ArticlePDF Available

Mutational Analyses Reveal Overall Topology and Functional Regions of NilB, a Bacterial Outer Membrane Protein Required for Host Association in a Model of Animal-Microbe Mutualism

American Society for Microbiology
Journal of Bacteriology
Authors:

Abstract and Figures

The gammaproteobacterium Xenorhabdus nematophila is a mutualistic symbiont that colonizes the intestine of the nematode Steinernema carpocapsae. nilB (nematode intestine localization) is essential for X. nematophila colonization of nematodes and is predicted to encode an integral outer membrane beta-barrel protein, but evidence supporting this prediction has not been reported. The function of NilB is not known, but when expressed with two other factors encoded by nilA and nilC, it confers upon noncognate Xenorhabdus spp. the ability to colonize S. carpocapsae nematodes. We present evidence that NilB is a surface-exposed outer membrane protein whose expression is repressed by NilR and growth in nutrient-rich medium. Bioinformatic analyses reveal that NilB is the only characterized member of a family of proteins distinguished by N-terminal region tetratricopeptide repeats (TPR) and a conserved C-terminal domain of unknown function (DUF560). Members of this family occur in diverse bacteria and are prevalent in the genomes of mucosal pathogens. Insertion and deletion mutational analyses support a beta-barrel structure model with an N-terminal globular domain, 14 transmembrane strands, and seven extracellular surface loops and reveal critical roles for the globular domain and surface loop 6 in nematode colonization. Epifluorescence microscopy of these mutants demonstrates that NilB is necessary at early stages of colonization. These findings are an important step in understanding the function of NilB and, by extension, its homologs in mucosal pathogens.
NilB protein expression is highest in stationary phase, repressed by NilR, and elevated by growth in defined medium. (A) Growth curves of X. nematophila nilR⁺ and nilR mutant strains in defined medium (DM) and LB. White boxes and white circles represent the nilR mutant in LB and defined medium, respectively. Gray boxes and gray circles represent nilR⁺ in LB and defined medium, respectively. The growth curves of both strains are identical in LB. In defined medium, the curves are similar for the two strains, except that the nilR mutant reaches a higher A600 than the nilR⁺ strain in stationary phase. (B) Western blots of NilB protein expression. nilR⁺ and nilR mutant strains were grown in LB or DM as indicated. Crude cell extracts were taken at 6, 12, 24, 36, 54, and 77 h, total protein content was normalized, and samples were loaded onto 12% SDS-PAGE gels for anti-FLAG Western blotting. The arrows indicate NilB, and NF contains extracts from the equivalent X. nematophila strains lacking the FLAG tag. (C) Quantitation of the blots in panel B. White boxes and white circles represent levels of NilB detected in the nilR mutant strain when grown over time in LB and defined medium, respectively. Gray boxes and gray circles represent levels of NilB detected in the nilR⁺ strain when grown over time in LB and defined medium, respectively. Levels of NilB in LB are similar for the two strains; low levels of NilB are detectable at 6, 12, and 24 h (log phase), and maximal expression is at 54 and 77 h (stationary phase). Levels of NilB in the nilR⁺ strain grown in defined medium are similar to levels of NilB in LB, except that the amount of NilB expressed is constant over time. Maximum NilB levels are produced in the nilR mutant strain grown in defined medium; low levels of NilB are detectable at 6, 12, and 24 h (log phase), with an increase in expression at 54 and 77 h (stationary phase). Quantitation was done with ImageQuant IQTL 7.0.
… 
NilB is a membrane-localized protein and is surface exposed. (A) X. nematophila cells expressing NilB-FLAG26 in a nilR mutant background (HGB1200) were pelleted, resuspended, and sonicated. Soluble and membrane fractions were separated by ultracentrifugation. The crude extract (lane 1), cleared lysate (lane 2), soluble fraction (lane 3), and membrane fraction (lane 4) were electrophoresed on a 12% SDS-polyacrylamide gel and transferred to a membrane. NilB protein (arrowhead) was detected using anti-FLAG antibody. (B) Whole X. nematophila nilR mutant cells expressing NilB-FLAG26 (HGB1200) were incubated with 38.4 μg/ml proteinase K (PK) (final concentration) for 1 to 30 min at 37°C. At various time points (indicated in minutes above each lane), samples were removed and reactions were stopped with the addition of 5 mM PMSF (final concentration) and kept on ice. SDS-PAGE loading dye was added to all the reaction mixtures, which were heated at 95°C for 10 min and electrophoresed on 12% SDS-PAGE gels. NilB (top panel) (anti-NilB) or NilC (bottom panel) (anti-NilC) was detected in Western blots using anti-FLAG or anti-NilC (15) antibody, respectively. Full-length NilB-FLAG26 protein (black arrowheads, No PK lanes) was present in the no-proteinase K control but not in cells lacking nilB (Δnil) (HGB1251). This species decreases in intensity after the addition of proteinase K, while smaller-molecular-mass proteins (black arrowheads, lanes 1 and 10) appear. M, Thermo Scientific Pierce prestained protein molecular mass marker (sizes in kDa are noted on the left).
… 
FLAG tag insertions distinguish whether NilB adopts a predicted 18-TM or 14-TM strand structure. Insertions were made in three N-terminal TM strands (before aa 68, aa 102, and aa 121) predicted only by the 18-TM strand model (D) and in two TM strands predicted by both models (before aa 379 and at aa 421); all mutations were made in nilR mutant cells. The effect of each mutation on nematode colonization (A) and protein production (B) was assessed by nematode sonication and immunoblotting with anti-FLAG, respectively. (A) Insertions before aa 68 or aa 121 reduce nematode colonization below wild-type levels but not to the level observed when nilB is absent (ETn7). Insertions before aa 102, aa 379, and aa 421 cause severe nematode colonization defects. *, P < 0.01, relative to ETn7 values (except with the aa 102 strain, for which P was 0.049). IJ, infective juvenile. (B) When FLAG insertions are in TM strands predicted only by the 18-TM strand model, NilB protein is detectable in whole cells by anti-FLAG Western blotting, but when insertions are in TM strands predicted by both models, NilB is not detectable. (C and D) Schematic of NilB's structure and the relative positions of FLAG tag insertions in the 14-TM (C) and 18-TM (D) strand models. Circles indicate that an insertion mutant colonized better than (white circles) or similarly to (shaded circles) a strain that lacked SR1. NF, Tn7/SR1 without a FLAG26 insertion in NilB; ETn7, Tn7 lacking SR1; FLAG26, Tn7/SR1 with a FLAG26 insertion in NilB (no deletions).
… 
This content is subject to copyright. Terms and conditions apply.
Mutational Analyses Reveal Overall Topology and Functional Regions
of NilB, a Bacterial Outer Membrane Protein Required for Host
Association in a Model of Animal-Microbe Mutualism
Archna Bhasin,
a
John M. Chaston,
b
and Heidi Goodrich-Blair
b
Department of Bacteriology, University of Wisconsin—Madison, Madison, Wisconsin, USA,
b
and Biology Department, Valdosta State University, Valdosta, Georgia, USA
a
The gammaproteobacterium Xenorhabdus nematophila is a mutualistic symbiont that colonizes the intestine of the nematode
Steinernema carpocapsae.nilB (nematode intestine localization) is essential for X. nematophila colonization of nematodes and
is predicted to encode an integral outer membrane beta-barrel protein, but evidence supporting this prediction has not been
reported. The function of NilB is not known, but when expressed with two other factors encoded by nilA and nilC, it confers
upon noncognate Xenorhabdus spp. the ability to colonize S. carpocapsae nematodes. We present evidence that NilB is a surface-
exposed outer membrane protein whose expression is repressed by NilR and growth in nutrient-rich medium. Bioinformatic
analyses reveal that NilB is the only characterized member of a family of proteins distinguished by N-terminal region tetratrico-
peptide repeats (TPR) and a conserved C-terminal domain of unknown function (DUF560). Members of this family occur in di-
verse bacteria and are prevalent in the genomes of mucosal pathogens. Insertion and deletion mutational analyses support a be-
ta-barrel structure model with an N-terminal globular domain, 14 transmembrane strands, and seven extracellular surface loops
and reveal critical roles for the globular domain and surface loop 6 in nematode colonization. Epifluorescence microscopy of
these mutants demonstrates that NilB is necessary at early stages of colonization. These findings are an important step in under-
standing the function of NilB and, by extension, its homologs in mucosal pathogens.
The mutualism between the gammaproteobacterium Xenorh-
abdus nematophila and the nematode Steinernema carpocapsae
is a model to understand the molecular mechanisms underlying
the initiation and maintenance of long-term beneficial animal-
microbe associations (13). The soil-dwelling infective third-stage
juvenile nematode (the infective juvenile) is colonized by a mon-
oculture of X. nematophila bacteria in a structure, known as the
receptacle (previously the vesicle), located at the anterior of the
nematode intestine (10, 21, 42, 58). The infective juvenile infects
the blood of insect hosts, releases the bacteria from the receptacle,
and together with the symbiont rapidly kills the insect host (22, 26,
32). Within the insect cadaver, the nematode reinitiates develop-
ment and reproduces through adult and juvenile stages that lack
the receptacle structure characteristic of the infective juvenile
stage. When nutrients derived from the insect cadaver are exhausted,
progeny juveniles reassociate with the symbiotic bacteria and develop
into the infective juvenile stage, which exits the insect cadaver to
search for a new host (reviewed in reference 50). X. nematophila’s
contributions to mutualism are that it is essential for both insect host
killing and S. carpocapsae nematode growth. In turn, X. nematophila
uses its nematode host as a vector into insects and as a protected
environment between infections (39, 43, 44).
Bacterial colonization of the infective juvenile nematode host
takes place in at least three stages: (i) entry of a few bacterial cells
into the receptacle, (ii) outgrowth of the initiating population to
30 to 200 bacterial cells, entirely filling the receptacle (37), and (iii)
persistence of the colonizing bacteria until the bacteria are re-
leased into an insect host by a recovering infective juvenile (21,
25). The process by which X. nematophila cells initially gain entry
into the infective juvenile receptacle and the point at which other
species are excluded have not been elucidated. Transposon screens
for colonization-defective X. nematophila mutants revealed genes
involved in both entry and outgrowth (30, 37). Outgrowth mu-
tants, including those with defects in methionine and threonine
biosynthesis, localize to the receptacle but do not replicate to fill it
(21, 38) and ultimately disappear from within the nematode pop-
ulation (38). In contrast, mutants that are defective in early stages
of colonization are never apparent within the receptacle, either
because they fail to enter or because they fail to survive immedi-
ately after entering. Mutations in any of three genes, nilA,-B, and
-C(nematode intestine localization), carried on a 3.5-kb locus
(called SR1), cause entry defects. nilA mutants colonize nema-
todes to approximately 2.5% of the level of the wild type, while
nilB and nilC mutants do not colonize above the level of detection
(0.005 CFU/infective juvenile) (30). Microscopic analysis demon-
strated that the partial colonization phenotype of a nilA mutant
occurs because a few nematodes are partially colonized, while the
rest are devoid of bacteria. This indicates that nilA mutants are
defective primarily in entry but also have outgrowth defects (17).
However, it remains unclear whether nilA mutations cause a de-
fect in adherence to specific tissues or in survival during coloniza-
tion initiation.
An early observation regarding the nil locus was that it is absent
from other species of Xenorhabdus (17, 30), suggesting that it may
be specifically required for colonization of S. carpocapsae nema-
todes. Indeed, when nilA,-B, and -Cwere introduced into the
Received 16 December 2011 Accepted 20 January 2012
Published ahead of print 27 January 2012
Address correspondence to Heidi Goodrich-Blair, hgblair@bact.wisc.edu.
A.B. and J.M.C. contributed equally to this work.
Supplemental material for this article may be found at http://jb.asm.org/.
Copyright © 2012, American Society for Microbiology. All Rights Reserved.
doi:10.1128/JB.06711-11
0021-9193/12/$12.00 Journal of Bacteriology p. 1763–1776 jb.asm.org 1763
chromosome of Xenorhabdus poinarii or Xenorhabdus bovienii,
these strains colonized the receptacles of S. carpocapsae nema-
todes, but the corresponding control strains lacking the nil locus
did not (17). Therefore, the nil locus not only is necessary for S.
carpocapsae nematode colonization by X. nematophila (30) but
also functions as a specificity determinant.
Online databases lack nil gene homologs of known function,
and understanding nil gene function therefore requires funda-
mental genetic and biochemical characterization. NilC is a 282-
amino-acid (aa) outer membrane-localized lipoprotein that is
predominantly oriented toward the periplasm (based on its resis-
tance to degradation by proteinase K in whole cells) (15). NilA is
predicted to be a small 90-aa inner membrane protein (30), and
NilB is predicted to encode a 466-aa protein that adopts a beta-
barrel structure in the outer membrane (Fig. 1) (30). The pre-
dicted localization of NilB to the outer membrane suggests that it
may interact directly with host surfaces or molecules, for example,
by binding to nematode surface structures or by transporting a
nutrient provided to the bacteria from the host.
To investigate the topology and function of NilB, we assessed
the colonization phenotypes of nilB mutants created by FLAG tag
insertions and domain deletions. These data reveal certain do-
mains critical for NilB function in colonization. Also, coupled
with in vivo imaging of colonizing bacteria within the nematode
host, our data suggest that NilB functions solely at entry into the
nematode receptacle. Unlike with another specificity factor, nilA,
we found no indication that nilB is necessary for outgrowth or
persistence within the nematode receptacle.
MATERIALS AND METHODS
Strains, plasmids, media, and growth conditions. X. nematophila strains
(Table 1) were grown aerobically at 30°C on LB agar plates containing
0.1% pyruvate or in the dark in LB broth in the presence of appropriate
antibiotics, unless stated otherwise. Final antibiotic concentrations used
were ampicillin at 100
g/ml, chloramphenicol at 30 (for E. coli)or15(X.
nematophila)
g/ml, erythromycin at 200
g/ml, kanamycin at 50
g/ml,
and streptomycin at 25 (E. coli) or 12.5 (X. nematophila)
g/ml. Defined
medium was made as described previously (41) except that agar and
amino acids were omitted and 0.68 g MgCl
2
(H
2
O)
6
liter
1
and 0.91%
glucose were added after the autoclaving.
An X. nematophila strain (HGB1251) that lacks nilR (encoding a re-
pressor of the nil locus) as well as the locus carrying nilA,nilB, and nilC,
symbiosis region 1 (SR1), was created by conjugating strain HGB1141 (E.
coli S17
pir containing pnilR16::Sm [16]) into the recipient strain
HGB777 (SR1-7::Kan) (15). To facilitate in vivo visualization of strains
FIG 1 NilB topology predicted by the PROFtmb (9) hidden Markov model. The model presented is similar to the model predicted by B2TMR-HMM (see Fig.
S2 in the supplemental material and reference 36). While the PROFtmb model predicts that the 130 N-terminal amino acids are periplasmic, TMBB-PRED
predicts that 100 of these amino acids form four additional TM domains (5, 6) (drawn in the periplasm in the same manner as TM domains and surface loops;
the black box denotes this region). Only mature NilB is pictured (as predicted by SignalP 3.0 [8]), and numbering is according to that of immature NilB. Black
arrows, FLAG tag insertions (before indicated amino acid); gray text, deletions (underlining distinguishes adjacent deletions); circles, the predicted TPR domain;
SL, surface loop; 18TMSL, surface loop unique to the 18-TM strand model.
Bhasin et al.
1764 jb.asm.org Journal of Bacteriology
colonizing nematodes, the green fluorescent protein (GFP) expression
plasmid JMC001 was created. Plasmid pURR25 was digested with SacI
and BglII (Promega) to isolate the 2.5-kb cat-lacZp-GFP fragment.
pECM20 (37) was also digested with SacI and BglII to remove the aphAp-
GFP region. The resulting 5-kb pECM20 fragment was ligated to the
2.5-kb cat-lacZp-GFP fragment using T4 DNA ligase (Promega) and elec-
troporated into E. coli S17-1
pir cells. A colony with green coloration,
visible by the naked eye, was frozen and used as a donor to transfer
pECM20-lacZp-GFP into HGB1251 to create X. nematophila
nilR16::Sm SR1-7::Km lacZp-GFP (HGB1429) or into HGB777 to cre-
ate X. nematophila SR1-7::Km lacZp-GFP (HGB1430).
NilB-FLAG insertion mutants were created by PCR amplification
(with Platinum PFX DNA polymerase; Invitrogen) of the entire pTn7/SR1
plasmid (15) with primers flanking the mutation site (see Table S1 in the
supplemental material). One primer contained an nilB-complementary se-
quence, the FLAG sequence (nucleotides, GATTACAAGGATGACGACGA
TAAG; amino acids, DYKDDDDK) and the NheI restriction site, while
the other primer contained an nilB-complementary sequence and the
TABLE 1 Xenorhabdus strains used in this study
Strain type Strain Parent strain Relevant characteristic(s)
Source or
reference
Insertionless
a
HGB007 NA
b
Sequenced ATCC 19061 X. nematophila wild type ATCC
HGB777 HGB007 (nilA-nilC)7::Km (SR1) 15
HGB1251 HGB777 (nilA-nilC)7::Km (SR1) nilR16::Sm This study
HGB1430 HGB777 (nilA-nilC)7::Km lacZp-GFP (from pJMC001) This study
HGB1429 HGB1251 (nilA-nilC)7::Km nilR16::Sm lacZp-GFP
(from pJMC001)
This study
Non-GFP-expressing nilB FLAG
insertion mutants
HGB778 HGB777 SR1 15
HGB1256 HGB777 SR1/nilB26-FLAG aa026 (nilB-FLAG26) This study
HGB1200 HGB1251 SR1/nilB26-FLAG aa026 (nilB-FLAG26) This study
HGB1805 HGB1251 SR1/nilB27-FLAG aa068 This study
HGB1201 HGB1251 SR1/nilB28-FLAG aa085 This study
HGB1806 HGB1251 SR1/nilB29-FLAG aa102 This study
HGB1807 HGB1251 SR1/nilB30-FLAG aa121 This study
HGB1203 HGB1251 SR1/nilB31-FLAG aa137 This study
HGB1254 HGB1251 SR1/nilB32-FLAG aa186 This study
HGB1204 HGB1251 SR1/nilB33-FLAG aa233 This study
HGB1253 HGB1251 SR1/nilB34-FLAG aa273 This study
HGB1206 HGB1251 SR1/nilB35-FLAG aa322 This study
HGB1260 HGB1251 SR1/nilB36-FLAG aa357 This study
HGB1808 HGB1251 SR1/nilB37-FLAG aa379 This study
HGB1207 HGB1251 SR1/nilB38-FLAG aa399 This study
HGB1809 HGB1251 SR1/nilB39-FLAG aa421 This study
HGB1208 HGB1251 SR1/nilB40-FLAG aa451 This study
HGB1324 HGB1251 SR1/nilB41-FLAG aa466 This study
HGB1255 HGB1251 SR1 This study
GFP-expressing nilB FLAG
insertion mutants
HGB1502 HGB1430 SR1/nilB34-FLAG aa270 This study
HGB1495 HGB1429 Empty This study
HGB1496 HGB1429 SR1 (wild type nilB without FLAG) This study
HGB1484 HGB1429 SR1/nilB26-FLAG aa026 (nilB-FLAG26) This study
HGB1487 HGB1429 SR1/nilB32-FLAG aa186 This study
HGB1489 HGB1429 SR1/nilB34-FLAG aa273 This study
HGB1490 HGB1429 SR1/nilB35-FLAG aa322 This study
HGB1492 HGB1429 SR1/nilB38-FLAG aa399 This study
GFP-expressing nilB-FLAG26
deletion mutants
HGB1460 HGB1429 SR1/nilB42-026–042 This study
HGB1461 HGB1429 SR1/nilB43-043–059 This study
HGB1462 HGB1429 SR1/nilB44-026–059 This study
HGB1463 HGB1429 SR1/nilB45-076–091 This study
HGB1464 HGB1429 SR1/nilB46-092–108 This study
HGB1465 HGB1429 SR1/nilB47-026–108 This study
HGB1510 HGB1429 SR1/nilB48-109–156 This study
HGB1511 HGB1429 SR1/nilB49-026–156 This study
HGB1466 HGB1429 SR1/nilB50-266–279 This study
HGB1467 HGB1429 SR1/nilB51-292–298 This study
HGB1468 HGB1429 SR1/nilB52-314–320 This study
HGB1469 HGB1429 SR1/nilB53-352–357 This study
HGB1470 HGB1429 SR1/nilB54-389-410 This study
a
All strains except HGB007, HGB777, HGB1251, HGB1430, and HGB1429 had att Tn7insertions.
b
NA, not applicable.
Topology and Functional Regions of X. nematophila NilB
April 2012 Volume 194 Number 7 jb.asm.org 1765
NheI restriction site. PCR products were cleaned using the QIAquick PCR
purification kit (Qiagen), digested with the restriction enzyme NheI at
37°C, ligated with T4 DNA ligase (Promega), and electroporated into E.
coli S17-1
pir cells.
With one exception, nilB in-frame deletion constructs were created by
PCR by amplifying the entire pAB001 plasmid (containing nilB26-
FLAG26), with primers oriented in the direction opposite to that of the
desired deletion site and containing 9 bp of the sequence upstream and 18
bp of the sequence downstream of the desired deletion site (see Table S1 in
the supplemental material), using Takara PrimeStar polymerase per the
manufacturer’s instructions and annealing at 55°C for 5 to 10 s. PCR
products were cleaned using a Zyppy PCR purification kit (Zymo Re-
search), digested with DpnI, reverse dialyzed on a 0.025-
m VSWP mem-
brane (Millipore) for 15 min, and electroporated into E. coli S17-1
pir
cells. The exception was that the nilB construct with a deletion of aa 26 to
156 (nilB26 –156 construct) could not be created using this approach, so
the NheI restriction site approach utilized for creating FLAG insertions
was adopted, resulting in insertion of duplicate NheI sites flanking the
FLAG26 insertion (AS-DYKDDDDK-AS). Enzymes used for creating de-
letions were also used to create the nilB26-156 construct and the con-
structs nilB-FLAG aa068 (in which the FLAG tag is inserted before aa 68),
nilB-FLAG aa102, nilB-FLAG aa121, nilB-FLAG aa379, and nilB-FLAG
aa421. The complete SR1 region of each Tn7-nilB mutant construct (plas-
mids are listed in Table 2) was sequenced to ensure that additional muta-
tions did not arise during PCR. Constructs were incorporated into the att
Tn7site of X. nematophila ATCC 19061 SR1 nilR (HGB1251), SR1
(HGB777), SR1 lacZp-GFP (HGB1430), or SR1 nilR lacZp-GFP
(HGB1429) by triparental mating as described previously (7). Verification
TABLE 2 Plasmids used in this study
Type of construct Plasmid Relevant properties Reference or source
pEVS107 oriR6K mobRP4; mobilizable suicide miniTn7-Erm delivery
vector; Erm
r
Km
r
52
pUX-BF13 Mobilizable Tn7transposition helper plasmid that expresses Tn7
transposase in trans;Ap
r
7
pTn7/SR1 Delivers miniTn7-Erm/SR1; used as a template for creation of
the NilB-FLAG insertion and deletion mutations
15
pJMC001 Plasmid pECM20 with cat-lacZp-GFP from pURR25 cloned into
the SacI and BglII sites, replacing the original gene encoding
GFP on this plasmid
This study
pECM20 GFP plasmid that contains a 614-bp chromosomal sequence of
X. nematophila ATCC 19061 that facilitates plasmid
integration at a site that does not interfere with nematode
colonization
37
pURR25 Tn7PA1/03/04gfpmut3; cat-lacZp-GFP source for pJMC001 D. Lies and D. Newmann
via T. Ciche
pTn7/SR1nilB-FLAG insertion
mutant constructs
pAB001 pTn7/SR1/nilB26-FLAG aa026 (nilB-FLAG26) This study
pAB002 pTn7/SR1/nilB28-FLAG aa085 This study
pAB003 pTn7/SR1/nilB31-FLAG aa137 This study
pAB004 pTn7/SR1/nilB32-FLAG aa186 This study
pAB005 pTn7/SR1/nilB33-FLAG aa233 This study
pAB006 pTn7/SR1/nilB34-FLAG aa273 This study
pAB007 pTn7/SR1/nilB35-FLAG aa322 This study
pAB008 pTn7/SR1/nilB36-FLAG aa357 This study
pAB009 pTn7/SR1/nilB38-FLAG aa399 This study
pAB010 pTn7/SR1/nilB40-FLAG aa451 This study
pAB011 pTn7/SR1/nilB41-FLAG aa466 This study
pJMC016 pTn7/SR1/nilB27-FLAG aa068 This study
pJMC017 pTn7/SR1/nilB29-FLAG aa102 This study
pJMC018 pTn7/SR1/nilB30-FLAG aa121 This study
pJMC019 pTn7/SR1/nilB37-FLAG aa379 This study
pJMC020 pTn7/SR1/nilB39-FLAG aa421 This study
pAB001 nilB-FLAG26 deletion
mutant constructs
pJMC002 pTn7/SR1/nilB42-026–042 This study
pJMC003 pTn7/SR1/nilB43-043–059 This study
pJMC004 pTn7/SR1/nilB44-026–059 This study
pJMC005 pTn7/SR1/nilB45-076–091 This study
pJMC006 pTn7/SR1/nilB46-092–108 This study
pJMC007 pTn7/SR1/nilB47-026–108 This study
pJMC014 pTn7/SR1/nilB48-109–156 This study
pJMC015 pTn7/SR1/nilB49-026–156 This study
pJMC008 pTn7/SR1/nilB50-266-279 This study
pJMC009 pTn7/SR1/nilB51-292–298 This study
pJMC010 pTn7/SR1/nilB52-314–320 This study
pJMC011 pTn7/SR1/nilB53-352–357 This study
pJMC012 pTn7/SR1/nilB54-389–410 This study
Bhasin et al.
1766 jb.asm.org Journal of Bacteriology
of insertion of each Tn7/nilB-FLAG or deletion construct at the att Tn7
site was verified by PCR as described previously (15). Since insertion or
deletion mutations were created in the context of the SR1 genomic region,
the mutant nilB alleles were expressed from the native nilB promoter and
regulatory regions. Constructs were introduced into a strain lacking SR1
so that no SR1 DNA sequence was present in duplicate. Therefore,
HGB777, -1251, -1429, and -1430 complemented with Tn7/SR1 lacking
any modifications to SR1 are considered wild-type controls in this study.
Fractionation, NilB detection by Western blotting, and glycosyla-
tion. Two Western blotting methods were used in this study. In method 1,
used for all experiments except where stated otherwise, strains were grown
as described above prior to harvesting of cell pellets. Pellets were washed
and resuspended in cold 1phosphate-buffered saline (PBS) and soni-
cated on ice using a Sonic Dismembrator with 5 bursts of 3 s each or until
the cells lysed. Lysed cells were spun at 13,000 rpm in a microcentrifuge to
remove cell debris. The total protein concentration of the cleared lysates
was determined by the Bradford assay (Bio-Rad), and equal amounts of
protein were heated for 10 min at 95°C in 1SDS-PAGE loading dye
prior to being loaded onto 12% SDS-PAGE gels. In other cases, the cleared
lysate was spun at 434,513 gin a Beckman TLA 100.2 rotor for 30 min
at 4°C for separation of the membrane and soluble fractions. The mem-
brane fraction was solubilized in 1PBS plus 0.5% Sarkosyl by stirring for
30 min. Fractions were heated for 10 min at 95°C in 1SDS-PAGE load-
ing dye prior to being loaded onto 12% SDS-PAGE gels. After transfer to
polyvinylidene difluoride (PVDF) membranes (Bio-Rad), membranes
were blocked for 1 h with 5% dry milk in 1Tris-buffered saline with
Tween 20 (TBS-T) at room temperature, washed, and incubated with
1:1,000-diluted anti-FLAG antibody (Cell Signaling) in 1TBS-T plus
5% bovine serum albumin (BSA) overnight at 4°C. The next day, mem-
branes were washed, incubated with 1:10,000-diluted horseradish perox-
idase (HRP)-linked anti-rabbit IgG (Cell Signaling) for1hatroom tem-
perature, and washed again prior to detection using the Amersham ECL
Plus Western blotting detection system (GE Healthcare) and the Storm/
Typhoon scanner or the LAS4010 biomolecular imager (ImageQuant)
according to the manufacturers’ instructions.
Western blotting method 2 was used for the experiments from which
results are shown in Fig. 4 and for which data are not shown. In this
method, strains were subcultured from LB at a 1:100 dilution into defined
medium for 24 h prior to harvesting of cell pellets. Pellets were washed
in an equal volume of PBS, concentrated 4-fold in PBS, and normalized to
their optical density at 600 nm (OD
600
) prior to being mixed with 2
SDS-PAGE loading dye, heated to 95°C for 10 min, and run at 100 V on a
7.5% SDS-PAGE gel. Gels were transferred to a PVDF membrane (Bio-
Rad) by a Trans-Blot SD semidry electrophoretic transfer cell (Bio-Rad)
set to transfer at 22 V for at least 15 min. Membranes were blocked over-
night in PBS plus 5% dry milk prior to being sequentially washed and
incubated with a 1:10,000 or 1:5,000 dilution of anti-FLAG antibody
(Sigma F7425) and HRP-conjugated anti-rabbit secondary antibody
(Pierce catalog number 32260) according to the manufacturers’ instruc-
tions. Blots were detected using the ECL Plus Western blotting substrate
(Pierce catalog number 32209) and XAR film (Kodak XAR catalog num-
ber 165-1454). Adobe Photoshop CS3 was used to alter the intensity and
contrast of scanned images of films, with all samples within an individual
panel (including the controls) manipulated in the same way.
The glycosylation state of NilB was determined using the GelCode
glycoprotein staining kit (Pierce). X. nematophila crude protein was ex-
tracted from all nilB-FLAG insertion mutants in the SR1 nilR back-
ground and electrophoresed as described above, except that transfer was
onto a nitrocellulose membrane (Bio-Rad) rather than a PVDF mem-
brane. Positive (horseradish peroxidase) and negative (soybean trypsin
inhibitor) control proteins were also loaded on the protein gel. Upon
transfer, the membrane was washed with acetic acid, oxidizing solution,
staining reagent, and reducing agent according to the manufacturer’s pro-
tocol.
Proteinase K assay. Strain HGB1200 (SR1 nilR Tn7/SR1nilB-
FLAG26) was grown at 30°C overnight in LB broth. Cells were pelleted
and then washed and resuspended in 50 mM Tris-HCl, pH 8, 1 mM
CaCl
2
. The resuspended whole cells were split into two aliquots. One
aliquot was treated with 0.1% Triton X-100 (final concentration) to break
open the cells. Both aliquots were treated with 38.4
g/ml proteinase K
(final concentration; Fermentas) for 0 to 30 min at 37°C. At various time
points, samples were removed and reactions were stopped with the addi-
tion of 5 mM PMSF (Sigma catalog number P-7626) and kept on ice.
SDS-PAGE loading dye was added to all the reaction mixtures, which were
heated at 95°C for 10 min, sonicated for8stoshear the DNA, and elec-
trophoresed on 12% SDS-PAGE gels. NilB and NilC (as a control) were
detected in Western blots using anti-FLAG and anti-NilC antibody (15),
respectively.
Growth curves and time courses of NilB protein levels. Strains
HGB1200 (SR1 nilR Tn7/SR1 nilB-FLAG26) and HGB1256 (SR1
Tn7/SR1 nilB-FLAG26) were streaked from 80°C onto LB pyruvate
plates with antibiotics. Colonies from these plates were inoculated into
liquid LB with antibiotics and grown at 30°C overnight in a shaking water
bath. Overnight cultures were subcultured 1:100 into 1 ml LB or defined
medium (no antibiotics) in the wells of a 24-well plate. Absorbance at 600
nm was measured at 1-h intervals using the SpectraMax M5
e
plate reader
(Molecular Devices). At 6, 12, 24, 36, 54, and 77 h, 0.5 ml of each strain in
either LB or defined medium was removed, microcentrifuged to remove
the medium, washed with 1cold PBS, resuspended in 1PBS, and
sonicated at approximately 14 W for 10 s. Sample protein concentrations
were assessed by the Bradford assay (Bio-Rad), and all samples were nor-
malized to each other. Samples at later time points were diluted in more
PBS than earlier time points (PBS resuspension volumes ranged from 37.5
to 150
l). After resuspension, 4SDS-PAGE loading dye was added and
the samples were stored at 20°C until Western blotting was performed.
Quantitation of blots was performed using the ImageQuant TL program
(GE Healthcare).
Nematode-bacterium cocultivations. The effect of FLAG insertions
in NilB on nematode colonization was tested by growing each mutant on
lipid agar (54) and inoculating greater than 200 sterile nematode eggs (55)
or surface-sterilized axenic infective juveniles onto each plate. The latter
were prepared by inoculating sterile nematode eggs onto a GFP-express-
ing strain of X. nematophila (HGB1430 SR1::Km lacZp-GFP) defective
in nematode colonization. Infective juveniles were harvested from White
traps and stored at 25°C until they were surface sterilized for 3 min in 0.5%
NaOCl (30), and then they were inoculated onto bacterial lawns. Infective
juveniles that developed on the FLAG-tagged mutant bacteria were har-
vested in White traps (56), and bacterial colonization of the nematode
host was assessed either by dilution plating surface-sterilized nematode
sonicates to calculate average numbers of CFU per infective juvenile (15)
or by analyzing nematode carriage of GFP-labeled bacteria by epifluores-
cence microscopy as described previously (37). For microscopy, nema-
todes were analyzed at a 100magnification and scored as either colo-
nized (bacteria were localized to the receptacle) or uncolonized (no
bacteria were observed within the receptacle). If greater than 30 in 500
nematodes in the population were colonized, nematodes were scored in-
dividually. If fewer than 30 nematodes in 500 were colonized, samples of
nematode populations were scanned for colonized nematodes in 24-well
plates. The number of colonized nematodes in each well was determined,
and the entire nematode population in each well was enumerated by di-
lution counting. Organisms in additional wells were counted until either
30 colonized nematodes or all nematode progeny from the inoculated
plate were counted (whichever occurred first). Data are reported as the
percentages of total nematodes that carried GFP-labeled bacteria.
For data presented in Fig. 5, sterile nematode eggs were added to
each lawn of bacteria and nematode sonications were performed on
samples from at least five different experiments performed at different
times. For Fig. 4, surface-sterilized axenic infective juveniles were
added to each lawn of bacteria. Nematode sonications were performed
Topology and Functional Regions of X. nematophila NilB
April 2012 Volume 194 Number 7 jb.asm.org 1767
on samples from two experiments performed at different times. For
the first experiment, two otherwise isogenic exconjugants were inoc-
ulated, and three replicate experiments were performed for each (six
replicate experiments, total) (data not shown). Colonization results
for each exconjugant pair were the same, so for the second experiment
(data shown), the results of two replicate samples for each of three
isogenic cultures of a single representative exconjugant were assessed.
The results of replicate samples were averaged to give biological-rep-
licate (i.e., isogenic strain) values, and averages and standard devia-
tions among biological-replicate values for the second experiment are
presented in Fig. 4. A linear mixed-effects model was applied to each
experiment individually, and similar statistical trends were observed
between the two experiments. For all other colonization tests, surface-
sterilized axenic infective juveniles were used to inoculate bacterial
lawns. Assays were performed as two experiments at different times
with two replicate samples for each of three biological replicates of
each strain. For graphical presentation, the results of replicate samples
were averaged to give biological-replicate values, and averages and
standard deviations among biological-replicate values for a single rep-
resentative experiment are presented. A linear mixed-effects model or
a generalized linear mixed-effects model was applied to raw values or
data after combination of the results of both experiments.
Statistical analysis. Statistical differences in Fig. 5 were assessed in
Microsoft Excel using a two-tailed heteroscedastic Student ttest. Other-
wise, differences between sonicated samples were performed using a lin-
ear mixed-effects model, and differences between samples analyzed by
microscopy were assessed using a generalized linear mixed-effects model.
Statistical tests were performed in R (48). All collected data points were
included in the analyses unless otherwise indicated, and the Pvalue cutoffs
are at least 0.05.
RESULTS
nilB represents a family of proteins found within symbionts of
animals. In 2002, it was reported that, while X. nematophila nilB is
absent from other Xenorhabdus species tested to date, it has 18
to 25% full-length sequence similarity (BLASTp) to proteins of
unknown function encoded by Neisseria meningitidis,Pasteurella
multocida,Haemophilus influenzae, and Moraxella catarrhalis (2,
30). A more recent BLASTX analysis probing the NCBI database
(December 2011) revealed additional proteins with 18 to 26%
sequence identity to NilB (cutoff E value of 10
5
) (see Table S2 in
the supplemental material). All BLASTX hits were encoded by
harmful or beneficial animal symbionts, including many Neisseria
and Actinobacillus spp., and isolates revealed by the Human Mi-
crobiome Project. Two species of Psychrobacter, normally found
in cold-adapted environments, appear on this list, and both are
also host associated. Psychrobacter sp. strain PRwf-1 is found on
fish skin and gills and was isolated in Puerto Rico (see http://ge-
nome.jgi-psf.org/psy_p/psy_p.home.html); Psychrobacter sp.
strain 1510 was isolated from blood as part of the Human Micro-
biome Project (NCBI accession number AFHU00000000). In the
enterobacteriaceae, which are well represented with sequenced
isolates, NilB homologs have been identified in only three species:
X. nematophila,Salmonella enterica subsp. arizonae, a reptile
pathogen, and Edwardsiella tarda, a vertebrate symbiont and fish
pathogen.
NilB has not been found in any Xenorhabdus species tested to
date other than X. nematophila (17, 30). We also did not observe
any Xenorhabdus proteins in our nilB BLASTX analysis. However,
we did identify a protein in the X. nematophila genome (14) with
sequence similarity to NilB homologs (this protein was identified
by reciprocally comparing low-scoring NilB homologs against the
X. nematophila genome [cutoff E value of 10
5
]). This X. nema-
tophila protein (XNC1_0074, YP_003710424) is found in the
same genomic context in X. nematophila,X. bovienii,Photorhab-
dus luminescens, and Photorhabdus asymbiotica (close relatives of
Xenorhabdus spp.) (see Fig. S1 in the supplemental material) (12,
26). We distinguish NilB homologs from NilB-like proteins based
on their identification by BLASTX: NilB homologs are classified as
those retrieved by nilB BLASTX (E value of 10
5
; 109 hits);
NilB-like proteins are those identified using XNC1_0074 BLASTX
(E value of 10
5
; 326 hits). There is substantial overlap between
these protein groups, since 80 proteins are identified by both nilB
and XNC1_0074 BLASTX queries.
NilB is predicted to be an outer membrane protein with 7 or
9 surface-exposed loops and two conserved domains. Prediction
programs indicate that NilB homologs and NilB-like proteins lo-
calize to the outer membrane and adopt beta-barrel structures (5,
6, 9) (Fig. 1 and data not shown). The N-terminal 25 amino acids
of X. nematophila NilB are predicted to encode a sec secretion
signal sequence that is cleaved prior to outer membrane localiza-
tion (8, 30). Of the online programs available for predicting the
structure of beta-barrels, hidden Markov models (HMM) have
the highest accuracy (4). Two HMM programs (PROFtmb [9] and
B2TMR-HMM [36]) each predict that NilB adopts an outer mem-
brane structure with 14 transmembrane (TM) strands, 7 surface
loops (SLs), and a 138- or 140-aa (postcleavage) N-terminal
periplasmic domain, respectively (Fig. 1). These structures differ
from the model predicted by the HMM program PRED-TMBB
(6), which predicts that the 350 C-terminal amino acids of NilB
comprise 18 TM strands and 9 SLs and that the N terminus is a
38-aa periplasmic domain (Fig. 1). All three programs predict
similar locations for the seven extracellular SLs (Fig. 1), but the
PROFtmb and B2TMR-HMM models predict that the two
additional SLs in the PRED-TMBB model make up a longer N-
terminal periplasmic domain.
The family of NilB/NilB-like proteins shares two conserved
domains, neither of which is unique to this family. At least one
copy of a 34-aa tetratricopeptide repeat (TPR)-like domain, pre-
dicted to fold into two packed
-helices (1, 3, 11), is found within
the N-terminal one-third of the proteins. TPR motifs are widely
distributed and occur within proteins involved in diverse func-
tions, including transcription, cell cycle control, and protein
transport complexes (19). The TPR domain is thought to mediate
protein-protein interactions (11). In NilB, a TPR domain was not
identified by a TPR prediction program (TPRpred; http://toolkit
.tuebingen.mpg.de/tprpred) using standard maximum cutoffs
(1e4), but a single TPR domain (aa 112 to 145) was identified by
lowering the stringency of the cutoff value (Pvalue 1.2e03).
Several NilB homologs tested (e.g., those of Aggregatibacter aphro-
philus,Actinobacillus minor, and Haemophilus parasuis) had mul-
tiple consecutive TPR domains, while some NilB homologs had
only a single copy (e.g., that of Mannheimia haemolytica). The X.
nematophila NilB-like protein XNC1_0074 contains four consec-
utive predicted TPR domains (data not shown). The second con-
served domain, DUF560/pfam04575 (35), spans the C-terminal
two-thirds of NilB homologs, essentially constituting the barrel
domain predicted by the 14-TM strand model. This domain of
unknown function is present in over 900 proteins, including all
NilB homologs evaluated. In some cases, the TPR and DUF560
domains overlap. No structure or function has been assigned to
the DUF560 domain.
Bhasin et al.
1768 jb.asm.org Journal of Bacteriology
NilB protein levels are elevated by nilR mutation and nutri-
ent-limiting conditions. The current working model for Xenorh-
abdus-Steinernema symbiosis is that nutrient limitation in the in-
sect cadaver is the signal for Xenorhabdus to colonize the
nematode intestine. This idea, combined with previous data dem-
onstrating that nilC transcription is upregulated during stationary
phase, led us to investigate whether NilB production is also growth
phase dependent (15). To monitor NilB protein levels, we at-
tempted to generate polyclonal antibodies against NilB but were
unsuccessful. Instead, we engineered a FLAG (DYKDDDDK)-
tagged version of NilB, in which the tag amino acids were inserted
immediately preceding amino acid 26 (predicted to be the first
amino acid in the mature protein after signal sequence cleavage).
FLAG-tagged nilB was provided in trans at the att Tn7site of an X.
nematophila strain lacking wild-type nilB. Since under laboratory
growth conditions NilR suppresses transcription of the nil genes
(16), we also tested NilB production in the presence and absence
of nilR. Neither the nilR mutation (16) nor the FLAG insertion in
NilB (see below; also, data not shown) negatively affects nematode
colonization (i.e., NilB function). To determine the effect of the
nilR mutation and growth phase on NilB protein levels, we mon-
itored growth and NilB-FLAG26 levels of nilR
(HGB1256) and
nilR mutant (HGB1200) strains in LB broth and defined medium.
nilR
and nilR mutant strains have similar growth curves in LB. In
defined medium, both strains grow similarly until a plateau is
reached at 25h(A
600
0.22), after which nilR mutant cells
continue to grow at a higher rate and ultimately reach a higher
A
600
(0.7) than do nilR
cells (A
600
,0.5) (Fig. 2A). The differ-
ence in the maximum levels of absorbance of the two strains was
not due to differences in viability (both strains yielded 1 10
9
CFU/ml at 77 h), differences in absorbance of the culture super-
natant, or gross differences in cell size (monitored by microscopy)
(data not shown).
At various times during growth in defined medium and LB,
crude cell extracts were taken and evaluated for NilB protein pro-
duction by Western blotting (Fig. 2C). As expected based on pre-
vious analyses of nilB-lacZ transcriptional fusions (16), quantifi-
cation of the Western blot revealed that NilB protein levels were
higher in the nilR mutant strain than in the nilR
strain (Fig. 2B).
Maximum levels of NilB production are found in the nilR mutant
strain grown in defined medium, with low-but-detectable
amounts of NilB produced at 6, 12, and 24 h (log phase), an inter-
mediate amount at 36 h, and maximal amounts at 54 and 77 h
(stationary phase) of growth in defined medium. In contrast, the
nilR
strain exhibited low and constant levels of NilB at all time
points of growth in defined medium (Fig. 2B and C). In LB me-
dium, NilB levels in both nilR
and nilR mutant strains increased
in stationary phase but never reached the levels observed in the
nilR mutant strain during growth in defined medium (Fig. 2B and
C). In some samples, another FLAG-reactive band was apparent.
This protein is not a variant of NilB, since it is present in the
no-FLAG (NF) control (Fig. 2C). The data shown in Fig. 2 indicate
FIG 2 NilB protein expression is highest in stationary phase, repressed by
NilR, and elevated by growth in defined medium. (A) Growth curves of X.
nematophila nilR
and nilR mutant strains in defined medium (DM) and LB.
White boxes and white circles represent the nilR mutant in LB and defined
medium, respectively. Gray boxes and gray circles represent nilR
in LB and
defined medium, respectively. The growth curves of both strains are identical
in LB. In defined medium, the curves are similar for the two strains, except that
the nilR mutant reaches a higher A
600
than the nilR
strain in stationary phase.
(B) Western blots of NilB protein expression. nilR
and nilR mutant strains
were grown in LB or DM as indicated. Crude cell extracts were taken at 6,
12, 24, 36, 54, and 77 h, total protein content was normalized, and samples
were loaded onto 12% SDS-PAGE gels for anti-FLAG Western blotting.
The arrows indicate NilB, and NF contains extracts from the equivalent X.
nematophila strains lacking the FLAG tag. (C) Quantitation of the blots in
panel B. White boxes and white circles represent levels of NilB detected in
the nilR mutant strain when grown over time in LB and defined medium,
respectively. Gray boxes and gray circles represent levels of NilB detected in
the nilR
strain when grown over time in LB and defined medium, respec-
tively. Levels of NilB in LB are similar for the two strains; low levels of
NilB are detectable at 6, 12, and 24 h (log phase), and maximal expression is at
54 and 77 h (stationary phase). Levels of NilB in the nilR
strain grown in
defined medium are similar to levels of NilB in LB, except that the amount of
NilB expressed is constant over time. Maximum NilB levels are produced in
the nilR mutant strain grown in defined medium; low levels of NilB are detect-
able at 6, 12, and 24 h (log phase), with an increase in expression at 54 and 77
h (stationary phase). Quantitation was done with ImageQuant IQTL 7.0.
Topology and Functional Regions of X. nematophila NilB
April 2012 Volume 194 Number 7 jb.asm.org 1769
that under laboratory growth conditions, NilB levels are highest
under stationary-phase and nutrient-limited conditions. Further-
more, NilR negatively impacts NilB levels in all growth phases,
perhaps indicating that the signal for NilR derepression is not
present in laboratory growth medium. Also, nilB is repressed by
other factors (e.g., Lrp [18]) that may contribute to the effects that
we observed here.
NilB is a surface-exposed outer membrane protein. To con-
firm predictions that NilB is localized to the outer membrane, we
monitored NilB-FLAG levels in X. nematophila cellular fractions.
NilB-FLAG detection by Western blotting was facilitated by pre-
paring extracts from the X. nematophila nilB-FLAG26 strain lack-
ing nilR, which represses nilB transcription (16) and protein levels
(Fig. 2). Using anti-FLAG antibody, NilB-FLAG26 (53 kDa) was
detected in whole cells, crude sonicates, and membrane fractions
isolated from X. nematophila but not in soluble, cleared X. nema-
tophila lysates (Fig. 3A), supporting predictions that NilB is local-
ized to the outer membrane and therefore likely adopts a beta-
barrel structure (33). To determine if any portion of NilB is
surface exposed, we treated whole cells expressing NilB-FLAG26
with proteinase K and then monitored NilB levels over time using
anti-FLAG immunoblotting. Levels of full-length NilB-FLAG in
whole cells decreased dramatically by 1 min after exposure to pro-
teinase K and were not detectable after 4 min of incubation, indi-
cating that at least some regions of NilB are surface exposed and
susceptible to cleavage (Fig. 3B). NilC, which is not sensitive to
proteinase K digestion (15), remained detectable over the same
time period (Fig. 3B). Both NilB-FLAG and NilC were sensitive to
proteinase K digestion in lysed cell extracts (data not shown).
After 1 min of incubation with proteinase K, concomitant
with the reduction in full-length NilB-FLAG, another band that
migrated slightly faster by SDS-PAGE than full-length mature
NilB-FLAG appeared (Fig. 3B), and it had an approximate molec-
ular mass of 48 kDa. Given that the FLAG tag is encoded at the
amino-terminus of the protein, the appearance of these bands
is consistent with proteinase K cleavage somewhere in the C-ter-
minal region of the protein. The estimated molecular masses of
NilB if it is cleaved in the middle of one of the three SLs at the
C-terminal end of the protein are 40 (SL5), 46 (SL6), and 51
(SL7) kDa. Based on this, initial proteinase K cleavage may occur
within or near SL6 (aa 387 to 410), the longest predicted surface-
exposed loop, which may therefore be particularly susceptible to
cleavage. The intensity of the band decreases progressively over
time of exposure to proteinase K, while a third band of 40 kDa
appears by 10 min of incubation with proteinase K, possibly rep-
resenting another cleavage product (perhaps within SL5).
FLAG tag insertions support a model of NilB with an
140-aa periplasmic domain and 7 surface-exposed loops. To
probe NilB structure and differentiate among bioinformatic to-
pology predictions (Fig. 1), we exploited the fact that insertions
within TM strands should destabilize the protein and prevent nor-
mal protein localization, while insertions within extracellular or
periplasmic domains are more likely to be tolerated (28, 49).
Based on HMM models predicted by PRED-TMBB, PROFtmb,
and B2TMR-HMM, we created five unique FLAG tag insertion
nilB mutants in addition to a nilB-FLAG26 strain to distinguish
between the HMM models predicting either 14 (PROFtmb or
B2TMR-HMM) or 18 (PRED-TMBB) TM strands (with 7 or 9
SLs, respectively) (Fig. 1).
We inserted FLAG tags into three regions predicted to be trans-
membrane strands in the 18-strand model but periplasmic strands
in the 14-strand model (before aa 68, aa 102, or aa 121). As a
control, we also inserted FLAG tags into two domains consistently
predicted by all HMM programs to be TMs (before aa 379 and aa
421, flanking SL6). As predicted, X. nematophila expressing NilB
FLAG insertions before aa 379 and aa 421 are colonization defec-
tive (Fig. 4A) and fail to yield detectable protein (Fig. 4B), indicat-
ing that these insertions destabilize the NilB protein. In contrast,
X. nematophila carrying NilB-FLAG variants before aa 68 or aa
121 colonize at levels significantly higher than in a strain lacking
NilB entirely (Fig. 4A). The difference in colonization phenotype
of a strain carrying NilB-FLAG aa102 and a strain lacking nilB
entirely was barely significant (P0.049) (Fig. 4A) and may in-
dicate that FLAG insertion at aa 102 disrupts NilB function. Re-
gardless, NilB-FLAG protein is detectable in strains expressing
NilB-FLAG insertions at aa 68, aa 102, and aa 121 (Fig. 4B), and
this is consistent with a periplasmic localization of these domains
as predicted in the 14-TM model (Fig. 1 and 4C) but not the
18-TM model (Fig. 4D).
Colonization function of nilB insertion mutants. We next
used FLAG tag insertions to assess which regions of NilB are nec-
essary for colonization, as well as for further verification of the
FIG 3 NilB is a membrane-localized protein and is surface exposed. (A) X.
nematophila cells expressing NilB-FLAG26 in a nilR mutant background
(HGB1200) were pelleted, resuspended, and sonicated. Soluble and mem-
brane fractions were separated by ultracentrifugation. The crude extract (lane
1), cleared lysate (lane 2), soluble fraction (lane 3), and membrane fraction
(lane 4) were electrophoresed on a 12% SDS-polyacrylamide gel and trans-
ferred to a membrane. NilB protein (arrowhead) was detected using anti-
FLAG antibody. (B) Whole X. nematophila nilR mutant cells expressing NilB-
FLAG26 (HGB1200) were incubated with 38.4
g/ml proteinase K (PK) (final
concentration) for 1 to 30 min at 37°C. At various time points (indicated in
minutes above each lane), samples were removed and reactions were stopped
with the addition of 5 mM PMSF (final concentration) and kept on ice. SDS-
PAGE loading dye was added to all the reaction mixtures, which were heated at
95°C for 10 min and electrophoresed on 12% SDS-PAGE gels. NilB (top panel)
(anti-NilB) or NilC (bottom panel) (anti-NilC) was detected in Western blots
using anti-FLAG or anti-NilC (15) antibody, respectively. Full-length NilB-
FLAG26 protein (black arrowheads, No PK lanes) was present in the no-pro-
teinase K control but not in cells lacking nilB (nil) (HGB1251). This species
decreases in intensity after the addition of proteinase K, while smaller-molec-
ular-mass proteins (black arrowheads, lanes 1 and 10) appear. M, Thermo
Scientific Pierce prestained protein molecular mass marker (sizes in kDa are
noted on the left).
Bhasin et al.
1770 jb.asm.org Journal of Bacteriology
14-TM strand model shown in Fig. 1, relative to other possible
topologies predicted by non-HMM algorithms (see Fig. S2 in the
supplemental material). To this end, we created 10 insertions at
regions of NilB predicted to be periplasmic or surface exposed
(Fig. 5C) in the 14-TM model (before aa 85, 137, 186, 233, 273,
322, 357, 399, 451, and 466), and at least one of which is within a
predicted TM domain in all of the other models that we identified
by various prediction programs (Fig. S2).
Nematode colonization phenotypes varied among the inser-
tion mutants. The insertion at aa 84 in the predicted periplasmic
domain completely abolishes colonization, measured as the aver-
age number of CFU/infective juvenile determined by nematode
sonication and dilution plating (Fig. 5A). Colonization by strains
carrying insertions at aa 322 (SL4), aa 399 (SL6), and aa 451 (SL7)
was detectable but significantly attenuated relative to that by a
non-FLAG-tagged wild-type strain, while strains carrying inser-
tions at aa 137, 186, 233, 273, 357, and 466 colonized at levels not
significantly different from that of the wild-type nilB control (Fig.
5A). To determine if the colonization defects of the insertion mu-
tants are due to a lack of NilB protein production, Western blot-
ting with anti-FLAG antibody was performed on whole-cell ex-
tracts from all mutants. NilB protein was detected at various
amounts and levels of mobility in X. nematophila strains express-
ing each nilB variant. The amount of NilB produced during labo-
ratory growth did not correlate with nematode colonization, since
FIG 4 FLAG tag insertions distinguish whether NilB adopts a predicted
18-TM or 14-TM strand structure. Insertions were made in three N-terminal
TM strands (before aa 68, aa 102, and aa 121) predicted only by the 18-TM
strand model (D) and in two TM strands predicted by both models (before aa
379 and at aa 421); all mutations were made in nilR mutant cells. The effect of
each mutation on nematode colonization (A) and protein production (B) was
assessed by nematode sonication and immunoblotting with anti-FLAG, re-
spectively. (A) Insertions before aa 68 or aa 121 reduce nematode colonization
below wild-type levels but not to the level observed when nilB is absent (ETn7).
Insertions before aa 102, aa 379, and aa 421 cause severe nematode coloniza-
tion defects. *, P0.01, relative to ETn7values (except with the aa 102 strain,
for which Pwas 0.049). IJ, infective juvenile. (B) When FLAG insertions are in
TM strands predicted only by the 18-TM strand model, NilB protein is detect-
able in whole cells by anti-FLAG Western blotting, but when insertions are in
TM strands predicted by both models, NilB is not detectable. (C and D) Sche-
matic of NilB’s structure and the relative positions of FLAG tag insertions in
the 14-TM (C) and 18-TM (D) strand models. Circles indicate that an inser-
tion mutant colonized better than (white circles) or similarly to (shaded cir-
cles) a strain that lacked SR1. NF, Tn7/SR1 without a FLAG26 insertion in
NilB; ETn7,Tn7lacking SR1; FLAG26, Tn7/SR1 with a FLAG26 insertion in
NilB (no deletions).
FIG 5 Identification of NilB regions necessary for colonization. FLAG tags
were inserted into predicted NilB surface loops, the predicted N-terminal
periplasmic domain, and the C terminus. The numbers indicate the amino
acid locations of the tag; for example, 26 indicates that the FLAG tag is directly
before (N-terminal to) amino acid 26 (Fig. 1). Mutants were made in nilR
mutant backgrounds. (A) X. nematophila NilB-FLAG mutants were assessed
for S. carpocapsae nematode colonization. Asterisks indicate colonization lev-
els significantly different (P0.05) from that of X. nematophila expressing
NilB without the FLAG tag (NF). (B) NilB protein production by the NilB-
FLAG mutants was assayed on whole cells by Western blotting with anti-FLAG
antibody. The abundances and SDS-PAGE migrations of the NilB-FLAG pro-
teins (bracket) are varied. (C) Schematic of NilB structure, including predicted
locations of FLAG tag insertions in the 14-TM strand model (Fig. 1). Shaded
circles represent insertions that abolished or attenuated nematode coloniza-
tion, while white circles represent insertions that do not significantly affect
colonization.
Topology and Functional Regions of X. nematophila NilB
April 2012 Volume 194 Number 7 jb.asm.org 1771
two strains with barely detectable levels of NilB (those with FLAG
inserted before aa 233 and aa 466) displayed colonization profi-
ciency (Fig. 5B). The migration distance of each mutant protein
also did not correlate with colonization proficiency or deficiency.
For example, NilB proteins with insertions at aa 137 and 186 mi-
grated differently than NilB protein with an insertion at aa 26, but
both strains carrying these nilB variants colonized the nematode
receptacle at levels not different from that of the wild type. The
cause of variability in NilB protein mobility is unknown, but a
similar phenomenon has been reported previously in some inser-
tion (including FLAG tag) mutagenesis studies (28, 49) but not
others (40). In studies where migration variability was observed,
differences were attributed to heat-modifiable stability (27). NilB-
FLAG mutants were analyzed for heat modifiability by either boil-
ing crude extracts for 15 min at 97°C or not boiling them; samples
were then electrophoresed by SDS-PAGE in a 4°C cold room and
Western blotted. No heat-dependent changes in mobility were
observed (data not shown). Alternatively, insertion mutants may
display differential detergent binding as a reflection of variable
protein-lipid or intraprotein interactions (46). When we tested if
the protein migration differences were due to variable NilB glyco-
sylation, results showed that in glycoprotein stains of whole-cell X.
nematophila extracts, no X. nematophila proteins were found to be
glycosylated (data not shown). Despite mobility differences, the
presence of NilB protein in the colonization-defective strains with
a FLAG insertion before aa 85, aa 322, aa 399, and aa 451 supports
the conclusion that these regions are critical for the function of
NilB.
nilB mutants are defective in the entry stage of nematode
colonization. To gain further insight into the effect of nilB-FLAG
mutations on nematode colonization, we assessed nematode col-
onization phenotypes in vivo using GFP-expressing nilB-FLAG
mutants and epifluorescence microscopy. We expected that par-
tial-colonization mutants would fall into one of two classes: out-
growth mutants, with nearly all nematodes carrying one or a few
bacteria in each receptacle, or entry or persistence mutants, with a
mixture of nematodes with fully colonized or completely empty
receptacles. We chose FLAG tag insertion constructs that cause
normal to high (with FLAG before aa 26 and aa 273), low (with
FLAG before aa 322 and aa 399), and intermediate (with FLAG
before aa 186) levels of colonization and incorporated them into
the att Tn7site of a GFP-expressing SR1 strain (HGB1430). The
percentage of nematodes within each population where the GFP-
labeled bacteria were observed in the receptacle was determined.
An oligo-colonization phenotype, in which a receptacle contained
one or a few bacterial cells (29, 38), was not observed for any of the
5 NilB FLAG tag mutants tested in this experiment (data not
shown). Further, nematode receptacles qualitatively appeared to
be full of bacteria or were empty (e.g., Fig. 6A), suggesting that
these nilB mutations affect entry, not outgrowth. This idea was
further addressed by quantifying the average bacterial load per
nematode through surface sterilization and sonication of 10,000
nematodes. Sonication (average bacterial load) and microscopy
(colonization frequency) data are presented normalized against a
strain carrying wild-type nilB without a FLAG insertion (Table 3).
Fold differences reveal good correlation between the outputs of
the two tests, suggesting that bacteria that gained access to the
receptacle fully colonized it and that the colonization defect of a
nilB mutant occurs prior to its gaining access to the receptacle.
This finding was not necessarily expected, since a similar compar-
ison of sonication and microscopy data for a nilA mutant revealed
a25-fold difference between nematode colonization frequency
(65% of nematodes) and bacterial load (2.5% of wild-type cells)
(17). Thus, the partial colonization defects observed for nilA and
nilB mutants are distinct. These findings raise the possibility that
the nil genes, which are together sufficient for nematode coloni-
zation, may act at distinct stages of nematode colonization.
Colonization function of nilB deletion mutants. NilB sur-
face-exposed loops and portions of the predicted N-terminal
periplasmic domain (Fig. 6D) were deleted to further assess the
importance of these regions in nematode colonization. Since mi-
croscopic analysis gave greater sensitivity and smaller errors be-
tween replicates than sonication assays, and results were generally
similar to those derived from sonication assays (Table 3), deletion
constructs were introduced into a GFP-expressing X. nematophila
background and colonization was assessed by epifluorescence mi-
croscopy. In each of these deletion mutants, nematode receptacles
appeared to be empty or fully colonized, consistent with our pre-
vious observations. Deletions in SL4 (of aa 314 to 320), SL5 (aa
352 to 357), and SL6 (aa 389 to 410) had significant negative yet
variable effects on nematode colonization (Fig. 6B), and deleting
SL6 caused the most severe colonization deficiency (equivalent to
a strain lacking nilB entirely). The finding that the SL3 (aa 266 to
279) deletion strain colonized greater than 50% of the nematodes
in the population emphasizes that some extracellular domains
play a more significant role in colonization than others.
Some regions of the predicted N-terminal globular domain (aa
26 to 156) are also important for nematode colonization (Fig. 6B).
Strains carrying deletions of aa 92 to 108 or aa 109 to 156 (the latter
encompassing the predicted TPR domain) reduce colonization to
levels equivalent to that of a strain lacking nilB, whereas deletions
of aa 76 to 91, 26 to 108, and 26 to 156 have a milder, yet signifi-
cant, colonization defect. In contrast, deleting aa 26 to 42 and aa
26 to 59 does not reduce colonization levels, and deleting aa 43 to
59 has only a mildly deleterious effect on nematode colonization
frequency.
Although NilB protein levels and mobilities on SDS-PAGE gels
vary for the different deletion mutants, the migration differences
correlate with the expected size change due to deletion of residues
(Fig. 6C). NilB is detectable for every deletion mutant except that
with a deletion of aa 292 to 298, consistent with the prediction that
these amino acids form a TM domain. As seen with the FLAG
insertion mutants, colonization levels do not correlate with NilB
protein levels. For example, NilB is relatively abundant in the col-
onization-defective SL6 strain (deletion of aa 389 to 410), in the
colonization-defective strain with a deletion of aa 76 to 91, and in
the colonization-proficient strain with a deletion of aa 26 to 42 but
not in the colonization-proficient strain with a deletion of aa 26 to
59. Therefore, low protein levels likely do not fully explain the
colonization defects of deletion mutants.
DISCUSSION
We have shown that NilB is a surface-exposed protein that is
poorly expressed under laboratory growth conditions. When nilB
expression is derepressed by deletion of nilR, NilB protein levels
are highest under nutrient-limiting conditions (e.g., in defined
medium and/or stationary phase) (Fig. 2), indicating that NilB
expression responds to nutrient conditions independently of
NilR. Because of this, we predict that, during nematode coloniza-
tion, NilB expression is triggered by multiple signals, both NilR
Bhasin et al.
1772 jb.asm.org Journal of Bacteriology
FIG 6 In-frame amino acid deletions identify structurally and functionally required domains of NilB. (A) Partial nematode colonization mutants fully
colonize few nematodes instead of partially colonizing many nematodes in the population. Representative images of nematodes grown on colonization
mutants are shown. Images from left to right are as follows: a 40magnification of 2 colonized (HGB1502) nematodes (the green dot is an epifluorescent
overlay) and 1 uncolonized nematode (the black arrow points to the location of an empty receptacle), a zoomed image of a colonized nematode
(HGB1502) with an epifluorescent overlay on a DIC image, and a zoomed image of an uncolonized nematode (HGB1429) with an epifluorescent overlay
on a DIC image, with the empty nematode receptacle highlighted (white outlining). Scale bars from left to right: 50
m, 5
m, and 5
m. (B) Small regions
in the N-terminal periplasmic domain were individually sequentially deleted (026042, 043– 059, 076 091, 092–108, 109 –156), along with
larger regions of this domain (026 –059, 026 –108, 026–156). Amino acids were also deleted from SL3, SL4, SL5, or SL6, as well as a region within the
predicted TM domain between SL3 and SL4 (as a negative control). Each of these deletions was created in the plasmid pTn7/SR1/nilB-FLAG26, and Tn7
integration at the att Tn7site of GFP-expressing X. nematophila ATCC 19061 SR1 nilR was confirmed. (A) Nematode colonization of each deletion
mutant assessed by microscopy. Pwas 0.05 for differences from values for the negative control (eTn7) (#) or positive control (NF) (*). Strain eTn7lacks
nilB, while NF contains nilB but no FLAG tag. (C) NilB-FLAG proteins of various abundances and mobilities are produced in whole cells by all deletion
mutants except that lacking aa 292 to 298. indicates detectable NilB, and indicates no detectable NilB protein. NF contains nilB but no FLAG tag.
FLAG26 contains the FLAG tag before aa 26 of nilB. (D) Schematics of NilB structures and deletions. The bars represent the N-terminal periplasmic
domain from aa 26 to 164. The length of the shaded area represents the size of the deletion. The color of the shading represents the effect of the deletion
on nematode colonization; black indicates no effect relative to that of the wild type, gray indicates an effect intermediate between those of the wild type
and the negative control (no NilB), and the cross-hatched shading indicates no significant difference from the effect of the negative control.
Topology and Functional Regions of X. nematophila NilB
April 2012 Volume 194 Number 7 jb.asm.org 1773
dependent and NilR independent. Since under laboratory growth
conditions NilR-dependent repression occurred regardless of
growth phase or medium, we suggest that the signal necessary for
NilR derepression is absent under these conditions and may be
specific to the nematode host. Analysis of the colonization pheno-
types of nilB mutant alleles revealed that NilB functions at the
stage of entry into the nematode receptacle, rather than during
later stages of growth and persistence, indicating that the signals
necessary for NilB expression likely are present at the very early
stages of colonization.
Our analysis of nilB mutants supports the model that NilB
adopts a 14-TM beta-barrel structure, based on the premise that
insertions within, or deletions of, TM domains prevent proper
localization and lead to protein degradation, while insertions or
deletions in other regions (e.g., those that are surface exposed) are
generally tolerated (23, 49). Insertions or deletions that prevented
detection of NilB protein were in TM strands (aa 292 to 298, 379,
and 421) predicted by the 14-TM model. In contrast, NilB protein
was detectable in mutants with insertions or deletions in regions
predicted to be TM in an 18-TM model but periplasmic in a
14-TM model (aa 68, 92 to 108, 102, 109 to 156, and 121). We also
show that NilB is membrane localized and has surface-exposed
regions, based on its sensitivity to proteinase K digestion in whole
cells (Fig. 3). The appearance of truncated, large-molecular-
weight FLAG-reactive proteins after proteinase K digestion is con-
sistent with cleavage of the C-terminal but not the N-terminal
(antibody-reactive) domain (Fig. 3), supporting the bioinformati-
cally predicted location of the N terminus in the periplasm. Taken
together, these data support a protein structure prediction model
where NilB has 7 SLs (SL1 to -7) and an 140-aa N-terminal
periplasmic domain. Below we discuss how these results provide
insight into possible mechanisms of NilB function in coloniza-
tion.
We identified several connections between NilB function
and nutrient availability/transport through our studies on NilB
expression, genomic context, and predicted topology. As
discussed above, NilB expression is elevated by nutrient limi-
tation during laboratory growth. Further, pathogens such as
Mannheimia haemolytica and Actinobacillus pleuropneumo-
niae, which derive iron from host-specific transferrin, lactofer-
rin, or hemoglobin molecules (51), encode NilB homologs near
iron binding (YP_002173032) or host transferrin utilization
(ZP_04464085) genes, respectively. This genomic context sug-
gests that NilB homologs, and perhaps NilB, may also function
in nutrient acquisition. Finally, our data suggest that NilB is a
14-TM beta-barrel protein with a large periplasmic N-terminal
domain. In many beta-barrel proteins, large periplasmic do-
mains function as plugs that help control beta-barrel pore per-
meability. For example, such topology is typical of small-mol-
ecule transporters (47, 53, 57) that depend on the inner
membrane protein TonB to provide the energy for nutrient
transport (e.g., iron, vitamin B
12
) (31). TonB-dependent trans-
porters include transferrin receptors (31), such as those en-
coded near nilB homologs. At first glance, NilB seems unlikely
to be a TonB-dependent protein since it lacks a canonical TonB
box, since most TonB-dependent transporters are 22-TM do-
main proteins, and since an X. nematophila tonB mutant does
not have a colonization defect (38). However, TonB box iden-
tities are varied among TonB-dependent transporters (34), and
the X. nematophila genome encodes two additional TonB-like
genes (13), one of which is oriented in tandem with the NilB-
like protein XNC1_0074 (see Fig. S1 in the supplemental ma-
terial). Taken together, the links between NilB and TonB de-
tailed above are consistent with a role for NilB in nutrient
transport. However, our data do not exclude the possibility of
an adhesin function for NilB, and adhesion and nutrient acqui-
sition models are not mutually exclusive.
Our work has revealed several critical regions of NilB necessary
for its function in nematode colonization, including SL6 and the
N-terminal TPR-like domain. Deletion of SL6 reduced nematode
colonization to levels observed when nilB was absent, but this
reduction in nematode colonization is not a result of NilB insta-
bility since NilB is detected in Western blots of both whole-cell
lysates (Fig. 5B and 6C) and membrane preparations (data not
shown). SL6 (28 residues) is the longest SL in NilB (9 residues
longer than the next longest, SL1), and in a full-length alignment
of all NilB homologs, NilB has more residues at SL6 than any other
NilB homolog (data not shown). The longer length of SL6 in NilB
than in NilB homologs may indicate the importance of SL6 resi-
dues in the specificity of NilB in S. carpocapsae association. Vari-
ous functions could be envisioned for this region, including di-
rectly binding to host surfaces or molecules. For example, in beta-
TABLE 3 Fold change in two independent experiments between bacterial load and nematode colonization frequency of nilB-FLAG insertion
mutants
FLAG
insertion
Expt 1 Expt 2
Relative bacterial
load (%)
a
Relative
frequency (%)
b
Fold
difference
c
Relative bacterial
load (%)
Relative
frequency (%)
Fold
difference
aa 26 55.882 89.490 1.60 119.156 102.007 1.17
aa 186 52.574 52.352 1.00 61.039 49.763 1.23
aa 273 106.985 96.734 1.11 99.675 91.455 1.09
aa 322 32.279 25.762 1.25 60.390 38.542 1.57
aa 399 1.397 0.960 1.46 4.789 4.721 1.01
eTn70.042 0.015 2.86 0.009 0.009 1.00
None 100.000 100.000 1.00 100.000 100.000 1.00
a
Bacterial load was assessed by sonication of 10,000 nematodes and dilution plating to determine bacterial counts. Data are presented as percentages of the value for cells expressing
wild-type nilB with no FLAG insertion.
b
Nematode colonization frequency was assessed by epifluorescence microscopy of nematodes grown on GFP-expressing bacteria. Data are presented as percentages of the value for
cells expressing wild-type nilB with no FLAG insertion.
c
Fold difference represents the ratio of relative colonization frequency to relative bacterial load. A ratio of 1 indicates that the colonization defect of the strain is explained by a
reduced frequency of colonization (e.g., initiation). A ratio of greater than 1 indicates that a defect in outgrowth contributes to observed defects in bacterial load.
Bhasin et al.
1774 jb.asm.org Journal of Bacteriology
barrel transporters, surface loops act by trapping and directing
transport of specific molecules through the pore and/or occluding
foreign molecules during transport as part of a molecular airlock
system (see, e.g., references 20 and 45).
Deletion of the TPR-like domain (109 –156) caused a coloni-
zation defect that could be ameliorated by concomitant deletion
of aa 26 to 108 (in Fig. 6B, compare infective juvenile colonization
of the 109 –156 mutant [0.008% of nematodes] versus that of
the 26 –156 mutant [0.8%]). Such a result could be explained if
the NilB periplasmic domain acts as a plug whose “open” and
“closed” states, or some other pore modification, are regulated
within the TPR region. Deletion of the domain necessary to tran-
sition to the “open” state would effectively block the pore and
prevent transport, whereas deletion of the entire plug might result
in a constitutively leaky pore that could function, albeit less effi-
ciently than wild-type NilB, in colonization. Alternatively, the
TPR-like domain may interact with other regions of the protein
(e.g., a pore) to impact the structure and function of surface-
exposed loops that may be necessary for adhesion, while deletion
of additional residues of the N-terminal domain may allow NilB to
adopt a more wild-type structure. Regardless, our data suggest
that the TPR-like domain is critical for NilB function, possibly due
to its role in mediating protein-protein interactions. In this re-
gard, it will be of interest to determine if NilB interacts with other
proteins and if such interactions depend on the TPR-like domain.
NilC, an outer membrane, periplasmically oriented lipoprotein
(15), is an obvious candidate for a NilB-interacting protein since it
is encoded on the same genetic locus as NilB (30), is repressed by
NilR (16), and is a species specificity factor for S. carpocapsae nem-
atodes (17).
Although, based on our data, we are unable to definitively con-
clude that NilB functions as a transporter or an adhesin, either
model warrants some speculation regarding the identity of its sub-
strate and whether this molecule is specific to S. carpocapsae. NilB
is a specificity determinant, and S. carpocapsae nematodes are col-
onized only by Xenorhabdus species that express nilB (17). Since
no defects in nilB mutant growth, secondary metabolism, or other
aspects of physiology have been observed in nilR
cells during
laboratory culture (30), including testing with the full panel of
Biolog metabolic phenotype arrays (PM1 to PM10; Biolog) (data
not shown), NilB may be involved in the metabolism of a nutrient
or molecule that is absent under laboratory conditions. Alterna-
tively, NilR repression of NilB under these conditions may pre-
clude identification of NilB-dependent phenotypes. Therefore,
further investigations of NilB-dependent laboratory growth phe-
notypes in a nilR mutant background are warranted. However,
intriguing hypotheses are that NilB is necessary to transport a
host-specific nutrient or adhere to a host-specific molecule and
that NilR derepression occurs only in the presence of this mole-
cule to prevent inappropriate expression of NilB (e.g., in the insect
host, where it may be immunogenic).
We previously suggested (13, 17) that species specificity in Xe-
norhabdus does not follow the paradigm of host range specificity
in the well-studied leguminous plant-Rhizobium symbioses. In
these symbioses, each rhizobium species produces a common sig-
naling molecule, the nod factor, and specificity is determined by
modifications of the core nod factor molecule (24). In contrast,
the host range specificity determinant NilB is lacking from other
Xenorhabdus species, suggesting that it is a novel, derived trait that
determines specificity in this bacterium. However, this idea is
challenged by the results presented here. Structural predictions of
NilB-like proteins, including one in X. bovienii, reveal similar to-
pologies despite relatively low amino acid sequence similarity
(18 to 25%), raising the possibility that NilB-like proteins with
structural (and perhaps functional) similarity are produced by
other Xenorhabdus species, but they have insufficiently low coding
sequence identity to be identified by PCR (30), Southern hybrid-
ization (17; J. Chaston and H. Goodrich-Blair, unpublished data),
or genome sequencing (14). Therefore, NilB-like proteins may
play similar roles among all Xenorhabdus species in mediating host
interactions and specificity, with NilB representing a highly di-
verged, horizontally acquired (17) member of this family. Subtle
differences in surface-exposed loops (e.g., SL6) may determine
specificity, much like nod factor modifications. Future studies
that determine the mechanism of action of NilB and species spec-
ificity factors in other Xenorhabdus spp. will help test if the para-
digm of specificity determined by “variations on a common
theme” pertains to Xenorhabdus-nematode interactions.
ACKNOWLEDGMENTS
This work was supported by grants awarded to H.G.-B. from the National
Science Foundation (NSF) (IOS-0950873) and the National Institutes of
Health (NIH) (GM059776). J.M.C. was supported by NIH National Re-
search Service Award T32 (grant AI55397, Microbes in Health and Dis-
ease) and an NSF Graduate Research Fellowship. A.B. was supported by
NIH grant F32 GM072342.
We are grateful to Tom Silhavy for helpful discussions and to anony-
mous reviewers, whose comments significantly improved the manuscript.
We are also grateful to Aaron Andersen for technical support and Xinxin
Yu and Yang Zhao in the CALS statistical consulting service (UW—Mad-
ison) for assistance with choosing and applying statistical models.
REFERENCES
1. Abe Y, et al. 2000. Structural basis of presequence recognition by the
mitochondrial protein import receptor Tom20. Cell 100:551–560.
2. Altschul SF, et al. 1997. Gapped BLAST and PSI-BLAST: a new genera-
tion of protein database search programs. Nucleic Acids Res. 25:3389
3402.
3. Andrade MA, Perez-Iratxeta C, Ponting CP. 2001. Protein repeats:
structures, functions, and evolution. J. Struct. Biol. 134:117–131.
4. Bagos PG, Liakopoulos TD, Hamodrakas SJ. 2005. Evaluation of meth-
ods for predicting the topology of beta-barrel outer membrane proteins
and a consensus prediction method. BMC Bioinformatics 6:7.
5. Bagos PG, Liakopoulos TD, Spyropoulos IC, Hamodrakas SJ. 2004. A
hidden Markov model method, capable of predicting and discriminating
beta-barrel outer membrane proteins. BMC Bioinformatics 5:29.
6. Bagos PG, Liakopoulos TD, Spyropoulos IC, Hamodrakas SJ. 2004.
PRED-TMBB: a web server for predicting the topology of beta-barrel
outer membrane proteins. Nucleic Acids Res. 32:W400–W404.
7. Bao Y, Lies DP, Fu H, Roberts GP. 1991. An improved Tn7-based system
for the single-copy insertion of cloned genes into chromosomes of Gram-
negative bacteria. Gene 109:167–168.
8. Bendtsen JD, Nielsen H, von Heijne G, Brunak S. 2004. Improved
prediction of signal peptides: SignalP 3.0. J. Mol. Biol. 340:783–795.
9. Bigelow HR, Petrey DS, Liu J, Przybylski D, Rost B. 2004. Predicting
transmembrane beta-barrels in proteomes. Nucleic Acids Res. 32:2566
2577.
10. Bird AF, Akhurst RJ. 1983. The nature of the intestinal vesicle in nema-
todes of the family Steinernematidae. Int. J. Parasitol. 13:599 606.
11. Blatch GL, Lassle M. 1999. The tetratricopeptide repeat: a structural
motif mediating protein-protein interactions. Bioessays 21:932–939.
12. Boemare NE, Akhurst RJ, Mourant RG. 1993. DNA relatedness between
Xenorhabdus spp. (Enterobacteriaceae), symbiotic bacteria of ento-
mopathogenic nematodes, and a proposal to transfer Xenorhabdus lumi-
nescens to a new genus, Photorhabdus, gen. nov. Int. J. Syst. Bacteriol.
43:249–255.
13. Chaston J, Goodrich-Blair H. 2010. Common trends in mutualism re-
Topology and Functional Regions of X. nematophila NilB
April 2012 Volume 194 Number 7 jb.asm.org 1775
vealed by model associations between invertebrates and bacteria. FEMS
Microbiol. Rev. 34:41–58.
14. Chaston JM, et al. 2011. The entomopathogenic bacterial endosymbionts
Xenorhabdus and Photorhabdus: convergent lifestyles from divergent ge-
nomes. PLoS One 6:e27909.
15. Cowles CE, Goodrich-Blair H. 2004. Characterization of a lipoprotein,
NilC, required by Xenorhabdus nematophila for mutualism with its nem-
atode host. Mol. Microbiol. 54:464 477.
16. Cowles CE, Goodrich-Blair H. 2006. nilR is necessary for co-ordinate
repression of Xenorhabdus nematophila mutualism genes. Mol. Microbiol.
62:760–771.
17. Cowles CE, Goodrich-Blair H. 2008. The Xenorhabdus nematophila nil-
ABC genes confer the ability of Xenorhabdus spp. to colonize Steinernema
carpocapsae nematodes. J. Bacteriol. 190:4121–4128.
18. Cowles KN, Cowles CE, Richards GR, Martens EC, Goodrich-Blair H.
2007. The global regulator Lrp contributes to mutualism, pathogenesis
and phenotypic variation in the bacterium Xenorhabdus nematophila.
Cell. Microbiol. 9:1311–1323.
19. D’Andrea LD, Regan L. 2003. TPR proteins: the versatile helix. Trends
Biochem. Sci. 28:655– 662.
20. Ferguson AD, et al. 2002. Structural basis of gating by the outer mem-
brane transporter FecA. Science 295:1715–1719.
21. Flores-Lara Y, Renneckar D, Forst S, Goodrich-Blair H, Stock P. 2007.
Influence of nematode age and culture conditions on morphological and
physiological parameters in the bacterial vesicle of Steinernema carpocap-
sae (Nematoda: Steinernematidae). J. Invertebr. Pathol. 95:110–118.
22. Forst S, Dowds B, Boemare N, Stackebrandt E. 1997. Xenorhabdus and
Photorhabdus spp.: bugs that kill bugs. Annu. Rev. Microbiol. 51:47–72.
23. Freudl R. 1989. Insertion of peptides into cell-surface-exposed areas of
the Escherichia coli OmpA protein does not interfere with export and
membrane assembly. Gene 82:229–236.
24. Garg N, Geetanjali. 2007. Symbiotic nitrogen fixation in legume nodules:
process and signaling. A review. Agron. Sustain. Dev. 27:59 68.
25. Goetsch M, Owen H, Goldman B, Forst S. 2006. Analysis of the PixA
inclusion body protein of Xenorhabdus nematophila. J. Bacteriol. 188:
2706–2710.
26. Goodrich-Blair H, Clarke DJ. 2007. Mutualism and pathogenesis in
Xenorhabdus and Photorhabdus: two roads to the same destination. Mol.
Microbiol. 64:260–268.
27. Hancock RE, Carey AM. 1979. Outer membrane of Pseudomonas aerugi-
nosa: heat-2-mercaptoethanol-modifiable proteins. J. Bacteriol. 140:902–
910.
28. Hay ID, Rehman ZU, Rehm BH. 2010. Membrane topology of outer
membrane protein AlgE, which is required for alginate production in
Pseudomonas aeruginosa. Appl. Environ. Microbiol. 76:1806–1812.
29. Herbert Tran EE, Andersen AW, Goodrich-Blair H. 2009. CpxRA in-
fluences Xenorhabdus nematophila colonization initiation and outgrowth
in Steinernema carpocapsae nematodes through regulation of the nil locus.
Appl. Environ. Microbiol. 75:4007–4014.
30. Heungens K, Cowles CE, Goodrich-Blair H. 2002. Identification of
Xenorhabdus nematophila genes required for mutualistic colonization of
Steinernema carpocapsae nematodes. Mol. Microbiol. 45:1337–1353.
31. Jarosik GP, Maciver I, Hansen EJ. 1995. Utilization of transferrin-bound
iron by Haemophilus influenzae requires an intact tonB gene. Infect.
Immun. 63:710–713.
32. Kaya HK, Gaugler R. 1993. Entomopathogenic nematodes. Annu. Rev.
Entomol. 38:181–206.
33. Knowles TJ, Scott-Tucker A, Overduin M, Henderson IR. 2009. Mem-
brane protein architects: the role of the BAM complex in outer membrane
protein assembly. Nat. Rev. Microbiol. 7:206–214.
34. Krewulak KD, Vogel HJ. 2011. TonB or not TonB: is that the question?
Biochem. Cell Biol. 89:87–97.
35. Marchler-Bauer A, et al. 2007. CDD: a conserved domain database for
interactive domain family analysis. Nucleic Acids Res. 35:D237–D240.
36. Martelli PL, Fariselli P, Krogh A, Casadio R. 2002. A sequence-profile-
based HMM for predicting and discriminating beta barrel membrane pro-
teins. Bioinformatics 18(Suppl 1):S46 –S53.
37. Martens EC, Heungens K, Goodrich-Blair H. 2003. Early colonization
events in the mutualistic association between Steinernema carpocapsae
nematodes and Xenorhabdus nematophila bacteria. J. Bacteriol. 185:3147–
3154.
38. Martens EC, Russell FM, Goodrich-Blair H. 2005. Analysis of Xenorh-
abdus nematophila metabolic mutants yields insight into stages of Steiner-
nema carpocapsae nematode intestinal colonization. Mol. Microbiol. 51:
28 45.
39. Morgan JAW, Kuntzelmann V, Tavernor S, Ousley MA, Winstanley C.
1997. Survival of Xenorhabdus nematophilus and Photorhabdus lumine-
scens in water and soil. J. Appl. Microbiol. 83:665–670.
40. Nguyen KA, et al. 2009. Verification of a topology model of PorT as an
integral outer-membrane protein in Porphyromonas gingivalis. Microbi-
ology 155:328–337.
41. Orchard SS, Goodrich-Blair H. 2004. Identification and functional char-
acterization of a Xenorhabdus nematophila oligopeptide permease. Appl.
Environ. Microbiol. 70:5621–5627.
42. Poinar GO. 1966. The presence of Achromobacter nematophilus in the
infective stage of a Neoaplectana sp. (Steinernematidae: Nematoda).
Nematologica 12:105–108.
43. Poinar GO, Jr. 1979. Nematodes for biological control of insects. CRC
Press, Boca Raton, FL.
44. Poinar GO, Jr, Thomas GM. 1967. The nature of Achromobacter nema-
tophilus as an insect pathogen. J. Invertebr. Pathol. 9:510–514.
45. Postle K, Kadner RJ. 2003. Touch and go: tying TonB to transport. Mol.
Microbiol. 49:869 882.
46. Rath A, Glibowicka M, Nadeau VG, Chen G, Deber CM. 2009. Deter-
gent binding explains anomalous SDS-PAGE migration of membrane
proteins. Proc. Natl. Acad. Sci. U. S. A. 106:1760–1765.
47. Ratledge C, Dover LG. 2000. Iron metabolism in pathogenic bacteria.
Annu. Rev. Microbiol. 54:881–941.
48. R Development Core Team. 2006. R: a language and environment for
statistical computing. R Foundation for Statistical Computing, Vienna,
Austria.
49. Rehm BH, Hancock RE. 1996. Membrane topology of the outer mem-
brane protein OprH from Pseudomonas aeruginosa: PCR-mediated site-
directed insertion and deletion mutagenesis. J. Bacteriol. 178:3346–3349.
50. Richards GR, Goodrich-Blair H. 2009. Masters of conquest and pillage:
Xenorhabdus nematophila global regulators control transitions from viru-
lence to nutrient acquisition. Cell. Microbiol. 11:1025–1033.
51. Schryvers AB, Gonzalez GC. 1990. Receptors for transferrin in patho-
genic bacteria are specific for the host’s protein. Can. J. Microbiol. 36:145–
147.
52. Stabb EV, Ruby EG. 2002. RP4-based plasmids for conjugation between
Escherichia coli and members of the Vibrionaceae. Methods Enzymol. 358:
413–426.
53. Usher KC, O
¨zkan E, Gardner KH, Deisenhofer J. 2001. The plug domain
of FepA, a TonB-dependent transport protein from Escherichia coli, binds
its siderophore in the absence of the transmembrane barrel domain. Proc.
Natl. Acad. Sci. U. S. A. 98:10676–10681.
54. Vivas EI, Goodrich-Blair H. 2001. Xenorhabdus nematophilus as a model
for host-bacterium interactions: rpoS is necessary for mutualism with
nematodes. J. Bacteriol. 183:4687–4693.
55. Volgyi A, Fodor A, Szentirmai A, Forst S. 1998. Phase variation in
Xenorhabdus nematophilus. Appl. Environ. Microbiol. 64:1188–1193.
56. White GF. 1927. A method for obtaining infective nematode larvae from
cultures. Science 66:302–303.
57. Wiener MC. 2005. TonB-dependent outer membrane transport: going
for Baroque? Curr. Opin. Struct. Biol. 15:394 400.
58. Wouts WM. 1980. Biology, life cycle, and redescription of Neoaplectana
bibionis Bovien, 1937 Nematoda: Steinernematidae. J. Nematol. 12:62–72.
Bhasin et al.
1776 jb.asm.org Journal of Bacteriology
... In contrast, the N-terminal ligand-binding domain is specific to hemophilin homologs, making these a distinct structural subgroup of T11SS cargo proteins. We established that, like the hemophilins X. nematophila HrpC and A. baumannii HphA (8,34,35), two other hemophilin homologs, H. haemolyticus Hpl and X. cabanillasii CrpC, rely on a T11SS secretor to reach the extracellular milieu. This establishes T11SS-dependence across all four hemophilin sequence subclusters tested and indicates T11SS-dependent secretion is a hallmark of the entire family. ...
... All strains, plasmids, and primers utilized in this study are described in Supplemental File 3. All cultures were grown in glucose minimal media (34), LB stored in the dark to prevent the formation of oxidative radicals (henceforth dark LB), or glucose minimal media supplemented with 1% dark LB. Plate-based cultures were grown on either LB supple mented with pyruvate to prevent the formation of reactive oxygen radicals (henceforth, LBP or glucose minimal plates) (34). ...
... All cultures were grown in glucose minimal media (34), LB stored in the dark to prevent the formation of oxidative radicals (henceforth dark LB), or glucose minimal media supplemented with 1% dark LB. Plate-based cultures were grown on either LB supple mented with pyruvate to prevent the formation of reactive oxygen radicals (henceforth, LBP or glucose minimal plates) (34). For plasmid-based expression, chemically competent E. coli strain BL21-DE3 (C43) were chosen for ease of transformation and their ability to tolerate expression of membrane proteins (52,53). ...
Article
Full-text available
Cellular life relies on enzymes that require metals, which must be acquired from extracellular sources. Bacteria utilize surface and secreted proteins to acquire such valuable nutrients from their environment. These include the cargo proteins of the type eleven secretion system (T11SS), which have been connected to host specificity, metal homeostasis, and nutritional immunity evasion. This Sec-dependent, Gram-negative secretion system is encoded by organisms throughout the phylum Proteobacteria, including human pathogens Neisseria meningitidis, Proteus mirabilis, Acinetobacter baumannii, and Haemophilus influenzae. Experimentally verified T11SS-dependent cargo include transferrin-binding protein B (TbpB), the hemophilin homologs heme receptor protein C (HrpC), hemophilin A (HphA), the immune evasion protein factor-H binding protein (fHbp), and the host symbiosis factor nematode intestinal localization protein C (NilC). Here, we examined the specificity of T11SS systems for their cognate cargo proteins using taxonomically distributed homolog pairs of T11SS and hemophilin cargo and explored the ligand binding ability of those hemophilin cargo homologs. In vivo expression in Escherichia coli of hemophilin homologs revealed that each is secreted in a specific manner by its cognate T11SS protein. Sequence analysis and structural modeling suggest that all hemophilin homologs share an N-terminal ligand-binding domain with the same topology as the ligand-binding domains of the Haemophilus haemolyticus heme binding protein (Hpl) and HphA. We term this signature feature of this group of proteins the hemophilin ligand-binding domain. Network analysis of hemophilin homologs revealed five subclusters and representatives from four of these showed variable heme-binding activities, which, combined with sequence-structure variation, suggests that hemophilins are diversifying in function. IMPORTANCE The secreted protein hemophilin and its homologs contribute to the survival of several bacterial symbionts within their respective host environments. Here, we compared taxonomically diverse hemophilin homologs and their paired Type 11 secretion systems (T11SS) to determine if heme binding and T11SS secretion are conserved characteristics of this family. We establish the existence of divergent hemophilin sub-families and describe structural features that contribute to distinct ligand-binding behaviors. Furthermore, we demonstrate that T11SS are specific for their cognate hemophilin family cargo proteins. Our work establishes that hemophilin homolog-T11SS pairs are diverging from each other, potentially evolving into novel ligand acquisition systems that provide competitive benefits in host niches.
... The mechanisms by which certain classes of proteins, including lipoproteins, are targeted to and oriented within the outer-membrane are still largely unknown. The newly described type XI secretion system (TXISS), comprising an outer membrane protein (OMP) containing a DUF560 (a domain of unknown function 560), is broadly distributed among proteobacteria and mediates translocation of lipoprotein and a soluble protein cargo across the outer membrane (Heungens et al., 2002;Bhasin et al., 2012;Hooda et al., 2017;Grossman et al., 2021). ...
... An X. nematophila TXISS OMP , NilB, is encoded near an outer membrane lipoprotein NilC on a locus known as Symbiosis Region 1 (SR1) (Heungens et al., 2002;Cowles and Goodrich-Blair, 2004;Bhasin et al., 2012). In X. nematophila, the SR1 locus, which encodes both nilB and nilC, is necessary and sufficient for normal levels of colonization of S. carpocapsae intestines (Heungens et al., 2002;Cowles and Goodrich-Blair, 2008;Chaston et al., 2013). ...
... Previous whole-cell protease-digestion data demonstrated periplasmic orientation of the lipoprotein NilC, based on the observation of protease resistance of NilC in the whole cell but not lysate samples of wild type X. nematophila (Cowles and Goodrich-Blair, 2004). Later, another protease digestion experiment to detect surface NilC was performed on an X. nematophila nilR mutant, in which the absence of the transcription factor NilR causes nilB and nilC expression to be derepressed (Cowles and Goodrich-Blair, 2006;Bhasin et al., 2012). In this analysis, slight shaving of NilC was detected in whole cells, indicating some surface exposure (Bhasin et al., 2012). ...
Article
Full-text available
The only known required component of the newly described Type XI secretion system (TXISS) is an outer membrane protein (OMP) of the DUF560 family. TXISSOMPs are broadly distributed across proteobacteria, but properties of the cargo proteins they secrete are largely unexplored. We report biophysical, histochemical, and phenotypic evidence that Xenorhabdus nematophila NilC is surface exposed. Biophysical data and structure predictions indicate that NilC is a two-domain protein with a C-terminal, 8-stranded β-barrel. This structure has been noted as a common feature of TXISS effectors and may be important for interactions with the TXISSOMP. The NilC N-terminal domain is more enigmatic, but our results indicate it is ordered and forms a β-sheet structure, and bioinformatics suggest structural similarities to carbohydrate-binding proteins. X. nematophila NilC and its presumptive TXISSOMP partner NilB are required for colonizing the anterior intestine of Steinernema carpocapsae nematodes: the receptacle of free-living, infective juveniles and the anterior intestinal cecum (AIC) in juveniles and adults. We show that, in adult nematodes, the AIC expresses a Wheat Germ Agglutinin (WGA)-reactive material, indicating the presence of N-acetylglucosamine or N-acetylneuraminic acid sugars on the AIC surface. A role for this material in colonization is supported by the fact that exogenous addition of WGA can inhibit AIC colonization by X. nematophila. Conversely, the addition of exogenous purified NilC increases the frequency with which X. nematophila is observed at the AIC, demonstrating that abundant extracellular NilC can enhance colonization. NilC may facilitate X. nematophila adherence to the nematode intestinal surface by binding to host glycans, it might support X. nematophila nutrition by cleaving sugars from the host surface, or it might help protect X. nematophila from nematode host immunity. Proteomic and metabolomic analyses of wild type X. nematophila compared to those lacking nilB and nilC revealed differences in cell wall and secreted polysaccharide metabolic pathways. Additionally, purified NilC is capable of binding peptidoglycan, suggesting that periplasmic NilC may interact with the bacterial cell wall. Overall, these findings support a model that NilB-regulated surface exposure of NilC mediates interactions between X. nematophila and host surface glycans during colonization. This is a previously unknown function for a TXISS.
... The DUF560 homolog NilB is a host association and species specificity factor in the nematode symbiont Xenorhabdus nematophila, a proteobacterium in the family Morganellaceae (12)(13)(14). A screen for X. nematophila mutants defective in colonizing Steinernema carpocapsae intestines revealed the nematode intestinal localization (nil) locus (14,15). The nil locus contains the genes nilB and nilC, each of which is independently necessary for colonization of nematodes. ...
... The nil locus contains the genes nilB and nilC, each of which is independently necessary for colonization of nematodes. Biochemical and bioinformatic analyses have established that NilC is an outer membrane-associated lipoprotein, and NilB is an outer membrane b-barrel in the DUF560 family with an ;140-amino-acid periplasmic N-terminal domain that contains tetratricopeptide repeats (15)(16)(17)(18). ...
... TXISS cluster according to environment. Using homology to NilB or Slam proteins, previous work identified a wide distribution of DUF560 proteins within mucosa-associated bacteria (9,14,15). To quantifiably delineate subfamilies within the TXISS, we generated a sequence similarity network (SSN) using the Enzyme Function Initiative toolset (EFI) (21)(22)(23) and annotated it to highlight environmental source or taxonomic grouping of microbes containing DUF560 homologs ( Fig. 2; see also Table S1 in the supplemental material). ...
Article
Full-text available
The microbial constituency of a host-associated microbiome emerges from a complex physical and chemical interplay of microbial colonization factors, host surface conditions, and host immunological responses. To fill unique niches within a host, bacteria encode surface and secreted proteins that enable interactions with and responses to the host and cooccurring microbes.
... DUF560 family proteins are present throughout Proteobacteria, and only a few members have been characterized [13,14,16,17]. One characterized DUF560 homolog, NilB, was identified as a host-association and species-specificity factor in the entomopathogenic nematode symbiont To begin to understand the range of functions of DUF560 proteins in biology, we assessed their distribution, genomic context, and relatedness. ...
... Using either NilB or Slam proteins as bait, previous work had identified a wide distribution of DUF560 proteins in Gram-negative bacteria, seemingly enriched in those associated with animal mucosal surfaces [14, 16,17]. We sought to gain more quantifiable information about whether sub-families of DUF560 homologs exist and whether DUF560 are enriched among mucosal symbionts. ...
... One example is X. nematophila XNC1_0075, encoded adjacent to the DUF560 homolog XNC1_0074 in a TonB-dependent heme-receptor protein/TonB genomic context (Hrp locus). XNC1_0075 is encoded adjacent to XNC1_0074 and in the same orientation, as previously described [17] and it includes a TbpB_B_D "lipoprotein-5" class domain (Fig. 5) [14,17]. However, XNC1_0075 has a SPI-type (rather than SPII) signal sequence and lacks the canonical lipobox necessary for lipidation so is predicted to be a secreted soluble protein. ...
Preprint
Full-text available
In host-associated bacteria, surface and secreted proteins mediate acquisition of nutrients, interactions with host cells, and specificity of tissue-localization. In Gram-negative bacteria, the mechanism by which many proteins cross or become tethered to the outer membrane remains unclear. The d omain of u nknown function (DUF)560 occurs in outer membrane proteins throughout Proteobacteria and has been implicated in host-bacteria interactions and lipoprotein surface exposure. We used sequence similarity networking to reveal three subfamilies of DUF560 homologs. One subfamily includes those DUF560 proteins experimentally characterized to date: NilB, a host-range determinant of the nematode-mutualist Xenorhabdus nematophila , and the s urface lipoprotein a ssembly m odulators Slam1 and Slam2, which facilitate msurface exposure of lipoproteins in Neisseria meningitidis (1, 2). We show that DUF560 proteins from a second subfamily facilitate secretion of soluble, non-lipidated proteins across the outer membrane. Using in silico analysis, we demonstrate that DUF560 gene complement correlates with bacterial environment at a macro level and host association at a species level. The DUF560 protein superfamily represents a newly characterized Gram-negative secretion system capable of lipoprotein surface exposure and soluble protein secretion with conserved roles in facilitating symbiosis. In light of these data, we propose that it be titled the type eleven s ecretion s ystem (TXISS). Importance The microbial constituents of a host associated microbiome are decided by a complex interplay of microbial colonization factors, host surface conditions, and host immunological responses. Filling such niches requires bacteria to encode an arsenal of surface and secreted proteins to effectively interact with the host and co-occurring microbes. Bioinformatic predictions of the localization and function of putative bacterial colonization factors are essential for assessing the potential of bacteria to engage in pathogenic, mutualistic, or commensal activities. This study uses publicly available genome sequence data, alongside experimental results from representative gene products from Xenorhabdus nematophila , to demonstrate a role for DUF560 family proteins in the secretion of bacterial effectors of host interactions. Our research delineates a broadly distributed family of proteins and enables more accurate predictions of the localization of colonization factors throughout Proteobacteria.
... In the complexes of S. carpocapsae and X. nematophila, nematode intestine localization (nil) factors A, B, and C are identified as molecular components that explain the specificity between the nematode and bacteria [10]. NilB and NilC are an outer membrane beta barrel protein and periplasmic lipoprotein, respectively [11,12], suggesting their roles in the molecular interactions with host gut epithelium. ...
... The different letters above standard deviation bars indicate significant differences among means at Type I error = 0.05 (LSD test). 12 ...
Preprint
Full-text available
An entomopathogenic nematode, Oscheius tipulae, was isolated from a soil sample. The identification of this species was supported by morphological and molecular markers. The nematode isolate exhibited pathogenicity against different target insects including lepidopteran, coleopteran, and dipteran insects. The virulence of this nematode was similar to that of a well-known entomopathogenic nematode, Steinernema carpocapsae, against the same insect targets. A comparative metagenomics analysis of these two nematode species predicted the existence of a combined total of 272 bacterial species in their intestines, of which 51 bacterial species were shared between the two nematode species. In particular, the common gut bacteria included several entomopathogenic bacteria including Xenorhabdus nematophila, which is known as a symbiotic bacterium to S. carpocapsae. The nematode virulence of O. tipulae to insects was enhanced by an addition of dexamethasone but suppressed by an addition of arachidonic acid, suggesting that the immune defenses of the target insects against the nematode infection is mediated by eicosanoids, which would be manipulated by the symbiotic bacteria of the nematode. Unlike S. carpocapsae, O. tipulae showed high virulence against dipteran insects including fruit flies, onion flies, and mosquitoes. O. tipulae showed particularly high control efficacies against the onion maggot, Delia platura, infesting the Welsh onion in the rhizosphere in both pot and field assays.
... The globular domain and surface loop 6 play a crucial role in the nematode colonization. Epifluorescence microscopy of these mutants revealed that NilB is necessary at early stages of colonization (Bhasin et al., 2012). (ii) Cj0561c (DUF2860) is a probable membrane fusion protein and contributes to intestinal colonization. ...
Article
Full-text available
OMPdb ( www.ompdb.org ) was introduced as a database for β-barrel outer membrane proteins from Gram-negative bacteria in 2011 and then included 69,354 entries classified into 85 families. The database has been updated continuously using a collection of characteristic profile Hidden Markov Models able to discriminate between the different families of prokaryotic transmembrane β-barrels. The number of families has increased ultimately to a total of 129 families in the current, second major version of OMPdb. New additions have been made in parallel with efforts to update existing families and add novel families. Here, we present the upgrade of OMPdb, which from now on aims to become a global repository for all transmembrane β-barrel proteins, both eukaryotic and bacterial.
... The pBlueXIS1_460109UpDn or pBlueXIS1_460115UpDn construct was cloned into a pKR100 suicide vector; the resulting pKRXIS1_460115 and pKRXIS1_460109 constructs (Table 1) were separately conjugated into the WT X. innexi using E. coli S-17 λpir donor strain. The resulting mutants were first verified by PCR amplification of nilB, which is a Xenorhabdus-specific gene [118]. The position of mutation was also confirmed by PCR amplification of the flanking regions of the inserted kanamycin cassette. ...
Article
Full-text available
Background: Xenorhabdus innexi is a bacterial symbiont of Steinernema scapterisci nematodes, which is a cricket specialist parasite and together the nematode and bacteria infect and kill crickets. Curiously, X. innexi expresses a potent extracellular mosquitocidal toxin activity in culture supernatants. We sequenced a draft genome of X. innexi and compared it to the genomes of related pathogens to elucidate the nature of specialization. Results: Using green fluorescent protein-expressing X. innexi we confirm previous reports using culture-dependent techniques that X. innexi colonizes its nematode host at low levels (~3–8 cells per nematode), relative to other Xenorhabdus-Steinernema associations. We found that compared to the well-characterized entomopathogenic nematode symbiont X. nematophila, X. innexi fails to suppress the insect phenoloxidase immune pathway and is attenuated for virulence and reproduction in the Lepidoptera Galleria mellonella and Manduca sexta, as well as the dipteran Drosophila melanogaster. To assess if, compared to other Xenorhabdus spp., X. innexi has a reduced capacity to synthesize virulence determinants, we obtained and analyzed a draft genome sequence. We found no evidence for several hallmarks of Xenorhabdus spp. toxicity, including Tc and Mcf toxins. Similar to other Xenorhabdus genomes, we found numerous loci predicted to encode non-ribosomal peptide/polyketide synthetases. Anti-SMASH predictions of these loci revealed one, related to the fcl locus that encodes fabclavines and zmn locus that encodes zeamines, as a likely candidate to encode the X. innexi mosquitocidal toxin biosynthetic machinery, which we designated Xlt. In support of this hypothesis, two mutants each with an insertion in an Xlt biosynthesis gene cluster lacked the mosquitocidal compound based on HPLC/MS analysis and neither produced toxin to the levels of the wild type parent. Conclusions: The X. innexi genome will be a valuable resource in identifying loci encoding new metabolites of interest, but also in future comparative studies of nematode-bacterial symbiosis and niche partitioning among bacterial pathogens.
... This lipoprotein, named NilC has been previously shown to be lipidated in vivo, present in the outer membrane and is important for host colonization (Cowles and Goodrich-Blair, 2004). The gene encoding NilB, a putative Slam homolog, is present next to the gene encoding NilC, and is shown to be required for host colonization (Bhasin et al., 2012). Our analysis suggests that NilC is a surface lipoprotein (SLP) that is dependent on the Slam homolog, NilB, for surface display. ...
Article
Full-text available
The surfaces of many Gram-negative bacteria are decorated with soluble proteins anchored to the outer membrane via an acylated N-terminus; these proteins are referred to as surface lipoproteins or SLPs. In Neisseria meningitidis, SLPs such as transferrin-binding protein B (TbpB) and factor-H binding protein (fHbp) are essential for host colonization and infection because of their essential roles in iron acquisition and immune evasion, respectively. Recently, we identified a family of outer membrane proteins called Slam (Surface lipoprotein assembly modulator) that are essential for surface display of neisserial SLPs. In the present study, we performed a bioinformatics analysis to identify 832 Slam related sequences in 638 Gram-negative bacterial species. The list included several known human pathogens, many of which were not previously reported to possess SLPs. Hypothesizing that genes encoding SLP substrates of Slams may be present in the same gene cluster as the Slam genes, we manually curated neighboring genes for 353 putative Slam homologs. From our analysis, we found that 185 (~52%) of the 353 putative Slam homologs are located adjacent to genes that encode a protein with an N-terminal lipobox motif. This list included genes encoding previously reported SLPs in Haemophilus influenzae and Moraxella catarrhalis, for which we were able to show that the neighboring Slams are necessary and sufficient to display these lipoproteins on the surface of Escherichia coli. To further verify the authenticity of the list of predicted SLPs, we tested the surface display of one such Slam-adjacent protein from Pasteurella multocida, a zoonotic pathogen. A robust Slam-dependent display of the P. multocida protein was observed in the E. coli translocation assay indicating that the protein is a Slam-dependent SLP. Based on multiple sequence alignments and domain annotations, we found that an eight-stranded beta-barrel domain is common to all the predicted Slam-dependent SLPs. These findings suggest that SLPs with a TbpB-like fold are found widely in Proteobacteria where they exist with their interaction partner Slam. In the future, SLPs found in pathogenic bacteria can be investigated for their role in virulence and may also serve as candidates for vaccine development.
Article
The surface of many Gram-negative bacteria contains lipidated protein molecules referred to as surface lipoproteins or SLPs. SLPs play critical roles in host immune evasion, nutrient acquisition and regulation of bacterial stress response, and have been extensively studied as vaccine antigens. The aim of this review is to summarize the recent studies that have investigated the biosynthetic and translocation pathways used by different bacterial species to deliver SLPs to the surface. We will specifically focus on Slam, a novel outer membrane protein first discovered in pathogenic Neisseria sp., that is involved in translocation of SLPs across the outer membrane.
Article
Full-text available
Ore mineral and host lithologies have been sampled with 89 oriented samples from 14 sites in the Naica District, northern Mexico. Magnetic parameters permit to charac- terise samples: saturation magnetization, density, low- high-temperature magnetic sus- ceptibility, remanence intensity, Koenigsberger ratio, Curie temperature and hystere- sis parameters. Rock magnetic properties are controlled by variations in titanomag- netite content and hydrothermal alteration. Post-mineralization hydrothermal alter- ation seems the major event that affected the minerals and magnetic properties. Curie temperatures are characteristic of titanomagnetites or titanomaghemites. Hysteresis parameters indicate that most samples have pseudo-single domain (PSD) magnetic grains. Alternating filed (AF) demagnetization and isothermal remanence (IRM) ac- quisition both indicate that natural and laboratory remanences are carried by MD-PSD spinels in the host rocks. The trend of NRM intensity vs susceptibility suggests that the carrier of remanent and induced magnetization is the same in all cases (spinels). The Koenigsberger ratio range from 0.05 to 34.04, indicating the presence of MD and PSD magnetic grains. Constraints on the geometry of the intrusive source body devel- oped in the model of the magnetic anomaly are obtained by quantifying the relative contributions of induced and remanent magnetization components.
Article
Full-text available
The monogeneric nematode families Steinernematidae (Steinernema) and Heterorhabditidae (Heterorhabditis), mutually associated with the pathogenic bacteria Xenorhabdus, are similar in their actions. The free-living, non-feeding infective juveniles possess attributes of both insect parasitoids or predators (they have chemoreceptors and are motile) and microbial pathogens (virulence, high reproductive capacity, numerical but no functional response to host population changes). Their potential as agents of biological control are accentuated by their dispersability using spray equipment, their compatibility with many pesticides, and their amenability to genetic selection. Major sections of this review are on: taxonomy; biology of the nematode-bacterium complex; host range (insects are killed so rapidly that highly adapted host-parasite relationships do not form); behaviour; ecology (including dispersal and host finding, survival, interspecific competition, and recycling and epizootiology); genetics; commercialisation; and efficacy. -P.J.Jarvis
Article
Full-text available
Most infective juveniles of DD-136 (Neoaplectana, Steinernematidae) were found to contain cells of Achromobacter nematophilus Poinar & Thomas in the ventricular portion of their intestinal lumen. In two instances, anterior intestinal cells of the infective juveniles were found to contain bacterial cells, presumably those of A. nematophilus. When the infective stage penetrated into the body cavity of a suitable host, the bacteria were released through the anus and multiplied rapidly in the host's body, resulting in a fatal septicemia.
Article
Full-text available
In the laboratory, mortality rates of the agromyzid leafminer larvae, Liriomyza trifolii (Burgess), ranged from 48 to 98% by 20 strains and/or species of steinernematid and heterorhabditid nematodes. In the greenhouse, abamectin provided superior control of larval leafminers, killing 100% of them as compared with Steinernema carpocapsae (Weiser) All strain (24 to 43% leafminer mortality) or S. carpocapsae Liriomyza-selected strain (8 to 44% leafminer mortality); the maximum relative humidity (r.h.) ranged between 81 and 91% and the minimum r.h. between 50 and 70%. In the foghouse under high r.h., the commercially available All strain and the Hawaiian isolate of S. feltiae (Filipjev) MG-14 strain caused 69 and 67% mean mortality, respectively. There was a significant correlation (P<0.01) between nematode mortality of leafminers and r.h., including the mean, standard deviation, and minimum r.h. during the 48 h after treatment. Average r.h. >92% with a standard deviation of <9% r.h. and a minimum of 72% r.h. provided S. carpocapsae All strain mortality rates of leafminers >65%. The major constraint against the use of nematodes against leafminers in the foliar environment is low r.h. The use of nematodes against L. trifolii can be successful if the r.h. remains high and if nematodes enter leafmines before desiccation, and the nematodes should be integrated with chemical insecticides such as abamectin to manage pesticide resistance in L. trifolii.
Article
The bacterium Achromobacter nematophilus, which has been isolated from two populations of nematodes belonging to the genus Neoaplectana, was found to be extremely lethal to larvae of Galleria mellonella when injected into the hemocoel. Although producing no obvious effects if fed per os, when one to three bacterial cells were introduced into the body cavity, the insects usually died. In nature, this bacterium occurs in the body of insects attacked by certain neoaplectanid nematodes and, between periods of insect infection, inside the gut of infective stage nematodes. The cells are very fastidious and appear to have a very transient existence in soil or water.
Article
The levels of DNA relatedness for a broad sample ofXenorhabdus strains isolated from different species of entomopathogenic nematodes (Steinernematidae and Heterorhabditidae) and from different geographical sources were estimated by the hydroxyapatite method. The level of DNA-DNA relatedness for the two phases of each isolate tested was not significantly different from lo%, demonstrating unequivocally that the phase variation demonstrated by all Xenorhabdus spp. is not due to contamination. The isolates of the described Xenorhabdus species coalesced into different DNA relatedness groups, confirming that Xenorhubdus nemuto- philus, Xenorhabdus bovienii, Xenorhubdus poinarii, and Xenorhabdus beddingii, defined on the basis of phenotypic differences, are valid species. The symbiont of Steinernemu intermedia also coalesced with the X. bovienii isolates. This was the only symbiont of seven recently described and unamed Steinernema spp. (including Steinernemu ritteri, Steinemem rara, and Steinernemu anomali) that formed a group with any of the previously described Xenorhabdus species; new species descriptions are required to accommodate the other taxa, but too few isolates were available to allow satisfactory descriptions of them. The DNA relatedness data also showed that the bacteria currently classified as Xenorhabdus luminescens are significantly different from all other Xenorhubdus strains. These data strongly support indications from previous studies of phenotypic characteristics, cellular fatty acids, and DNA relatedness that X. luminescens should be classified as a separate genus, A new genus, Photorhabdus, with an amended description of the type species, Photorhabdus luminescens, is proposed.
Article
Strains of Xenorhabdus nematophilus and Photorhabdus luminescenswere genetically marked with kanamycin resistance and the xylE gene to aid theirdetection in water and soil. Following release in river water, cells declined to undetectable levelsin 6 d. In sterile river water, this decline was enhanced with cells detectable for only 2 d. In sterileMilli-Q purified water, the decline was slower than in either sterile or non-sterile river water.Survival in soil was also restricted with cells only detectable for 7 d. These experiments indicatedthat both X. nematophilus and P. luminescens have limited survival orcompetitive abilities in these environments. The faster decline of populations in sterile river waterwas unexpected, and the possible formation of specialized survival stages was investigated. Insterile water, a non-culturable but viable population of cells was detected, indicating that cellsmay survive longer than anticipated in the environment and remain undetectable using standardmicrobiological methods. The implications of this work to the use of these strains in biologicalcontrol and the release of genetically-modified micro-organisms is discussed.