HRm camera operating through Axio- IMAGE software and was superimposed over the grayscale reference image.
Vision software (version 4.4). Immune cell reconstitution of Rag1 KO mice. The adaptive immune system
Quantitative histological studies. (i) Goblet cell enumeration. PAS- and he- was reconstituted in Rag1 KO mice using splenic and mesenteric lymph node
matoxylin-stained sections prepared at various time points were photographed, (MLN) populations of T and B lymphocytes as previously described (66). In
and the total numbers of epithelial cells and PAS-positive cells were determined brief, wild-type immunocompetent mice were euthanized, and their spleens and
for 20 to 30 longitudinally sectioned crypts per section. The number of goblet MLNs were aseptically removed. Spleens and MLNs were placed in RPMI
cells was expressed as the total number of PAS-positive cells per 100 epithelial medium with 10% fetal bovine serum, mashed to a pulp with the rubber end of
cells. Phenotypically mature goblet cells were assessed based on the intensity of the plunger from a 1.0-ml syringe, and then forced through a 70-m-pore-size
staining, the size of the apical region, the location on the crypt base-to-surface filter (BD Biosciences), generating a single-cell suspension. Cells were spun
axis, and morphology, similar to the method described by Katz and coworkers down and resuspended in red blood cell lysis buffer (155 mM NH4Cl, 1 mM
(30). KHO3, 0.01 mM Na2EDTA-2H2O; pH 7.4) for 5 min to lyse the red blood cells.
(ii) Quantification of infected crypts. For crypt infection studies involving Following two washes with RPMI medium, cells were pelleted and then resus-
double immunostaining for Muc2 and Tir, crypts exhibiting goblet cell depletion pended in PBS. Cells were then counted, and viability was analyzed by trypan
were defined as crypts with less than three Muc2-positive phenotypically mature blue exclusion. Recipient Rag1 KO mice were then inoculated via the tail vein
goblet cells and with little Muc2 in the crypt lumen. Crypts that did not exhibit with 2 108 viable mononuclear cells. Mice were left for 6 weeks and then tested
goblet cell depletion were defined as crypts with three or more strongly Muc2- for the success of reconstitution by staining colonic tissue sections with isolated
positive phenotypically mature goblet cells and an intense secreted Muc2 signal lymphoid follicles for the presence of T lymphocytes using the marker CD3 and
in the crypt lumen. Positively infected crypts were defined as crypts that were for the presence of B lymphocytes using the marker B220.
positive for Tir staining on the cells of the surface epithelium and on cells in the Statistical analysis. Statistical significance was calculated by using either a
upper one-third of colonic crypts. two-tailed Student t test or the Mann-Whitney t test as indicated, with assistance
TEM. Mouse colons were immersed for 3 h in a fixative containing 0.1 M from GraphPad Prism software (version 4.00; GraphPad Software, San Diego,
sodium cacodylate, 1.5% paraformaldehyde, and 1.5% glutaraldehyde (pH 7.3). CA) (www.graphpad.com). A P value of 0.05 was considered significant. The
Following fixation, the material was postfixed for 1 h on ice in 1% osmium results were expressed as means and standard errors of the means or as means
tetroxide in 0.1 M sodium cacodylate (pH 7.3) and stained with 0.1% uranyl and standard deviations as indicated below.
acetate. The material was dehydrated using an ascending alcohol series, followed
by incubation in propylene oxide. The blocks were then left in a 1:1 solution of
propylene oxide-Polybed overnight. The material was embedded in 100% Poly-
bed, and the resin was polymerized at 60C for 24 h. Sections were viewed and
RESULTS
photographed using a Philips 300 electron microscope operated at 60 kV.
Depletion of mucus-containing goblet cells correlates with
In vivo imaging: bacterial strains and generation of bioluminescent C. roden-
tium. Bioluminescent strains of C. rodentium were constructed by introducing peak bacterial colonization during C. rodentium infection. Pre-
plasmid pT7 (E. A. Meighen, Department of Biochemistry, McGill University) vious studies have indicated that C. rodentium infection causes
carrying the entire lux operon from Photorhabdus luminescens. Bioluminescent goblet cell depletion within the colons of infected mice (38);
colonies were selected on Luria broth agar plates supplemented with 100 g/ml however, the timing of this pathology and how the changes are
of ampicillin and were screened using a model 1420 Victor3V multilabel counter
(PerkinElmer). For in vivo tissue imaging of bioluminescent C. rodentium at 6
related to C. rodentium colonization have not been addressed.
and 10 days p.i., mice were anesthetized and then euthanized. The colons and To begin, we examined the time course and spatial character-
ceca were removed and opened lengthwise so that the lumen was exposed, and istics of C. rodentium colonization in C57BL/6 mice (Fig. 1).
VOL. 76, 2008 GOBLET CELLS AND C. RODENTIUM INFECTION 799

stain (Fig. 2F), we found that there was a substantially greater


reduction in the size of this population (P 0.002) (Table 2).
As the infection progressed, the hypotrophic goblet cell phe-
notype was still predominant at day 14 p.i., but the number of
large mucin-filled goblet cells began to rebound again by day
21 p.i. (data not shown). Thus, our data indicate that C. ro-
dentium infection leads to a reduction in PAS staining, poten-
tially due to a decrease in mucin glycoprotein content of goblet
cells, rather than an actual loss of the goblet cell lineage. As a
result, we observed only a modest decrease in the total number
of goblet cells, and the dramatic reduction in the large goblet
cells that appeared to be mature explained the previously de-
FIG. 1. C. rodentium infection peaks at day 10 p.i. in C57BL/6 mice. scribed goblet cell depletion.
Each symbol indicates the mean for three independent infections, each Goblet cell-specific gene expression is reduced during infec-
with three mice per time point. The error bars indicate standard
tion. While infection led to a decrease in overall PAS staining
deviations.
and a reduction in the number of phenotypically mature goblet
cells, it remained unclear whether these effects solely reflected
a reduction in mucin expression and production, which directly
When total pathogen burdens were assessed, C. rodentium was impacts goblet cell morphology (70), or whether they were also
found to heavily colonize the colon by day 6 p.i., with the accompanied by altered production of additional goblet cell-
number of bacteria peaking at day 10 p.i. By day 14 p.i, the specific mediators. To assess this, the expression of the major
number of C. rodentium cells began to drop, and the infection goblet cell mucin Muc2, as well as Tff3, was assayed by semi-
essentially cleared by day 21 p.i., with only a few thousand quantitative RT-PCR. As assessed by densitometry of end
bacteria remaining (Fig. 1 and data not shown). In addition, point PCR products shown in Fig. 3A, there were slight but
consistent with recent studies of C. rodentium infection dynam- nonsignificant reductions in expression of Muc2 and Tff3
ics by other groups (76), we found that bioluminescent C. mRNA by day 6 p.i., but there were notably larger and signif-
rodentium cells were predominantly localized in the cecum and icant decreases in expression of Muc2 and especially Tff3
distal half of the colon, and little infection was detected in the mRNA by day 10 p.i. (Fig. 3A, graph). To determine whether
proximal colon at days 6 and 10 p.i. (not shown). C. rodentium infection impacted Muc2 and Tff3 protein expres-
We next examined whether the localization of C. rodentium sion, we performed immunoperoxidase staining for Muc2 and
influenced colonic pathology and the reported goblet cell de- Tff3 of formalin-fixed paraffin-embedded sections of distal co-
pletion by comparing tissue sections taken from the distal and lons from control and infected mice. In control mice, Muc2
proximal colons of uninfected and infected mice at days 6 and protein expression and Tff3 protein expression were abundant
10 p.i. Using the PAS staining technique to stain goblet cell in goblet cells, although the expression patterns were distinct.
mucins, only minimal changes in PAS staining were observed While Muc2-positive cells were observed from the base to the
in any region of the cecum and colon at day 6 pi (data not luminal surface of colonic crypts (Fig. 3B, upper left panel),
shown), and there was little overt change in PAS staining of Tff3-positive goblet cells were confined predominantly to the
goblet cells in the cecum (data not shown) or in the proximal upper half of the crypts (Fig. 3B, lower left panel), in agree-
colon (Fig. 2A and B) at day 10 p.i. However, a dramatic ment with previous reports (52). However, consistent with our
reduction in PAS staining was observed in tissue sections of results showing that infection causes down-regulation of Muc2
samples taken from the distal colon at day 10 p.i. (Fig. 2C and and Tff3 mRNA, we observed a dramatic reduction in staining
D). While a small number of crypts still contained phenotyp- for both the Muc2 and Tff3 proteins at day 10 p.i. throughout
ically mature mucin-filled goblet cells, characterized by a swol- crypts having the depleted goblet cell phenotype (Fig. 3B,
len (PAS-stained) apical region and a tapered basolateral com- upper and lower right panels, respectively), although there was
partment containing the nucleus, in most colonic crypts there still faint Muc2 staining in scattered crypt cells. Thus, our data
were almost none of these cells (Fig. 2E). However, examina- show that infection leads not only to reduced mucin (Muc2)
tion under higher magnification revealed that these tissues expression and production, resulting in decreased PAS stain-
contained many cells that exhibited weak PAS staining in their ing, but also to reduced production of the goblet cell-specific
apical compartments (Fig. 2D). Compared to the cells that mediator Tff3; taken together, the results suggest that goblet
stained strongly with PAS stain (Fig. 2E), these weakly stained cell function is profoundly altered during infection. Consider-
cells were smaller and exhibited a more columnar morphology ing that this pathology was most pronounced in regions of the
(Fig. 2F), reminiscent of the hypotrophic goblet cell pheno- colon heavily infected by C. rodentium, we next addressed
type observed in the interleukin-10 (IL-10)-deficient mouse whether direct infection might play a role in this modulation of
model of spontaneous colitis (41). When we determined the goblet cell function.
number of PAS-positive cells per 100 epithelial cells at day C. rodentium directly interacts with goblet cells in vivo. It
10 p.i., we observed a modest but statistically significant de- remains unclear whether C. rodentium is able to infect cells
crease in the size of the total PAS-positive population (P other than enterocytes in vivo, such as goblet cells. To address
0.036) compared to uninfected controls. However, when we this question, we first assessed whether C. rodentium cells in
determined the number of cells possessing the phenotypically the infected colons were close to or in contact with goblet cells.
mature goblet cell morphology that stained strongly with PAS Here, we focused on day 6 p.i., when colonization in the distal
800 BERGSTROM ET AL. INFECT. IMMUN.

FIG. 2. Depletion of mucus-containing goblet cells occurs in the distal colon and is pronounced when the C. rodentium burdens peak. (A to
D) PAS staining of formalin-fixed paraffin-embedded tissue sections obtained from the proximal (A and B) and distal (C and D) colons of both
nave (uninfected) mice (A and C) and mice at day 10 p.i (B and D). There was a reduction in overall PAS staining within many crypts in distal
colons at day 10 p.i., although scattered PAS-positive cells were still evident (arrows in panel D). Bars 100 m. DIS, distal colon; PROX,
proximal colon. (E and F) Representative images of PAS-positive cells at day 10 p.i. exhibiting a large mucin-filled goblet cell morphology (arrows
in panel E) or a mucin-depleted columnar morphology (arrows in panel F). The images are representative of at least three independent infections
with two or three mice per time point.

colon was established but phenotypically distinct goblet cells where Muc2-positive goblet cells also resided. To confirm the
were still relatively abundant (Table 2). To identify any poten- colocalization of Tir with Muc2-positive cells, we examined
tial C. rodentium-goblet cell interactions, we a performed a tissues at a higher magnification and found positive Tir staining
dual immunostaining analysis to look for colocalization of on the apical surface of enterocytes, as well as on cells that
Muc2-positive goblet cells and the locus of enterocyte efface- were positive for Muc2 (Fig. 4B). Next, we examined whether
ment-encoded virulence factor Tir, which is expressed on the C. rodentium cells were adherent to colonic goblet cells using
apical surface of cells only following direct infection by C. TEM. TEM analysis showed that C. rodentium cells were ad-
rodentium (13). As shown in Fig. 4A, Tir staining was found on herent to the apical plasma membrane of goblet cells (Fig. 4C
the surface epithelium and progressed down some crypts, and D). Although goblet cells have a morphology and function
VOL. 76, 2008 GOBLET CELLS AND C. RODENTIUM INFECTION 801

TABLE 2. Goblet cell enumeration in the distal colon of C57BL/6 that are very distinct from the morphology and function of
mice following C. rodentium infectiona their absorptive enterocytic counterparts, they still contain in-
No. of phenotypically tact, albeit shortened and less dense, microvilli (58). Notably,
Total no. of PAS-positive
mature goblet cells/ however, C. rodentium-associated goblet cells also exhibited
Days p.i. cells/100 epithelial cells
100 epithelial cells
(avg SD) evidence of microvillar effacement, a hallmark of A/E lesion
(avg SD)
formation (Fig. 4C). Interestingly, we also occasionally ob-
Uninfected control 32.9 7.9 17.9 2.8
6 28.1 7.3 14.0 4.6 served bacteria within the apical granule mass of goblet cells,
10 21.6 7.3b 7.7 3.2b suggesting that C. rodentium may be internalized by some gob-
a
C. rodentium infection results in depletion of phenotypically mature goblet
let cells (Fig. 4C). In contrast, when mice were infected with
cells in the distal colon during times corresponding to peak infection. Enumer- escN C. rodentium, which lacks a functional T3SS, no C.
ation was performed using PAS- and hematoxylin-stained tissue sections of distal rodentium cells were found to be adherent to goblet cells, nor
colons by determining the number of distinctly PAS-positive cells per 100 epi-
thelial cells, differentiating the cells that had a mature goblet phenotype. The was any effacement of goblet cell (or enterocyte) microvilli
data are values from three independent infections, with each infection group observed in these mice (Fig. 4E). To determine the frequency
containing three mice per time point.
b
P 0.05, as determined by the Mann-Whitney t test. at which goblet cells were infected, we determined the propor-

FIG. 3. C. rodentium infection results in reduction of Muc2 and Tff3 gene expression. (A) Representative RT-PCR analysis of Muc2 and Tff3
gene expression in distal colonic tissues of nave C57BL/6 mice or mice at days 6 and 10 p.i. -Actin was used as a loading control. Differences
in levels of expression relative to the levels in nave (uninfected) mice after normalization to -actin were determined by densitometric analysis,
as shown in the graph on the right. The bars indicate the means for three independently infected mice per time point. The error bars indicate the
standard errors of the means. Asterisk, P 0.05; ns, not significant (P 0.05), as determined by Students t test. (B) Immunoperoxidase staining
for Muc2 (upper panels) and Tff3 (lower panels) protein (brown) in serial sections of distal colons from either nave mice (left panels) or mice
at day 10 p.i. (right panels). In control mice, Muc2-positive cells are present from the crypt base to the luminal surface, whereas Tff3 expression
is restricted to the top half of the crypts (arrows). The arrowheads indicate cells that are coexpressing the Muc2 and Tff3 proteins. Note the lack
of staining in intact crypts at day 10 p.i. (arrows in left panels). Bar 50 m.
FIG. 4. C. rodentium directly infects colonic goblet cells in vivo. (A) Immunofluorescence staining for the C. rodentium translocated effector Tir (red),
goblet cell-specific Muc2 (green), and DNA (blue) in distal colonic tissues at day 6 p.i. Tir staining is present in Muc2-positive cells (arrowheads) and
Muc2-negative cells within the surface epithelium, progressing down the length of some crypts (arrow). (B) Magnified image of Tir (red) staining (arrows)
on the apical surface of cells staining strongly for Muc2 (green) in the apical compartment and exhibiting distinct goblet cell morphology. GA, goblet cell
apical compartment; GN, goblet cell nucleus; E, enterocyte; L, lumen. Original magnification, 1,000. (C to E) TEM micrographs of distal colons of
C57BL/6 mice taken at day 7 (C) and day 11 (D) following infection with wild-type C. rodentium and at day 7 following infection with escN C. rodentium
(E). C. rodentium is in direct contact with goblet cells in panels C and D (black arrows). In panel C, effacement of microvilli (black arrows) on a goblet
cell and internalization of bacteria into the apical granule mass of the goblet cell (white arrow) can be seen. In panel E, no infection of goblet cells and
intact microvilli (arrow) can be seen following infection with escN C. rodentium. (F) Graph showing the proportions of Muc2-positive (goblet) cells and
Muc2-negative (nongoblet) cells of the surface epithelia that were positive for Tir staining at day 6 p.i. The bars indicate the averages for three mice from
two independent infections. The error bars indicate the standard errors of the means. Asterisk, P 0.05, as determined by Students t test.

802
VOL. 76, 2008 GOBLET CELLS AND C. RODENTIUM INFECTION 803

FIG. 5. C. rodentium associates with crypts that are strongly positive for Muc2 at day 10 p.i. (A) Dual immunofluorescence staining for C.
rodentium Tir (red) and Muc2 (green) in a distal colon at day 10 p.i. Tir staining (white arrow) is associated with crypts strongly positive for Muc2
(white arrowhead), but little if any Tir staining (yellow arrow) is associated with crypts that are weakly positive for Muc2 (yellow arrowhead). Note
the frequent colocalization of Tir with Muc2-postive cells and luminal Muc2. (B) Proportions of crypts exhibiting depletion of Muc2-positive goblet
cells that are positive for Tir compared to crypts that are not depleted of Muc2-postive goblet cells. The bars indicate the averages for 30 crypts
counted within colonic sections from two individual mice, and the error bars indicate standard deviations.

tions of Muc2-positive (goblet) cells and Muc2-negative (en- C. rodentium predominantly associates with crypts that do
terocyte, nongoblet) cells that were positive for Tir staining on not exhibit goblet cell depletion. Our next goal was to examine
the surface epithelium, where most C. rodentium cells were the interactions of C. rodentium with colonic crypts at day
localized. Our results showed that there was a moderate but 10 p.i. in order to determine whether C. rodentium could di-
significant difference between the proportion of infected rectly mediate the loss of mature goblet cells. If this were the
Muc2-positive cells and the proportion of infected Muc2-neg- case, we would expect to find C. rodentium associated with
ative cells, in that fewer than 50% of Muc2-positive cells were hypotrophic goblet cells, which we tested by immunofluores-
positive for Tir staining, whereas just over 60% of Muc2- cently staining for both Tir and Muc2. Strikingly, only minimal
negative cells were positive for Tir staining (Fig. 4F). These Tir staining was associated with crypts exhibiting hypotrophic
results indicate that along with colonocytes, goblet cells are goblet cells, and no C. rodentium cells were identified in prox-
subject to direct infection by A/E pathogens in vivo, although imity to these cells (Fig. 5A). To clarify the relationship be-
Muc2-negative cells are more frequently observed to be in- tween C. rodentium infection and goblet cell depletion, we
fected. identified individual colonic crypts that were infected by C.
804 BERGSTROM ET AL. INFECT. IMMUN.

rodentium (i.e., Tir positive) and, as described in Materials and 10 p.i., and the staining patterns were similar to those observed
Methods, differentiated between crypts that exhibited a weak in uninfected colons (Fig. 7A). In addition, there was not a
overall signal for Muc2, indicative of the goblet cell depletion significant decrease in Muc2 or Tff3 gene expression at day
phenotype, and crypts that displayed a strong signal for Muc2, 10 p.i. compared with C57BL/6 mice (Fig. 7B). To determine
indicating that goblet cell depletion did not occur. Interest- how the preservation of mature goblet cells is related to the
ingly, we found that the proportion of crypts that were infected number of bacteria, we quantified the bacterial burdens at days
and exhibited goblet cell depletion was more than threefold 6 and 10 p.i. Consistent with previous reports (66), Rag1 KO
lower than the proportion of crypts that did not display an mice had greater bacterial burdens than C57BL/6 mice at day
overt loss of goblet cells (Fig. 5B). This trend was even more 10 p.i. (Fig. 6F), which indicates that the maintenance of goblet
dramatic at day 14 p.i., when in virtually all crypts Muc2 ex- cell number and function in the Rag1 KO mice was not due to
pression was markedly reduced and there was little if any reduced numbers of bacteria in these mice and confirms that
associated Tir staining (data not shown). These results suggest direct C. rodentium infection plays a minor role in the overall
that there is a surprising inverse relationship between C. ro- goblet cell depletion observed in C57BL/6 mice. These results
dentium infection of crypts and goblet cell depletion. It should strongly suggested that T and/or B lymphocytes were involved
be noted that we did observe one or two crypts per colonic in the loss of mature goblet cells during C. rodentium infection.
section that were heavily invaded by C. rodentium and were Proinflammatory cytokine expression in C57BL/6 and Rag1
strongly Tir positive from the surface epithelium to the crypt KO mice during C. rodentium infection. Because proinflamma-
base, yet were also depleted of Muc2-positive cells. However, tory cytokines, such as tumor necrosis factor alpha (TNF-),
unlike the typical goblet cell-depleted crypts, these crypts in- have been implicated in causing goblet cell depletion in the
variably looked atrophic and/or necrotic, suggesting that the small intestine in a Salmonella enterica serovar Typhimurium
loss of Muc2 staining in these crypts was due to destruction of ligated ileal loop model (1), we compared TNF- gene expres-
the crypt epithelium rather than to modulation of goblet cell sion in C57BL/6 mice and TNF- gene expression in Rag1 KO
function. These data therefore demonstrate that aside from mice to begin to address which host molecules are responsible
occasional crypt destruction, the widespread goblet cell deple- for the loss of mature goblet cells in the colon during infection.
tion seen in colonic crypts during C. rodentium infection is As shown in Fig. 8, TNF- expression was induced in both
independent of direct bacterial contact or infection of the strains at day 10 p.i. but was induced to a greater extent in
altered goblet cells. C57BL/6 mice; however, the difference did not reach statistical
Rag1 KO mice do not suffer goblet cell depletion during C. significance (P 0.0727). These results indicated that cyto-
rodentium infection. The studies described above suggested kines besides TNF- are probably involved in the observed
that goblet cell depletion is not a result of direct infection, changes in colonic goblet cells during C. rodentium infection.
since it is frequently observed in uninfected crypts. When al- We have previously shown that gamma interferon expression is
ternative mechanisms were considered, the adaptive immune lower in Rag1 KO mice than in C57BL/6 mice during infection
response was shown to modulate goblet cell function during (66), but recent studies have shown that IL-17A (IL-17), a
infection by intestinal helminths (35, 48). Moreover, as previ- proinflammatory T-cell-derived cytokine that has been re-
ous studies have found that a robust adaptive immune re- ported to directly affect intestinal epithelial function (54), is
sponse to C. rodentium infection occurs by day 10 p.i. (26), we also upregulated in mice with C. rodentium-induced colitis
tested whether the host adaptive immune response mediated (42). Therefore, we compared IL-17 mRNA levels in C57BL/6
the observed reduction in the number of mature goblet cells and Rag1 KO mice. As expected, we found that in infected
and Muc2 and Tff3 expression. Thus, Rag1 KO mice (lacking C57BL/6 mice, IL-17 mRNA levels were significantly increased
mature T and B lymphocytes) were infected, and their goblet and were more than 100-fold greater than the levels in unin-
cell responses were analyzed at day 10 p.i. As observed with fected mice; in contrast, IL-17 expression in Rag1 KO mice did
C57BL/6 mice, mucosal thickening associated with epithelial not increase to levels greater than those in uninfected C57BL/6
hyperplasia was observed in Rag1 KO mice, consistent with control mice (Fig. 8), despite the greater numbers of bacteria
previous reports (66). However, compared with infected in the Rag1 KO mice (Fig. 6F). These results, coupled with
C57BL/6 mice (Fig. 6B), no decrease in the number of mature those of previous studies (66), show that effector T-cell cyto-
goblet cells was observed in the Rag1 KO mice at day 10 p.i., kines are highly expressed at times when goblet cell depletion
as assessed by PAS staining (Fig. 6D and 6E); in fact, there was is apparent and may play a direct or indirect role in functional
a modest trend toward increased numbers of PAS-positive cells modulation of goblet cells during infection.
at day 10 p.i. compared to uninfected controls (Fig. 6E). More- Adoptive transfer of T and B lymphocytes rescues the goblet
over, the majority of goblet cells in the Rag1 KO mice, at both cell depletion phenotype in Rag1 KO mice. While the results
the base and surface of the crypts, had a highly differentiated described above indicated that a functional adaptive immune
appearance and contained full mucin-filled apical granule system was important for meditating the observed goblet cell
masses that were stained strongly with PAS stain (Fig. 6D). depletion phenotype, we tested this hypothesis directly by re-
However, as observed for immunocompetent mice in the stud- constituting Rag1 KO mice with T and B lymphocytes isolated
ies mentioned above, we occasionally observed heavily infected from spleens and MLNs of C57BL/6 mice (using PBS as a
and atrophic crypts that exhibited reduced PAS staining, (Fig. negative control) and subsequently challenging them with C.
6D). The maintenance of overall goblet cell function within the rodentium. At day 10 p.i., PBS-treated (nonreconstituted) mice
colons of infected Rag1 KO mice was further confirmed by or mice reconstituted with T and B lymphocytes were eutha-
immunostaining for the Muc2 and Tff3 proteins. High levels of nized, and the distal colons were used for histological assess-
both proteins were detected in infected Rag1 KO mice at day ment of goblet cells via PAS staining, as well as the presence of
VOL. 76, 2008 GOBLET CELLS AND C. RODENTIUM INFECTION 805

FIG. 6. C. rodentium infection does not result in goblet cell depletion phenotype in Rag1 KO mice. (A to D) Histology of PAS- and
hematoxylin-stained distal colonic sections taken from (A) nave C57BL/6 mice, (B) C57BL/6 mice at day 10 p.i., (C) nave Rag1 KO mice, and
(D) Rag1 KO mice at day 10 p.i. The mature PAS-positive cell population is still large in Rag1 KO mice at day 10 p.i. Heavily infected
mucin-depleted atrophic crypts are indicated by arrowheads. Original magnification, 100. Bar 100 m. (E) Quantitation of goblet cells in distal
colons from nave C57BL/6 and Rag1 KO mice or from infected mice at day 10 p.i. The bars indicate the mean numbers of goblet cells per 100
epithelial cells counted in PAS- and hematoxylin-stained sections of the distal colon. The error bars indicate standard errors of the means for at
least three independent infections with one to three mice per time point. Asterisk, P 0.05. (F) Comparison of numbers of bacteria in whole colons
of C57BL/6 mice (filled bars) and Rag1 KO mice (open bars) at day 10 p.i. The bars indicate the means from three independent infections, each
with two or three mice per time point. The error bars indicate the standard deviations. Asterisk, P 0.05, as determined by the Mann-Whitney
t test.

infiltrating lymphocytes via CD3 (T-cell) and B220 (B-cell) within these tissues revealed that while B220-positive cells
staining. Consistent with the hypothesis that T and/or B cells were concentrated mainly in isolated lymphoid follicles found
mediate the goblet cell depletion phenotype, we observed re- only in reconstituted mice (data not shown), the goblet cell
duced overall PAS staining within crypts of the reconstituted depletion seen in the reconstituted mice was associated with a
mice compared with crypts of nonreconstituted mice (Fig. 9A prominent population of CD3-positive cells in the mucosa and
and B). Examination of the T- and B-lymphocyte populations submucosa of the reconstituted mice (Fig. 9C) but not the
806 BERGSTROM ET AL. INFECT. IMMUN.

FIG. 7. Muc2 and Tff3 are abundantly expressed in Rag1 KO mice following C. rodentium infection. (A) Immunohistochemical staining for
Muc2 (upper panels) and Tff3 (lower panels) in the distal colons of nave mice (right panels) and infected mice (left panels). Note the abundance
of Tff3 and Muc2 in crypts that are evidently undergoing hyperplasia, with staining patterns similar to those of nave mice (arrows). Bars 50 m.
Original magnification, 200. (B) Quantitative RT-PCR analysis of Muc2 and Tff3 gene expression in distal colons from nave and infected
C57BL/6 and Rag1 KO mice. Each bar indicates the mean relative expression for four mice from three independent infections compared to nave
mice of the same strain, which were assigned an arbitrary expression value of 1.00. Genes encoding Muc2 and Tff3 were both expressed at
significantly higher levels in Rag1 KO mice than in C57BL/6 mice at day 10 p.i. All samples were normalized to -actin. The error bars indicate
the standard errors of the means. Asterisk, P 0.05, as determined by the Mann-Whitney t test.

nonreconstituted mice (Fig. 9D). Moreover, the reduced PAS significant reduction in Tff3 mRNA expression compared to
staining and the large CD3-positive cell population in distal infected nonreconstituted mice (Fig. 9E). Lastly, similar to the
colons of infected reconstituted mice at day 10 p.i. were also results for infected C57BL/6 mice, we observed significantly
accompanied by a reduction in Muc2 mRNA expression and a greater expression of IL-17 mRNA and a slight, nonsignificant
VOL. 76, 2008 GOBLET CELLS AND C. RODENTIUM INFECTION 807

state of goblet cell function that occurs in the distal colon


during infection. As initially suggested by Makkink et al. for
another model of colitis showing goblet cell depletion (41), the
reduced Muc2 protein expression may have been directly re-
sponsible for the hypotrophic goblet cell phenotype that we
observed, since Muc2 is the main morphological determinant
of goblet cell morphology (41, 70). In fact, mice genetically
deficient in this mucin lack phenotypically distinct goblet cells,
yet they are strongly reactive for other goblet cell markers, like
Tff3 (70). However, in addition to reduced Muc2 levels, we also
found a marked reduction in Tff3 protein levels in C57BL/6
FIG. 8. Cytokine gene expression during C. rodentium infection:
mice but not in Rag1 KO mice. Because Tff3 is thought to be
quantitative RT-PCR analysis of TNF- and IL-17 expression in distal
colons taken from C57BL/6 and Rag1 KO at day 10 following C. expressed primarily by mature goblet cells (51), this suggests
rodentium infection. Induction of TNF- was moderately increased in that the modulation of goblet cell function may reflect immune
both strains at day 10 p.i.; the levels in C57BL/6 mice were increased, system-mediated impairment of the ability of immature goblet
but the difference was not significant (P 0727). However, IL-17 gene cells to fully differentiate into mature goblet cells.
expression was significantly increased in infected C57BL/6 mice com-
pared to the minimal IL-17 expression observed in infected Rag1 KO The biological consequences of the functional modulation of
mice. All samples were normalized to -actin. The bars indicate the goblet cells are currently unclear. In some respects it is para-
means for eight C57BL/6 mice and four Rag1 KO mice from three doxical that the host immune system reduces expression of
independent infections. The error bars indicate the standard errors of Muc2 and Tff3, considering the protective roles that these
the means. Asterisk, P 0.05, as determined by the Mann-Whitney t
goblet cell-derived proteins play in the intestine. Muc2 was
test.
recently shown to be important in maintaining overall mu-
cosal homeostasis, ultimately suppressing spontaneous tu-
mor growth (70), as well as colitis development (68). In vitro
increase in TNF- expression in reconstituted mice compared studies have reported that intestinal mucins prevent EPEC
to infected nonreconstituted mice at day 10 p.i. (Fig. 9F). attachment (40, 56) to epithelial cells, as well as bacterial
Together, these results show that the adaptive immune re- translocation (17) across epithelial cell monolayers. Moreover,
sponse, presumably through the actions of T lymphocytes, in mice lacking Tff3 the ability to heal colonic injury induced by
plays a central role in regulating goblet cell gene expression, the cytotoxic agent dextran sodium sulfate is impaired, and the
with downstream effects on goblet cell function and morphol- mice suffer an exaggerated and fatal colitis as a result (45).
ogy during infection by a noninvasive enteric pathogen. Thus, down-regulating both these genes could compromise the
host defense when an animal is challenged by an enteric bac-
DISCUSSION terial pathogen.
On the other hand, the impact of the loss of these goblet
While goblet cell depletion during C. rodentium infection cell-derived factors may be minimal or even beneficial to the
has been reported previously (38), our studies are the first host during infection with enteric bacteria. For example, sim-
studies to directly characterize this pathology and specifically ilar to commensal species (57), enteropathogenic bacteria such
address the underlying mechanisms. We show here that during as Yersinia enterocolitica have been demonstrated to use car-
C. rodentium-induced colitis, there is a significant reduction in bohydrate-laden mucins as a food source (43), and S. enterica
the size of the mature goblet cell population in the distal colon serovar Typhimurium is thought to bind to intestinal mucins to
during periods corresponding to heavy pathogen burden, and facilitate colonization (73). Thus, reducing mucin production
this is associated with reduced expression of the goblet cell- might be important for reducing energy sources for the patho-
specific genes encoding Muc2 and Tff3 at the mRNA level and, genic bacteria, as well as for reducing potential anchoring sites
histologically, at the protein level. We also demonstrate that in required for initial colonization. Indeed, as C. rodentium con-
addition to colonocytes, goblet cells represent a target of C. stitutes approximately 90% of the bacterial flora at the peak of
rodentium infection, as the bacteria directly interact with and infection (38), reducing a potential nutrient source, such as
infect goblet cells. However, we show that only a portion of mucins, may inhibit pathogen growth. In this regard, the robust
goblet cells in the murine colon are infected, and it is likely the mature goblet cell population observed in infected Rag1 KO
host adaptive immune system that mediates the majority of the mice may facilitate the increased bacterial burdens that are
goblet cell depletion that occurs during C. rodentium infection. observed within the colons of these mice and perhaps even the
The relationship between the histological changes in the previously described impaired clearance of the pathogen (66).
goblet cell population and the alterations in goblet cell-specific The mechanisms underlying the immune system-driven loss
gene expression seen in this model was intriguing and provided of mature Muc2- and Tff3-expressing goblet cells are currently
insight into how the immune system modulates goblet cells unclear; however, goblet cell depletion has also been observed
during infection. While infection did lead to a significant re- in human colonic tissues in association with hyperproliferation
duction in the total number of PAS-stained mucin-containing of colonic crypts, in a manner dependent upon activation of T
cells relative to other non-carbohydrate-producing cells, this in cells within the lamina propria (15). While infection-induced
itself is unlikely to account for the dramatic reduction in Muc2 alterations in the turnover of the colonic epithelium may be
and Tff3 immunostaining that was observed. Rather, the prev- involved in the observed goblet cell depletion, we observed
alence of hypotrophic PAS-positive cells may reflect an altered that both C57BL/6 mice and Rag1 KO mice showed evidence
808 BERGSTROM ET AL. INFECT. IMMUN.

FIG. 9. Adaptive transfer of T and B cells into Rag1 KO mice restored the goblet cell depletion phenotype during C. rodentium infection. (A
and B) PAS staining of distal colons removed on day 10 p.i. from Rag1 KO mice that were reconstituted with T and B lymphocytes from nave
C57BL/6 mice (Recon) (A) or with PBS (nonreconstituted control) (B). The arrows indicate intact crypts exhibiting reduced PAS staining and
numerous hypotrophic goblet cells. The asterisks indicate crypts magnified in the insets. Note that reduced PAS staining in PBS-treated mice
(B) was associated mainly with crypts that were severely disrupted due to bacterial overload (arrowheads), which was not observed in infected
reconstituted mice. Original magnification, 100. Bar 100 m. (C and D) Anti-CD3 staining showing greater numbers of infiltrating
CD3-positive cells in the lamina propria and submucosa of reconstituted Rag1 KO mice (arrows) (C) than in the lamina propria and submucosa
of PBS-treated mice (D). Original magnification, 200. Bar 50 m. (E and F) Quantitative RT-PCR analysis of expression of genes encoding
Muc2 and Tff3 (E) or TNF- and IL-17 (F) in distal colons of reconstituted or nonreconstituted (PBS-treated) mice at day 10 p.i. All samples were
normalized to -actin. The bars indicate the means for seven reconstituted mice and three nonreconstituted mice from two independent infections.
The error bars indicate the standard errors of the means. Asterisk, P 0.05, as determined by the Mann-Whitney t test.
VOL. 76, 2008 GOBLET CELLS AND C. RODENTIUM INFECTION 809

of colonic hyperplasia during infection, consistent with previ- functional modulation by the host immune system during in
ous reports by workers in our lab (66) and other groups (55). vivo infection by this A/E bacterial pathogen. As the host can
We also found that the numbers of epithelial cells within elon- utilize goblet cells for protection against an array of challenges,
gated crypts were not significantly different in the two mouse understanding how the host modulates goblet cells during A/E
strains following infection (unpublished observations). Indeed, and other bacterial challenges should help us determine what
the relationship between crypt hyperplasia and goblet cell de- role these important cell types play during infectious colitis and
pletion phenotypes observed in other models of intestinal in- during maladaptive responses against normal microflora, such
flammation appears to be complex: for example, the immune as those observed during human inflammatory bowel disease.
system-mediated pathology that occurs during murine hel-
minth infections results in both crypt and goblet cell hyperpla- ACKNOWLEDGMENTS
sia (2, 34, 35). Moreover, in SAMP1/YitFc mice which develop
We gratefully acknowledge Jan Dekker for providing the anti-mu-
spontaneous Crohns disease-like ileitis (72), inflammation- rine Muc2 antisera, Daniel Podolsky for providing the anti-Tff3 anti-
induced crypt elongation is associated with the expansion of sera, and Wanyin Deng for providing the anti-C. rodentium Tir anti-
secretory lineages, including Paneth and goblet cells (72). sera. We also thank Mehran Ghoreishi for help with the reconstitution
Taken together, these studies suggest that the mechanisms experiments and Sharon Edwards for help with the immunohistochem-
underlying the depletion of mature goblet cells during C. ro- istry.
This study was supported by operating grants from the Canadian
dentium infection may involve processes that are independent Institutes for Health Research and the Crohns and Colitis Foundation
of, but coordinated with, the induction of crypt hyperplasia. of Canada to B.A.V. B.A.V. is the Children with Intestinal and Liver
The concept of immunomodulation of goblet cells during Disorders (CHILD) Foundation Research Scholar, a Michael Smith
enteric bacterial infection is intriguing in light of the role that Foundation for Health Research Scholar, and the Canada Research
Chair in Pediatric Gastroenterology. K.S.B.B. was supported by an
the immune system plays in modulating goblet cell function MSFHR graduate scholarship. J.A.G. is a CAG/CIHR/AstraZeneca
when it is faced with other intestinal challenges. For example, and MSFHR Postdoctoral Fellow.
the robust Th2 response that typically accompanies intestinal
nematode infections induces goblet cell hyperplasia (34) and REFERENCES
goblet cell-specific gene expression, which are thought to con- 1. Arnold, J. W., G. R. Klimpel, and D. W. Niesel. 1993. Tumor necrosis factor
tribute to host defense (3, 34). In contrast, we observed loss of (TNF) regulates intestinal mucus production during salmonellosis. Cell.
Immunol. 151:336344.
the mature goblet cell phenotype. Given that Th1 responses 2. Artis, D., C. S. Potten, K. J. Else, F. D. Finkelman, and R. K. Grencis. 1999.
(26) and, more recently, Th17 responses (42) have been linked Trichuris muris: host intestinal epithelial cell hyperproliferation during
to C. rodentium infection, it is possible that these polarized chronic infection is regulated by interferon-gamma. Exp. Parasitol. 92:144
153.
T-helper-type responses are specifically responsible for medi- 3. Artis, D., M. L. Wang, S. A. Keilbaugh, W. He, M. Brenes, G. P. Swain, P. A.
ating the loss of the goblet cell phenotype. In this regard, while Knight, D. D. Donaldson, M. A. Lazar, H. R. P. Miller, G. A. Schad, P. Scott,
and G. D. Wu. 2004. RELM/FIZZ2 is a goblet cell-specific immune-effector
we observed increased IL-17 expression during infection in molecule in the gastrointestinal tract Proc. Natl. Acad. Sci. USA 101:13596
immunocompetent mice compared to Rag1 KO mice, we have 13600.
in previous studies observed a similar trend for gamma inter- 4. Boyle, E. C., N. F. Brown, and B. B. Finlay. 2006. Salmonella enterica serovar
Typhimurium effectors SopB, SopE, SopE2 and SipA disrupt tight junction
feron expression (66). Therefore, further studies are needed to structure and function. Cell. Microbiol. 8:19461957.
address which T-cell subsets and potentially which cytokine(s) 5. Brand, S., F. Beigel, T. Olszak, K. Zitzmann, S. T. Eichhorst, J.-M. Otte, J.
are specifically responsible for the modulation of goblet cell Diebold, H. Diepolder, B. Adler, C. J. Auernhammer, B. Goke, and J. Dam-
bacher. 2005. IL-28A and IL-29 mediate antiproliferative and antiviral sig-
function during C. rodentium infection. nals in intestinal epithelial cells and murine CMV infection increases colonic
In addition, our report of direct infection of goblet cells in IL-28A expression. Am. J. Physiol. Gastrointest. Liver Physiol. 289:G960
G968.
vivo by C. rodentium reflects a novel and intriguing host-patho- 6. Celli, J., W. Deng, and B. B. Finlay. 2000. Enteropathogenic Escherichia coli
gen interaction in the intestine that may have consequences for (EPEC) attachment to epithelial cells: exploiting the host cell cytoskeleton
local colonization by this pathogen. Given that C. rodentium, from the outside. Cell. Microbiol. 2:19.
7. Clarke, S. C., R. D. Haigh, P. P. E. Freestone, and P. H. Williams. 2003.
like EPEC and EHEC, can subvert enterocyte function, it is Virulence of enteropathogenic Escherichia coli, a global pathogen. Clin.
tempting to speculate that these pathogens can also subvert the Microbiol. Rev. 16:365378.
function of intestinal goblet cells; however, this possibility has 8. Cliffe, L. J., N. E. Humphreys, T. E. Lane, C. S. Potten, C. Booth, and R. K.
Grencis. 2005. Accelerated intestinal epithelial cell turnover: a new mecha-
yet to be explored in detail. Furthermore, the evidence of nism of parasite expulsion. Science 308:14631465.
possible bacterial internalization within goblet cells is intrigu- 9. Colony, P. C., and R. D. Specian. 1987. Endocytosis and vesicular traffic in
fetal and adult colonic goblet cells. Anat. Rec. 218:365372.
ing, although its significance and frequency of occurrence re- 10. Dean, P., and B. Kenny. 2004. Intestinal barrier dysfunction by enteropatho-
main unclear. While it seems counterintuitive that cells as genic Escherichia coli is mediated by two effector molecules and a bacterial
specialized for secretion as goblet cells could be involved in surface protein. Mol. Microbiol. 54:665675.
11. Deng, W., Y. Li, B. A. Vallance, and B. B. Finlay. 2001. Locus of enterocyte
uptake of luminal contents, it is interesting to note that rodent effacement from Citrobacter rodentium: sequence analysis and evidence for
colonic goblet cells have been observed to internalize their horizontal transfer among attaching and effacing pathogens. Infect. Immun.
apical membrane along with experimentally injected luminal 69:63236335.
12. Deng, W., J. L. Puente, S. Gruenheid, Y. Li, B. A. Vallance, A. Vazquez, J.
cationic ferritin via endocytosis (9). Clearly, the interactions Barba, J. A. Ibarra, P. ODonnell, P. Metalnikov, K. Ashman, S. Lee, D.
between enteric pathogens like C. rodentium and goblet cells Goode, T. Pawson, and B. B. Finlay. 2004. Dissecting virulence: systematic
and functional analyses of a pathogenicity island. Proc. Natl. Acad. Sci. USA
reflect a dynamic host-bacterium interaction, with the goblet 101:35973602.
cells exposed to A/E effectors and C. rodentium directly ex- 13. Deng, W., B. A. Vallance, Y. Li, J. L. Puente, and B. B. Finlay. 2003.
posed to proteins secreted by the goblet cells. Citrobacter rodentium translocated intimin receptor (Tir) is an essential
virulence factor needed for actin condensation, intestinal colonization and
In conclusion, we demonstrate here that although goblet colonic hyperplasia in mice. Mol. Microbiol. 48:95115.
cells are infected by C. rodentium, they are also subject to 14. Deplancke, B., and H. R. Gaskins. 2001. Microbial modulation of innate
810 BERGSTROM ET AL. INFECT. IMMUN.

defense: goblet cells and the intestinal mucus layer. Am. J. Clin. Nutr. 38. Luperchio, S. A., and D. B. Schauer. 2001. Molecular pathogenesis of
73:1131S1141. Citrobacter rodentium and transmissible murine colonic hyperplasia. Mi-
15. Evans, C. M., A. D. Phillips, J. A. Walker-Smith, and T. T. MacDonald. 1992. crobes Infect. 3:333340.
Activation of lamina propria T cells induces crypt epithelial proliferation and 39. Ma, C., M. E. Wickham, J. A. Guttman, W. Deng, J. Walker, K. L. Madsen,
goblet cell depletion in cultured human fetal colon. Gut 33:230235. K. Jacobson, W. A. Vogl, B. B. Finlay, and B. A. Vallance. 2006. Citrobacter
16. Frankel, G., A. D. Phillips, I. Rosenshine, G. Dougan, J. B. Kaper, and S. rodentium infection causes both mitochondrial dysfunction and intestinal
Knutton. 1998. Enteropathogenic and enterohaemorrhagic Escherichia coli: epithelial barrier disruption in vivo: role of mitochondrial associated protein
more subversive elements. Mol. Microbiol. 30:911921. (Map). Cell. Microbiol. 8:16691686.
17. Gork, A. S., N. Usui, E. Ceriati, R. A. Drongowski, M. D. Epstein, A. G. 40. Mack, D. R., S. Michail, S. Wei, L. McDougall, and M. A. Hollingsworth.
Coran, and C. M. Harmon. 1999. The effect of mucin on bacterial translo- 1999. Probiotics inhibit enteropathogenic E. coli adherence in vitro by in-
cation in I-407 fetal and Caco-2 adult enterocyte cultured cell lines. Pediatr. ducing intestinal mucin gene expression. Am. J. Physiol. Gastrointest. Liver
Surg. Int. 15:155159. Physiol. 276:G941G950.
18. Gruenheid, S., I. Sekirov, N. A. Thomas, W. Deng, P. ODonnell, D. Goode, 41. Makkink, M. K., N. M. J. Schwerbrock, M. Mahler, J. A. Boshuizen, I. B.
Y. Li, E. A. Frey, N. F. Brown, P. Metalnikov, T. Pawson, K. Ashman, and Renes, M. Cornberg, H. J. Hedrich, A. W. C. Einerhand, H. A. Buller, S.
B. B. Finlay. 2004. Identification and characterization of NleA, a non-LEE- Wagner, M. L. Enss, and J. Dekker. 2002. Fate of goblet cells in experimen-
encoded type III translocated virulence factor of enterohaemorrhagic Esch- tal colitis. Dig. Dis. Sci. 47:22862297.
erichia coli O157:H7. Mol. Microbiol. 51:12331249. 42. Mangan, P. R., L. E. Harrington, D. B. OQuinn, W. S. Helms, D. C. Bullard,
19. Guttman, J. A., Y. Li, M. E. Wickham, W. Deng, A. W. Vogl, and B. B. Finlay. C. O. Elson, R. D. Hatton, S. M. Wahl, T. R. Schoeb, and C. T. Weaver. 2006.
2006. Attaching and effacing pathogen-induced tight junction disruption in Transforming growth factor- induces development of the TH17 lineage.
vivo. Cell. Microbiol. 8:634645. Nature 441:231234.
20. Guttman, J. A., F. N. Samji, Y. Li, W. Deng, A. Lin, and B. B. Finlay. 2007. 43. Mantle, M., and C. Rombough. 1993. Growth in and breakdown of purified
Aquaporins contribute to diarrhoea caused by attaching and effacing bacte- rabbit small intestinal mucin by Yersinia enterocolitica. Infect. Immun. 61:
rial pathogens. Cell. Microbiol. 9:131141. 41314138.
21. Guttman, J. A., F. N. Samji, Y. Li, A. W. Vogl, and B. B. Finlay. 2006. 44. Maresca, M., D. Miller, S. Quitard, P. Dean, and B. Kenny. 2005. Entero-
Evidence that tight junctions are disrupted due to intimate bacterial contact pathogenic Escherichia coli (EPEC) effector-mediated suppression of anti-
and not inflammation during attaching and effacing pathogen infection in microbial nitric oxide production in a small intestinal epithelial model sys-
vivo. Infect. Immun. 74:60756084. tem. Cell. Microbiol. 7:17491762.
22. Handa, Y., M. Suzuki, K. Ohya, H. Iwai, N. Ishijima, A. J. Koleske, Y. Fukui, 45. Mashimo, H., D.-C. Wu, D. K. Podolsky, and M. C. Fishman. 1996. Impaired
and C. Sasakawa. 2007. Shigella IpgB1 promotes bacterial entry through the defense of intestinal mucosa in mice lacking intestinal trefoil factor. Science
ELMO-Dock180 machinery. Nat. Cell Biol. 9:121128. 274:262265.
23. Happel, K. I., E. A. Lockhart, C. M. Mason, E. Porretta, E. Keoshkerian, 46. Moncada, D. M., S. J. Kammanadiminti, and K. Chadee. 2003. Mucin and
A. R. Odden, S. Nelson, and A. J. Ramsay. 2005. Pulmonary interleukin-23 Toll-like receptors in host defense against intestinal parasites. Trends Para-
gene delivery increases local T-cell immunity and controls growth of Myco- sitol. 19:305311.
bacterium tuberculosis in the lungs. Infect. Immun. 73:57825788. 47. Nagai, T., A. Abe, and C. Sasakawa. 2005. Targeting of enteropathogenic
24. Heller, F., P. Florian, C. Bojarski, J. Richter, M. Christ, B. Hillenbrand, J. Escherichia coli EspF to host mitochondria is essential for bacterial patho-
Mankertz, A. Gitter, N. Burgel, M. Fromm, M. Zeitz, I. Fuss, W. Strober, genesis: critical role of the 16th leucine residue in EspF. J. Biol. Chem.
and J. D. Schulzke. 2005. Interleukin-13 is the key effector Th2 cytokine in 280:29983011.
ulcerative colitis that affects epithelial tight junctions, apoptosis, and cell 48. Nair, M. G., K. J. Guild, and D. Artis. 2006. Novel effector molecules in type
restitution. Gastroenterology 129:550564. 2 inflammation: lessons drawn from helminth infection and allergy. J. Im-
25. Higgins, L. M., G. Frankel, I. Connerton, N. S. Goncalves, G. Dougan, and munol. 177:13931399.
T. T. MacDonald. 1999. Role of bacterial intimin in colonic hyperplasia and
49. Nataro, J. P., and J. B. Kaper. 1998. Diarrheagenic Escherichia coli. Clin.
inflammation. Science 285:588591.
Microbiol. Rev. 11:142201.
26. Higgins, L. M., G. Frankel, G. Douce, G. Dougan, and T. T. MacDonald.
50. Nougayrede, J.-P., P. J. Fernandes, and M. S. Donnenberg. 2003. Adhesion
1999. Citrobacter rodentium infection in mice elicits a mucosal Th1 cytokine
of enteropathogenic Escherichia coli to host cells. Cell. Microbiol. 5:359
response and lesions similar to those in murine inflammatory bowel disease.
372.
Infect. Immun. 67:30313039.
51. Podolsky, D. K., K. Lynch-Devaney, J. L. Stow, P. Oates, B. Murgue, M.
27. Hinz, M., H. Schwegler, C. E. Chwieralski, G. Laube, R. Linke, W. Pohle,
DeBeaumont, B. E. Sands, and Y. R. Mahida. 1993. Identification of human
and W. Hoffmann. 2004. Trefoil factor family (TFF) expression in the mouse
intestinal trefoil factor. Goblet cell-specific expression of a peptide targeted
brain and pituitary: changes in the developing cerebellum. Peptides 25:827
for apical secretion. J. Biol. Chem. 268:66946702.
832.
28. Kaper, J. B. 1998. Enterohemorrhagic Escherichia coli. Curr. Opin. Micro- 52. Renes, I. B., M. Verburg, D. J. Van Nispen, J. A. Taminiau, H. A. Buller, J.
biol. 1:103108. Dekker, and A. W. Einerhand. 2002. Epithelial proliferation, cell death, and
29. Karam, S. M. 1999. Lineage commitment and maturation of epithelial cells gene expression in experimental colitis: alterations in carbonic anhydrase I,
in the gut. Front. Biosci. 4:D286D298. mucin MUC2, and trefoil factor 3 expression. Int. J. Colorectal Dis. 17:317
326.
30. Katz, J. P., N. Perreault, B. G. Goldstein, C. S. Lee, P. A. Labosky, V. W.
Yang, and K. H. Kaestner. 2002. The zinc-finger transcription factor Klf4 is 53. Sachdev, H. P., V. Chadha, V. Malhotra, A. Verghese, and R. K. Puri. 1993.
required for terminal differentiation of goblet cells in the colon. Develop- Rectal histopathology in endemic Shigella and Salmonella diarrhea. J. Pe-
ment 129:2619S2628. diatr. Gastroenterol. Nutr. 16:3338.
31. Kelly, M., E. Hart, R. Mundy, O. Marches, S. Wiles, L. Badea, S. Luck, M. 54. Schwartz, S., J. F. Beaulieu, and F. M. Ruemmele. 2005. Interleukin-17 is a
Tauschek, G. Frankel, R. M. Robins-Browne, and E. L. Hartland. 2006. potent immuno-modulator and regulator of normal human intestinal epithe-
Essential role of the type III secretion system effector NleB in colonization lial cell growth. Biochem. Biophys. Res. Commun. 337:505509.
of mice by Citrobacter rodentium. Infect. Immun. 74:23282337. 55. Simmons, C. P., S. Clare, M. Ghaem-Maghami, T. K. Uren, J. Rankin, A.
32. Kenny, B., S. Ellis, A. D. Leard, J. Warawa, H. Mellor, and M. A. Jepson. Huett, R. Goldin, D. J. Lewis, T. T. MacDonald, R. A. Strugnell, G. Frankel,
2002. Co-ordinate regulation of distinct host cell signalling pathways by and G. Dougan. 2003. Central role for B lymphocytes and CD4 T cells in
multifunctional enteropathogenic Escherichia coli effector molecules. Mol. immunity to infection by the attaching and effacing pathogen Citrobacter
Microbiol. 44:10951107. rodentium. Infect. Immun. 71:50775086.
33. Khan, M. A., C. Ma, L. A. Knodler, Y. Valdez, C. M. Rosenberger, W. Deng, 56. Smith, C. J., J. B. Kaper, and D. R. Mack. 1995. Intestinal mucin inhibits
B. B. Finlay, and B. A. Vallance. 2006. Toll-like receptor 4 contributes to adhesion of human enteropathogenic Escherichia coli to HEp-2 cells. J. Pe-
colitis development but not to host defense during Citrobacter rodentium diatr. Gastroenterol. Nutr. 21:269276.
infection in mice. Infect. Immun. 74:25222536. 57. Sonnenburg, J. L., J. Xu, D. D. Leip, C.-H. Chen, B. P. Westover, J. Weather-
34. Khan, W. I., P. Blennerhasset, C. Ma, K. I. Matthaei, and S. M. Collins. ford, J. D. Buhler, and J. I. Gordon. 2005. Glycan foraging in vivo by an
2001. Stat6 dependent goblet cell hyperplasia during intestinal nematode intestine-adapted bacterial symbiont. Science 307:19551959.
infection. Parasite Immunol. 23:3942. 58. Specian, R. D., and M. R. Neutra. 1981. The surface topography of the
35. Khan, W. I., and S. M. Collins. 2004. Immune-mediated alteration in gut colonic crypt in rabbit and monkey. Am. J. Anat. 160:461472.
physiology and its role in host defence in nematode infection. Parasite 59. Specian, R. D., and M. G. Oliver. 1991. Functional biology of intestinal
Immunol. 26:319326. goblet cells. Am. J. Physiol. Cell Physiol. 260:C183C193.
36. Kindon, H., C. Pothoulakis, L. Thim, K. Lynch-Devaney, and D. Podolsky. 60. Steinberg, S. E., J. G. Banwell, J. H. Yardley, G. T. Keusch, and T. R.
1995. Trefoil peptide protection of intestinal epithelial barrier function: Hendrix. 1975. Comparison of secretory and histological effects of shigella
cooperative interaction with mucin glycoprotein. Gastroenterology 109:516 and cholera enterotoxins in rabbit jejunum. Gastroenterology 68:309317.
523. 61. Stevens, J. M., E. E. Galyov, and M. P. Stevens. 2006. Actin-dependent
37. Lambert, M. E., P. F. Schofield, A. G. Ironside, and B. K. Mandal. 1979. movement of bacterial pathogens. Nat. Rev. Microbiol. 4:91101.
Campylobacter colitis. Br. Med. J. i:857859. 62. Sugawara, I., H. Yamada, C. Li, S. Mizuno, O. Takeuchi, and S. Akira. 2003.
VOL. 76, 2008 GOBLET CELLS AND C. RODENTIUM INFECTION 811

Mycobacterial infection in TLR2 and TLR6 knockout mice. Microbiol. Im- 1999. Gastrointestinal expression and partial cDNA cloning of murine Muc2.
munol. 47:327336. Am. J. Physiol. Gastrointest Liver Physiol. 276:G115G124.
63. Taupin, D., and D. K. Podolsky. 2003. Trefoil factors: initiators of mucosal 70. Velcich, A., W. Yang, J. Heyer, A. Fragale, C. Nicholas, S. Viani, R. Kucher-
healing. Nat. Rev. Mol. Cell Biol. 4:721732. lapati, M. Lipkin, K. Yang, and L. Augenlicht. 2002. Colorectal cancer in
64. Taupin, D. R., K. Kinoshita, and D. K. Podolsky. 2000. Intestinal trefoil mice genetically deficient in the mucin Muc2. Science 295:17261729.
factor confers colonic epithelial resistance to apoptosis. Proc. Natl. Acad. 71. Reference deleted.
Sci. USA 97:799804. 72. Vidrich, A., J. M. Buzan, S. Barnes, B. K. Reuter, K. Skaar, C. Ilo, F.
65. Vallance, B. A., W. Deng, M. De Grado, C. Chan, K. Jacobson, and B. B. Cominelli, T. Pizarro, and S. M. Cohn. 2005. Altered epithelial cell lineage
Finlay. 2002. Modulation of inducible nitric oxide synthase expression by the allocation and global expansion of the crypt epithelial stem cell population
attaching and effacing bacterial pathogen Citrobacter rodentium in infected are associated with ileitis in SAMP1/YitFc mice. Am. J. Pathol. 166:1055
mice. Infect. Immun. 70:64246435.
1067.
66. Vallance, B. A., W. Deng, L. A. Knodler, and B. B. Finlay. 2002. Mice lacking
73. Vimal, D., M. Khullar, S. Gupta, and N. Ganguly. 2000. Intestinal mucins:
T and B lymphocytes develop transient colitis and crypt hyperplasia yet suffer
the binding sites for Salmonella typhimurium. Mol. Cell. Biochem. 204:107
impaired bacterial clearance during Citrobacter rodentium infection. Infect.
Immun. 70:20702081. 117.
67. Vallance, B. A., and B. B. Finlay. 2000. Exploitation of host cells by entero- 74. Wales, A. D., M. J. Woodward, and G. R. Pearson. 2005. Attaching-effacing
pathogenic Escherichia coli. Proc. Natl. Acad. Sci. USA 97:87998806. bacteria in animals. J. Comp. Pathol. 132:126.
68. Van der Sluis, M., B. A. E. De Koning, A. C. J. M. De Bruijn, A. Velcich, 75. Wang, F., W. V. Graham, Y. Wang, E. D. Witkowski, B. T. Schwarz, and J. R.
J. P. P. Meijerink, J. B. Van Goudoever, H. A. Buller, J. Dekker, I. Van Turner. 2005. Interferon- and tumor necrosis factor- synergize to induce
Seuningen, I. B. Renes, and A. W. C. Einerhand. 2006. Muc2-deficient mice intestinal epithelial barrier dysfunction by up-regulating myosin light chain
spontaneously develop colitis, indicating that MUC2 is critical for colonic kinase expression. Am. J. Pathol. 166:409419.
protection. Gastroenterology 131:117129. 76. Wiles, S., K. M. Pickard, K. Peng, T. T. MacDonald, and G. Frankel. 2006.
69. Van Klinken, B. J.-W., A. W. C. Einerhand, L. A. Duits, M. K. Makkink, In vivo bioluminescence imaging of the murine pathogen Citrobacter roden-
K. M. A. J. Tytgat, I. B. Renes, M. Verburg, H. A. Buller, and J. Dekker. tium. Infect. Immun. 74:53915396.

Editor: J. F. Urban, Jr.

You might also like