SlideShare a Scribd company logo
1 of 12
Download to read offline
Bacteriophage adhering to mucus provide
a non–host-derived immunity
Jeremy J. Barra,1
, Rita Auroa
, Mike Furlana
, Katrine L. Whitesona
, Marcella L. Erbb
, Joe Poglianob
, Aleksandr Stotlanda
,
Roland Wolkowicza
, Andrew S. Cuttinga
, Kelly S. Dorana
, Peter Salamonc
, Merry Youled
, and Forest Rohwera
a
Department of Biology, San Diego State University, San Diego, CA 92182; b
Division of Biological Sciences, University of California, San Diego, CA 92093;
c
Department of Mathematics and Statistics, San Diego State University, San Diego, CA 92182; and d
Rainbow Rock, Ocean View, HI 96737
Edited by Richard E. Lenski, Michigan State University, East Lansing, MI, and approved April 18, 2013 (received for review March 28, 2013)
Mucosal surfaces are a main entry point for pathogens and the
principal sites of defense against infection. Both bacteria and
phage are associated with this mucus. Here we show that phage-
to-bacteria ratios were increased, relative to the adjacent envi-
ronment, on all mucosal surfaces sampled, ranging from cnidarians
to humans. In vitro studies of tissue culture cells with and without
surface mucus demonstrated that this increase in phage abun-
dance is mucus dependent and protects the underlying epithelium
from bacterial infection. Enrichment of phage in mucus occurs via
binding interactions between mucin glycoproteins and Ig-like
protein domains exposed on phage capsids. In particular, phage
Ig-like domains bind variable glycan residues that coat the mucin
glycoprotein component of mucus. Metagenomic analysis found
these Ig-like proteins present in the phages sampled from many
environments, particularly from locations adjacent to mucosal
surfaces. Based on these observations, we present the bacterio-
phage adherence to mucus model that provides a ubiquitous, but
non–host-derived, immunity applicable to mucosal surfaces. The
model suggests that metazoan mucosal surfaces and phage co-
evolve to maintain phage adherence. This benefits the metazoan
host by limiting mucosal bacteria, and benefits the phage through
more frequent interactions with bacterial hosts. The relationships
shown here suggest a symbiotic relationship between phage and
metazoan hosts that provides a previously unrecognized antimi-
crobial defense that actively protects mucosal surfaces.
symbiosis | host-pathogen | virus | immunoglobulin | immune system
Mucosal surfaces are the primary zones where animals meet
their environment, and thus also the main points of entry
for pathogenic microorganisms. The mucus layer is heavily col-
onized by bacteria, including many symbionts that contribute
additional genetic and metabolic potential to the host (1, 2).
Bacterial symbionts associated with a variety of other host sur-
faces also provide goods and services, e.g., nutrients (3–6), bio-
luminescence (7, 8), and antibiotics (9, 10). These resident
symbionts benefit from increased nutrient availability (5, 11–13),
as well as the opportunity for both vertical transmission and in-
creased dissemination (14–16).
Within the mucus, the predominant macromolecules are the
large (up to 106
–109
Da) mucin glycoproteins. The amino acid
backbone of these proteins incorporates tandem repeats of ex-
posed hydrophobic regions alternating with blocks bearing ex-
tensive O-linked glycosylation (17). Hundreds of variable,
branched, negatively charged glycan chains extend 0.5–5 nm
from the peptide core outward into the surrounding environment
(17, 18). Other proteins, DNA, and cellular debris also are
present. Continual secretion and shedding of mucins maintain
a protective mucus layer from 10–700 μm thick depending on
species and body location (19–22).
By offering both structure and nutrients, mucus layers com-
monly support higher bacterial concentrations than the sur-
rounding environment (11, 23). Of necessity, hosts use a variety of
mechanisms to limit microbial colonization (24–27). Secretions
produced by the underlying epithelium influence the composition
of this microbiota (12, 27, 28). When invaded by pathogens, the
epithelium may respond by increased production of antimicrobial
agents, hypersecretion of mucin, or alteration of mucin glycosyl-
ation patterns to subvert microbial attachment (29–31).
Also present in the mucus environment are bacteriophage
(phage), the most common and diverse biological entities. As
specific bacterial predators, they increase microbial diversity
through Red Queen/kill-the-winner dynamics (32, 33). Many
phages establish conditional symbiotic relationships with their
bacterial hosts through lysogeny. As integrated prophages, they
often express genes that increase host fitness or virulence (34–
36) and protect their host from lysis by related phages. As free
phage, they aid their host strain by killing related competing
strains (37–39). Phages participate, along with their bacterial
hosts, in tripartite symbioses with metazoans that affect meta-
zoan fitness (40–43). However, no direct symbiotic interactions
between phage and metazoans are known.
Recently, Minot et al. (44) showed that phages in the human
gut encode a population of hypervariable proteins. For 29 hyper-
variable regions, evidence indicated that hypervariability was con-
ferred by targeted mutagenesis through a reverse transcription
mechanism (44, 45). Approximately half of these encoded proteins
possessed the C-type lectin fold previously found in the major
tropism determinant protein at the tip of the Bordetella phage BPP-
1 tail fibers (46); six others contained Ig-like domains. These Ig-like
proteins, similar to antibodies and T-cell receptors, can accom-
modate large sequence variation (>1013
potential alternatives)
(47). Ig-like domains also are displayed in the structural proteins of
many phage (48, 49). That most of these displayed Ig-like domains
are dispensable for phage growth in the laboratory (45, 49) led to
the hypothesis that they aid adsorption to their bacterial prey under
environmental conditions (49). The possible role and function of
these hypervariable proteins remain to be clarified.
Here, we show that phage adhere to mucus and that this as-
sociation reduces microbial colonization and pathology. In vitro
studies demonstrated that this adherence was mediated by the
interaction between displayed Ig-like domains of phage capsid
proteins and glycan residues, such as those in mucin glyco-
proteins. Homologs of these Ig-like domains are encoded by
phages from many environments, particularly those adjacent to
mucosal surfaces. We propose the bacteriophage adherence
to mucus (BAM) model whereby phages provide a non–host-
derived antimicrobial defense on the mucosal surfaces of diverse
metazoan hosts.
Author contributions: J.J.B. and F.R. designed research; J.J.B., R.A., K.L.W., M.L.E., J.P.,
A.S.C., and P.S. performed research; J.J.B., K.L.W., M.L.E., J.P., A.S., R.W., A.S.C., and
K.S.D. contributed new reagents/analytic tools; J.J.B., R.A., M.F., K.L.W., A.S., R.W., P.S.,
M.Y., and F.R. analyzed data; and J.J.B., M.Y., and F.R. wrote the paper.
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
Freely available online through the PNAS open access option.
Data deposition: Raw glycan array data are available from the Consortium for Functional
Glycomics (accession no. 2621).
See Commentary on page 10475.
1
To whom correspondence should be addressed. E-mail: jeremybarr85@gmail.com.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
1073/pnas.1305923110/-/DCSupplemental.
www.pnas.org/cgi/doi/10.1073/pnas.1305923110 PNAS | June 25, 2013 | vol. 110 | no. 26 | 10771–10776
MICROBIOLOGYSEECOMMENTARY
Results
Phage Adhere to Mucus. Our preliminary investigations of mucosal
surfaces suggested that phage concentrations in the mucus layer
were elevated compared with the surrounding environment.
Here, we used epifluorescence microscopy to count the phage
and bacteria in mucus sampled from a diverse range of mucosal
surfaces (e.g., sea anemones, fish, human gum), and in each
adjacent environment (SI Materials and Methods and Fig. S1).
Comparing the calculated phage-to-bacteria ratios (PBRs)
showed that PBRs in metazoan-associated mucus layers were on
average 4.4-fold higher than those in the respective adjacent
environment (Fig. 1A). The PBRs on these mucus surfaces
ranged from 21:1 to 87:1 (average, 39:1), compared with 3:1 to
20:1 for the surrounding milieus (average, 9:1; n = 9, t = 4.719;
***P = 0.0002). Earlier investigations of phage abundance in
marine environments reported that phage typically outnumber
bacteria by an order of magnitude (50–52), but here we dem-
onstrate that this margin was significantly larger in metazoan-
associated mucus surface layers.
To determine whether this enrichment was dependent on the
presence of mucus rather than some general properties of the
cell surface (e.g., charge), phage adherence was tested with tissue
culture (TC) cells with and without surface mucus (SI Materials
and Methods). In these assays, T4 phage were washed across
confluent cell monolayers for 30 min, after which nonadherent
phage were removed by repeated washings and the adherent
phage quantified by epifluorescence microscopy. Two mucus-
producing cell lines were used: T84 (human colon epithelial
cells) and A549 (human lung epithelial cells). For these cells,
mucin secretion was stimulated by pretreatment with phorbol
12-myristate 13-acetate (53, 54). Comparison of the T84 cells
with the non–mucus-producing Huh-7 human hepatocyte cell
line showed that T4 phage adhered significantly more to the
mucus-producing T84 cells (Fig. 1B; n > 18, t = 8.366; ****P <
0.0001). To demonstrate the mucus dependence of this adher-
ence, the mucus layer was chemically removed from A549 cells
by N-acetyl-L-cysteine (NAC) treatment (55) (Fig. S2). This
significantly reduced the number of adherent phage to levels
similar to those observed with non–mucus-producing cell lines
(Fig. 1B; n > 40, t = 9.561; ****P < 0.0001). We also created an
A549 shRNA mucus knockdown cell line (MUC–
), reducing
mucus production in A549, and a nonsense shRNA control
(shControl; Figs. S3 and S4). Again, T4 phage adhered signifi-
cantly more to the mucus-producing cells (Fig. 1B; n > 37, t =
7.673; ****P < 0.0001).
Although mucin glycoproteins are the predominant component
of mucus, other macromolecular components also are present, any
of which might be involved in the observed phage adherence. We
developed a modified top agar assay to determine whether phage
adhered to a specific macromolecular component of mucus. Plain
agar plates and agar plates coated with 1% (wt/vol) mucin, DNA,
or protein were prepared. That concentration was chosen because
it is at the low end of the range of physiological mucin concen-
trations (56). T4 phage suspensions were incubated on the plates
for 30 min, after which the phage suspension was decanted to
remove unbound particles. The plates then were overlaid with
a top agar containing Escherichia coli hosts and incubated over-
night. The number of adherent phage was calculated from the
number of plaque-forming units (pfu) observed. Significantly more
T4 phage adhered to the 1% mucin-coated agar surface (Fig. 1C;
n = 12, t = 5.306; ****P < 0.0001). Combined, these three assays
show that phage adhere to mucin glycoproteins.
Phage Adherence and Bacterial Infection. The mucus layer is an
optimal environment for microbial growth, providing structure as
well as nutrients in the form of diverse, mucin-associated glycans.
To limit this growth, the metazoan host retards microbial coloni-
zation by diverse antimicrobial mechanisms (24–27). Does the
increased number of adherent phage found on mucosal surfaces
also reduce microbial colonization? To answer this, bacterial at-
tachment to mucus-producing and non–mucus-producing TC cells
was assayed both with and without pretreatment of the cells with
the mucus-adherent phage T4. Here confluent monolayers of TC
cells were overlaid with T4 phage for 30 min, washed to remove
nonadherent phage, and then incubated with E. coli for 4 h. Cells
then were scraped from the plates and the attached bacteria were
fluorescently stained and counted by epifluorescence microscopy.
Phage pretreatment of mucus-producing TC cell lines (T84, A549)
significantly decreased subsequent bacterial attachment (Fig. 2A;
T84: n > 30, t = 32.05, ****P < 0.0001; A549: n > 30, t = 36.85,
****P < 0.0001). Phage pretreatment of non–mucus-producing
cells (Huh-7; MUC–
, an A549 mucus knockdown strain) had
a much smaller effect on bacterial attachment.
To determine whether this reduced bacterial attachment de-
pended on bacterial lysis and the production of progeny viruses,
we repeated these experiments using an amber mutant T4 phage
(T4 am43–
44–
). When infecting wild-type E. coli, this phage pro-
duces no infective progeny virions, but infection of the E. coli
amber suppressor strain SupD yields infective virions. For these
experiments, mucus-producing A549 cells were pretreated with
amber mutant T4 am43–
44–
phage and then incubated with either
wild-type or the amber suppressor strain E. coli (Fig. 2B). Bacte-
rial attachment was reduced by more than four orders of magni-
tude when phage could replicate and thereby increase the number
of infective virions within the mucus (n = 8, ****P < 0.0001).
Comparatively, when no phage replication occurred in the mucus,
there was no observable change in bacterial colonization and
fewer phages were detected (n = 8, *P = 0.0227). These results
show that pretreatment of a mucosal surface with phage reduces
adherence of a bacterial pathogen and that this protection is
mediated by continued phage replication in the mucus.
To test whether the observed reduced bacterial adherence was
accompanied by reduced pathology of the underlying TC cells,
mucus-producing A549 and non–mucus-producing MUC–
TC cells
were exposed to E. coli overnight, either with or without a 30-min
pretreatment with T4 phage. Infection was quantified as the per-
centage of cell death. Adherence of phage effectively protected
the mucus-producing cells against the subsequent bacterial chal-
lenge (Fig. 2C; n = 12, ****P < 0.0001). Phage pretreatment
showed a reduced protection to the non–mucus-producing MUC–
Fig. 1. Phage adhere to cell-associated mucus layers and mucin glycopro-
tein. (A) PBR for diverse mucosal surfaces and the adjacent environment. On
average, PBRs for mucosal surfaces were 4.4-fold greater than for the ad-
jacent environment (n = 9, t = 4.719, ***P = 0.0002, unpaired t test). (B)
Phage adherence to TC cell monolayers, with and without surface mucus
(unpaired t tests). (Left) Non–mucus-producing Huh-7 liver hepatocyte cells
and mucus-producing T84 colon epithelial cells (n > 18, t = 8.366, ****P <
0.0001). (Center) Mucus-producing A549 lung epithelial cells with and
without treatment with NAC, a mucolytic agent (n > 40, t = 9.561, ****P <
0.0001). (Right) Mucus-producing shRNA control A549 cells (shControl) and
mucus knockdown (MUC–
) A549 cells (n > 37, t = 7.673, ****P < 0.0001). (C)
Phage adherence to uncoated agar plates and agar coated with mucin, DNA,
or protein (n = 12, t = 5.306, ****P < 0.0001, unpaired t test).
10772 | www.pnas.org/cgi/doi/10.1073/pnas.1305923110 Barr et al.
cells, decreasing cell death only twofold. Evaluating the impor-
tance of mucus production for effective protection, we found that
phage pretreatment of mucus-producing A549 cells resulted in a
3.6-fold greater reduction in cell death (n = 12, *P = 0.0181) than
the same pretreatment of the mucin knockdown MUC–
cells.
Role of Capsid Ig-Like Domains in Phage Adherence. Minot et al. (44)
recently reported that phage communities associated with the
human gut encode a diverse array of hypervariable proteins,
including some with hypervariable Ig-like domains. Four Ig-like
domains are found in highly antigenic outer capsid protein
(Hoc), a T4 phage structural protein of which 155 copies are
displayed on the capsid surface (57, 58). Based on this, and given
that most Ig-like domains function in recognition and adhesion
processes, we hypothesized that the T4 Hoc protein might me-
diate the adherence of T4 phage to mucus. To test this, we
performed three experiments. First, we compared the adherence
of hoc+
T4 phage and a hoc–
mutant to mucin-, DNA-, and
protein-coated agar plates to an uncoated agar control using the
modified top agar assay (see above). Relative to plain agar, the
adherence of hoc+
T4 phage increased 4.1-fold for mucin-coated
agar (n > 11, t = 3.977, ***P = 0.0007), whereas adherence in-
creased only slightly for agar coated with DNA (1.1-fold) or
protein (1.2-fold; Fig. 3A). Unlike the hoc+
T4 phage, the hoc–
phage did not adhere preferentially to the mucin-coated agar,
but instead showed 1.2-, 1.2-, and 1.1-fold increased adherence
for mucin, DNA, and protein coatings, respectively. To ensure
that none of the macromolecules directly affected phage in-
fectivity, hoc+
and hoc–
T4 phage were incubated in 1% (wt/vol)
solutions of mucin, DNA, or protein. Phage suspensions were
combined with E. coli top agar as described above and layered
over uncoated agar plates. The results confirmed that the mac-
romolecules did not alter phage infectivity (Fig. S5). To provide
further evidence that the mucin adherence was dependent on the
capsid displayed Ig-like domains rather than some other prop-
erty of T4 phage, we repeated the modified top agar assay using
Ig+
and Ig–
T3 phage. As with T4, the Ig-like domains of T3 are
displayed on the surface of the major capsid protein (49). Results
indicated a similar increase in adherence to mucin for the Ig+
, but
not the Ig–
, T3 phage (Fig. S6). Thus, adherence of these phage to
mucus requires the Ig-like protein domains.
Second, a competition assay using hoc+
and hoc−
T4 phage
and mucus-producing TC cells was performed to demonstrate
the role of mucin in phage adherence. Phage suspended in mucin
solutions ranging from 0% to 5% (wt/vol) were washed over
confluent layers of mucus-producing A549 TC cells; phage ad-
herence then was assayed as above. Adherence of hoc+
T4
phage, but not of hoc–
T4 phage, was reduced by mucin com-
petition in a concentration-dependent manner (Fig. 3B).
Third, interaction of the Hoc protein domains displayed on
the capsid surface with mucin glycoproteins was hypothesized
to affect the rate of diffusion of T4 virions in mucus. To eval-
uate this, we used multiple-particle tracking (MPT) to quantify
transport rates of phage particles in buffer and in mucin suspen-
sions. The ensemble average effective diffusivity (Deff) calculated
at a time scale of 1 s for both hoc+
and hoc–
T4 phage in buffer was
compared against that in 1% (wt/vol) mucin suspensions (SI
Materials and Methods). Both hoc+
and hoc–
phage diffuse rapidly
through buffer (Fig. 3C). Whereas hoc–
phage diffused in 1%
mucin at the same rate as in buffer, the mucin decreased the
diffusion rate for hoc+
phage particles eightfold. Thus, all three
of these experimental approaches supported our hypothesis that
Fig. 2. Effect of phage adsorption on subsequent bacterial infection of
epithelial cells. (A) Bacterial attachment to mucus-producing (T84 and A549)
and non–mucus-producing (Huh-7, MUC–
) TC cells, with and without phage
pretreatment. T4 phage pretreatment significantly decreased subsequent
bacterial adherence to mucus-producing TC cell lines (T84: n > 30, t = 32.05,
****P < 0.0001; A549: n > 30, t = 36.85, ****P < 0.0001; unpaired t tests).
Less dramatic shifts were seen for non–mucus-producing cells (Huh-7: n > 30,
t = 2.72, **P = 0.0098; MUC–
: n > 30, t = 3.52, ***P = 0.0007; unpaired t
tests). (B) Mucus-producing A549 cells were pretreated with T4 am43–
44–
phage (Materials and Methods) and then incubated for 4 h with either wild-
type (wt) or amber-suppressor (SupD) E. coli. Phage replication in the SupD E.
coli strain significantly reduced bacterial colony-forming units (CFU) in the
mucus (n = 8, ****P < 0.0001, Tukey’s two-way ANOVA) and increased
phage-forming units (PFU) relative to the no-phage replication wt E. coli
(n = 8, *P = 0.0227). (C) Mortality of mucus-producing (A549) and mucus
knockdown (MUC–
) A549 lung epithelial cells following overnight in-
cubation with E. coli. Phage pretreatment completely protected mucus-
producing A549 cells from bacterial challenge (n = 12, ****P < 0.0001,
Tukey’s one-way ANOVA); protection of MUC–
cells was 3.1-fold less (n = 12,
*P = 0.0181). ns, not significant.
Fig. 3. Effect of Hoc protein on phage–mucin interactions. (A) Adherence
of hoc+
and hoc–
T4 phage to agar coated with mucin, DNA, or protein
reported as an increase relative to plain agar controls (n > 11, t = 3.977,
***P = 0.0007, unpaired t test). (B) Competitive effect of mucin on phage
adherence when hoc+
and hoc–
T4 phage in 0–5% (wt/vol) mucin solution
(1× PBS) were washed over mucus-producing A549 cells (n = 25 per sample).
(C) Diffusion of fluorescence-labeled hoc+
(Left) and hoc–
(Right) T4 phage in
buffer and 1% mucin as determined by MPT. Mucin hindered diffusion of
hoc+
T4 phage but not hoc–
phage (10 analyses per sample, trajectories of
n > 100 particles for each analysis; error bars represent SE).
Barr et al. PNAS | June 25, 2013 | vol. 110 | no. 26 | 10773
MICROBIOLOGYSEECOMMENTARY
the Hoc proteins displayed on the T4 phage capsid interact
with mucin.
Phage Capsid Ig-Like Domains Interact with Glycans. It is known that
∼25% of sequenced tailed dsDNA phages (Caudovirales) en-
code structural proteins with predicted Ig-like domains (48).
A search of publicly available viral metagenomes for homologs
of the Ig-like domains of the T4 Hoc protein yielded numerous
viral Ig-like domains from a variety of environments (Fig. 4A).
These domains were more likely to be found in samples collected
directly from mucus (e.g., sputum samples) or from an environ-
ment adjacent to a mucosal surface (e.g., intestinal lumen, oral
cavity). All homologs displayed high structural homology (Phyre2
confidence score average, 96 ± 5%) with a plant-sugar binding
domain known for its promiscuous carbohydrate binding speci-
ficity (SI Materials and Methods and Table S1), suggesting an in-
teraction between these Ig-like domains and glycans.
Mucins are complex glycoproteins with highly variable glycan
groups exposed to the environment. To investigate whether
Hoc interacts with glycans and, if so, to determine whether it
interacts with a specific glycan or with a diverse array of glycans,
we assayed phage adherence to microarrays printed with 610
mammalian glycans. The hoc+
T4 phage adhered to many di-
verse glycans and showed a preference for the O-linked glycan
residues typically found in mucin glycoproteins (Fig. 4B, SI
Materials and Methods, and Table S2). The hoc–
T4 phage
exhibited significantly lower affinity for all tested glycans. This
indicates that Hoc mediates interactions between T4 phage and
varied glycan residues.
Discussion
In diverse metazoans, body surfaces that interact with the envi-
ronment are covered by a protective layer of mucus. Because
these mucus layers provide favorable habitats for bacteria, they
serve as the point of entry for many pathogens and support large
populations of microbial symbionts. Also present are diverse
phages that prey on specific bacterial hosts. Moreover, phage
concentrations in mucus are elevated relative to the surrounding
environment (an average 4.4-fold increase for a diverse sample
of invertebrate and vertebrate metazoans; Fig. 1A). The in-
creased concentration of lytic phage on mucosal surfaces pro-
vides a previously unrecognized metazoan immune defense
affected by phage lysis of incoming bacteria.
Working with a model system using T4 phage and various TC
cell lines, we demonstrated that the increased concentration of
phage on mucosal surfaces is mediated by weak binding inter-
actions between the variable Ig-like domains on the T4 phage
capsid and mucin-displayed glycans. The Ig protein fold is well
known for its varied but essential roles in the vertebrate immune
response and cell adhesion. Ig-like domains also are present in
approximately one quarter of the sequenced genomes of tailed
DNA phages, the Caudovirales (48). Notably, these domains
were found only in virion structural proteins and typically are
displayed on the virion surface. Thus, they were postulated to
bind to bacterial surface carbohydrates during infection (48, 49).
However, mucin glycoproteins, the predominant macromolecu-
lar constituent of mucus, display hundreds of variable glycan
chains to the environment that offer potential sites for binding by
phage Ig-like proteins. Furthermore, we speculate that phage use
the variability of the Ig-like protein scaffold (supporting >1013
potential alternatives) to adapt to the host’s ever-changing pat-
terns of mucin glycosylation.
The presence of an Ig-like protein (Hoc) displayed on the
capsid of T4 phage significantly slowed the diffusion of the phage
in mucin solutions. In vivo, similar phage binding to mucin gly-
cans would increase phage residence time in mucus layers. Be-
cause bacterial concentrations typically are enriched in mucus
(Fig. S1), we predict that mucus-adherent phage are more likely
to encounter bacteria, potentially increasing their replicative
success. If so, phage Ig-like domains that bind effectively to the
mucus layer would be under positive selection. Likely, Hoc and
other phage proteins with Ig-like domains interact with other
glycans with different ramifications, as well (49, 58).
Previous metagenomic studies documented the ubiquity and
diversity of bacteria and phage within mucus-associated envi-
ronments (e.g., human gut, human respiratory tract, corals) (52,
59–64). Known also were some of the essential but adaptable
services provided by symbiotic bacteria in these environments
(65). However, only recently have efforts been made to in-
vestigate the dynamic influences of phage within host-associated
ecosystems (37, 44, 66). In this work, we used an in vitro model
system to demonstrate a mechanism of phage adherence to the
mucus layers that shield metazoan cells from the environment.
Furthermore, adherent phage protected the underlying epithelial
cells from bacterial infection. Based on these observations and
previous research, we proposed the BAM model of immunity, in
which the adherence of phage to mucosal surfaces yields a non–
host-derived, antimicrobial defense. According to this model
(summarized in Fig. 5), the mucus layer, already considered part
of the innate immune system and known to provide physical
and biochemical antimicrobial defenses (18, 27, 67), also
accumulates phage.
The model system we used involved a single lytic phage and
host bacterium; the situation in vivo undoubtedly is more com-
plex. Within the mucosal layer reside diverse bacterial lineages
and predictably an even greater diversity of phage strains, both
enmeshed within complex phage–bacterial infection networks
and engaged in a dynamic arms race (68, 69). These and other
factors lower the probability that any given phage–bacterium
encounter will result in a successful infection. The time di-
mension adds further complexity. The mucus layer is dynamic.
Mucins are secreted continually by the underlying epithelium
while mucus is sloughed continually from the outer surface. As
a result, there is an ongoing turnover of both the bacterial and
phage populations in the mucus layer. Driven by kill-the-winner
dynamics, the population of phage types that can infect the
dominant bacterial types present will cycle along with the pop-
ulations of their hosts. Through such mechanisms, we envision
that adherent lytic phages provide a dynamic and adaptable
defense for their metazoan hosts—a unique example of a meta-
zoan–phage symbiosis.
We posit that BAM immunity reduces bacterial pathogenesis
and provides a previously unrecognized, mucosal immunity. This
has far-reaching implications for numerous fields, such as human
immunity, gastroenterology, coral disease, and phage therapy.
Meanwhile, key questions remain. For instance, what role do
temperate phages play in the dynamics of BAM immunity?
When integrated into the bacterial chromosome as prophages,
Fig. 4. Hoc-mediated glycan binding and Hoc-related phylogeny. (A) Phy-
logenetic tree of sequences from viral metagenomes with high-sequence
homology to Ig-like domains. Many of the identified homologs are from
mucus-associated environments (e.g., human feces, sputum). Also included
are the Hoc protein of T4 phage and the hypervariable Ig-like domains
previously obtained by deep sequencing of phage DNA from the human gut
(44). The scale bar represents an estimated 0.5 amino acid substitutions per
site. See SI Materials and Methods for methods. (B) Binding of fluorescence-
stained hoc+
and hoc–
T4 phage to a microarray of 610 mammalian glycans.
Normalized relative fluorescence units (RFU) were calculated from mean
fluorescence minus background binding.
10774 | www.pnas.org/cgi/doi/10.1073/pnas.1305923110 Barr et al.
they protect their bacterial hosts from infection by related phages;
as free phages, they infect and kill sensitive related bacterial
strains that compete with their bacterial hosts (37–39). Both
mechanisms may benefit their metazoan host by contributing to
the maintenance of a selected commensal mucosal microbiota.
These possibilities remain to be investigated. Likewise, in vivo
investigations are needed to characterize the bacterial and phage
diversity present and the consequent effects on BAM immunity.
As of now, the relationships shown here open an arena for im-
munological study, introduce a phage–metazoan symbiosis, and
recognize the key role of the world’s most abundant biological
entities in the metazoan immune system.
Materials and Methods
Bacterial Strains, Phage Stocks, TC Cell Lines, and Growth Conditions. E. coli
1024 strain was used for all E. coli experiments and was grown in LB (10 g
tryptone, 5 g yeast extract, 10 g NaCl, in 1 L dH2O) at 37 °C overnight. E. coli
amber-suppressor strain SupD strain CR63 was used as a host for amber
mutant phage and grown as above. Bacteriophage T4 was used at ∼109
pfu·mL–1
. Hoc–
T4 phage were kindly supplied by Prof. Venigalla Rao (58), The
Catholic University of America, Washington, D.C. T3 am10 Ig–
amber mutant
phage were kindly supplied by Prof. Ian J. Molineux (70), University of Texas,
Austin, TX. T4 replication-negative 43–
(amE4332: DNA polymerase) 44–
(amN82: subunit of polymerase clamp holder) amber mutant phage were
kindly supplied by Prof. Kenneth Kreuzer (71), Duke University School of
Medicine, Durham, NC. The human tumorigenic colon epithelial cell line, T84,
was obtained from the American Type Culture Collection (ATCC) and cultured
in DMEM/F12-K media with 5% FBS and 100 μg·mL–1
penicillin–streptomycin
(PS). The human tumorigenic lung epithelial cell line A549 was kindly sup-
plied by Prof. Kelly Doran, San Diego State University, San Diego, CA and
cultured in F12-K media with 10% FBS, 100 μg·mL–1
PS. The human tumor-
igenic liver epithelial cell line Huh7 was kindly supplied by Prof. Roland
Wolkowicz, San Diego State University, San Diego, CA, and cultured in F12-K
media with 10% FBS, 100 μg·mL–1
PS. All TC cell lines initially were grown in
50 mL Primaria Tissue Culture Flasks (Becton Dickinson) at 37 °C and 5% CO2.
Phage Adherence to Mucus-Associated Macromolecules. LB agar plates were
coated with 1 mL of 1% (wt/vol) of one of the following in 1× PBS: type III
porcine stomach mucin, DNA from salmon testes, or BSA (all three from
Sigma–Aldrich) and then allowed to dry. Stocks of hoc+
and hoc–
T4 phage
(109
pfu·mL−1
) were serially diluted to 1 × 10−7
and 1 × 10−8
per milliliter in
LB, and a 5-mL aliquot of each dilution was washed across the plates for 30
min at 37 °C on an orbital shaker. After the phage suspensions were dec-
anted from the plates, the plates were shaken twice to remove excess liquid
and dried. Each plate then was layered with 1 mL of overnight E. coli culture
(109
mL–1
) in 3 mL of molten top agar and incubated overnight at 37 °C. The
number of adherent phage was calculated from the number of plaque-
forming units observed multiplied by the initial phage dilution. To de-
termine whether mucus macromolecules directly affected phage infectivity,
hoc+
and hoc–
T4 phage (109
pfu·mL−1
) were serially diluted as described
above into 1 mL LB solutions containing 1% (wt/vol) mucin, DNA, or BSA.
After incubation for 30 min at 37 °C, the phage suspensions were combined
with E. coli top agar as described above and layered over uncoated agar
plates (Fig. S5).
Phage Treatment of TC Cells. TC cells were washed twice with 5 mL of serum-
free media to remove residual antibiotics, layered with 2 mL of serum-free
media containing T4 phage (107
or 109
mL–1
), and incubated at 37 °C and 5%
CO2 for 30 min. Cells then were washed five times with 5 mL of serum-free
media to remove nonadherent phage.
Phage Adherence to TC Cells. TC cells were treated with phage (109
mL–1
; see
above), then scraped from plates using Corning Cell Scrapers (Sigma–
Aldrich). Adherent phage were counted by epifluorescence microscopy as
described above.
Bacterial Adherence to TC Cells With/Without Phage Pretreatment. TC cells with
or without pretreatment with T4 phage (107
mL–1
) were layered with 2 mL
serum-free media containing E. coli (107
mL–1
), incubated at 37 °C and 5% CO2
for 4 h, and then washed five times with 5 mL serum-free medium to remove
nonadherent phage and bacteria. Cells were scraped from plates, and ad-
herent phage and bacteria were counted by either epifluorescence micros-
copy, as described above, or colony-forming and plaque-forming units. Then,
100 μL of a relevant dilution was spread onto an agar plate and incubated
overnight at 37 °C, and the number of adherent bacteria was calculated from
the colony-forming units observed multiplied by the initial dilution. Plaque-
forming units were counted by a top agar assay as described above.
TC Cell Death from Bacterial Infection. Mucus-producing A549 and MUC–
A549
TC cells were grown to confluence. T4 phage were cleaned using Amicon
50-kDa centrifugal filters (Millipore) and saline magnesium buffer (SM) to
remove bacterial lysis products. Cells, with or without T4 phage pre-
treatment (107
mL–1
), were incubated with E. coli (107
mL–1
) overnight. Af-
terward, TC cells were recovered from the plates by trypsin/EDTA solution
(Invitrogen). Cells were pelleted by centrifugation and resuspended in 1×
PBS. Dead cells were identified by staining with 1 mg/mL of propidium io-
dide (Invitrogen). Samples then were analyzed on a FACSCanto II flow
cytometer (BD Biosciences) with excitation at 488 nm and emission detected
through a 670 long pass filter. The forward scatter threshold was set at
5,000, and a total of 10,000 events were collected for each sample.
Mucin Competition Assay. Mucus-producing A549 TC cells were grown to
confluence. Hoc+
and hoc–
T4 phage (109
mL–1
) were diluted into mucin
solutions ranging between 0% and 5% (wt/vol) in 1× PBS then washed over
TC cells for 30 min at 37 °C and 5% CO2. Cells were washed five times with
5 mL serum-free media to remove nonadherent phage, scraped from plates,
and adherent phage were quantified as described above.
Graphing and Statistics. Graphing and statistical analyses were performed
using GraphPad Prism 6 (GraphPad Software). All error bars represent 5–95%
confidence intervals. The midline represents the median and the mean for
box plots and bar plots, respectively.
Fig. 5. The BAM model. (1) Mucus is produced and se-
creted by the underlying epithelium. (2) Phage bind vari-
able glycan residues displayed on mucin glycoproteins via
variable capsid proteins (e.g., Ig-like domains). (3) Phage
adherence creates an antimicrobial layer that reduces bac-
terial attachment to and colonization of the mucus, which
in turn lessens epithelial cell death. (4) Mucus-adherent
phage are more likely to encounter bacterial hosts, thus are
under positive selection for capsid proteins that enable
them to remain in the mucus layer. (5) Continual sloughing of
the outer mucus provides a dynamic mucosal environment.
Barr et al. PNAS | June 25, 2013 | vol. 110 | no. 26 | 10775
MICROBIOLOGYSEECOMMENTARY
ACKNOWLEDGMENTS. This work was supported by National Institutes of
Health (NIH) Grants R01: GM095384, GM073898, and R21: AI094534 from
the National Institute of General Medical Sciences. The authors thank the
Protein-Glycan Interaction Resource at Emory University School of Medicine,
Atlanta, GA (funded by NIH Grant GM98791), for support of the glycan
microarray analyses. The authors acknowledge the San Diego State Univer-
sity (SDSU) Flow Cytometry Core Facility and the SDSU Electron Microscopy
Facility for assistance with sample analysis.
1. Bäckhed F, Ley RE, Sonnenburg JL, Peterson DA, Gordon JI (2005) Host-bacterial
mutualism in the human intestine. Science 307(5717):1915–1920.
2. Dethlefsen L, McFall-Ngai M, Relman DA (2007) An ecological and evolutionary
perspective on human-microbe mutualism and disease. Nature 449(7164):811–818.
3. Clay K, Holah J (1999) Fungal endophyte symbiosis and plant diversity in successional
fields. Science 285(5434):1742–1745.
4. Douglas AE (1989) Mycetocyte symbiosis in insects. Biol Rev Camb Philos Soc 64(4):
409–434.
5. Hooper LV, Midtvedt T, Gordon JI (2002) How host-microbial interactions shape the
nutrient environment of the mammalian intestine. Annu Rev Nutr 22(1):283–307.
6. Hosokawa T, Koga R, Kikuchi Y, Meng X-Y, Fukatsu T (2010) Wolbachia as a bacter-
iocyte-associated nutritional mutualist. Proc Natl Acad Sci USA 107(2):769–774.
7. Nyholm SV, McFall-Ngai MJ (2004) The winnowing: Establishing the squid-vibrio
symbiosis. Nat Rev Microbiol 2(8):632–642.
8. Ruby EG (1996) Lessons from a cooperative, bacterial-animal association: The Vibrio
fischeri-Euprymna scolopes light organ symbiosis. Annu Rev Microbiol 50(1):591–624.
9. Currie CR, Scott JA, Summerbell RC, Malloch D (1999) Fungus-growing ants use an-
tibiotic-producing bacteria to control garden parasites. Nature 398(6729):701–704.
10. Kaltenpoth M, Göttler W, Herzner G, Strohm E (2005) Symbiotic bacteria protect wasp
larvae from fungal infestation. Curr Biol 15(5):475–479.
11. Martens EC, Chiang HC, Gordon JI (2008) Mucosal glycan foraging enhances fitness
and transmission of a saccharolytic human gut bacterial symbiont. Cell Host Microbe
4(5):447–457.
12. Sonnenburg JL, et al. (2005) Glycan foraging in vivo by an intestine-adapted bacterial
symbiont. Science 307(5717):1955–1959.
13. Berry D, et al. (2013) Host-compound foraging by intestinal microbiota revealed by
single-cell stable isotope probing. Proc Natl Acad Sci USA 110(12):4720–4725.
14. Sachs JL, Skophammer RG, Regus JU (2011) Evolutionary transitions in bacterial
symbiosis. Proc Natl Acad Sci USA 108(Suppl 2):10800–10807.
15. Stouthamer R, Breeuwer JA, Hurst GD (1999) Wolbachia pipientis: Microbial manip-
ulator of arthropod reproduction. Annu Rev Microbiol 53(1):71–102.
16. Chow J, Lee SM, Shen Y, Khosravi A, Mazmanian SK (2010) Host-bacterial symbiosis in
health and disease. Adv Immunol 107:243–274.
17. Cone RA (2009) Barrier properties of mucus. Adv Drug Deliv Rev 61(2):75–85.
18. Linden SK, Sutton P, Karlsson NG, Korolik V, McGuckin MA (2008) Mucins in the
mucosal barrier to infection. Mucosal Immunol 1(3):183–197.
19. Clunes MT, Boucher RC (2007) Cystic fibrosis: The mechanisms of pathogenesis of an
inherited lung disorder. Drug Discov Today Dis Mech 4(2):63–72.
20. Strugala V, Allen A, Dettmar PW, Pearson JP (2003) Colonic mucin: Methods of
measuring mucus thickness. Proc Nutr Soc 62(1):237–243.
21. Garren M, Azam F (2012) Corals shed bacteria as a potential mechanism of resilience
to organic matter enrichment. ISME J 6(6):1159–1165.
22. Button B, et al. (2012) A periciliary brush promotes the lung health by separating the
mucus layer from airway epithelia. Science 337(6097):937–941.
23. Poulsen LK, et al. (1994) Spatial distribution of Escherichia coli in the mouse large
intestine inferred from rRNA in situ hybridization. Infect Immun 62(11):5191–5194.
24. Phalipon A, et al. (2002) Secretory component: A new role in secretory IgA-mediated
immune exclusion in vivo. Immunity 17(1):107–115.
25. Raj PA, Dentino AR (2002) Current status of defensins and their role in innate and
adaptive immunity. FEMS Microbiol Lett 206(1):9–18.
26. Vaishnava S, et al. (2011) The antibacterial lectin RegIII{gamma} promotes the spatial
segregation of microbiota and host in the intestine. Sci STKE 334(6053):255.
27. Schluter J, Foster KR (2012) The evolution of mutualism in gut microbiota via host
epithelial selection. PLoS Biol 10(11):e1001424.
28. Hooper LV, Xu J, Falk PG, Midtvedt T, Gordon JI (1999) A molecular sensor that allows
a gut commensal to control its nutrient foundation in a competitive ecosystem. Proc
Natl Acad Sci USA 96(17):9833–9838.
29. Gerken TA (2004) Kinetic modeling confirms the biosynthesis of mucin core 1 (β-Gal(1-3)
α-GalNAc-O-Ser/Thr) O-glycan structures are modulated by neighboring glycosylation
effects. Biochemistry 43(14):4137–4142.
30. Jentoft N (1990) Why are proteins O-glycosylated? Trends Biochem Sci 15(8):291–294.
31. Schulz BL, et al. (2007) Glycosylation of sputum mucins is altered in cystic fibrosis
patients. Glycobiology 17(7):698–712.
32. Rodriguez-Brito B, et al. (2010) Viral and microbial community dynamics in four
aquatic environments. ISME J 4(6):739–751.
33. Thingstad T, Lignell R (1997) Theoretical models for the control of bacterial growth
rate, abundance, diversity and carbon demand. Aquat Microb Ecol 13:19–27.
34. Groisman EA, Ochman H (1993) Cognate gene clusters govern invasion of host epi-
thelial cells by Salmonella typhimurium and Shigella flexneri. EMBO J 12(10):
3779–3787.
35. Johansen BK, Wasteson Y, Granum PE, Brynestad S (2001) Mosaic structure of Shiga-
toxin-2-encoding phages isolated from Escherichia coli O157:H7 indicates frequent
gene exchange between lambdoid phage genomes. Microbiology 147(Pt 7):
1929–1936.
36. Willner D, et al. (2011) Metagenomic detection of phage-encoded platelet-binding
factors in the human oral cavity. Proc Natl Acad Sci USA 108(Suppl 1):4547–4553.
37. Duerkop BA, Clements CV, Rollins D, Rodrigues JLM, Hooper LV (2012) A composite
bacteriophage alters colonization by an intestinal commensal bacterium. Proc Natl
Acad Sci USA 109(43):17621–17626.
38. Furuse K, et al. (1983) Bacteriophage distribution in human faeces: Continuous survey
of healthy subjects and patients with internal and leukaemic diseases. J Gen Virol
64(Pt 9):2039–2043.
39. Weinbauer MG (2004) Ecology of prokaryotic viruses. FEMS Microbiol Rev 28(2):
127–181.
40. Clokie MR, Millard AD, Letarov AV, Heaphy S (2011) Phages in nature. Bacteriophage
1(1):31–45.
41. Moran NA, Degnan PH, Santos SR, Dunbar HE, Ochman H (2005) The players in
a mutualistic symbiosis: Insects, bacteria, viruses, and virulence genes. Proc Natl Acad
Sci USA 102(47):16919–16926.
42. Oliver KM, Degnan PH, Hunter MS, Moran NA (2009) Bacteriophages encode factors
required for protection in a symbiotic mutualism. Science 325(5943):992–994.
43. Roossinck MJ (2011) The good viruses: Viral mutualistic symbioses. Nat Rev Microbiol
9(2):99–108.
44. Minot S, Grunberg S, Wu GD, Lewis JD, Bushman FD (2012) Hypervariable loci in the
human gut virome. Proc Natl Acad Sci USA 109(10):3962–3966.
45. McMahon SA, et al. (2005) The C-type lectin fold as an evolutionary solution for
massive sequence variation. Nat Struct Mol Biol 12(10):886–892.
46. Medhekar B, Miller JF (2007) Diversity-generating retroelements. Curr Opin Microbiol
10(4):388–395.
47. Halaby DM, Mornon JPE (1998) The immunoglobulin superfamily: an insight on its
tissular, species, and functional diversity. J Mol Evol 46(4):389–400.
48. Fraser JS, Yu Z, Maxwell KL, Davidson AR (2006) Ig-like domains on bacteriophages: A
tale of promiscuity and deceit. J Mol Biol 359(2):496–507.
49. Fraser JS, Maxwell KL, Davidson AR (2007) Immunoglobulin-like domains on bacte-
riophage: Weapons of modest damage? Curr Opin Microbiol 10(4):382–387.
50. Fuhrman JA (1999) Marine viruses and their biogeochemical and ecological effects.
Nature 399(6736):541–548.
51. Danovaro R, Serresi M (2000) Viral density and virus-to-bacterium ratio in deep-sea
sediments of the Eastern Mediterranean. Appl Environ Microbiol 66(5):1857–1861.
52. Breitbart M, et al. (2002) Genomic analysis of uncultured marine viral communities.
Proc Natl Acad Sci USA 99(22):14250–14255.
53. Hong DH, Petrovics G, Anderson WB, Forstner J, Forstner G (1999) Induction of mucin
gene expression in human colonic cell lines by PMA is dependent on PKC-e. Am J
Physiol 277(5 Pt 1):G1041–G1047.
54. Forstner G, Zhang Y, McCool D, Forstner J (1993) Mucin secretion by T84 cells: Stim-
ulation by PKC, Ca2+, and a protein kinase activated by Ca2+ ionophore. Am J Physiol
264(6 Pt 1):G1096–G1102.
55. Alemka A, et al. (2010) Probiotic colonization of the adherent mucus layer of
HT29MTXE12 cells attenuates Campylobacter jejuni virulence properties. Infect Im-
mun 78(6):2812–2822.
56. Lieleg O, Vladescu I, Ribbeck K (2010) Characterization of particle translocation
through mucin hydrogels. Biophys J 98(9):1782–1789.
57. Sathaliyawala T, et al. (2010) Functional analysis of the highly antigenic outer capsid
protein, Hoc, a virus decoration protein from T4-like bacteriophages. Mol Microbiol
77(2):444–455.
58. Fokine A, et al. (2011) Structure of the three N-terminal immunoglobulin domains of
the highly immunogenic outer capsid protein from a T4-like bacteriophage. J Virol
85(16):8141–8148.
59. Reyes A, et al. (2010) Viruses in the faecal microbiota of monozygotic twins and their
mothers. Nature 466(7304):334–338.
60. Marhaver KL, Edwards RA, Rohwer F (2008) Viral communities associated with healthy
and bleaching corals. Environ Microbiol 10(9):2277–2286.
61. Willner D, et al. (2009) Metagenomic analysis of respiratory tract DNA viral commu-
nities in cystic fibrosis and non-cystic fibrosis individuals. PLoS ONE 4(10):e7370.
62. Wegley L, Edwards R, Rodriguez-Brito B, Liu H, Rohwer F (2007) Metagenomic analysis
of the microbial community associated with the coral Porites astreoides. Environ
Microbiol 9(11):2707–2719.
63. Willner D, et al. (2012) Case studies of the spatial heterogeneity of DNA viruses in the
cystic fibrosis lung. Am J Respir Cell Mol Biol 46(2):127–131.
64. Eckburg PB, et al. (2005) Diversity of the human intestinal microbial flora. Science
308(5728):1635–1638.
65. McFall-Ngai M, et al. (2013) Animals in a bacterial world, a new imperative for the life
sciences. Proc Natl Acad Sci USA 110(9):3229–3236.
66. Minot S, et al. (2011) The human gut virome: Inter-individual variation and dynamic
response to diet. Genome Res 21(10):1616–1625.
67. Lieleg O, Lieleg C, Bloom J, Buck CB, Ribbeck K (2012) Mucin biopolymers as broad-
spectrum antiviral agents. Biomacromolecules 13(6):1724–1732.
68. Labrie SJ, Samson JE, Moineau S (2010) Bacteriophage resistance mechanisms. Nat Rev
Microbiol 8(5):317–327.
69. Weitz JS, et al. (2013) Phage-bacteria infection networks. Trends Microbiol 21(2):82–91.
70. Condreay JP, Wright SE, Molineux IJ (1989) Nucleotide sequence and complementa-
tion studies of the gene 10 region of bacteriophage T3. J Mol Biol 207(3):555–561.
71. Benson KH, Kreuzer KN (1992) Plasmid models for bacteriophage T4 DNA replication:
Requirements for fork proteins. J Virol 66(12):6960–6968.
10776 | www.pnas.org/cgi/doi/10.1073/pnas.1305923110 Barr et al.
Supporting Information
Barr et al. 10.1073/pnas.1305923110
SI Materials and Methods
Mucus Sample Collection. Mucus samples were collected directly
from the surface of organisms using a syringe, swab, or custom
suction device. Environmental samples were collected as close to
the mucus sample as possible, typically within 30–50 cm of the
mucosal surface. Specific organism details are as follows:
Sea anemones were sampled from tidal rock pools at Ocean
Beach, San Diego, CA. Surface mucus was collected by a custom
suction device that dislodges surface mucus using a stream of
0.02 μm filtered seawater; the environmentalg sample was sea-
water collected directly above the anemone.
Hard corals were sampled at the Birch Aquarium, San Diego,
CA. Surface mucus was collected by syringe directly from coral
surfaces; environmental water samples were collected directly
above the coral.
The polychaete, along with surrounding water, was collected at
Scripps Pier, San Diego, CA, and carefully transported to the
laboratory in a container. Surface mucus was collected via syringe,
and the environmental sample was seawater from the container.
Teleost surface mucus was sampled at the Birch Aquarium, San
Diego, CA. Surface mucus was collected by custom suction de-
vice; the environmental water sample was collected directly above
the teleost within its tank.
Human gum mucus was sampled from a male subject with no
current pathology/disease. Surface mucus was collected by swab;
the environmental sample was expectorated saliva. Consent was
obtained for all human samples collected under the San Diego
State University Institutional Review Board #2121.
Mouse intestine was excised from a healthy mouse. Surface
mucus was collected by cutting open the intestine, washing the
mucosal surface with 0.02 μm-filtered PBS buffer, then scraping
off the mucus layer; the environmental sample was collected from
the intestinal lumen directly adjacent to the sampled mucosa. All
animal experiments were approved by the Committee on the Use
and Care of Animals (SDSU, APF #10-08-024D) and performed
using accepted veterinary standards.
Bacterial and Phage Counts from Mucus and Environmental Samples.
Samples of mucus and the adjacent environment were collected
directly from nine evolutionarily diverse mucosal surfaces (Fig.
S1). Samples were transported and maintained on ice until pro-
cessed. All samples were fixed overnight in 0.5% glutaraldehyde at
4 °C, then incubated in 6.5 mM DTT at 37 °C for 1 h to assist
mucus degradation. A 1–100-μL aliquot was diluted with 2 mL of
0.02 μm SM buffer (100 mM NaCl, 8 mM MgSO4, 50 mM Tris·Cl,
in dH2O), briefly mixed, then filtered onto a 0.02-μm Anodisc
polycarbonate filter (Whatman). Filters were stained with 10×
SYBR Gold, washed, and visualized on a Zeiss epifluorescence
microscope. For each sample, 20–30 images were taken for both
bacteria and virus-like particles. Images were analyzed using Im-
age-Pro Plus 5.1 software (MediaCybernetics). Counts of bac-
teria and virus-like particles (referred to as “phage” throughout
the text) per milliliter were made as previously described (1).
Tissue Culture Cells and Mucus Reduction. Monolayers of various
mucus-producing and non–mucus-producing tissue culture (TC)
cells were grown to confluence in six-well Multiwell tissue culture
plates (Becton Dickinson). (i) Mucus-producing TC cells were
exposed to 1 μg/mL of a phorbol ester, phorbol-12-myristate
13-acetate (Sigma–Aldrich) in the culture media overnight to
stimulate the mucin secretory response (2). (ii) The mucolytic
agent N-acetyl-L-cysteine (NAC; Sigma–Aldrich) was used to
chemically remove mucus from A549 TC cells (60 mM NAC in
serum-free media for 1 h with agitation) (3). Mucus depletion
was confirmed using periodic acid-Schiff–Alcian blue (PAS/AB)
(Fig. S2). (iii) A mucus-knockdown (MUC–
) A549 cell line was
produced by transduction of A549 cells with GIPZ Lentiviral
Human MUC1 shRNA and TRIPZ Inducible Lentiviral Hu-
man MUC5AC shRNA as target vectors; an shControl A549 cell
line was produced using the GIPZ Nonsilencing Lentiviral
shRNA Control as a control vector (Thermo Scientific).
Knockdown of mucus production in the MUC–
cell line was
confirmed by Western blot analysis and PAS/AB (Sigma–Aldrich;
Figs. S3 and S4).
Transfection and Selection of A549 TC Mucus-Negative Clones. A549
cells were transduced with GIPZ Lentiviral Human MUC1 shRNA
and TRIPZ Inducible Lentiviral Human MUC5AC shRNA as
target vectors or GIPZ Nonsilencing Lentiviral shRNA Control as
a control vector (Thermo Scientific) according to the manufacturer’s
instructions. Viral particles were produced by transfecting HEK
293T cells with a combination of plasmids containing 2 μg of
packaging vector pCMV d8.2 containing the gag-pol proteins of
HIV-1, 3 μg of the transfer vectors containing the LTRs of HIV-1,
3 μg of vesicular stomatitis virus envelope glycoprotein plasmid, and
1.5 μg of pci-HIV-1 viral protein R accessory protein plasmid.
Growth medium was replaced 24 h post transfection, and viral
supernatant was collected 48 and 72 h after transfection and then
filtered through 0.45-μm polytetrafluoroethylene (PTFE) filters
(Pall Corporation). A549 cells were seeded into six-well culture
plates and grown to ∼70% confluence. The cells then were washed
twice in serum-free media before being incubated overnight in
1 mL of growth media and 1 mL of virus-containing media con-
taining 5 μg/mL of polybrene. The transduced cells subsequently
were washed and cultured for 24 h in complete medium with
2 μg/mL of doxycycline to induce expression of shRNA. Cells
then were sorted using a BD FACSAria (BD Biosciences) at the
San Diego State University Flow Cytometry Facility. A 100-μm
nozzle was used at a sheath pressure of 20 psi. Excitation source
was a 488-nm laser and emissions were collected using 530/30
band pass (BP) and 585/42 BP filters for GFP and red fluores-
cent protein, respectively. Between 10,000 and 300,000 cells were
sorted for each population and collected in a 5-mL tube with 250
μL of FBS. The efficiency of MUC1-MUC5AC knockdown was
confirmed by Western blot analysis and PAS/AB (Sigma–Al-
drich), a stain for mucus-like substances.
Western Blot Analysis. Expression of MUC1 (a membrane-tethered
mucin) and MUC5AC (a secreted gel-forming mucin) was ex-
amined by Western blot analysis. MUC–, shControl, and native
A549 cell lines were grown to confluence and then lysed using
radio-immunoprecipitation assay (RIPA) buffer (Thermo Sci-
entific) containing 2 mM Na3VO4, 100 mM NaF, 10 mM sodium
pyrophosphate, 1 mM PMSF, and protease inhibitor mixture
(Millipore). Aliquots containing 50 μg of total protein were
subjected to SDS/PAGE, and the protein bands were transferred
to a polyvinylidene difluoride membrane (Sigma–Aldrich). Mem-
branes were blocked with 5% nonfat milk in Tris-buffered saline
containing 0.05% Tween 20 at room temperature for 1 h and then
incubated overnight at 4 °C with mouse anti-human MUC1
monoclonal antibody (clone S.854.6; Thermo Scientific), mouse
anti-human MUC5AC monoclonal antibody (clone 2H7; Sigma–
Aldrich), and rabbit anti-human GAPDH antibody (Millipore).
After three washes, membranes were incubated for 1 h at room
Barr et al. www.pnas.org/cgi/content/short/1305923110 1 of 6
temperature with anti-mouse or anti-rabbit IgG horseradish
peroxidase-linked, species-specific, whole antibody (Fisher Sci-
entific). Immunoreactivity was visualized and band intensity was
normalized to the constitutively expressed GAPDH protein.
Multiple Particle Tracking. Assays were performed in plastic well
chambers mounted on glass slides that had been coated with poly
(dimethylsiloxane) to prevent phage adherence. Five microliters
of 109
mL–1
SYBR Gold-labeled phage suspensions was added to
50 μL of 1% (wt/vol) mucin solution in 1× PBS buffer. Trajec-
tories of fluorescently labeled phage were observed using a
DeltaVision Spectris Model DV4 deconvolution microscope
(Applied Precision) equipped with a 100× Olympus PlanApo
1.4 lens. Movies were captured using SoftWoRx 5.0.0 (Applied
Precision): 100-ms temporal resolution for 30 s, 10 analyses per
sample, n > 100 particle trajectories per analysis. Trajectories
were analyzed with the ParticleTracker plugin for ImageJ (4).
The coordinates of phage particle centroids were transformed
into time-averaged mean square displacements: <Δr2
(τ)> =
<Δx2
+ Δy2
>, from which effective diffusivities (<Deff>) were
calculated; Deff = <Δr2
(τ)>/(4 τ) (5, 6).
Glycan Microarray. Phage binding to glycans was assayed using
printed mammalian glycan microarrays (version 5.1, Consortium
for Functional Glycomics Core) containing 610 glycan targets.
Samples of highly antigenic outer capsid protein (hoc+
) T4 phage,
hoc–
T4 phage, and buffer controls were applied to separate
glycan microarray slides. Each slide received 35 μL of sample, 35
μL of binding buffer (Tris saline with 2 mM Ca2+
, 2 mM Mg2+
,
1% bovine serum albumin (BSA), and 0.05% Tween 20), and
a coverslip. Slides first were incubated for 1 h at room temper-
ature and washed with binding buffer. Slides then were in-
cubated in SYBR Gold fluorescence dye (diluted 1:10,000
in binding buffer) for 1 h under a coverslip at room tempera-
ture, washed, dried, and immediately scanned in a PerkinElmer
ProScanArray microarray scanner using an excitation wavelength
of 488 nm. ImaGene software (BioDiscovery, Inc) was used
to quantify fluorescence. Normalized relative fluorescence unit
(RFU) values reported are the average (after subtraction of
background buffer fluorescence) from six spots for each glycan
represented on the array.
Phylogenetic Analysis of Ig-Like Domains. The SEED database (www.
theseed.org) collection of Ig-like polycystic kidney disease (PKD)
protein families (Pfam) (PF00801) and the T4 Hoc sequence were
searched against the 124 viral metagenomic datasets contained in the
My Metagenome (MyMg) database (http://edwards.sdsu.edu/cgi-bin/
mymgdb/show.cgi) using tBLASTn (PubMed accession numbers:
16336043, 17620602, 19156205, 19816605, 20547834, 17921274,
18441115, 19892985, 19555373, 20573248, 20631792, 21167942.79,
21193730.87, 21219518.96, 21245307.04, 21271095.12, 21296883.2,
21322671.29, 21348459.37, 21374247.45, 21400035.53; MG-Rast
IDs: 21167942.79, 21193730.87, 21219518.96, 21245307.04,
21271095.12, 21296883.2, 21322671.29, 21348459.37, 21374247.45,
21400035.53).
Sequences with an e value of less than 1e-5 to Ig-like domains
were retrieved. ORFs were called from the metagenome reads using
Artemis (Wellcome Trust Sanger Institute); their position in the
FASTA file is shown in Table S2. ORFs that were 60 bp long with
40% tBLASTn identity to T4 Hoc or a member of the PKD Ig
Pfam were retained. The six contigs containing Ig-like hypervariable
domains from the published study by Minot et al. (7) were down-
loaded from the National Center for Biotechnology Information
(NCBI). Identical sequences were collapsed using the Trie clus-
tering method implemented in Qiime (8). The resulting unique
sequences were mapped to the position-specific scoring matrix for
the PKD Ig Pfam (PF00801) using hmmalign (9). The hmmalign
trimming function was used; sequences that were dominated by
gaps after alignment were removed. A maximum likelihood tree
was generated from the aligned unique sequences using FastTree
version 2.1.1 SSE3 and viewed in MEGA 5. Environmental data for
the metagenomes were obtained from the MyMg database. In a
separate analysis, structural homology of these same sequences
to a carbohydrate-binding protein (10) was determined using the
Phyre2 structural homology prediction pipeline (www.sbg.bio.ic.
ac.uk/phyre2/html/help.cgi).
1. Patel A, et al. (2007) Virus and prokaryote enumeration from planktonic aquatic
environments by epifluorescence microscopy with SYBR Green I. Nat Protoc 2(2):269–276.
2. Forstner G, Zhang Y, McCool D, Forstner J (1993) Mucin secretion by T84 cells:
Stimulation by PKC, Ca2+, and a protein kinase activated by Ca2+ ionophore. Am J
Physiol 264(6 Pt 1):G1096–G1102.
3. Alemka A, et al. (2010) Probiotic colonization of the adherent mucus layer of
HT29MTXE12 cells attenuates Campylobacter jejuni virulence properties. Infect Immun
78(6):2812–2822.
4. Sbalzarini IF, Koumoutsakos P (2005) Feature point tracking and trajectory analysis for
video imaging in cell biology. J Struct Biol 151(2):182–195.
5. Suh J, Dawson M, Hanes J (2005) Real-time multiple-particle tracking: Applications to
drug and gene delivery. Adv Drug Deliv Rev 57(1):63–78.
6. Lai SK, et al. (2007) Rapid transport of large polymeric nanoparticles in fresh
undiluted human mucus. Proc Natl Acad Sci USA 104(5):1482–1487.
7. Minot S, Grunberg S, Wu GD, Lewis JD, Bushman FD (2012) Hypervariable loci in the
human gut virome. Proc Natl Acad Sci USA 109(10):3962–3966.
8. Caporaso JG, et al. (2010) QIIME allows analysis of high-throughput community
sequencing data. Nat Methods 7(5):335–336.
9. Eddy SR (2009) A new generation of homology search tools based on probabilistic
inference. Genome Inform 23(1):205–211.
10. Najmudin S, et al. (2006) Xyloglucan is recognized by carbohydrate-binding modules
that interact with β-glucan chains. J Biol Chem 281(13):8815–8828.
Barr et al. www.pnas.org/cgi/content/short/1305923110 2 of 6
Fig. S1. Epifluorescence counts of phage and bacteria from diverse environments and mucosa. (Left to right) Invertebrates: Actiniaria sp., Acropora sp.,
Echinopora sp., Oxypora sp., Capnela sp., and Phyllodoce sp. Vertebrates: Paralichthys sp., Homo sapiens, and Mus musculis. Error bars represent ±SD with n > 25.
Fig. S2. Mucolytic treatment of mucus-producing A549 cells. Mucus removal from A549 lung epithelial cells by NAC treatment was assessed by PAS/AB stain,
which stains mucus-like substances pink/purple. Scale bars represent 100 μm.
Fig. S3. Growth and mucus production of A549 and siRNA knockdown cell lines. Shown are mucus-producing A549 lung epithelial TC cells, mucus-producing
nonsense shRNA control A549 cell line (shControl), and non–mucus-producing MUC1 and MUC5AC shRNA knockdown A549 cell line (MUC–
) after 2 and 4 d in
culture. Mucus production was assessed on day 5 by PAS/AB stain, which stains mucus-like substances pink/purple. Scale bars represent 100 μm.
Barr et al. www.pnas.org/cgi/content/short/1305923110 3 of 6
Fig. S4. Western blot analysis of MUC1 and MUC5AC in total cell lysates of A549 lung epithelial cell knockdowns. Lysates of confluent cell layers were
separated by SDS/PAGE and then immunoblotted with anti-MUC1 and anti-MUC5AC antibodies. Shown are the MUC–
knockdown cell line, the nonsilencing
shControl control cell line, and native A549 cells. GAPDH was used as an intracellular protein control.
Fig. S5. Surface-free control for the assay of phage adherence to mucus-associated macromolecules. Both hoc+
and hoc–
T4 phage (109
pfu·mL−1
) were serially
diluted to 1 × 10−7
and 1 × 10−8
, and then incubated in 1% (wt/vol) solutions of mucin, DNA, or protein in 1 mL LB for 30 min at 37 °C. Each incubation mixture
was then mixed with Escherichia coli top agar and layered over plain agar plates. Resulting plaque-forming unit (PFU) counts showed that infectivity of hoc+
and hoc–
T4 phage was not significantly altered in the presence of the macromolecules used in the phage adherence assays (mucin, DNA, and BSA protein).
Fig. S6. Adherence of Ig+
and Ig–
T3 phage to mucin. Phage adherence assays to mucin-coated agar plates were performed as described in SI Materials and
Methods, except that the Ig+
and Ig–
T3 phage (1011
pfu·mL−1
) were serially diluted to 1 × 10−9
and 1 × 10−10
pfu·ml−1
. The resultant PFU counts of adherent
phage showed that Ig+
T3 phage adhered to mucin-coated agar plates significantly more than to the plain agar control plates (n = 6, t = 4.443, **P = 0.0012,
unpaired t test), whereas there was no significant increase in adherence for the Ig–
T3 phage. ns, not significant.
Barr et al. www.pnas.org/cgi/content/short/1305923110 4 of 6
Table S1. Phyre2 structural homology of Ig-like proteins encoded by viral metagenomes
Phyre2 analysis ORF, bp
No. Environment PDB ID
Confidence,
%
Identity,
% Start Stop Length Database Sequence identifier
1 Sputum 2C26-A 99.8 29 29 277 248 MyMg d7c74d66ea493c0c1fca41f718d22125_16271_279
2 Sputum 2C26-A 99.9 36 280 555 275 MyMg d7c74d66ea493c0c1fca41f718d22125_64742_278
3 Sputum 2C26-A 99.7 32 837 1,066 229 MyMg 03ca0e6ad90102ab264cf521ed58209e_112029_232
4 Sputum 2C26-A 99.7 26 2,951 3,184 233 MyMg 03ca0e6ad90102ab264cf521ed58209e_52923_234
5 Sputum 2C26-A 99.7 32 3,417 3,654 237 MyMg 2da3ea31d1b30a11b0f080a5b91b9df2_256934_240
6 Sputum 2C26-A 99.7 31 4,189 4,416 227 MyMg 03ca0e6ad90102ab264cf521ed58209e_134692_228
7 Freshwater 2C26-A 99.8 29 3,935 4,188 253 MyMg 6f8c77f72920950139dc6b3520cf86b7_69438_254
8 Oral 2C26-A 99.8 29 2,437 2,715 278 MyMg f44d959b723905a049b0334f19668e5c_48421_279
9 Sputum 2C26-A 99.8 30 2,716 2,949 233 MyMg 875f11dbd745609c9d3e12c5b3b5636a_66237_235
10 Sputum 2C26-A 99.9 30 16,236 16,491 255 MyMg 88f61453e560016a0e2a238351d7292b_109758_257
11 Sputum 2C26-A 99.9 29 15,508 15,766 258 MyMg ebdf0605a03616ab168eddf68ca506e1_121095_259
12 Marine* 1E07-A 99.8 13 3 800 797 MyMg 05d2b5884d248d570fe8a2c0d390c97c_3911_800
13 Human feces 2C26-A 88.8 23 1 246 245 MyMg f58fff76a10b642883986ee0e1a30514_10941_246
14 Human feces 2C26-A 94.7 31 248 484 236 MyMg f58fff76a10b642883986ee0e1a30514_22104_238
15 Human feces 2C26-A 86.8 22 720 960 240 MyMg f58fff76a10b642883986ee0e1a30514_10554_243
16 Human feces 2C26-A 86 29 1,445 1,682 237 MyMg f58fff76a10b642883986ee0e1a30514_32210_238
17 Human feces 2C26-A 85.5 32 2,400 2,637 237 MyMg f58fff76a10b642883986ee0e1a30514_22379_238
18 Human feces 2C26-A 90.1 26 3,122 3,361 239 MyMg f58fff76a10b642883986ee0e1a30514_6620_240
19 Human feces 2C26-A 98.1 11 574 NCBI gi 377806168 gb AFB75876.1
20 Human feces 2C26-A 99 15 429 NCBI gi 377806248 gb AFB75953.1
21 Human feces 2C26-A 99.6 18 529 NCBI gi 377806350 gb AFB76049.1
All Ig-like domain homologs shown in Fig. 4A displayed high structural homology with a promiscuous carbohydrate-binding domain [Protein Data Bank (PDB) 2C26].
The Hoc homolog from a marine sample (no. 12, column 1) displayed high structural homology with 1E07-A as well as several other immune proteins. bp, base pairs.
*Hoc homolog.
Barr et al. www.pnas.org/cgi/content/short/1305923110 5 of 6
Table S2. Glycan microarray analysis of T4 and hoc–
phage displayed in Fig. 4A
Glycan no. Structure Linkage T4 RFU T4 %CV hoc–
RFU hoc–
%CV
609 GlcNAcb1-3Fuca -N(CH3)-O-(CH2)2-NH2 4,921 3 394 16
610 Galb1-3GalNAcb1-4(Neu5Aca2-8Neu5Aca2-8Neu5Aca2-3)
Galb1-4Glcb
-N(CH3)-O-(CH2)2-NH2 4,685 11 472 13
573 Neu5Aca2-8Neu5Aca2-3Galb1-3GalNAcb1-4(Neu5Aca2-3)
Galb1-4Glc
-N(CH3)-O-(CH2)2-NH2 4,161 1 316 11
608 Neu5Aca2-6Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-2Mana1-
6(Neu5Aca2-6Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-
2Mana1-3)Manb1-4GlcNAcb1-4GlcNAcb
Asparagine 4,685 4 497 17
145 Galb1-3GalNAcb1-4Galb1-4Glcb CH2CH2CH2NH2 5,823 3 565 7
195 Glca1-4Glcb CH2CH2CH2NH2 5,845 1 544 3
514 GalNAcb1-4(6S)GlcNAc CH2CH2CH2NH2 4,518 6 568 17
287 Neu5Gca CH2CH2CH2NH2 4,182 3 356 9
119 Gala1-4(Fuca1-2)Galb1-4GlcNAcb CH2CH2CH2NH2 4,157 3 331 8
336 GalNAca1-3(Fuca1-2)Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-
3Galb1-4GlcNAcb
CH2CH2NH2 7,670 31 472 51
217 Manb1-4GlcNAcb CH2CH2NH2 4,766 3 647 24
144 Galb1-3GalNAcb1-4(Neu5Aca2-3)Galb1-4Glcb CH2CH2NH2 4,797 4 589 26
517 Galb1-4(6P)GlcNAcb CH2CH2NH2 4,751 11 526 20
218 Neu5Aca2-3Galb1-4GlcNAcb1-3Galb1-4(Fuca1-3)GlcNAcb CH2CH2NH2 4,737 3 322 10
334 GalNAcb1-3Gala1-4Galb1-4GlcNAcb1-3Galb1-4Glcb CH2CH2NH2 4,261 4 499 8
143 Galb1-3GalNAcb1-3Gala1-4Galb1-4Glcb CH2CH2NH2 4,236 1 451 13
581 Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-
3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-2Mana1-6(Galb1-
4GlcNAcb1-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-
3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-2Mana1-3)
Manb1-4GlcNAcb1-4(Fuca1-6)GlcNAcb
EN or NK 5,465 6 262 7
360 Fuca1-2Galb1-3GlcNAcb1-2Mana1-6(Fuca1-2Galb1-
3GlcNAcb1-2Mana1-3)Manb1-4GlcNAcb1-4GlcNAcb
GENR 4,914 2 482 12
588 Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-
3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-6(Galb1-
4GlcNAcb1-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-
3Galb1-4GlcNAcb1-3Galb1-4GlcNAb1-2)Mana1-6(Galb1-
4GlcNAcb1-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-
3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-2Mana1-3)
Manb1-4GlcNAcb1-4(Fuca1-6)GlcNAcb
KVANKT 3,926 2 154 5
470 Glca1-4Glca1-4Glca1-4Glcb NHCOCH2NH 7,521 10 1,288 17
516 (4S)GalNAcb NHCOCH2NH 6,148 2 464 3
359 KDNa2-3Galb1-3GalNAca Threonine (O-linked glycan) 7,484 17 1,080 24
471 Neu5Aca2-3Galb1-4GlcNAcb1-6(Neu5Aca2-3Galb1-
4GlcNAcb1-3)GalNAca
Threonine (O-linked glycan) 5,877 6 585 2
491 Neu5Aca2-3Galb1-3GlcNAcb1-6GalNAca Threonine (O-linked glycan) 4,755 2 394 8
596 Neu5Aca2-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-6
(Neu5Aca2-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-3)
GalNAca
Threonine (O-linked glycan) 4,703 14 242 7
595 GlcNAcb1-3Galb1-4GlcNAcb1-6(GlcNAcb1-3Galb1-
4GlcNAcb1-3)GalNAca
Threonine (O-linked glycan) 4,653 1 392 9
605 Neu5Aca2-6Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-6
(Neu5Aca2-6Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-3)
GalNAca
Threonine (O-linked glycan) 4,078 1 410 10
480 Neu5Aca2-6Galb1-4GlcNAcb1-6GalNAca Threonine (O-linked glycan) 4,143 5 301 20
592 Neu5Aca2-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-
3GalNAca
Threonine (O-linked glycan) 4,200 11 266 11
“Glycan no.” indicates the glycan ID number used on the Consortium for Functional Glycomics Version 5.1 microarray. “Linkage” denotes the chemical
linkage joining the glycan to the macromolecule. Bold threonine linkages represent O-linked glycan residues likely to be associated with mucin glycoproteins.
Barr et al. www.pnas.org/cgi/content/short/1305923110 6 of 6

More Related Content

What's hot

What's hot (20)

EVE 161 Winter 2018 Class 17
EVE 161 Winter 2018 Class 17EVE 161 Winter 2018 Class 17
EVE 161 Winter 2018 Class 17
 
EVE 161 Winter 2018 Class 14
EVE 161 Winter 2018 Class 14EVE 161 Winter 2018 Class 14
EVE 161 Winter 2018 Class 14
 
EVE 161 Winter 2018 Class 13
EVE 161 Winter 2018 Class 13EVE 161 Winter 2018 Class 13
EVE 161 Winter 2018 Class 13
 
Microbial Phylogenomics (EVE161) Class 5
Microbial Phylogenomics (EVE161) Class 5Microbial Phylogenomics (EVE161) Class 5
Microbial Phylogenomics (EVE161) Class 5
 
Beiko cms final
Beiko cms finalBeiko cms final
Beiko cms final
 
Gene sharing in microbes: good for the individual, good for the community?
Gene sharing in microbes: good for the individual, good for the community?Gene sharing in microbes: good for the individual, good for the community?
Gene sharing in microbes: good for the individual, good for the community?
 
Microbial Metagenomics and Human Health
Microbial Metagenomics and Human HealthMicrobial Metagenomics and Human Health
Microbial Metagenomics and Human Health
 
Metagenomics by microbiology dept. panjab university2018copy
Metagenomics by microbiology dept. panjab university2018copyMetagenomics by microbiology dept. panjab university2018copy
Metagenomics by microbiology dept. panjab university2018copy
 
Metagenomics: An overview
Metagenomics: An overviewMetagenomics: An overview
Metagenomics: An overview
 
1-s2.0-S0923250811001872-main
1-s2.0-S0923250811001872-main1-s2.0-S0923250811001872-main
1-s2.0-S0923250811001872-main
 
Clinical Metagenomics for Rapid Detection of Enteric Pathogens and Characteri...
Clinical Metagenomics for Rapid Detection of Enteric Pathogens and Characteri...Clinical Metagenomics for Rapid Detection of Enteric Pathogens and Characteri...
Clinical Metagenomics for Rapid Detection of Enteric Pathogens and Characteri...
 
Beiko dcsi2013
Beiko dcsi2013Beiko dcsi2013
Beiko dcsi2013
 
Targeted RNA Sequencing, Urban Metagenomics, and Astronaut Genomics
Targeted RNA Sequencing, Urban Metagenomics, and Astronaut GenomicsTargeted RNA Sequencing, Urban Metagenomics, and Astronaut Genomics
Targeted RNA Sequencing, Urban Metagenomics, and Astronaut Genomics
 
Microbial Phylogenomics (EVE161) Class 7: rRNA PCR and Major Groups
Microbial Phylogenomics (EVE161) Class 7: rRNA PCR and Major Groups Microbial Phylogenomics (EVE161) Class 7: rRNA PCR and Major Groups
Microbial Phylogenomics (EVE161) Class 7: rRNA PCR and Major Groups
 
the others our biased perspective
the others our biased perspectivethe others our biased perspective
the others our biased perspective
 
UC Davis EVE161 Lecture 9 by @phylogenomics
UC Davis EVE161 Lecture 9 by @phylogenomicsUC Davis EVE161 Lecture 9 by @phylogenomics
UC Davis EVE161 Lecture 9 by @phylogenomics
 
American Gut Project presentation at Masaryk University
American Gut Project presentation at Masaryk UniversityAmerican Gut Project presentation at Masaryk University
American Gut Project presentation at Masaryk University
 
metagenomics
metagenomicsmetagenomics
metagenomics
 
Eugene Koonin for Knowledge Stream
Eugene Koonin for Knowledge StreamEugene Koonin for Knowledge Stream
Eugene Koonin for Knowledge Stream
 
Chp%3 a10.1007%2f978 81-322-2283-5-31
Chp%3 a10.1007%2f978 81-322-2283-5-31Chp%3 a10.1007%2f978 81-322-2283-5-31
Chp%3 a10.1007%2f978 81-322-2283-5-31
 

Similar to PNAS-2013-Barr-10771-6

Abstract_Xmeeting_Laux
Abstract_Xmeeting_LauxAbstract_Xmeeting_Laux
Abstract_Xmeeting_Laux
Marcele Laux
 
Jeffrey Noland Publication 2013
Jeffrey Noland Publication 2013Jeffrey Noland Publication 2013
Jeffrey Noland Publication 2013
Jeffrey Noland
 
fmicb-10-01923 August 20, 2019 Time 1756 # 1ORIGINAL RES
fmicb-10-01923 August 20, 2019 Time 1756 # 1ORIGINAL RESfmicb-10-01923 August 20, 2019 Time 1756 # 1ORIGINAL RES
fmicb-10-01923 August 20, 2019 Time 1756 # 1ORIGINAL RES
ShainaBoling829
 

Similar to PNAS-2013-Barr-10771-6 (20)

Soil microbial diversity
Soil microbial diversitySoil microbial diversity
Soil microbial diversity
 
P gingivalis
P gingivalis P gingivalis
P gingivalis
 
Ayyappan et al., PLoS One
Ayyappan et al., PLoS OneAyyappan et al., PLoS One
Ayyappan et al., PLoS One
 
Abstract_Xmeeting_Laux
Abstract_Xmeeting_LauxAbstract_Xmeeting_Laux
Abstract_Xmeeting_Laux
 
Rapid Impact Assessment of Climatic and Physio-graphic Changes on Flagship G...
Rapid Impact Assessment of Climatic and Physio-graphic Changes  on Flagship G...Rapid Impact Assessment of Climatic and Physio-graphic Changes  on Flagship G...
Rapid Impact Assessment of Climatic and Physio-graphic Changes on Flagship G...
 
Smbe 2013 talk
Smbe 2013 talkSmbe 2013 talk
Smbe 2013 talk
 
Biofilms
BiofilmsBiofilms
Biofilms
 
Marine Host-Microbiome Interactions: Challenges and Opportunities
Marine Host-Microbiome Interactions: Challenges and OpportunitiesMarine Host-Microbiome Interactions: Challenges and Opportunities
Marine Host-Microbiome Interactions: Challenges and Opportunities
 
The transformational role of polymerase chain reaction (pcr) in environmental...
The transformational role of polymerase chain reaction (pcr) in environmental...The transformational role of polymerase chain reaction (pcr) in environmental...
The transformational role of polymerase chain reaction (pcr) in environmental...
 
José Luis Martínez - Simposio Microbiología: Transmisión
José Luis Martínez - Simposio Microbiología: TransmisiónJosé Luis Martínez - Simposio Microbiología: Transmisión
José Luis Martínez - Simposio Microbiología: Transmisión
 
Austin Andrology
Austin AndrologyAustin Andrology
Austin Andrology
 
Rick Stevens: Prospects for a Systematic Exploration of Earths Microbial Dive...
Rick Stevens: Prospects for a Systematic Exploration of Earths Microbial Dive...Rick Stevens: Prospects for a Systematic Exploration of Earths Microbial Dive...
Rick Stevens: Prospects for a Systematic Exploration of Earths Microbial Dive...
 
bats bacterioma.pdf
bats bacterioma.pdfbats bacterioma.pdf
bats bacterioma.pdf
 
Biofilms
BiofilmsBiofilms
Biofilms
 
PhD thesis_Opt
PhD thesis_OptPhD thesis_Opt
PhD thesis_Opt
 
Jeffrey Noland Publication 2013
Jeffrey Noland Publication 2013Jeffrey Noland Publication 2013
Jeffrey Noland Publication 2013
 
Phylogeny of Bacterial and Archaeal Genomes Using Conserved Genes: Supertrees...
Phylogeny of Bacterial and Archaeal Genomes Using Conserved Genes: Supertrees...Phylogeny of Bacterial and Archaeal Genomes Using Conserved Genes: Supertrees...
Phylogeny of Bacterial and Archaeal Genomes Using Conserved Genes: Supertrees...
 
fmicb-10-01923 August 20, 2019 Time 1756 # 1ORIGINAL RES
fmicb-10-01923 August 20, 2019 Time 1756 # 1ORIGINAL RESfmicb-10-01923 August 20, 2019 Time 1756 # 1ORIGINAL RES
fmicb-10-01923 August 20, 2019 Time 1756 # 1ORIGINAL RES
 
Enzybiotics New era of antimicrobials.ppt
Enzybiotics New era of antimicrobials.pptEnzybiotics New era of antimicrobials.ppt
Enzybiotics New era of antimicrobials.ppt
 
10.1007@978-981-13-8315-1.pdf
10.1007@978-981-13-8315-1.pdf10.1007@978-981-13-8315-1.pdf
10.1007@978-981-13-8315-1.pdf
 

PNAS-2013-Barr-10771-6

  • 1. Bacteriophage adhering to mucus provide a non–host-derived immunity Jeremy J. Barra,1 , Rita Auroa , Mike Furlana , Katrine L. Whitesona , Marcella L. Erbb , Joe Poglianob , Aleksandr Stotlanda , Roland Wolkowicza , Andrew S. Cuttinga , Kelly S. Dorana , Peter Salamonc , Merry Youled , and Forest Rohwera a Department of Biology, San Diego State University, San Diego, CA 92182; b Division of Biological Sciences, University of California, San Diego, CA 92093; c Department of Mathematics and Statistics, San Diego State University, San Diego, CA 92182; and d Rainbow Rock, Ocean View, HI 96737 Edited by Richard E. Lenski, Michigan State University, East Lansing, MI, and approved April 18, 2013 (received for review March 28, 2013) Mucosal surfaces are a main entry point for pathogens and the principal sites of defense against infection. Both bacteria and phage are associated with this mucus. Here we show that phage- to-bacteria ratios were increased, relative to the adjacent envi- ronment, on all mucosal surfaces sampled, ranging from cnidarians to humans. In vitro studies of tissue culture cells with and without surface mucus demonstrated that this increase in phage abun- dance is mucus dependent and protects the underlying epithelium from bacterial infection. Enrichment of phage in mucus occurs via binding interactions between mucin glycoproteins and Ig-like protein domains exposed on phage capsids. In particular, phage Ig-like domains bind variable glycan residues that coat the mucin glycoprotein component of mucus. Metagenomic analysis found these Ig-like proteins present in the phages sampled from many environments, particularly from locations adjacent to mucosal surfaces. Based on these observations, we present the bacterio- phage adherence to mucus model that provides a ubiquitous, but non–host-derived, immunity applicable to mucosal surfaces. The model suggests that metazoan mucosal surfaces and phage co- evolve to maintain phage adherence. This benefits the metazoan host by limiting mucosal bacteria, and benefits the phage through more frequent interactions with bacterial hosts. The relationships shown here suggest a symbiotic relationship between phage and metazoan hosts that provides a previously unrecognized antimi- crobial defense that actively protects mucosal surfaces. symbiosis | host-pathogen | virus | immunoglobulin | immune system Mucosal surfaces are the primary zones where animals meet their environment, and thus also the main points of entry for pathogenic microorganisms. The mucus layer is heavily col- onized by bacteria, including many symbionts that contribute additional genetic and metabolic potential to the host (1, 2). Bacterial symbionts associated with a variety of other host sur- faces also provide goods and services, e.g., nutrients (3–6), bio- luminescence (7, 8), and antibiotics (9, 10). These resident symbionts benefit from increased nutrient availability (5, 11–13), as well as the opportunity for both vertical transmission and in- creased dissemination (14–16). Within the mucus, the predominant macromolecules are the large (up to 106 –109 Da) mucin glycoproteins. The amino acid backbone of these proteins incorporates tandem repeats of ex- posed hydrophobic regions alternating with blocks bearing ex- tensive O-linked glycosylation (17). Hundreds of variable, branched, negatively charged glycan chains extend 0.5–5 nm from the peptide core outward into the surrounding environment (17, 18). Other proteins, DNA, and cellular debris also are present. Continual secretion and shedding of mucins maintain a protective mucus layer from 10–700 μm thick depending on species and body location (19–22). By offering both structure and nutrients, mucus layers com- monly support higher bacterial concentrations than the sur- rounding environment (11, 23). Of necessity, hosts use a variety of mechanisms to limit microbial colonization (24–27). Secretions produced by the underlying epithelium influence the composition of this microbiota (12, 27, 28). When invaded by pathogens, the epithelium may respond by increased production of antimicrobial agents, hypersecretion of mucin, or alteration of mucin glycosyl- ation patterns to subvert microbial attachment (29–31). Also present in the mucus environment are bacteriophage (phage), the most common and diverse biological entities. As specific bacterial predators, they increase microbial diversity through Red Queen/kill-the-winner dynamics (32, 33). Many phages establish conditional symbiotic relationships with their bacterial hosts through lysogeny. As integrated prophages, they often express genes that increase host fitness or virulence (34– 36) and protect their host from lysis by related phages. As free phage, they aid their host strain by killing related competing strains (37–39). Phages participate, along with their bacterial hosts, in tripartite symbioses with metazoans that affect meta- zoan fitness (40–43). However, no direct symbiotic interactions between phage and metazoans are known. Recently, Minot et al. (44) showed that phages in the human gut encode a population of hypervariable proteins. For 29 hyper- variable regions, evidence indicated that hypervariability was con- ferred by targeted mutagenesis through a reverse transcription mechanism (44, 45). Approximately half of these encoded proteins possessed the C-type lectin fold previously found in the major tropism determinant protein at the tip of the Bordetella phage BPP- 1 tail fibers (46); six others contained Ig-like domains. These Ig-like proteins, similar to antibodies and T-cell receptors, can accom- modate large sequence variation (>1013 potential alternatives) (47). Ig-like domains also are displayed in the structural proteins of many phage (48, 49). That most of these displayed Ig-like domains are dispensable for phage growth in the laboratory (45, 49) led to the hypothesis that they aid adsorption to their bacterial prey under environmental conditions (49). The possible role and function of these hypervariable proteins remain to be clarified. Here, we show that phage adhere to mucus and that this as- sociation reduces microbial colonization and pathology. In vitro studies demonstrated that this adherence was mediated by the interaction between displayed Ig-like domains of phage capsid proteins and glycan residues, such as those in mucin glyco- proteins. Homologs of these Ig-like domains are encoded by phages from many environments, particularly those adjacent to mucosal surfaces. We propose the bacteriophage adherence to mucus (BAM) model whereby phages provide a non–host- derived antimicrobial defense on the mucosal surfaces of diverse metazoan hosts. Author contributions: J.J.B. and F.R. designed research; J.J.B., R.A., K.L.W., M.L.E., J.P., A.S.C., and P.S. performed research; J.J.B., K.L.W., M.L.E., J.P., A.S., R.W., A.S.C., and K.S.D. contributed new reagents/analytic tools; J.J.B., R.A., M.F., K.L.W., A.S., R.W., P.S., M.Y., and F.R. analyzed data; and J.J.B., M.Y., and F.R. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. Freely available online through the PNAS open access option. Data deposition: Raw glycan array data are available from the Consortium for Functional Glycomics (accession no. 2621). See Commentary on page 10475. 1 To whom correspondence should be addressed. E-mail: jeremybarr85@gmail.com. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1305923110/-/DCSupplemental. www.pnas.org/cgi/doi/10.1073/pnas.1305923110 PNAS | June 25, 2013 | vol. 110 | no. 26 | 10771–10776 MICROBIOLOGYSEECOMMENTARY
  • 2. Results Phage Adhere to Mucus. Our preliminary investigations of mucosal surfaces suggested that phage concentrations in the mucus layer were elevated compared with the surrounding environment. Here, we used epifluorescence microscopy to count the phage and bacteria in mucus sampled from a diverse range of mucosal surfaces (e.g., sea anemones, fish, human gum), and in each adjacent environment (SI Materials and Methods and Fig. S1). Comparing the calculated phage-to-bacteria ratios (PBRs) showed that PBRs in metazoan-associated mucus layers were on average 4.4-fold higher than those in the respective adjacent environment (Fig. 1A). The PBRs on these mucus surfaces ranged from 21:1 to 87:1 (average, 39:1), compared with 3:1 to 20:1 for the surrounding milieus (average, 9:1; n = 9, t = 4.719; ***P = 0.0002). Earlier investigations of phage abundance in marine environments reported that phage typically outnumber bacteria by an order of magnitude (50–52), but here we dem- onstrate that this margin was significantly larger in metazoan- associated mucus surface layers. To determine whether this enrichment was dependent on the presence of mucus rather than some general properties of the cell surface (e.g., charge), phage adherence was tested with tissue culture (TC) cells with and without surface mucus (SI Materials and Methods). In these assays, T4 phage were washed across confluent cell monolayers for 30 min, after which nonadherent phage were removed by repeated washings and the adherent phage quantified by epifluorescence microscopy. Two mucus- producing cell lines were used: T84 (human colon epithelial cells) and A549 (human lung epithelial cells). For these cells, mucin secretion was stimulated by pretreatment with phorbol 12-myristate 13-acetate (53, 54). Comparison of the T84 cells with the non–mucus-producing Huh-7 human hepatocyte cell line showed that T4 phage adhered significantly more to the mucus-producing T84 cells (Fig. 1B; n > 18, t = 8.366; ****P < 0.0001). To demonstrate the mucus dependence of this adher- ence, the mucus layer was chemically removed from A549 cells by N-acetyl-L-cysteine (NAC) treatment (55) (Fig. S2). This significantly reduced the number of adherent phage to levels similar to those observed with non–mucus-producing cell lines (Fig. 1B; n > 40, t = 9.561; ****P < 0.0001). We also created an A549 shRNA mucus knockdown cell line (MUC– ), reducing mucus production in A549, and a nonsense shRNA control (shControl; Figs. S3 and S4). Again, T4 phage adhered signifi- cantly more to the mucus-producing cells (Fig. 1B; n > 37, t = 7.673; ****P < 0.0001). Although mucin glycoproteins are the predominant component of mucus, other macromolecular components also are present, any of which might be involved in the observed phage adherence. We developed a modified top agar assay to determine whether phage adhered to a specific macromolecular component of mucus. Plain agar plates and agar plates coated with 1% (wt/vol) mucin, DNA, or protein were prepared. That concentration was chosen because it is at the low end of the range of physiological mucin concen- trations (56). T4 phage suspensions were incubated on the plates for 30 min, after which the phage suspension was decanted to remove unbound particles. The plates then were overlaid with a top agar containing Escherichia coli hosts and incubated over- night. The number of adherent phage was calculated from the number of plaque-forming units (pfu) observed. Significantly more T4 phage adhered to the 1% mucin-coated agar surface (Fig. 1C; n = 12, t = 5.306; ****P < 0.0001). Combined, these three assays show that phage adhere to mucin glycoproteins. Phage Adherence and Bacterial Infection. The mucus layer is an optimal environment for microbial growth, providing structure as well as nutrients in the form of diverse, mucin-associated glycans. To limit this growth, the metazoan host retards microbial coloni- zation by diverse antimicrobial mechanisms (24–27). Does the increased number of adherent phage found on mucosal surfaces also reduce microbial colonization? To answer this, bacterial at- tachment to mucus-producing and non–mucus-producing TC cells was assayed both with and without pretreatment of the cells with the mucus-adherent phage T4. Here confluent monolayers of TC cells were overlaid with T4 phage for 30 min, washed to remove nonadherent phage, and then incubated with E. coli for 4 h. Cells then were scraped from the plates and the attached bacteria were fluorescently stained and counted by epifluorescence microscopy. Phage pretreatment of mucus-producing TC cell lines (T84, A549) significantly decreased subsequent bacterial attachment (Fig. 2A; T84: n > 30, t = 32.05, ****P < 0.0001; A549: n > 30, t = 36.85, ****P < 0.0001). Phage pretreatment of non–mucus-producing cells (Huh-7; MUC– , an A549 mucus knockdown strain) had a much smaller effect on bacterial attachment. To determine whether this reduced bacterial attachment de- pended on bacterial lysis and the production of progeny viruses, we repeated these experiments using an amber mutant T4 phage (T4 am43– 44– ). When infecting wild-type E. coli, this phage pro- duces no infective progeny virions, but infection of the E. coli amber suppressor strain SupD yields infective virions. For these experiments, mucus-producing A549 cells were pretreated with amber mutant T4 am43– 44– phage and then incubated with either wild-type or the amber suppressor strain E. coli (Fig. 2B). Bacte- rial attachment was reduced by more than four orders of magni- tude when phage could replicate and thereby increase the number of infective virions within the mucus (n = 8, ****P < 0.0001). Comparatively, when no phage replication occurred in the mucus, there was no observable change in bacterial colonization and fewer phages were detected (n = 8, *P = 0.0227). These results show that pretreatment of a mucosal surface with phage reduces adherence of a bacterial pathogen and that this protection is mediated by continued phage replication in the mucus. To test whether the observed reduced bacterial adherence was accompanied by reduced pathology of the underlying TC cells, mucus-producing A549 and non–mucus-producing MUC– TC cells were exposed to E. coli overnight, either with or without a 30-min pretreatment with T4 phage. Infection was quantified as the per- centage of cell death. Adherence of phage effectively protected the mucus-producing cells against the subsequent bacterial chal- lenge (Fig. 2C; n = 12, ****P < 0.0001). Phage pretreatment showed a reduced protection to the non–mucus-producing MUC– Fig. 1. Phage adhere to cell-associated mucus layers and mucin glycopro- tein. (A) PBR for diverse mucosal surfaces and the adjacent environment. On average, PBRs for mucosal surfaces were 4.4-fold greater than for the ad- jacent environment (n = 9, t = 4.719, ***P = 0.0002, unpaired t test). (B) Phage adherence to TC cell monolayers, with and without surface mucus (unpaired t tests). (Left) Non–mucus-producing Huh-7 liver hepatocyte cells and mucus-producing T84 colon epithelial cells (n > 18, t = 8.366, ****P < 0.0001). (Center) Mucus-producing A549 lung epithelial cells with and without treatment with NAC, a mucolytic agent (n > 40, t = 9.561, ****P < 0.0001). (Right) Mucus-producing shRNA control A549 cells (shControl) and mucus knockdown (MUC– ) A549 cells (n > 37, t = 7.673, ****P < 0.0001). (C) Phage adherence to uncoated agar plates and agar coated with mucin, DNA, or protein (n = 12, t = 5.306, ****P < 0.0001, unpaired t test). 10772 | www.pnas.org/cgi/doi/10.1073/pnas.1305923110 Barr et al.
  • 3. cells, decreasing cell death only twofold. Evaluating the impor- tance of mucus production for effective protection, we found that phage pretreatment of mucus-producing A549 cells resulted in a 3.6-fold greater reduction in cell death (n = 12, *P = 0.0181) than the same pretreatment of the mucin knockdown MUC– cells. Role of Capsid Ig-Like Domains in Phage Adherence. Minot et al. (44) recently reported that phage communities associated with the human gut encode a diverse array of hypervariable proteins, including some with hypervariable Ig-like domains. Four Ig-like domains are found in highly antigenic outer capsid protein (Hoc), a T4 phage structural protein of which 155 copies are displayed on the capsid surface (57, 58). Based on this, and given that most Ig-like domains function in recognition and adhesion processes, we hypothesized that the T4 Hoc protein might me- diate the adherence of T4 phage to mucus. To test this, we performed three experiments. First, we compared the adherence of hoc+ T4 phage and a hoc– mutant to mucin-, DNA-, and protein-coated agar plates to an uncoated agar control using the modified top agar assay (see above). Relative to plain agar, the adherence of hoc+ T4 phage increased 4.1-fold for mucin-coated agar (n > 11, t = 3.977, ***P = 0.0007), whereas adherence in- creased only slightly for agar coated with DNA (1.1-fold) or protein (1.2-fold; Fig. 3A). Unlike the hoc+ T4 phage, the hoc– phage did not adhere preferentially to the mucin-coated agar, but instead showed 1.2-, 1.2-, and 1.1-fold increased adherence for mucin, DNA, and protein coatings, respectively. To ensure that none of the macromolecules directly affected phage in- fectivity, hoc+ and hoc– T4 phage were incubated in 1% (wt/vol) solutions of mucin, DNA, or protein. Phage suspensions were combined with E. coli top agar as described above and layered over uncoated agar plates. The results confirmed that the mac- romolecules did not alter phage infectivity (Fig. S5). To provide further evidence that the mucin adherence was dependent on the capsid displayed Ig-like domains rather than some other prop- erty of T4 phage, we repeated the modified top agar assay using Ig+ and Ig– T3 phage. As with T4, the Ig-like domains of T3 are displayed on the surface of the major capsid protein (49). Results indicated a similar increase in adherence to mucin for the Ig+ , but not the Ig– , T3 phage (Fig. S6). Thus, adherence of these phage to mucus requires the Ig-like protein domains. Second, a competition assay using hoc+ and hoc− T4 phage and mucus-producing TC cells was performed to demonstrate the role of mucin in phage adherence. Phage suspended in mucin solutions ranging from 0% to 5% (wt/vol) were washed over confluent layers of mucus-producing A549 TC cells; phage ad- herence then was assayed as above. Adherence of hoc+ T4 phage, but not of hoc– T4 phage, was reduced by mucin com- petition in a concentration-dependent manner (Fig. 3B). Third, interaction of the Hoc protein domains displayed on the capsid surface with mucin glycoproteins was hypothesized to affect the rate of diffusion of T4 virions in mucus. To eval- uate this, we used multiple-particle tracking (MPT) to quantify transport rates of phage particles in buffer and in mucin suspen- sions. The ensemble average effective diffusivity (Deff) calculated at a time scale of 1 s for both hoc+ and hoc– T4 phage in buffer was compared against that in 1% (wt/vol) mucin suspensions (SI Materials and Methods). Both hoc+ and hoc– phage diffuse rapidly through buffer (Fig. 3C). Whereas hoc– phage diffused in 1% mucin at the same rate as in buffer, the mucin decreased the diffusion rate for hoc+ phage particles eightfold. Thus, all three of these experimental approaches supported our hypothesis that Fig. 2. Effect of phage adsorption on subsequent bacterial infection of epithelial cells. (A) Bacterial attachment to mucus-producing (T84 and A549) and non–mucus-producing (Huh-7, MUC– ) TC cells, with and without phage pretreatment. T4 phage pretreatment significantly decreased subsequent bacterial adherence to mucus-producing TC cell lines (T84: n > 30, t = 32.05, ****P < 0.0001; A549: n > 30, t = 36.85, ****P < 0.0001; unpaired t tests). Less dramatic shifts were seen for non–mucus-producing cells (Huh-7: n > 30, t = 2.72, **P = 0.0098; MUC– : n > 30, t = 3.52, ***P = 0.0007; unpaired t tests). (B) Mucus-producing A549 cells were pretreated with T4 am43– 44– phage (Materials and Methods) and then incubated for 4 h with either wild- type (wt) or amber-suppressor (SupD) E. coli. Phage replication in the SupD E. coli strain significantly reduced bacterial colony-forming units (CFU) in the mucus (n = 8, ****P < 0.0001, Tukey’s two-way ANOVA) and increased phage-forming units (PFU) relative to the no-phage replication wt E. coli (n = 8, *P = 0.0227). (C) Mortality of mucus-producing (A549) and mucus knockdown (MUC– ) A549 lung epithelial cells following overnight in- cubation with E. coli. Phage pretreatment completely protected mucus- producing A549 cells from bacterial challenge (n = 12, ****P < 0.0001, Tukey’s one-way ANOVA); protection of MUC– cells was 3.1-fold less (n = 12, *P = 0.0181). ns, not significant. Fig. 3. Effect of Hoc protein on phage–mucin interactions. (A) Adherence of hoc+ and hoc– T4 phage to agar coated with mucin, DNA, or protein reported as an increase relative to plain agar controls (n > 11, t = 3.977, ***P = 0.0007, unpaired t test). (B) Competitive effect of mucin on phage adherence when hoc+ and hoc– T4 phage in 0–5% (wt/vol) mucin solution (1× PBS) were washed over mucus-producing A549 cells (n = 25 per sample). (C) Diffusion of fluorescence-labeled hoc+ (Left) and hoc– (Right) T4 phage in buffer and 1% mucin as determined by MPT. Mucin hindered diffusion of hoc+ T4 phage but not hoc– phage (10 analyses per sample, trajectories of n > 100 particles for each analysis; error bars represent SE). Barr et al. PNAS | June 25, 2013 | vol. 110 | no. 26 | 10773 MICROBIOLOGYSEECOMMENTARY
  • 4. the Hoc proteins displayed on the T4 phage capsid interact with mucin. Phage Capsid Ig-Like Domains Interact with Glycans. It is known that ∼25% of sequenced tailed dsDNA phages (Caudovirales) en- code structural proteins with predicted Ig-like domains (48). A search of publicly available viral metagenomes for homologs of the Ig-like domains of the T4 Hoc protein yielded numerous viral Ig-like domains from a variety of environments (Fig. 4A). These domains were more likely to be found in samples collected directly from mucus (e.g., sputum samples) or from an environ- ment adjacent to a mucosal surface (e.g., intestinal lumen, oral cavity). All homologs displayed high structural homology (Phyre2 confidence score average, 96 ± 5%) with a plant-sugar binding domain known for its promiscuous carbohydrate binding speci- ficity (SI Materials and Methods and Table S1), suggesting an in- teraction between these Ig-like domains and glycans. Mucins are complex glycoproteins with highly variable glycan groups exposed to the environment. To investigate whether Hoc interacts with glycans and, if so, to determine whether it interacts with a specific glycan or with a diverse array of glycans, we assayed phage adherence to microarrays printed with 610 mammalian glycans. The hoc+ T4 phage adhered to many di- verse glycans and showed a preference for the O-linked glycan residues typically found in mucin glycoproteins (Fig. 4B, SI Materials and Methods, and Table S2). The hoc– T4 phage exhibited significantly lower affinity for all tested glycans. This indicates that Hoc mediates interactions between T4 phage and varied glycan residues. Discussion In diverse metazoans, body surfaces that interact with the envi- ronment are covered by a protective layer of mucus. Because these mucus layers provide favorable habitats for bacteria, they serve as the point of entry for many pathogens and support large populations of microbial symbionts. Also present are diverse phages that prey on specific bacterial hosts. Moreover, phage concentrations in mucus are elevated relative to the surrounding environment (an average 4.4-fold increase for a diverse sample of invertebrate and vertebrate metazoans; Fig. 1A). The in- creased concentration of lytic phage on mucosal surfaces pro- vides a previously unrecognized metazoan immune defense affected by phage lysis of incoming bacteria. Working with a model system using T4 phage and various TC cell lines, we demonstrated that the increased concentration of phage on mucosal surfaces is mediated by weak binding inter- actions between the variable Ig-like domains on the T4 phage capsid and mucin-displayed glycans. The Ig protein fold is well known for its varied but essential roles in the vertebrate immune response and cell adhesion. Ig-like domains also are present in approximately one quarter of the sequenced genomes of tailed DNA phages, the Caudovirales (48). Notably, these domains were found only in virion structural proteins and typically are displayed on the virion surface. Thus, they were postulated to bind to bacterial surface carbohydrates during infection (48, 49). However, mucin glycoproteins, the predominant macromolecu- lar constituent of mucus, display hundreds of variable glycan chains to the environment that offer potential sites for binding by phage Ig-like proteins. Furthermore, we speculate that phage use the variability of the Ig-like protein scaffold (supporting >1013 potential alternatives) to adapt to the host’s ever-changing pat- terns of mucin glycosylation. The presence of an Ig-like protein (Hoc) displayed on the capsid of T4 phage significantly slowed the diffusion of the phage in mucin solutions. In vivo, similar phage binding to mucin gly- cans would increase phage residence time in mucus layers. Be- cause bacterial concentrations typically are enriched in mucus (Fig. S1), we predict that mucus-adherent phage are more likely to encounter bacteria, potentially increasing their replicative success. If so, phage Ig-like domains that bind effectively to the mucus layer would be under positive selection. Likely, Hoc and other phage proteins with Ig-like domains interact with other glycans with different ramifications, as well (49, 58). Previous metagenomic studies documented the ubiquity and diversity of bacteria and phage within mucus-associated envi- ronments (e.g., human gut, human respiratory tract, corals) (52, 59–64). Known also were some of the essential but adaptable services provided by symbiotic bacteria in these environments (65). However, only recently have efforts been made to in- vestigate the dynamic influences of phage within host-associated ecosystems (37, 44, 66). In this work, we used an in vitro model system to demonstrate a mechanism of phage adherence to the mucus layers that shield metazoan cells from the environment. Furthermore, adherent phage protected the underlying epithelial cells from bacterial infection. Based on these observations and previous research, we proposed the BAM model of immunity, in which the adherence of phage to mucosal surfaces yields a non– host-derived, antimicrobial defense. According to this model (summarized in Fig. 5), the mucus layer, already considered part of the innate immune system and known to provide physical and biochemical antimicrobial defenses (18, 27, 67), also accumulates phage. The model system we used involved a single lytic phage and host bacterium; the situation in vivo undoubtedly is more com- plex. Within the mucosal layer reside diverse bacterial lineages and predictably an even greater diversity of phage strains, both enmeshed within complex phage–bacterial infection networks and engaged in a dynamic arms race (68, 69). These and other factors lower the probability that any given phage–bacterium encounter will result in a successful infection. The time di- mension adds further complexity. The mucus layer is dynamic. Mucins are secreted continually by the underlying epithelium while mucus is sloughed continually from the outer surface. As a result, there is an ongoing turnover of both the bacterial and phage populations in the mucus layer. Driven by kill-the-winner dynamics, the population of phage types that can infect the dominant bacterial types present will cycle along with the pop- ulations of their hosts. Through such mechanisms, we envision that adherent lytic phages provide a dynamic and adaptable defense for their metazoan hosts—a unique example of a meta- zoan–phage symbiosis. We posit that BAM immunity reduces bacterial pathogenesis and provides a previously unrecognized, mucosal immunity. This has far-reaching implications for numerous fields, such as human immunity, gastroenterology, coral disease, and phage therapy. Meanwhile, key questions remain. For instance, what role do temperate phages play in the dynamics of BAM immunity? When integrated into the bacterial chromosome as prophages, Fig. 4. Hoc-mediated glycan binding and Hoc-related phylogeny. (A) Phy- logenetic tree of sequences from viral metagenomes with high-sequence homology to Ig-like domains. Many of the identified homologs are from mucus-associated environments (e.g., human feces, sputum). Also included are the Hoc protein of T4 phage and the hypervariable Ig-like domains previously obtained by deep sequencing of phage DNA from the human gut (44). The scale bar represents an estimated 0.5 amino acid substitutions per site. See SI Materials and Methods for methods. (B) Binding of fluorescence- stained hoc+ and hoc– T4 phage to a microarray of 610 mammalian glycans. Normalized relative fluorescence units (RFU) were calculated from mean fluorescence minus background binding. 10774 | www.pnas.org/cgi/doi/10.1073/pnas.1305923110 Barr et al.
  • 5. they protect their bacterial hosts from infection by related phages; as free phages, they infect and kill sensitive related bacterial strains that compete with their bacterial hosts (37–39). Both mechanisms may benefit their metazoan host by contributing to the maintenance of a selected commensal mucosal microbiota. These possibilities remain to be investigated. Likewise, in vivo investigations are needed to characterize the bacterial and phage diversity present and the consequent effects on BAM immunity. As of now, the relationships shown here open an arena for im- munological study, introduce a phage–metazoan symbiosis, and recognize the key role of the world’s most abundant biological entities in the metazoan immune system. Materials and Methods Bacterial Strains, Phage Stocks, TC Cell Lines, and Growth Conditions. E. coli 1024 strain was used for all E. coli experiments and was grown in LB (10 g tryptone, 5 g yeast extract, 10 g NaCl, in 1 L dH2O) at 37 °C overnight. E. coli amber-suppressor strain SupD strain CR63 was used as a host for amber mutant phage and grown as above. Bacteriophage T4 was used at ∼109 pfu·mL–1 . Hoc– T4 phage were kindly supplied by Prof. Venigalla Rao (58), The Catholic University of America, Washington, D.C. T3 am10 Ig– amber mutant phage were kindly supplied by Prof. Ian J. Molineux (70), University of Texas, Austin, TX. T4 replication-negative 43– (amE4332: DNA polymerase) 44– (amN82: subunit of polymerase clamp holder) amber mutant phage were kindly supplied by Prof. Kenneth Kreuzer (71), Duke University School of Medicine, Durham, NC. The human tumorigenic colon epithelial cell line, T84, was obtained from the American Type Culture Collection (ATCC) and cultured in DMEM/F12-K media with 5% FBS and 100 μg·mL–1 penicillin–streptomycin (PS). The human tumorigenic lung epithelial cell line A549 was kindly sup- plied by Prof. Kelly Doran, San Diego State University, San Diego, CA and cultured in F12-K media with 10% FBS, 100 μg·mL–1 PS. The human tumor- igenic liver epithelial cell line Huh7 was kindly supplied by Prof. Roland Wolkowicz, San Diego State University, San Diego, CA, and cultured in F12-K media with 10% FBS, 100 μg·mL–1 PS. All TC cell lines initially were grown in 50 mL Primaria Tissue Culture Flasks (Becton Dickinson) at 37 °C and 5% CO2. Phage Adherence to Mucus-Associated Macromolecules. LB agar plates were coated with 1 mL of 1% (wt/vol) of one of the following in 1× PBS: type III porcine stomach mucin, DNA from salmon testes, or BSA (all three from Sigma–Aldrich) and then allowed to dry. Stocks of hoc+ and hoc– T4 phage (109 pfu·mL−1 ) were serially diluted to 1 × 10−7 and 1 × 10−8 per milliliter in LB, and a 5-mL aliquot of each dilution was washed across the plates for 30 min at 37 °C on an orbital shaker. After the phage suspensions were dec- anted from the plates, the plates were shaken twice to remove excess liquid and dried. Each plate then was layered with 1 mL of overnight E. coli culture (109 mL–1 ) in 3 mL of molten top agar and incubated overnight at 37 °C. The number of adherent phage was calculated from the number of plaque- forming units observed multiplied by the initial phage dilution. To de- termine whether mucus macromolecules directly affected phage infectivity, hoc+ and hoc– T4 phage (109 pfu·mL−1 ) were serially diluted as described above into 1 mL LB solutions containing 1% (wt/vol) mucin, DNA, or BSA. After incubation for 30 min at 37 °C, the phage suspensions were combined with E. coli top agar as described above and layered over uncoated agar plates (Fig. S5). Phage Treatment of TC Cells. TC cells were washed twice with 5 mL of serum- free media to remove residual antibiotics, layered with 2 mL of serum-free media containing T4 phage (107 or 109 mL–1 ), and incubated at 37 °C and 5% CO2 for 30 min. Cells then were washed five times with 5 mL of serum-free media to remove nonadherent phage. Phage Adherence to TC Cells. TC cells were treated with phage (109 mL–1 ; see above), then scraped from plates using Corning Cell Scrapers (Sigma– Aldrich). Adherent phage were counted by epifluorescence microscopy as described above. Bacterial Adherence to TC Cells With/Without Phage Pretreatment. TC cells with or without pretreatment with T4 phage (107 mL–1 ) were layered with 2 mL serum-free media containing E. coli (107 mL–1 ), incubated at 37 °C and 5% CO2 for 4 h, and then washed five times with 5 mL serum-free medium to remove nonadherent phage and bacteria. Cells were scraped from plates, and ad- herent phage and bacteria were counted by either epifluorescence micros- copy, as described above, or colony-forming and plaque-forming units. Then, 100 μL of a relevant dilution was spread onto an agar plate and incubated overnight at 37 °C, and the number of adherent bacteria was calculated from the colony-forming units observed multiplied by the initial dilution. Plaque- forming units were counted by a top agar assay as described above. TC Cell Death from Bacterial Infection. Mucus-producing A549 and MUC– A549 TC cells were grown to confluence. T4 phage were cleaned using Amicon 50-kDa centrifugal filters (Millipore) and saline magnesium buffer (SM) to remove bacterial lysis products. Cells, with or without T4 phage pre- treatment (107 mL–1 ), were incubated with E. coli (107 mL–1 ) overnight. Af- terward, TC cells were recovered from the plates by trypsin/EDTA solution (Invitrogen). Cells were pelleted by centrifugation and resuspended in 1× PBS. Dead cells were identified by staining with 1 mg/mL of propidium io- dide (Invitrogen). Samples then were analyzed on a FACSCanto II flow cytometer (BD Biosciences) with excitation at 488 nm and emission detected through a 670 long pass filter. The forward scatter threshold was set at 5,000, and a total of 10,000 events were collected for each sample. Mucin Competition Assay. Mucus-producing A549 TC cells were grown to confluence. Hoc+ and hoc– T4 phage (109 mL–1 ) were diluted into mucin solutions ranging between 0% and 5% (wt/vol) in 1× PBS then washed over TC cells for 30 min at 37 °C and 5% CO2. Cells were washed five times with 5 mL serum-free media to remove nonadherent phage, scraped from plates, and adherent phage were quantified as described above. Graphing and Statistics. Graphing and statistical analyses were performed using GraphPad Prism 6 (GraphPad Software). All error bars represent 5–95% confidence intervals. The midline represents the median and the mean for box plots and bar plots, respectively. Fig. 5. The BAM model. (1) Mucus is produced and se- creted by the underlying epithelium. (2) Phage bind vari- able glycan residues displayed on mucin glycoproteins via variable capsid proteins (e.g., Ig-like domains). (3) Phage adherence creates an antimicrobial layer that reduces bac- terial attachment to and colonization of the mucus, which in turn lessens epithelial cell death. (4) Mucus-adherent phage are more likely to encounter bacterial hosts, thus are under positive selection for capsid proteins that enable them to remain in the mucus layer. (5) Continual sloughing of the outer mucus provides a dynamic mucosal environment. Barr et al. PNAS | June 25, 2013 | vol. 110 | no. 26 | 10775 MICROBIOLOGYSEECOMMENTARY
  • 6. ACKNOWLEDGMENTS. This work was supported by National Institutes of Health (NIH) Grants R01: GM095384, GM073898, and R21: AI094534 from the National Institute of General Medical Sciences. The authors thank the Protein-Glycan Interaction Resource at Emory University School of Medicine, Atlanta, GA (funded by NIH Grant GM98791), for support of the glycan microarray analyses. The authors acknowledge the San Diego State Univer- sity (SDSU) Flow Cytometry Core Facility and the SDSU Electron Microscopy Facility for assistance with sample analysis. 1. Bäckhed F, Ley RE, Sonnenburg JL, Peterson DA, Gordon JI (2005) Host-bacterial mutualism in the human intestine. Science 307(5717):1915–1920. 2. Dethlefsen L, McFall-Ngai M, Relman DA (2007) An ecological and evolutionary perspective on human-microbe mutualism and disease. Nature 449(7164):811–818. 3. Clay K, Holah J (1999) Fungal endophyte symbiosis and plant diversity in successional fields. Science 285(5434):1742–1745. 4. Douglas AE (1989) Mycetocyte symbiosis in insects. Biol Rev Camb Philos Soc 64(4): 409–434. 5. Hooper LV, Midtvedt T, Gordon JI (2002) How host-microbial interactions shape the nutrient environment of the mammalian intestine. Annu Rev Nutr 22(1):283–307. 6. Hosokawa T, Koga R, Kikuchi Y, Meng X-Y, Fukatsu T (2010) Wolbachia as a bacter- iocyte-associated nutritional mutualist. Proc Natl Acad Sci USA 107(2):769–774. 7. Nyholm SV, McFall-Ngai MJ (2004) The winnowing: Establishing the squid-vibrio symbiosis. Nat Rev Microbiol 2(8):632–642. 8. Ruby EG (1996) Lessons from a cooperative, bacterial-animal association: The Vibrio fischeri-Euprymna scolopes light organ symbiosis. Annu Rev Microbiol 50(1):591–624. 9. Currie CR, Scott JA, Summerbell RC, Malloch D (1999) Fungus-growing ants use an- tibiotic-producing bacteria to control garden parasites. Nature 398(6729):701–704. 10. Kaltenpoth M, Göttler W, Herzner G, Strohm E (2005) Symbiotic bacteria protect wasp larvae from fungal infestation. Curr Biol 15(5):475–479. 11. Martens EC, Chiang HC, Gordon JI (2008) Mucosal glycan foraging enhances fitness and transmission of a saccharolytic human gut bacterial symbiont. Cell Host Microbe 4(5):447–457. 12. Sonnenburg JL, et al. (2005) Glycan foraging in vivo by an intestine-adapted bacterial symbiont. Science 307(5717):1955–1959. 13. Berry D, et al. (2013) Host-compound foraging by intestinal microbiota revealed by single-cell stable isotope probing. Proc Natl Acad Sci USA 110(12):4720–4725. 14. Sachs JL, Skophammer RG, Regus JU (2011) Evolutionary transitions in bacterial symbiosis. Proc Natl Acad Sci USA 108(Suppl 2):10800–10807. 15. Stouthamer R, Breeuwer JA, Hurst GD (1999) Wolbachia pipientis: Microbial manip- ulator of arthropod reproduction. Annu Rev Microbiol 53(1):71–102. 16. Chow J, Lee SM, Shen Y, Khosravi A, Mazmanian SK (2010) Host-bacterial symbiosis in health and disease. Adv Immunol 107:243–274. 17. Cone RA (2009) Barrier properties of mucus. Adv Drug Deliv Rev 61(2):75–85. 18. Linden SK, Sutton P, Karlsson NG, Korolik V, McGuckin MA (2008) Mucins in the mucosal barrier to infection. Mucosal Immunol 1(3):183–197. 19. Clunes MT, Boucher RC (2007) Cystic fibrosis: The mechanisms of pathogenesis of an inherited lung disorder. Drug Discov Today Dis Mech 4(2):63–72. 20. Strugala V, Allen A, Dettmar PW, Pearson JP (2003) Colonic mucin: Methods of measuring mucus thickness. Proc Nutr Soc 62(1):237–243. 21. Garren M, Azam F (2012) Corals shed bacteria as a potential mechanism of resilience to organic matter enrichment. ISME J 6(6):1159–1165. 22. Button B, et al. (2012) A periciliary brush promotes the lung health by separating the mucus layer from airway epithelia. Science 337(6097):937–941. 23. Poulsen LK, et al. (1994) Spatial distribution of Escherichia coli in the mouse large intestine inferred from rRNA in situ hybridization. Infect Immun 62(11):5191–5194. 24. Phalipon A, et al. (2002) Secretory component: A new role in secretory IgA-mediated immune exclusion in vivo. Immunity 17(1):107–115. 25. Raj PA, Dentino AR (2002) Current status of defensins and their role in innate and adaptive immunity. FEMS Microbiol Lett 206(1):9–18. 26. Vaishnava S, et al. (2011) The antibacterial lectin RegIII{gamma} promotes the spatial segregation of microbiota and host in the intestine. Sci STKE 334(6053):255. 27. Schluter J, Foster KR (2012) The evolution of mutualism in gut microbiota via host epithelial selection. PLoS Biol 10(11):e1001424. 28. Hooper LV, Xu J, Falk PG, Midtvedt T, Gordon JI (1999) A molecular sensor that allows a gut commensal to control its nutrient foundation in a competitive ecosystem. Proc Natl Acad Sci USA 96(17):9833–9838. 29. Gerken TA (2004) Kinetic modeling confirms the biosynthesis of mucin core 1 (β-Gal(1-3) α-GalNAc-O-Ser/Thr) O-glycan structures are modulated by neighboring glycosylation effects. Biochemistry 43(14):4137–4142. 30. Jentoft N (1990) Why are proteins O-glycosylated? Trends Biochem Sci 15(8):291–294. 31. Schulz BL, et al. (2007) Glycosylation of sputum mucins is altered in cystic fibrosis patients. Glycobiology 17(7):698–712. 32. Rodriguez-Brito B, et al. (2010) Viral and microbial community dynamics in four aquatic environments. ISME J 4(6):739–751. 33. Thingstad T, Lignell R (1997) Theoretical models for the control of bacterial growth rate, abundance, diversity and carbon demand. Aquat Microb Ecol 13:19–27. 34. Groisman EA, Ochman H (1993) Cognate gene clusters govern invasion of host epi- thelial cells by Salmonella typhimurium and Shigella flexneri. EMBO J 12(10): 3779–3787. 35. Johansen BK, Wasteson Y, Granum PE, Brynestad S (2001) Mosaic structure of Shiga- toxin-2-encoding phages isolated from Escherichia coli O157:H7 indicates frequent gene exchange between lambdoid phage genomes. Microbiology 147(Pt 7): 1929–1936. 36. Willner D, et al. (2011) Metagenomic detection of phage-encoded platelet-binding factors in the human oral cavity. Proc Natl Acad Sci USA 108(Suppl 1):4547–4553. 37. Duerkop BA, Clements CV, Rollins D, Rodrigues JLM, Hooper LV (2012) A composite bacteriophage alters colonization by an intestinal commensal bacterium. Proc Natl Acad Sci USA 109(43):17621–17626. 38. Furuse K, et al. (1983) Bacteriophage distribution in human faeces: Continuous survey of healthy subjects and patients with internal and leukaemic diseases. J Gen Virol 64(Pt 9):2039–2043. 39. Weinbauer MG (2004) Ecology of prokaryotic viruses. FEMS Microbiol Rev 28(2): 127–181. 40. Clokie MR, Millard AD, Letarov AV, Heaphy S (2011) Phages in nature. Bacteriophage 1(1):31–45. 41. Moran NA, Degnan PH, Santos SR, Dunbar HE, Ochman H (2005) The players in a mutualistic symbiosis: Insects, bacteria, viruses, and virulence genes. Proc Natl Acad Sci USA 102(47):16919–16926. 42. Oliver KM, Degnan PH, Hunter MS, Moran NA (2009) Bacteriophages encode factors required for protection in a symbiotic mutualism. Science 325(5943):992–994. 43. Roossinck MJ (2011) The good viruses: Viral mutualistic symbioses. Nat Rev Microbiol 9(2):99–108. 44. Minot S, Grunberg S, Wu GD, Lewis JD, Bushman FD (2012) Hypervariable loci in the human gut virome. Proc Natl Acad Sci USA 109(10):3962–3966. 45. McMahon SA, et al. (2005) The C-type lectin fold as an evolutionary solution for massive sequence variation. Nat Struct Mol Biol 12(10):886–892. 46. Medhekar B, Miller JF (2007) Diversity-generating retroelements. Curr Opin Microbiol 10(4):388–395. 47. Halaby DM, Mornon JPE (1998) The immunoglobulin superfamily: an insight on its tissular, species, and functional diversity. J Mol Evol 46(4):389–400. 48. Fraser JS, Yu Z, Maxwell KL, Davidson AR (2006) Ig-like domains on bacteriophages: A tale of promiscuity and deceit. J Mol Biol 359(2):496–507. 49. Fraser JS, Maxwell KL, Davidson AR (2007) Immunoglobulin-like domains on bacte- riophage: Weapons of modest damage? Curr Opin Microbiol 10(4):382–387. 50. Fuhrman JA (1999) Marine viruses and their biogeochemical and ecological effects. Nature 399(6736):541–548. 51. Danovaro R, Serresi M (2000) Viral density and virus-to-bacterium ratio in deep-sea sediments of the Eastern Mediterranean. Appl Environ Microbiol 66(5):1857–1861. 52. Breitbart M, et al. (2002) Genomic analysis of uncultured marine viral communities. Proc Natl Acad Sci USA 99(22):14250–14255. 53. Hong DH, Petrovics G, Anderson WB, Forstner J, Forstner G (1999) Induction of mucin gene expression in human colonic cell lines by PMA is dependent on PKC-e. Am J Physiol 277(5 Pt 1):G1041–G1047. 54. Forstner G, Zhang Y, McCool D, Forstner J (1993) Mucin secretion by T84 cells: Stim- ulation by PKC, Ca2+, and a protein kinase activated by Ca2+ ionophore. Am J Physiol 264(6 Pt 1):G1096–G1102. 55. Alemka A, et al. (2010) Probiotic colonization of the adherent mucus layer of HT29MTXE12 cells attenuates Campylobacter jejuni virulence properties. Infect Im- mun 78(6):2812–2822. 56. Lieleg O, Vladescu I, Ribbeck K (2010) Characterization of particle translocation through mucin hydrogels. Biophys J 98(9):1782–1789. 57. Sathaliyawala T, et al. (2010) Functional analysis of the highly antigenic outer capsid protein, Hoc, a virus decoration protein from T4-like bacteriophages. Mol Microbiol 77(2):444–455. 58. Fokine A, et al. (2011) Structure of the three N-terminal immunoglobulin domains of the highly immunogenic outer capsid protein from a T4-like bacteriophage. J Virol 85(16):8141–8148. 59. Reyes A, et al. (2010) Viruses in the faecal microbiota of monozygotic twins and their mothers. Nature 466(7304):334–338. 60. Marhaver KL, Edwards RA, Rohwer F (2008) Viral communities associated with healthy and bleaching corals. Environ Microbiol 10(9):2277–2286. 61. Willner D, et al. (2009) Metagenomic analysis of respiratory tract DNA viral commu- nities in cystic fibrosis and non-cystic fibrosis individuals. PLoS ONE 4(10):e7370. 62. Wegley L, Edwards R, Rodriguez-Brito B, Liu H, Rohwer F (2007) Metagenomic analysis of the microbial community associated with the coral Porites astreoides. Environ Microbiol 9(11):2707–2719. 63. Willner D, et al. (2012) Case studies of the spatial heterogeneity of DNA viruses in the cystic fibrosis lung. Am J Respir Cell Mol Biol 46(2):127–131. 64. Eckburg PB, et al. (2005) Diversity of the human intestinal microbial flora. Science 308(5728):1635–1638. 65. McFall-Ngai M, et al. (2013) Animals in a bacterial world, a new imperative for the life sciences. Proc Natl Acad Sci USA 110(9):3229–3236. 66. Minot S, et al. (2011) The human gut virome: Inter-individual variation and dynamic response to diet. Genome Res 21(10):1616–1625. 67. Lieleg O, Lieleg C, Bloom J, Buck CB, Ribbeck K (2012) Mucin biopolymers as broad- spectrum antiviral agents. Biomacromolecules 13(6):1724–1732. 68. Labrie SJ, Samson JE, Moineau S (2010) Bacteriophage resistance mechanisms. Nat Rev Microbiol 8(5):317–327. 69. Weitz JS, et al. (2013) Phage-bacteria infection networks. Trends Microbiol 21(2):82–91. 70. Condreay JP, Wright SE, Molineux IJ (1989) Nucleotide sequence and complementa- tion studies of the gene 10 region of bacteriophage T3. J Mol Biol 207(3):555–561. 71. Benson KH, Kreuzer KN (1992) Plasmid models for bacteriophage T4 DNA replication: Requirements for fork proteins. J Virol 66(12):6960–6968. 10776 | www.pnas.org/cgi/doi/10.1073/pnas.1305923110 Barr et al.
  • 7. Supporting Information Barr et al. 10.1073/pnas.1305923110 SI Materials and Methods Mucus Sample Collection. Mucus samples were collected directly from the surface of organisms using a syringe, swab, or custom suction device. Environmental samples were collected as close to the mucus sample as possible, typically within 30–50 cm of the mucosal surface. Specific organism details are as follows: Sea anemones were sampled from tidal rock pools at Ocean Beach, San Diego, CA. Surface mucus was collected by a custom suction device that dislodges surface mucus using a stream of 0.02 μm filtered seawater; the environmentalg sample was sea- water collected directly above the anemone. Hard corals were sampled at the Birch Aquarium, San Diego, CA. Surface mucus was collected by syringe directly from coral surfaces; environmental water samples were collected directly above the coral. The polychaete, along with surrounding water, was collected at Scripps Pier, San Diego, CA, and carefully transported to the laboratory in a container. Surface mucus was collected via syringe, and the environmental sample was seawater from the container. Teleost surface mucus was sampled at the Birch Aquarium, San Diego, CA. Surface mucus was collected by custom suction de- vice; the environmental water sample was collected directly above the teleost within its tank. Human gum mucus was sampled from a male subject with no current pathology/disease. Surface mucus was collected by swab; the environmental sample was expectorated saliva. Consent was obtained for all human samples collected under the San Diego State University Institutional Review Board #2121. Mouse intestine was excised from a healthy mouse. Surface mucus was collected by cutting open the intestine, washing the mucosal surface with 0.02 μm-filtered PBS buffer, then scraping off the mucus layer; the environmental sample was collected from the intestinal lumen directly adjacent to the sampled mucosa. All animal experiments were approved by the Committee on the Use and Care of Animals (SDSU, APF #10-08-024D) and performed using accepted veterinary standards. Bacterial and Phage Counts from Mucus and Environmental Samples. Samples of mucus and the adjacent environment were collected directly from nine evolutionarily diverse mucosal surfaces (Fig. S1). Samples were transported and maintained on ice until pro- cessed. All samples were fixed overnight in 0.5% glutaraldehyde at 4 °C, then incubated in 6.5 mM DTT at 37 °C for 1 h to assist mucus degradation. A 1–100-μL aliquot was diluted with 2 mL of 0.02 μm SM buffer (100 mM NaCl, 8 mM MgSO4, 50 mM Tris·Cl, in dH2O), briefly mixed, then filtered onto a 0.02-μm Anodisc polycarbonate filter (Whatman). Filters were stained with 10× SYBR Gold, washed, and visualized on a Zeiss epifluorescence microscope. For each sample, 20–30 images were taken for both bacteria and virus-like particles. Images were analyzed using Im- age-Pro Plus 5.1 software (MediaCybernetics). Counts of bac- teria and virus-like particles (referred to as “phage” throughout the text) per milliliter were made as previously described (1). Tissue Culture Cells and Mucus Reduction. Monolayers of various mucus-producing and non–mucus-producing tissue culture (TC) cells were grown to confluence in six-well Multiwell tissue culture plates (Becton Dickinson). (i) Mucus-producing TC cells were exposed to 1 μg/mL of a phorbol ester, phorbol-12-myristate 13-acetate (Sigma–Aldrich) in the culture media overnight to stimulate the mucin secretory response (2). (ii) The mucolytic agent N-acetyl-L-cysteine (NAC; Sigma–Aldrich) was used to chemically remove mucus from A549 TC cells (60 mM NAC in serum-free media for 1 h with agitation) (3). Mucus depletion was confirmed using periodic acid-Schiff–Alcian blue (PAS/AB) (Fig. S2). (iii) A mucus-knockdown (MUC– ) A549 cell line was produced by transduction of A549 cells with GIPZ Lentiviral Human MUC1 shRNA and TRIPZ Inducible Lentiviral Hu- man MUC5AC shRNA as target vectors; an shControl A549 cell line was produced using the GIPZ Nonsilencing Lentiviral shRNA Control as a control vector (Thermo Scientific). Knockdown of mucus production in the MUC– cell line was confirmed by Western blot analysis and PAS/AB (Sigma–Aldrich; Figs. S3 and S4). Transfection and Selection of A549 TC Mucus-Negative Clones. A549 cells were transduced with GIPZ Lentiviral Human MUC1 shRNA and TRIPZ Inducible Lentiviral Human MUC5AC shRNA as target vectors or GIPZ Nonsilencing Lentiviral shRNA Control as a control vector (Thermo Scientific) according to the manufacturer’s instructions. Viral particles were produced by transfecting HEK 293T cells with a combination of plasmids containing 2 μg of packaging vector pCMV d8.2 containing the gag-pol proteins of HIV-1, 3 μg of the transfer vectors containing the LTRs of HIV-1, 3 μg of vesicular stomatitis virus envelope glycoprotein plasmid, and 1.5 μg of pci-HIV-1 viral protein R accessory protein plasmid. Growth medium was replaced 24 h post transfection, and viral supernatant was collected 48 and 72 h after transfection and then filtered through 0.45-μm polytetrafluoroethylene (PTFE) filters (Pall Corporation). A549 cells were seeded into six-well culture plates and grown to ∼70% confluence. The cells then were washed twice in serum-free media before being incubated overnight in 1 mL of growth media and 1 mL of virus-containing media con- taining 5 μg/mL of polybrene. The transduced cells subsequently were washed and cultured for 24 h in complete medium with 2 μg/mL of doxycycline to induce expression of shRNA. Cells then were sorted using a BD FACSAria (BD Biosciences) at the San Diego State University Flow Cytometry Facility. A 100-μm nozzle was used at a sheath pressure of 20 psi. Excitation source was a 488-nm laser and emissions were collected using 530/30 band pass (BP) and 585/42 BP filters for GFP and red fluores- cent protein, respectively. Between 10,000 and 300,000 cells were sorted for each population and collected in a 5-mL tube with 250 μL of FBS. The efficiency of MUC1-MUC5AC knockdown was confirmed by Western blot analysis and PAS/AB (Sigma–Al- drich), a stain for mucus-like substances. Western Blot Analysis. Expression of MUC1 (a membrane-tethered mucin) and MUC5AC (a secreted gel-forming mucin) was ex- amined by Western blot analysis. MUC–, shControl, and native A549 cell lines were grown to confluence and then lysed using radio-immunoprecipitation assay (RIPA) buffer (Thermo Sci- entific) containing 2 mM Na3VO4, 100 mM NaF, 10 mM sodium pyrophosphate, 1 mM PMSF, and protease inhibitor mixture (Millipore). Aliquots containing 50 μg of total protein were subjected to SDS/PAGE, and the protein bands were transferred to a polyvinylidene difluoride membrane (Sigma–Aldrich). Mem- branes were blocked with 5% nonfat milk in Tris-buffered saline containing 0.05% Tween 20 at room temperature for 1 h and then incubated overnight at 4 °C with mouse anti-human MUC1 monoclonal antibody (clone S.854.6; Thermo Scientific), mouse anti-human MUC5AC monoclonal antibody (clone 2H7; Sigma– Aldrich), and rabbit anti-human GAPDH antibody (Millipore). After three washes, membranes were incubated for 1 h at room Barr et al. www.pnas.org/cgi/content/short/1305923110 1 of 6
  • 8. temperature with anti-mouse or anti-rabbit IgG horseradish peroxidase-linked, species-specific, whole antibody (Fisher Sci- entific). Immunoreactivity was visualized and band intensity was normalized to the constitutively expressed GAPDH protein. Multiple Particle Tracking. Assays were performed in plastic well chambers mounted on glass slides that had been coated with poly (dimethylsiloxane) to prevent phage adherence. Five microliters of 109 mL–1 SYBR Gold-labeled phage suspensions was added to 50 μL of 1% (wt/vol) mucin solution in 1× PBS buffer. Trajec- tories of fluorescently labeled phage were observed using a DeltaVision Spectris Model DV4 deconvolution microscope (Applied Precision) equipped with a 100× Olympus PlanApo 1.4 lens. Movies were captured using SoftWoRx 5.0.0 (Applied Precision): 100-ms temporal resolution for 30 s, 10 analyses per sample, n > 100 particle trajectories per analysis. Trajectories were analyzed with the ParticleTracker plugin for ImageJ (4). The coordinates of phage particle centroids were transformed into time-averaged mean square displacements: <Δr2 (τ)> = <Δx2 + Δy2 >, from which effective diffusivities (<Deff>) were calculated; Deff = <Δr2 (τ)>/(4 τ) (5, 6). Glycan Microarray. Phage binding to glycans was assayed using printed mammalian glycan microarrays (version 5.1, Consortium for Functional Glycomics Core) containing 610 glycan targets. Samples of highly antigenic outer capsid protein (hoc+ ) T4 phage, hoc– T4 phage, and buffer controls were applied to separate glycan microarray slides. Each slide received 35 μL of sample, 35 μL of binding buffer (Tris saline with 2 mM Ca2+ , 2 mM Mg2+ , 1% bovine serum albumin (BSA), and 0.05% Tween 20), and a coverslip. Slides first were incubated for 1 h at room temper- ature and washed with binding buffer. Slides then were in- cubated in SYBR Gold fluorescence dye (diluted 1:10,000 in binding buffer) for 1 h under a coverslip at room tempera- ture, washed, dried, and immediately scanned in a PerkinElmer ProScanArray microarray scanner using an excitation wavelength of 488 nm. ImaGene software (BioDiscovery, Inc) was used to quantify fluorescence. Normalized relative fluorescence unit (RFU) values reported are the average (after subtraction of background buffer fluorescence) from six spots for each glycan represented on the array. Phylogenetic Analysis of Ig-Like Domains. The SEED database (www. theseed.org) collection of Ig-like polycystic kidney disease (PKD) protein families (Pfam) (PF00801) and the T4 Hoc sequence were searched against the 124 viral metagenomic datasets contained in the My Metagenome (MyMg) database (http://edwards.sdsu.edu/cgi-bin/ mymgdb/show.cgi) using tBLASTn (PubMed accession numbers: 16336043, 17620602, 19156205, 19816605, 20547834, 17921274, 18441115, 19892985, 19555373, 20573248, 20631792, 21167942.79, 21193730.87, 21219518.96, 21245307.04, 21271095.12, 21296883.2, 21322671.29, 21348459.37, 21374247.45, 21400035.53; MG-Rast IDs: 21167942.79, 21193730.87, 21219518.96, 21245307.04, 21271095.12, 21296883.2, 21322671.29, 21348459.37, 21374247.45, 21400035.53). Sequences with an e value of less than 1e-5 to Ig-like domains were retrieved. ORFs were called from the metagenome reads using Artemis (Wellcome Trust Sanger Institute); their position in the FASTA file is shown in Table S2. ORFs that were 60 bp long with 40% tBLASTn identity to T4 Hoc or a member of the PKD Ig Pfam were retained. The six contigs containing Ig-like hypervariable domains from the published study by Minot et al. (7) were down- loaded from the National Center for Biotechnology Information (NCBI). Identical sequences were collapsed using the Trie clus- tering method implemented in Qiime (8). The resulting unique sequences were mapped to the position-specific scoring matrix for the PKD Ig Pfam (PF00801) using hmmalign (9). The hmmalign trimming function was used; sequences that were dominated by gaps after alignment were removed. A maximum likelihood tree was generated from the aligned unique sequences using FastTree version 2.1.1 SSE3 and viewed in MEGA 5. Environmental data for the metagenomes were obtained from the MyMg database. In a separate analysis, structural homology of these same sequences to a carbohydrate-binding protein (10) was determined using the Phyre2 structural homology prediction pipeline (www.sbg.bio.ic. ac.uk/phyre2/html/help.cgi). 1. Patel A, et al. (2007) Virus and prokaryote enumeration from planktonic aquatic environments by epifluorescence microscopy with SYBR Green I. Nat Protoc 2(2):269–276. 2. Forstner G, Zhang Y, McCool D, Forstner J (1993) Mucin secretion by T84 cells: Stimulation by PKC, Ca2+, and a protein kinase activated by Ca2+ ionophore. Am J Physiol 264(6 Pt 1):G1096–G1102. 3. Alemka A, et al. (2010) Probiotic colonization of the adherent mucus layer of HT29MTXE12 cells attenuates Campylobacter jejuni virulence properties. Infect Immun 78(6):2812–2822. 4. Sbalzarini IF, Koumoutsakos P (2005) Feature point tracking and trajectory analysis for video imaging in cell biology. J Struct Biol 151(2):182–195. 5. Suh J, Dawson M, Hanes J (2005) Real-time multiple-particle tracking: Applications to drug and gene delivery. Adv Drug Deliv Rev 57(1):63–78. 6. Lai SK, et al. (2007) Rapid transport of large polymeric nanoparticles in fresh undiluted human mucus. Proc Natl Acad Sci USA 104(5):1482–1487. 7. Minot S, Grunberg S, Wu GD, Lewis JD, Bushman FD (2012) Hypervariable loci in the human gut virome. Proc Natl Acad Sci USA 109(10):3962–3966. 8. Caporaso JG, et al. (2010) QIIME allows analysis of high-throughput community sequencing data. Nat Methods 7(5):335–336. 9. Eddy SR (2009) A new generation of homology search tools based on probabilistic inference. Genome Inform 23(1):205–211. 10. Najmudin S, et al. (2006) Xyloglucan is recognized by carbohydrate-binding modules that interact with β-glucan chains. J Biol Chem 281(13):8815–8828. Barr et al. www.pnas.org/cgi/content/short/1305923110 2 of 6
  • 9. Fig. S1. Epifluorescence counts of phage and bacteria from diverse environments and mucosa. (Left to right) Invertebrates: Actiniaria sp., Acropora sp., Echinopora sp., Oxypora sp., Capnela sp., and Phyllodoce sp. Vertebrates: Paralichthys sp., Homo sapiens, and Mus musculis. Error bars represent ±SD with n > 25. Fig. S2. Mucolytic treatment of mucus-producing A549 cells. Mucus removal from A549 lung epithelial cells by NAC treatment was assessed by PAS/AB stain, which stains mucus-like substances pink/purple. Scale bars represent 100 μm. Fig. S3. Growth and mucus production of A549 and siRNA knockdown cell lines. Shown are mucus-producing A549 lung epithelial TC cells, mucus-producing nonsense shRNA control A549 cell line (shControl), and non–mucus-producing MUC1 and MUC5AC shRNA knockdown A549 cell line (MUC– ) after 2 and 4 d in culture. Mucus production was assessed on day 5 by PAS/AB stain, which stains mucus-like substances pink/purple. Scale bars represent 100 μm. Barr et al. www.pnas.org/cgi/content/short/1305923110 3 of 6
  • 10. Fig. S4. Western blot analysis of MUC1 and MUC5AC in total cell lysates of A549 lung epithelial cell knockdowns. Lysates of confluent cell layers were separated by SDS/PAGE and then immunoblotted with anti-MUC1 and anti-MUC5AC antibodies. Shown are the MUC– knockdown cell line, the nonsilencing shControl control cell line, and native A549 cells. GAPDH was used as an intracellular protein control. Fig. S5. Surface-free control for the assay of phage adherence to mucus-associated macromolecules. Both hoc+ and hoc– T4 phage (109 pfu·mL−1 ) were serially diluted to 1 × 10−7 and 1 × 10−8 , and then incubated in 1% (wt/vol) solutions of mucin, DNA, or protein in 1 mL LB for 30 min at 37 °C. Each incubation mixture was then mixed with Escherichia coli top agar and layered over plain agar plates. Resulting plaque-forming unit (PFU) counts showed that infectivity of hoc+ and hoc– T4 phage was not significantly altered in the presence of the macromolecules used in the phage adherence assays (mucin, DNA, and BSA protein). Fig. S6. Adherence of Ig+ and Ig– T3 phage to mucin. Phage adherence assays to mucin-coated agar plates were performed as described in SI Materials and Methods, except that the Ig+ and Ig– T3 phage (1011 pfu·mL−1 ) were serially diluted to 1 × 10−9 and 1 × 10−10 pfu·ml−1 . The resultant PFU counts of adherent phage showed that Ig+ T3 phage adhered to mucin-coated agar plates significantly more than to the plain agar control plates (n = 6, t = 4.443, **P = 0.0012, unpaired t test), whereas there was no significant increase in adherence for the Ig– T3 phage. ns, not significant. Barr et al. www.pnas.org/cgi/content/short/1305923110 4 of 6
  • 11. Table S1. Phyre2 structural homology of Ig-like proteins encoded by viral metagenomes Phyre2 analysis ORF, bp No. Environment PDB ID Confidence, % Identity, % Start Stop Length Database Sequence identifier 1 Sputum 2C26-A 99.8 29 29 277 248 MyMg d7c74d66ea493c0c1fca41f718d22125_16271_279 2 Sputum 2C26-A 99.9 36 280 555 275 MyMg d7c74d66ea493c0c1fca41f718d22125_64742_278 3 Sputum 2C26-A 99.7 32 837 1,066 229 MyMg 03ca0e6ad90102ab264cf521ed58209e_112029_232 4 Sputum 2C26-A 99.7 26 2,951 3,184 233 MyMg 03ca0e6ad90102ab264cf521ed58209e_52923_234 5 Sputum 2C26-A 99.7 32 3,417 3,654 237 MyMg 2da3ea31d1b30a11b0f080a5b91b9df2_256934_240 6 Sputum 2C26-A 99.7 31 4,189 4,416 227 MyMg 03ca0e6ad90102ab264cf521ed58209e_134692_228 7 Freshwater 2C26-A 99.8 29 3,935 4,188 253 MyMg 6f8c77f72920950139dc6b3520cf86b7_69438_254 8 Oral 2C26-A 99.8 29 2,437 2,715 278 MyMg f44d959b723905a049b0334f19668e5c_48421_279 9 Sputum 2C26-A 99.8 30 2,716 2,949 233 MyMg 875f11dbd745609c9d3e12c5b3b5636a_66237_235 10 Sputum 2C26-A 99.9 30 16,236 16,491 255 MyMg 88f61453e560016a0e2a238351d7292b_109758_257 11 Sputum 2C26-A 99.9 29 15,508 15,766 258 MyMg ebdf0605a03616ab168eddf68ca506e1_121095_259 12 Marine* 1E07-A 99.8 13 3 800 797 MyMg 05d2b5884d248d570fe8a2c0d390c97c_3911_800 13 Human feces 2C26-A 88.8 23 1 246 245 MyMg f58fff76a10b642883986ee0e1a30514_10941_246 14 Human feces 2C26-A 94.7 31 248 484 236 MyMg f58fff76a10b642883986ee0e1a30514_22104_238 15 Human feces 2C26-A 86.8 22 720 960 240 MyMg f58fff76a10b642883986ee0e1a30514_10554_243 16 Human feces 2C26-A 86 29 1,445 1,682 237 MyMg f58fff76a10b642883986ee0e1a30514_32210_238 17 Human feces 2C26-A 85.5 32 2,400 2,637 237 MyMg f58fff76a10b642883986ee0e1a30514_22379_238 18 Human feces 2C26-A 90.1 26 3,122 3,361 239 MyMg f58fff76a10b642883986ee0e1a30514_6620_240 19 Human feces 2C26-A 98.1 11 574 NCBI gi 377806168 gb AFB75876.1 20 Human feces 2C26-A 99 15 429 NCBI gi 377806248 gb AFB75953.1 21 Human feces 2C26-A 99.6 18 529 NCBI gi 377806350 gb AFB76049.1 All Ig-like domain homologs shown in Fig. 4A displayed high structural homology with a promiscuous carbohydrate-binding domain [Protein Data Bank (PDB) 2C26]. The Hoc homolog from a marine sample (no. 12, column 1) displayed high structural homology with 1E07-A as well as several other immune proteins. bp, base pairs. *Hoc homolog. Barr et al. www.pnas.org/cgi/content/short/1305923110 5 of 6
  • 12. Table S2. Glycan microarray analysis of T4 and hoc– phage displayed in Fig. 4A Glycan no. Structure Linkage T4 RFU T4 %CV hoc– RFU hoc– %CV 609 GlcNAcb1-3Fuca -N(CH3)-O-(CH2)2-NH2 4,921 3 394 16 610 Galb1-3GalNAcb1-4(Neu5Aca2-8Neu5Aca2-8Neu5Aca2-3) Galb1-4Glcb -N(CH3)-O-(CH2)2-NH2 4,685 11 472 13 573 Neu5Aca2-8Neu5Aca2-3Galb1-3GalNAcb1-4(Neu5Aca2-3) Galb1-4Glc -N(CH3)-O-(CH2)2-NH2 4,161 1 316 11 608 Neu5Aca2-6Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-2Mana1- 6(Neu5Aca2-6Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1- 2Mana1-3)Manb1-4GlcNAcb1-4GlcNAcb Asparagine 4,685 4 497 17 145 Galb1-3GalNAcb1-4Galb1-4Glcb CH2CH2CH2NH2 5,823 3 565 7 195 Glca1-4Glcb CH2CH2CH2NH2 5,845 1 544 3 514 GalNAcb1-4(6S)GlcNAc CH2CH2CH2NH2 4,518 6 568 17 287 Neu5Gca CH2CH2CH2NH2 4,182 3 356 9 119 Gala1-4(Fuca1-2)Galb1-4GlcNAcb CH2CH2CH2NH2 4,157 3 331 8 336 GalNAca1-3(Fuca1-2)Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1- 3Galb1-4GlcNAcb CH2CH2NH2 7,670 31 472 51 217 Manb1-4GlcNAcb CH2CH2NH2 4,766 3 647 24 144 Galb1-3GalNAcb1-4(Neu5Aca2-3)Galb1-4Glcb CH2CH2NH2 4,797 4 589 26 517 Galb1-4(6P)GlcNAcb CH2CH2NH2 4,751 11 526 20 218 Neu5Aca2-3Galb1-4GlcNAcb1-3Galb1-4(Fuca1-3)GlcNAcb CH2CH2NH2 4,737 3 322 10 334 GalNAcb1-3Gala1-4Galb1-4GlcNAcb1-3Galb1-4Glcb CH2CH2NH2 4,261 4 499 8 143 Galb1-3GalNAcb1-3Gala1-4Galb1-4Glcb CH2CH2NH2 4,236 1 451 13 581 Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1- 3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-2Mana1-6(Galb1- 4GlcNAcb1-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1- 3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-2Mana1-3) Manb1-4GlcNAcb1-4(Fuca1-6)GlcNAcb EN or NK 5,465 6 262 7 360 Fuca1-2Galb1-3GlcNAcb1-2Mana1-6(Fuca1-2Galb1- 3GlcNAcb1-2Mana1-3)Manb1-4GlcNAcb1-4GlcNAcb GENR 4,914 2 482 12 588 Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1- 3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-6(Galb1- 4GlcNAcb1-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1- 3Galb1-4GlcNAcb1-3Galb1-4GlcNAb1-2)Mana1-6(Galb1- 4GlcNAcb1-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1- 3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-2Mana1-3) Manb1-4GlcNAcb1-4(Fuca1-6)GlcNAcb KVANKT 3,926 2 154 5 470 Glca1-4Glca1-4Glca1-4Glcb NHCOCH2NH 7,521 10 1,288 17 516 (4S)GalNAcb NHCOCH2NH 6,148 2 464 3 359 KDNa2-3Galb1-3GalNAca Threonine (O-linked glycan) 7,484 17 1,080 24 471 Neu5Aca2-3Galb1-4GlcNAcb1-6(Neu5Aca2-3Galb1- 4GlcNAcb1-3)GalNAca Threonine (O-linked glycan) 5,877 6 585 2 491 Neu5Aca2-3Galb1-3GlcNAcb1-6GalNAca Threonine (O-linked glycan) 4,755 2 394 8 596 Neu5Aca2-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-6 (Neu5Aca2-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-3) GalNAca Threonine (O-linked glycan) 4,703 14 242 7 595 GlcNAcb1-3Galb1-4GlcNAcb1-6(GlcNAcb1-3Galb1- 4GlcNAcb1-3)GalNAca Threonine (O-linked glycan) 4,653 1 392 9 605 Neu5Aca2-6Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-6 (Neu5Aca2-6Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1-3) GalNAca Threonine (O-linked glycan) 4,078 1 410 10 480 Neu5Aca2-6Galb1-4GlcNAcb1-6GalNAca Threonine (O-linked glycan) 4,143 5 301 20 592 Neu5Aca2-3Galb1-4GlcNAcb1-3Galb1-4GlcNAcb1- 3GalNAca Threonine (O-linked glycan) 4,200 11 266 11 “Glycan no.” indicates the glycan ID number used on the Consortium for Functional Glycomics Version 5.1 microarray. “Linkage” denotes the chemical linkage joining the glycan to the macromolecule. Bold threonine linkages represent O-linked glycan residues likely to be associated with mucin glycoproteins. Barr et al. www.pnas.org/cgi/content/short/1305923110 6 of 6