ABC Pseudomonas

download ABC Pseudomonas

of 14

Transcript of ABC Pseudomonas

  • 8/12/2019 ABC Pseudomonas

    1/14

    Molecular Microbiology (2003) 49(4), 905918 doi:10.1046/j.1365-2958.2003.03615.x

    2003 Blackwell Publishing Ltd

    Blackwell Science, LtdOxford, UKMMIMolecular Microbiology 1365-2958Blackwell Publishing Ltd, 2003494905918Original ArticleP.fluorescens biofilm formationS.M. Hinsa, M. Espinosa-Urgel, J. L. Ramos and G. A. OToole

    Accepted 7 May, 2003. *For correspondence. E-mail

    [email protected]; Tel. (+1) 603 650 1248; Fax (+1)603 650 1318.

    Transition from reversible to irreversible attachmentduring biofilm formation by Pseudomonas fluorescens

    WCS365 requires an ABC transporter and a large

    secreted protein

    Shannon M. Hinsa,

    1

    Manuel Espinosa-Urgel,

    2

    Juan L. Ramos

    2

    and George A. OToole

    1

    *

    1

    Department of Microbiology and Immunology, Dartmouth

    Medical School, Hanover, NH, USA.

    2

    Department of Plant Biochemistry and Molecular and

    Cell Biology, Estacion Experimental del Zaidin. CSIC.,

    Profesor Albareda, 1., 18008 Granada, Spain.

    SummaryWe report the identification of an ATP-binding cas-

    sette (ABC) transporter and an associated large cell-

    surface protein that are required for biofilm formation

    by Pseudomonas fluorescens

    WCS365. The genes

    coding for these proteins are designated lap

    for l

    arge

    a

    dhesion p

    rotein. The LapA protein, with a predicted

    molecular weight of ~~~~

    900 kDa, is found to be loosely

    associated with the cell surface and present in the

    culture supernatant. The LapB, LapC and LapE pro-

    teins are predicted to be the cytoplasmic membrane-

    localized ATPase, membrane fusion protein and outer

    membrane protein component, respectively, of anABC transporter. Consistent with this prediction,

    LapE, like other members of this family, is localized

    to the outer membrane. We propose that the lapEBC

    -

    encoded ABC transporter participates in the secre-

    tion of LapA, as strains with mutations in the lapEBC

    genes do not have detectable LapA associated with

    the cell surface or in the supernatant. The lap

    genes

    are conserved among environmental pseudomonads

    such as P. putida

    KT2440, P. fluorescens

    PfO1 and P.

    fluorescens

    WCS365, but are absent from pathogenic

    pseudomonads such as P. aeruginosa

    and P. syrin-

    gae

    . The wild-type strain of P. fluorescens

    WCS365

    and its lap

    mutant derivatives were assessed for their

    biofilm forming ability in static and flow systems. The

    lap

    mutant strains are impaired in an early step in

    biofilm formation and are unable to develop the

    mature biofilm structure seen for the wild-type bacte-

    rium. Time-lapse microscopy studies determined that

    the lap

    mutants are unable to progress from revers-

    ible (or transient) attachment to the irreversible

    attachment stage of biofilm development. The lap

    mutants were also found to be defective in attachment

    to quartz sand, an abiotic surface these organisms

    likely encounter in the environment.

    Introduction

    In natural settings, bacteria are most often found associ-ated with surfaces in communities known as biofilms, and

    not in the planktonic state (Costerton et al

    ., 1995; Davey

    and OToole, 2000). The formation of biofilms by

    pseudomonads has been proposed to occur as a series

    of regulated steps (OToole et al

    ., 2000a). First, flagellar-

    mediated motility may be required for a bacterium to swim

    toward a surface and to initiate reversible (or transient)

    attachment (Korber et al

    ., 1994; OToole and Kolter,

    1998a,b). A subpopulation of transiently attached bacteria

    become irreversibly attached to the surface to first form a

    monolayer, which is followed by the formation of small

    microcolonies (Zobell, 1943; Marshall et al

    ., 1971; vanLoosdrecht et al

    ., 1990; Jensen et al

    ., 1992; Fletcher,

    1996). The microcolonies develop into a mature biofilm

    with an architecture that is typically characterized by mac-

    rocolonies separated by fluid-filled channels (Tolker-

    Nielsen et al

    ., 2000). It is believed that these channels

    transport nutrients and oxygen to the bacteria and aid in

    waste removal (Costerton et al

    ., 1995; Davey and OToole

    2000). Other characteristics of a mature biofilm include

    production of an exopolysaccharide matrix and increased

    antimicrobial resistance (Costerton et al

    ., 1995; Mah and

    OToole, 2001).

    Pseudomonas fluorescens

    WCS365, a natural soil iso-

    late that is employed as a biological control agent against

    plant pathogenic fungi (Geels and Schippers, 1983;

    Simons et al

    ., 1996), has been used to study the molec-

    ular genetic basis of biofilm formation. Previous work has

    shown that a site-specific recombinase, a two component

    regulatory system, the synthesis of certain amino acids,

    the O-antigen of lipopolysaccharide, and type IV pili are

    important for P. fluorescens

    WCS365 to colonize tomato

    roots (Simons et al

    ., 1997; Dekkers et al

    ., 1998a,b,c;

  • 8/12/2019 ABC Pseudomonas

    2/14

    906

    S. M. Hinsa, M. Espinosa-Urgel, J. L. Ramos and G. A. OToole

    2003 Blackwell Publishing Ltd, Molecular Microbiology

    , 49

    , 905918

    Camacho-Carbajal, 2001). However, in their natural envi-

    ronment, bacteria are also likely to adhere to abiotic sur-

    faces such as soil particles.

    Transposon generated mutations that render P. fluore-

    scens

    WCS365 defective for attachment to a variety of

    abiotic surfaces, both hydrophobic (plastic) and hydro-

    philic (glass) were identified previously (OToole and

    Kolter, 1998b). Here we report the characterization of one

    class of these biofilm-defective mutants. We have

    identified an ATP-binding cassette transporter and a large,

    cell-surface associated protein that are required for P.

    fluorescens

    biofilm formation on abiotic surfaces in both

    static and flow cell systems. These genes are also

    required for robust biofilm formation on quartz sand, which

    serves as a model for surfaces typically encountered by

    soil pseudomonads. Our analyses suggest that the genes

    encoding this ABC transporter are conserved among

    sequenced soil pseudomonads, but are absent from

    pathogenic Pseudomonas

    strains. We discuss possible

    roles for this ABC transporter and the cell-surface asso-

    ciated protein in biofilm development.

    Results

    Initial molecular characterization of mutants defective in

    biofilm formation

    Previous experiments had identified a set of transposon

    mutations in P. fluorescens

    that render these strains

    unable to form a biofilm (OToole and Kolter, 1998b). The

    biofilm-defective mutant strains fell into two broad classes

    based on their ability to be rescued by changing growth

    conditions. Class I mutants could be rescued by growthin medium supplemented with certain amino acids,

    organic acids, and/or exogenous iron. Class II mutants

    were unable to make a biofilm under any growth condition

    tested (OToole and Kolter, 1998b). The studies here focus

    on this second class of mutants.

    We identified the genes disrupted by the transposon

    insertion in Class II mutant strains by determining the

    DNA sequence flanking the transposon, either through

    arbitrary-primed PCR or by sequencing the region adja-

    cent to a cloned transposon fragment (OToole et al

    .,

    1999). These DNA sequences were then compared with

    sequence from the P. putida

    KT2440 and P. fluorescens

    PfO1 genome projects. Because the P. putida

    KT2440

    genome is annotated (Nelson et al

    ., 2002) it was used to

    predict open reading frames as well as to assign putative

    functions to the genes disrupted in the P. fluorescens

    WCS365 mutants.

    Eight transposon mutants were analysed, and based on

    extensive sequencing of the DNA flanking the transposon

    insertions (Fig. 1A and data not shown), all transposons

    were found to map to genes located in close proximity to

    each other on the chromosome. Four independent trans-

    poson insertions mapped to an open reading frame des-

    ignated lapA

    . This open reading frame had been initially

    identified in strain mus-24

    , a transposon mutant of P.

    putida

    KT2440 defective in adhesion to corn seeds

    (Espinosa-Urgel et al

    ., 2000). The four transposon inser-

    tions mapping to lapA

    in P. fluorescens

    WCS365 (

    lapA18,

    51, 53

    and 62

    ) are located close to the 5

    end of the gene

    (corresponding to Domain 2 of the protein, Fig. 1B),

    whereas the insertion in P. putida

    mutant mus-24

    is close

    to the 3

    end of the gene (corresponding to Domain 4 of

    the protein, Fig. 1B). The remaining four mutations anal-

    ysed mapped to an adjacent gene cluster we have desig-

    nated lapEBC

    . The lapB

    gene is defined by one

    transposon insertion just 5

    of the start codon (

    lapB84

    )

    and a second transposon located in the middle of the

    gene (

    lapB52

    ). The lapC

    gene (

    lapC87

    ) and the lapE

    gene

    (

    lapE83

    ) are each defined by one transposon insertion,

    the lapC

    insertion is just 5

    of the start codon of lapC

    ,

    while the lapE

    insertion is in the middle of the gene. The

    gene order as shown in Fig. 1A was confirmed in P. fluo-rescens

    WCS365 by either sequencing across the junc-

    tions of genes or using PCR with primers whose design

    was based on the P. fluorescens

    WCS365 DNA sequence

    (data not shown).

    The lap

    genes are required for biofilm formation

    To further define the role of the lap

    genes in biofilm for-

    mation, and to assess the defects in the various lap

    mutant alleles, we carried out the detailed analysis of

    biofilm formation.

    To characterize the kinetics of biofilm formation weperformed a time-course study. The extent of biofilm

    formation was determined by measuring crystal violet

    (CV)-stained biomass accumulating on the walls of the

    microtitre dish over 24 h (OToole et al

    ., 1999). The wild-

    type strain reaches maximum biofilm formation in this

    assay by 10 h after which the extent of biofilm formed

    decreases, then remains steady, until the end of the assay

    period at 24 h (Fig. 2A). The lapB52

    and lapE83

    mutants

    were deficient in attachment over the entire 24 h period

    (Fig. 3A); similar biofilm results were seen for the lapA51

    and lapC87

    mutants (data not shown). These data dem-

    onstrated that the lap

    mutant strains were not simply

    delayed in the initiation of biofilm formation. The plank-

    tonic growth of all these lap

    mutants was identical to the

    wild type (not shown). These results are consistent with

    those for the P. putida lapA (mus-24

    ) mutant that is also

    unable to form a biofilm on plastic or glass, and has no

    planktonic growth defect (Espinosa-Urgel et al

    ., 2000).

    Several approaches were utilized to demonstrate that

    the mutations identified above caused the defects in bio-

    film formation. First, independent mutations in the lap

  • 8/12/2019 ABC Pseudomonas

    3/14

    P. fluorescens biofilm formation

    907

    2003 Blackwell Publishing Ltd, Molecular Microbiology

    , 49

    , 905918

    genes conferred identical phenotypes in several assays,

    providing strong genetic evidence that these mutations

    were responsible for the observed biofilm defects. Sec-

    ond, generalized transduction was utilized to mobilize one

    or more alleles in each lap

    gene into a wild-type genetic

    background. Of the over 40 transductants assayed, all

    conferred the documented antibiotic resistance and bio-

    film phenotypes of the parental strains, demonstrating

    100% linkage between the transposon insertions and the

    observed phenotypes. Finally, numerous attempts were

    made to clone the lapEBC

    region into various vectors

    (pGEM, pUCP18, pSMC32 and pME6000), however, we

    were unable to successfully clone this locus. These data

    suggest that providing the lapEBC

    genes in multiple cop-

    ies may be toxic to the cell. Despite the inability to perform

    complementation assays, the genetic data presented here

    demonstrate that the lap

    genes are required for biofilm

    formation by P. fluorescens

    WCS365.

    Monitoring early attachment in a static system with

    phase-contrast microscopy

    Microscopic analysis of plastic tabs confirmed the results

    of the CV assay presented above. To visualize the attach-

    ment during the early stages of biofilm formation, the wild-

    type and mutant strains lapA51

    and lapB84were allowed

    Fig. 1. Analysis of the lapgenes, the lapchromosomal region and flanking genes.A. Organization of the lapchromosomal region. Shown are the lapgenes and, where known, the predicted flanking genes. The purple and greenarrows represent genes coding for probable regulators, yellow and brown arrows represent genes coding for hypothetical proteins, and the aquaarrows represent genes coding for a putative deoxygenase. The vertical broken line indicates a gene not adjacent to the lapregion on thechromosome. The organization of the lapregions in P. fluorescensWCS365 and P. fluorescensPfO1 is similar. The lapregion of both P. fluorescens

    strains is similar to that of P. putidaKT2440, except the P. putida lapgenes are inverted in relation to the flanking ORFs and lapEis separatedfrom the rest of the lapgenes.B. The LapA protein. The structure of LapA, its four domains, and significant features are shown. The yellow arrows represent lapAmutants inP. fluorescensWCS365, whereas the red arrow represents the lapA(mus-24) mutant in P. putidaKT2440 (Espinosa-Urgel et al., 2000). Thecircles in the fourth domain represent putative calcium-binding domains.

    C. Consensus sequence of repeats domains. Shown at the top of this panel is the consensus sequence of the 100 amino acid repeats of Domain2. In blue are the positions that vary in one of the repeats, the red residues correspond to amino acids that vary in two to four of the repeats,and all the other residues (black) are identical in all repeats. The consensus sequence for Domain 3 is shown in the lower portion of this panel.The black residues are conserved in 85% of the repeats, the blue in 6585%, and red indicates the amino acid shown is found in 3065% of the

    repeats at that position.

  • 8/12/2019 ABC Pseudomonas

    4/14

    908 S. M. Hinsa, M. Espinosa-Urgel, J. L. Ramos and G. A. OToole

    2003 Blackwell Publishing Ltd, Molecular Microbiology, 49, 905918

    to attach to plastic tabs for up to 5 h (described in the

    Experimental procedures) then examined by phase-

    contrast microscopy. At early time points (less than 1 h)

    there was no difference in attachment between the wildand the lap mutants. For example, in a representative

    experiment assessing attachment at 40 min, an average

    of 134 (40) wild-type cells were attached per field viewed

    by phase-contrast microscopy compared to 135 (34)

    lapCmutant bacteria per field (eight fields were analysed

    for each strain). However, by 5 h postinoculation, there

    was a clear difference in biofilm formation between the

    wild-type and lapmutants. The wild type attached to the

    surface and formed organized microcolonies comprised

    of hundreds of cells by 5 h (Fig. 2B). In contrast, the

    lapA51and lapB84strains formed only small clusters of

    bacteria (520 cells), but did not establish the larger

    microcolonies typical of the wild type. Similar results were

    obtained for the lapE83 and lapC87 mutants (data not

    shown).

    To determine if lapmutants that attached to the tabs at

    later time-points were due to a secondary mutation that

    rescued the lapmutant defect, these cells were removed

    from the tab by sonication and tested for biofilm formation.

    Upon retesting, all the lapmutant cells removed from the

    tabs still had a biofilm formation deficiency identical to that

    of the original strains tested (data not shown). Therefore,

    the residual adherence displayed by the lapmutants was

    unlikely to be caused by the accumulation of a second,

    compensatory mutation.

    Analysis of biofilm formation in a flow cell

    To analyse development of a mature biofilm, we grew the

    wild type and the mutants in a flow cell system. Flow cells

    provide a constant influx of fresh nutrients, thus sustaining

    the continued development of the biofilm over many days.

    Biofilms grown in the flow cell form the characteristic

    architecture comprised of large macrocolonies sur-

    rounded by fluid-filled channels after 12 days of

    incubation.

    We examined biofilm formation at 1, 2 and 3 days. As

    shown in Fig. 3A, the wild type bacteria established a

    dense monolayer of cells by day 1, and then began devel-

    oping microcolonies and eventually macrocolonies by

    days 2 and 3. The macrocolonies formed by the wild-type

    strain were visible by the naked eye throughout the flowchamber by day 3 (Fig. 3C). In contrast, the lapB52,

    lapA51,and lapE83mutants showed a severe defect in

    attachment to the surface on day 1 (Fig. 3A). The surface

    of the flow cell was only sparsely covered at this time

    point. By day 2, the mutant bacteria had formed some

    small microcolonies, along the edge of the flow cell, which

    is generally subjected to slower medium flow. The few

    microcolonies formed on day 2 continue to develop and

    by day 3 formed some small macrocolonies. Similar

    results to those observed for the lap mutants shown in

    Fig. 3 were also observed for lapC87 (data not shown).

    The lapA (mus-24) mutant of P. putida KT2440 is alsodefective in biofilm formation in a flow cell system with an

    architecture similar to the P. fluorescens WCS365 lap

    mutants (data not shown).

    Quantitative analysis of biofilm structure

    The images in Fig. 3A show a striking difference in the

    architecture of the biofilms formed by the wild-type strain

    and the lap mutants. To quantify the biofilm formed by

    each strain we utilized the COMSTAT program (Heydorn

    et al., 2000a). COMSTATconverts the digital information of

    images (i.e. shown in Fig. 3) into quantitative parameters

    representing various aspects of biofilm architecture.

    The information required for calculating the quantitative

    parameters of the biofilm is acquired by obtaining optical

    sections in the z-plane of the biofilms, followed by decon-

    volution of these images with the OpenLab software

    package. Biofilms grown for two days in the flow cell were

    chosen in order to capture the initial stages of biofilm

    maturation. Twelve image series in the z-plane (a z-

    series) for each strain effectively captured the heteroge-

    Fig. 2. Monitoring early biofilm formation.A. The kinetics of biofilm formation. This figure illustrates formation

    of a biofilm at the airmedium interface over a 24 h period. Thesurface attached cells were stained with crystal violet, the stain sol-ubilized in ethanol and the absorbance at 550 nm determined (Y-

    axis). Legend: wild type, open squares; lapB52, open circles; lapE83,filled diamonds.B. Monitoring early biofilm formation by phase-contrast microscopy.To visualize the bacterial attachment to plastic, bacteria were incu-

    bated in the presence of a plastic tab (~3 3 mm) for 5 h at roomtemperature, the tabs washed, and attachment observed usingphase-contrast microscopy at 1400magnification.

  • 8/12/2019 ABC Pseudomonas

    5/14

    P. fluorescens biofilm formation 909

    2003 Blackwell Publishing Ltd, Molecular Microbiology, 49, 905918

    neous structure of each biofilm. Each z-series involved

    capturing an image every 0.5 mm starting at the attach-

    ment substratum and moving up to a final height of 50 mm

    above the surface. The results of the COMSTAT analysis

    are displayed in Table 1. The average thickness of the

    biofilm formed by the wild type (10.8 mm) was ~ 30-fold

    greater than the biofilm formed by the mutant (0.37 mm).

    The bio-volume analysis also demonstrated a similar

    trend with the average volume of wild type at 9.93 mm3

    m-1m2and the mutant registering 0.33 mm3m-1 m2. Further-

    more, the wild type covered approximately 34% of the

    substratum while the mutant occupied only 5%. The sur-

    face to volume ratio was calculated as a measure of the

    amount of biomass exposed to media. The wild type aver-

    aged 0.17 mm2 m-1m3 and the mutant 0.64 mm2m-1m3.

    Large structured biofilms like those formed by the wild

    type tend to produce a lower surface to volume ratio than

    do individual cells. These quantitative data are consistent

    with the biofilm formation phenotypes revealed by the

    microscopic images of initial attachment and later biofilm

    development for wild-type and mutant strains.

    Fig. 3.Monitoring biofilm development in flowcells.A. Flow cell-grown biofilms. The biofilm formedby wild type, lapA51, lapB52and lapE83strainsat days 1, 2 and 3 is shown. Day 1 shows top-

    down, phase-contrast images at 1400magni-fication. The day 2 and 3 images are top-down,epifluorescent micrographs at 230magnification.

    B. Enlarged view of initial attachment. Phase-

    contrast images captured every 3 s (at 4 h postinoculation) show individual cells attaching tothe surface for the lapC87mutant strain and forthe wild type. The blue arrow points to a bacte-

    rium that is standing on end, and viewed fromend-on appears as a dot. This bacterium rep-resents a cell that is in the initial reversibleattachment phase. The red arrows point to a

    bacterium that moves back and forth laterallywhile one end remains attached. In contrast,the wild-type bacteria are irreversibly attachedto the surface and thus remain fixed in place.C. Macroscopic macrocolonies. Shown is a

    picture of the flow cell chamber at day 3. Thewild-type strain has filled the chamber withmacrocolonies that are visible to the naked eye.Few or no macrocolonies are visible in the

    channels of the flow cell containing mutantstrains.

    Table 1.COMSTAT: Quantitative analysis of biofilm structure.

    Parameter measured Wild type lapC87

    Average thickness (mm) 10.8 (8.50) 0.37 (0.33)Bio-volume (mm3m-1m2) 9.93 (7.54) 0.33 (0.2)Substratum coverage (%) 33.6 (17.6) 4.9 (3.4)Surface to volume ratio (mm2m-1m3) 0.17 (0.10) 0.64 (0.41)

    The standard deviation for each value is shown in parentheses.

  • 8/12/2019 ABC Pseudomonas

    6/14

    910 S. M. Hinsa, M. Espinosa-Urgel, J. L. Ramos and G. A. OToole

    2003 Blackwell Publishing Ltd, Molecular Microbiology, 49, 905918

    The lapmutants are defective for the transition from

    reversible to irreversible attachment during biofilm

    development

    Microscopy observations from the static and flow cell

    experiments suggested that the lapmutants were defec-

    tive in an early stage of biofilm development. To further

    characterize the biofilm formation defect of the lap

    mutants we compared the bacterial attachment of the

    lapCmutant to the wild type over the first 8 h of biofilm

    development in a flow cell using time-lapse, phase-

    contrast microscopy. Images were acquired every 3 s over

    a period of 5 min. A series of representative images illus-

    trating attachment of the wild type and the lapCmutant

    are shown in Fig. 3B, and the time-lapse movies have

    been posted on the web (http://www.dartmouth.edu/

    ~gotoole/hinsamovies/hinsamovies.html).

    Both the lapC mutant and the wild type are able to

    anchor one pole of the cell to the surface (polar attach-

    ment). Bacteria attached in this fashion are known as

    reversibly attached because they can readily detach fromthe surface. The wild-type bacteria eventually become

    firmly anchored to the surface along the long axis of the

    cell in a process referred to as irreversible attachment

    (Zobell, 1943; Marshall et al., 1971; van Loosdrecht et al.,

    1990; Jensen et al., 1992; Fletcher, 1996). The bottom

    panel of Fig. 3B shows a field of approximately 10 wild-

    type bacteria which have adhered and remain unmoving

    over the course of the 12 s in which these images were

    captured. Time-lapse microscopy over a period of 5 min

    shows that the wild type is capable of prolonged stable

    interactions with the surface (see web site above). These

    data suggest that the wild-type strain is able to makestable interactions with the surface even under conditions

    of flow. Furthermore, a majority of the wild-type bacteria

    are attached to the surface across the long axis of their

    cell body (which serves here as the functional definition

    of irreversible attachment).

    In contrast to the development observed for the wild-

    type strain, the majority of lapCmutant bacteria appear

    unable to progress to the irreversible attachment phase of

    biofilm formation. As shown in Fig. 3B and the web sup-

    plement, many of the lapC mutant bacteria are still

    anchored by their pole at this time-point and can still be

    observed moving, spinning rapidly, and/or frequently

    detaching from the cell surface. After extended incubation

    in the flow cell, the lapC mutant is able to form small

    microcolonies, possibly at sites where a cell was occa-

    sionally able to tightly attach during initial colonization.

    Similar results were observed for the other lapmutants

    (data not shown).

    We also quantified the extent of reversible vs. irrevers-

    ible attachment for the wild type and the lapCmutant. By

    8 h postinoculation 94% (4.5%) of the wild-type bacteria

    were irreversibly attached to the surface. That is, these

    bacteria were attached by the long axis of the cell body

    and did not move over the course of the 5 min time period

    in which the images for the time-lapse movies were

    acquired. Only 6% (4.5%) of wild-type cells were

    attached by one pole and continued to move during this

    period. In contrast, for the lapCmutant, only 12% (4%)

    of were irreversibly attached, whereas 88% (4%) were

    attached by one pole and continued to move for at least

    some period during this 5 min interval.

    Sequence analysis of the lapgenes

    To begin to elucidate the mechanism underlying the bio-

    film defect of the lapmutants, we performed a detailed

    analysis of the predicted proteins encoded by these

    genes. The complete predicted lapA gene, as deduced

    from the genome sequence of P. putidaKT2440, is 26 kb

    long and would correspond to one of the largest bacterial

    proteins (8682 amino acids) with an estimated molecular

    weight of ~888 kDa and a predicted pI of 4.1. Thus, wehave named this gene lapA, for large adhesion protein

    (lapa is also the Spanish name for limpet, a mollusk that

    lives on seashore rocks and sticks firmly to the rock sur-

    face when disturbed). At this stage of the sequencing

    project of P. fluorescens it has not been possible to

    assemble the complete lapAgene, however, as detailed

    below, the size and structure of the protein appears to be

    very similar to its P. putidacounterpart.

    Four domains can be clearly distinguished in LapA, two

    of them being composed of long multiple repeats, which

    constitute more than three-quarters of the total length of

    the protein (Fig. 1C). Because no significantly similar pro-teins of known function could be found in the databases

    when LapA was compared as a whole we analysed each

    domain separately. Domain 1, encompassing the first 277

    amino acids, contains a predicted non-cleavable N-

    terminal signal sequence and a transmembrane region

    (PSORTprogram, http://psort.ims.u-tokyo.ac.jp). The func-

    tion of Domain 1 is unknown, but shows some limited

    sequence similarity (29% identity) with the N-terminal part

    of the RTX toxin of the fish pathogen Aeromonas salmo-

    nicida(GenBank accession #AF218037). The role of the

    N-terminal domain of this RTX toxin has not been eluci-

    dated. Domain 2, from amino acids 2781178, comprises

    nine quasi-perfect repeats of 100 aa stretches (Fig. 1C).

    Sequence similarity was found with the surface protein

    Bap, from Staphylococcus aureus(Cucarella et al., 2001),

    a protein that contains 13 nearly identical repeats of 86

    amino acids and is involved in biofilm formation by S.

    aureus. Domain 2 also shows structural similarity with the

    outer surface protein A of Borrelia burgdorferi(predicted

    with 123D+, http://123d.ncifcrf.gov/; (Alexandrov et al.,

    1995). Separated from Domain 2 by 18 aa is Domain 3,

    http://www.dartmouth.edu/http://psort.ims.u-tokyo.ac.jp/http://123d.ncifcrf.gov/http://123d.ncifcrf.gov/http://psort.ims.u-tokyo.ac.jp/http://www.dartmouth.edu/
  • 8/12/2019 ABC Pseudomonas

    7/14

    P. fluorescens biofilm formation 911

    2003 Blackwell Publishing Ltd, Molecular Microbiology, 49, 905918

    which is also a large repetitive region, spanning 6400 aa,

    organized in 29 imperfect repeats of 218225 aa

    (Fig. 1C). Some sequence similarity was found with the

    surface-associated adhesin CshA of the Gram-positive

    oral bacterium Streptococcus gordonii (McNab et al.,

    1994). CshA also shows a repetitive structure (13 repeats

    of 101 amino acids) and is an essential element for oral

    cavity colonization, participating in co-aggregation of S.

    gordonii with another oral microorganism, Actinomyces

    naeslundii (McNab et al., 1994; 1999). Domain 4 (1087

    aa) contains several Ca2+-binding motifs similar to those

    identified in haemolysins and other secreted proteins

    known to participate in bacterialeukaryotic interactions

    (Economou et al., 1990). This sequence analysis sug-

    gests that LapA may be a cell surface protein working as

    a multifunctional adhesin.

    Based on sequence analysis, we hypothesize that the

    lapEBC genes code for an ABC transporter (Fath and

    Kolter, 1993; Young and Holland, 1999; Dassa and

    Bouige, 2001). ABC transporters involved in export gen-

    erally are composed of three separate components aninner membrane anchored ATPase, a membrane fusion

    protein and an outer membrane protein (Dassa and

    Bouige, 2001). LapB is the predicted inner membrane

    protein of 74 kDa, with several predicted transmembrane

    regions and a C-terminal ATPase domain containing the

    canonical Walker box motifs characteristic of this family.

    LapC is predicted to be 50 kDa, contains a single pre-

    dicted transmembrane domain, shows similarity to toxin

    secretion proteins of the HlyD family, and is therefore

    proposed to be the membrane fusion protein. LapE is a

    48 kDa protein predicted to localize to the outer mem-

    brane and is similar to AggA (50% identity and 71% sim-ilarity over the 750 bp we have sequenced), a protein that

    was described as a factor involved in agglutination and

    adherence in Pseudomonas putidastrain Corvallis (Buell

    and Anderson, 1992). Furthermore, LapE contains a

    domain that is predicted to function as either an outer

    membrane efflux protein domain or a TolC-like domain

    (NCBI Conserved Domain Search). This TolC domain is

    characteristic of the PRT-HLY family of ABC transporters

    that are involved in protein export in prokaryotes (Dassa

    and Bouige, 2001). Most proteins exported by this sub-

    family of ABC transporters contain a series of glycine-rich

    repeats that forms a calcium-binding site (Baumann et al.,

    1993; Young and Holland, 1999; Dassa and Bouige,

    2001). This so-called repeats in toxin or RTX calcium-

    binding domain has been identified in LapA. Members of

    the PRT-HLY family of proteins can be further distin-

    guished based on their C-terminal sequence. LapA is

    likely a member of the HLY subfamily, because it lacks the

    signature extreme C-terminal motif DXXV (where X is a

    hydrophobic residue) of the PRT subfamily members (Let-

    offe and Wandersman, 1992; Ghigo and Wandersman,

    1994; Duong et al., 1996). Thus, we predicted that LapB

    is the ATP-binding element, LapC is the membrane fusion

    component and LapE is the outer membrane component

    of an ABC transporter responsible for the export of LapA.

    Localization of LapE and LapA

    The sequence analyses presented above allowed us to

    develop a model in which the LapA protein serves as an

    adhesin to firmly anchor these bacteria to a surface (i.e.

    promote irreversible attachment). Furthermore, the iden-

    tification of the LapE, LapB and LapC proteins as compo-

    nents of an ABC transporter suggested a mechanism by

    which LapA could be transported out of the cell.

    To gain insight into the role of the LapA and LapE

    proteins in irreversible attachment, we examined the local-

    ization of these proteins. The LapA protein was predicted

    to be localized to the outer membrane or cell-surface

    based on its sequence similarity to known proteins. Frac-

    tionation of P. fluorescens inner and outer membranes,

    followed by Western blot analysis, did not detect LapA ineither of these fractions (Fig. 4A, lanes labelled IM and

    OM). We next tested if LapA was secreted from the cell

    and/or was loosely associated with the cell surface. To test

    for a loose association of the protein to the surface of the

    cell, bacteria were grown as described in the Experimen-

    tal proceduresand 20 ml of the culture was centrifuged,

    then resuspended in a small volume of buffer resulting in

    a 50-fold concentration of the bacteria. These concen-

    trated cells were vortexed for five seconds, the suspension

    re-centrifuged, and an aliquot of the resulting supernatant

    (designated S2) was analysed by Western blot with anti-

    LapA antibodies. As shown in Fig. 4A, lane S2, a largemolecular weight band was detected for the wild-type

    strain that is absent from the lapAmutant. The exact size

    of the band is difficult to estimate as it runs significantly

    larger than the largest size marker which runs at

    ~190 kDa. An identically sized band was also detected in

    10-fold concentrated spent supernatant from the wild type

    but was absent from the lapAmutant strain supernatant.

    Therefore, LapA appears to be found both in the cell

    supernatant and in a loose association with the bacterial

    cell surface, but not in the outer membrane.

    To ensure that the centrifugation and vortexing proce-

    dures were not rupturing the cell membrane, we also

    probed these same fractions with antibody to LapE, a

    predicted outer membrane protein (Fig. 4B, top panel).

    LapE was not detected in either supernatant fraction, nor

    was it detected associated with the IM fraction (Fig. 4B).

    A band corresponding to the molecular weight of LapE

    (48 kDa) was detected in the OM fraction, but was absent

    from the lapE83 mutant OM fraction. A weakly cross-

    reacting band that runs at a molecular weight slightly

    greater than LapE was present in all samples.

  • 8/12/2019 ABC Pseudomonas

    8/14

    912 S. M. Hinsa, M. Espinosa-Urgel, J. L. Ramos and G. A. OToole

    2003 Blackwell Publishing Ltd, Molecular Microbiology, 49, 905918

    Sequence analysis of the LapE, LapB and LapC pro-

    teins showed similarity with ABC transporter components,

    and therefore suggested a possible role for this putative

    ABC transporter system in the secretion of the LapA pro-

    tein. To address this hypothesis, we tested for the pres-

    ence of cell surface-associated LapA (the S2 fraction) in

    the wild type and the lapEBCmutant strains. As shown in

    Fig. 4C (top panel) LapA was detected in the preparation

    from the wild type, but could not be detected in the S2

    supernatant fraction of any of the lapEBC mutants. Fur-

    thermore, no LapA was detected in the supernatant of the

    lapEBC mutants (data not shown). These observations

    suggest the ABC transporter may be responsible for the

    delivery of LapA to the exterior of the cell. As expected,

    no LapA was detected in the lapA51mutant. In contrast,

    LapE was present in the OM fractions of all of the lap

    mutant strains except for the lapE83 mutant (Fig. 4C,

    lower panel).

    Conservation of the lap genes among pseudomonads

    Comparisons were performed between the lapAchromo-

    somal region of P. putida KT2440 and the equivalent

    regions in the incomplete genome sequences of P.

    fluorescens PfO1 (http://www.Jgi.doe.gov/JGI_microbial/

    html/pseudomonas/pseudo_homepage.html) and P. syrin-

    gae (http://www.tigr.org), the finished genome sequence

    of P. aeruginosa(http://www. Pseudomonas.com), and the

    DNA sequence we obtained from strain P. fluorescens

    WCS365. The results of these analyses are shown in

    Fig. 1 and Table 2.

    The organization of the lapAEBCregion is very similar

    in P. fluorescens PfO1 and P. fluorescens WCS365

    (Fig. 1A). In P. putidaKT2440, lapE (aggA) is located on

    a different part of the chromosome from the rest of the lap

    genes and is associated with a different ABC transporter

    (not shown). The lapAand lapBCgenes are organized in

    a similar fashion to P. fluorescens PfO1, however, they

    appear in an inverted orientation relative to the flankinggenes. In spite of these differences in the chromosomal

    arrangement of the genes, LapA seems to play a similar

    role in P. fluorescens and P. putida. The lapA (mus-24)

    mutant of P. putida KT2440 is also defective in biofilm

    formation in static conditions (Espinosa-Urgel et al., 2000)

    and in a flow cell system, with a biofilm architecture similar

    to the P. fluorescens WCS365 lapA mutants (data not

    shown). LapA is not present in P. aeruginosaor P. syrin-

    gaebut proteins with some similarity to those encoded by

    the lapEBCgenes are present, although at different chro-

    mosomal locations. It is worth noting that even though

    lapA is not present in P. aeruginosa, the gene clustershowing the highest degree of similarity lapEBCgenes is

    also associated with a putative large outer membrane

    protein of ~2500 amino acids (not shown).

    The lapgenes are required for the colonization of

    quartz sand

    It has been previously reported that the lapmutants of P.

    fluorescensare deficient for attachment to polyvinylchlo-

    Fig. 4. Protein localization studies.A. Localization of LapA. A Western blot developed with antibody to

    LapA was performed on the following fractions: (i) the supernatant ofthe 50-fold concentrated, resuspended and vortexed cells (S2); (ii)the TCA-precipitated, 10-fold concentrated, cell-free supernatant(S1); (iii) the inner membrane fraction (IM), and (iv) the outer mem-brane fraction (OM) of these cells. The results of the Western analysis

    on fractions from the wild type (top panel) and lapA51mutant (lowerpanel) are shown.B. Localization of LapE. A Western blot using the fractions describedabove was developed with the LapE antibody. The fractions from the

    wild type (top panel) and lapE83mutant (lower panel) are shown.C. Detection of LapA and LapE in the lap mutant strains. The S2supernatants of the wild type and lapmutant strains were analysedfor the presence of LapA (top panel). The OM fractions of the wildtype and lapmutant strains were analysed for the presence of LapE

    (lower panel). In all experiments, cells from a ~16 h culture grown inminimal, citrate supplemented medium were used to prepare eachfraction and proteins were resolved on gradient polyacrylamide gels(415%).

    http://www.jgi.doe.gov/JGI_microbial/http://www.tigr.org/http://www/http://www/http://www/http://www.tigr.org/http://www.jgi.doe.gov/JGI_microbial/
  • 8/12/2019 ABC Pseudomonas

    9/14

    P. fluorescens biofilm formation 913

    2003 Blackwell Publishing Ltd, Molecular Microbiology, 49, 905918

    ride, polypropylene, polystyrene and borosilicate glass

    (OToole and Kolter, 1998). However, P. fluorescens is

    primarily a soil microorganism, therefore its natural sub-

    strate is most likely a variety of sands and soils. Therefore,

    we tested the wild type and the lapmutant strains for their

    ability to attach to a surface this bacterium may encounter

    in its natural soil environment, namely, quartz sand. We

    implemented assays to follow sand colonization visually

    and quantitatively.

    To visualize bacteria attached to sand, GFP-labelled

    bacteria were allowed to attach to the sand for 5 h and

    then the sand was washed and the attached bacteria

    visualized by epifluorescent microscopy. Figure 5A illus-

    trates the wild type and the lapB52mutant attached to a

    Table 2. Sequence similarities of Lap proteins.

    P. putidaKT2440 P. fluorescensWCS365 P. fluorescens PfO1 P. aeruginosaPAO1 P. syringae

    LapA (PP0168) 39%/52% 41%/53%b

    8682 aaa (491 aa) (4785 aa)c

    LapB (PP0167) PA1876d

    ATPase 86%/93% 85%/92% 37%/58% 30%/52%718 aa (322 aa) 722 aa 723 aa 613 aa

    LapC (PP0166) PA1877

    MFP 76%/88% 80%/90% 39%/57% 26%/47%452 aa (140 aa) 455 aa 395 aa 526 aa

    LapE (PP4519)e PA1875OM protein 50%/71% 53%/71% 22%/42% 24%/42%452 aa (163 aa) 451 aa 425 aa 479 aa

    a.The number of amino acids (aa) known from the genome sequence of P. putidaKT4220. The gene designation assigned in the P. putidagenome project are given in parentheses (http://www.tigr.org).

    b.Per cent identity and percentage similarity was determined by comparing each complete or partial ORF to the predicted aa sequence of theputative homologue determined from the complete P. putidaKT4220 genome sequence.

    c.Per cent identity/similarity corresponds to alignments with partial (in parentheses) or complete ORFs obtained from each respective genomeproject.

    d.The PA designations refer to ORF numbers from the P. aeruginosagenome project (http://www.Pseudomonas.com).e.LapE corresponds to the previously identified AggA protein (see text), but is annotated as TolC in the P. putidagenome.

    Abbreviations: OM, outer membrane; MFP, membrane fusion protein; ATPase, cytoplasmic membrane-localized ATPase

    Fig. 5.Sand attachment.A. Visualizing bacterial attachment to sand. The

    wild type and lapB52strains were allowed to

    attach to quartz sand for 5 h. The sand waswashed to remove unattached bacteria and epi-fluorescent microscopy used to visualize

    attached bacteria at 1400magnification.B. Quantifying attachment of individual strainsto sand. The colony forming units (CFU) pergram of sand (Y-axis) is plotted for the wild typeand lapB52mutant. The initial inoculum of the

    wild type and mutant was identical at ~1 109CFU ml-1.C. Quantifying competitive attachment to sand.Equal numbers of wild type and lapB52mutant

    bacteria (a total of ~1 109CFU ml-1) weremixed and inoculated onto quartz sand. The percent of each strain attached to the sand particle(Y-axis) was determined as described in theExperimental procedures.

    http://www.tigr.org/http://www.pseudomonas.com/http://www.pseudomonas.com/http://www.tigr.org/
  • 8/12/2019 ABC Pseudomonas

    10/14

    914 S. M. Hinsa, M. Espinosa-Urgel, J. L. Ramos and G. A. OToole

    2003 Blackwell Publishing Ltd, Molecular Microbiology, 49, 905918

    grain of sand. The wild-type strain efficiently colonizes the

    sand particle, whereas the lapB52strain is much reduced

    in its ability to colonize. Similar results were seen for the

    lapC87and lapA62mutants (data not shown).

    To quantify the number of cells attached to the sand

    particles, ~1 109 bacteria were allowed to adhere to

    sand as described above. The number of surface-

    associated bacteria was determined by vortexing and

    sonicating the sand to remove attached cells, and the

    number of bacteria attached to the sand was deter-

    mined by dilution plating (see Experimental procedures

    for details). As shown in Fig. 5B, there is a 10-fold differ-

    ence in attachment between wild type and the lapB52

    mutant. Similar results were seen with the lapA51,

    lapC87and lapE83mutants (data not shown). To deter-

    mine if competition for attachment or complementation

    would affect this outcome, we mixed equal amounts of

    wild-type and mutant bacteria (a total of ~1 109 cells)

    and allowed them to attach to the sand. After rinsing

    and vortexing/sonication to remove the attached bacte-

    ria, we found that approximately 80% of the bacteriaattached to the sand were the wild-type strain (Fig. 5C).

    Similar results were seen for lapA51, lapC87 and

    lapE83 (data not shown). These data also indicate that

    the wild-type strain is unable to rescue the biofilm for-

    mation defect(s) of the lapmutants.

    Discussion

    Cell-to-surface interaction events mark the early steps in

    the development of a mature biofilm. It has long been

    proposed that early attachment events first involve

    reversible attachment to a surface, marked by the tran-sient interactions of one pole of the bacterium with a

    substratum (Zobell, 1943; Marshall et al., 1971; van

    Loosdrecht et al., 1990; Jensen et al., 1992; Fletcher,

    1996). Following up on these previous studies, Sauer

    et al. (2002) made detailed observations of early biofilm

    development, and also observed reversible and subse-

    quent irreversible attachment steps by P. aeruginosa. In

    the studies presented here, microscopic analysis of

    early attachment events in P. fluorescensWCS365 dem-

    onstrated that this pseudomonad also undergoes the

    same two step early attachment pathway. In the wild-

    type strain, cells undergo transient polar attachment fol-

    lowed by subsequent undefined events that lead to a

    so-called irreversible attachment. The lapmutant strains

    are capable of initial attachment to a degree that is

    indistinguishable from the wild-type strain, thus they

    appear to have no defects in reversible attachment.

    However, the lap mutants appear to be unable to

    progress normally to the irreversible attachment step of

    biofilm development. This phenotype is most clearly

    observed in time-lapse movies obtained from flow cell

    studies of the wild type and lap mutant strains. Our

    studies show that LapA and the lap-encoded ABC trans-

    porter, which is required for LapA to be exported from

    the cell, are required for irreversible attachment. To our

    knowledge, this is the first report of genetic determi-

    nants that are necessary for, and define, the transition

    from reversible to irreversible attachment.

    Despite the apparent defect in irreversible attachment,

    the lapmutants can eventually form small clusters of bac-

    teria on the surface, however, these mutant bacteria are

    unable to develop the architecture observed for the wild-

    type bacteria. We demonstrated that the attachment of

    these few lapmutant cells to the surface was not due to

    a secondary mutation, therefore the lapmutants may use

    an undefined pathway to irreversibly attach to a surface

    after prolonged incubation. Microcolonies may eventually

    form as a consequence of cell-to-cell adherence (suggest-

    ing that the Lapgene products may not be required for

    these interactions) and/or cell division.

    What role does the putative lapEBC-encoded ABC

    transporter and the associated LapA protein play in biofilmdevelopment? One possible role for LapE, a predicted

    outer membrane protein with sequence similarity to the

    AggA adhesin (Buell and Anderson, 1992), may be as an

    attachment factor required for early biofilm development.

    However, fractionation of the lap mutants and Western

    analysis revealed that the LapE protein is localized to the

    outer membrane in all strains but the lapEmutant, sug-

    gesting that proper localization of this outer membrane

    protein is not sufficient for biofilm formation. In contrast,

    any mutation within the lapEBCcluster resulted in the loss

    of any detectable LapA associated with the cell surface or

    in the cell supernatant. These data are consistent with amodel in which the lapEBC-encoded ABC transporter is

    required for export of LapA outside of the cell. Particularly

    intriguing is the identification of a putative signal sequence

    at the N-terminus of LapA, which is not typical of proteins

    transported by ABC systems (Young and Holland, 1999;

    Dassa and Bouige, 2001). One possibility is that this

    secretion signal is cryptic and not typically utilized for

    transport of LapA. Another possibility is that LapA is trans-

    ported into the periplasm by the Sec-dependent transport

    system, and then delivered outside of the cell by the ABC

    transporter. The detailed analysis of the mechanism by

    which LapA exits the cell awaits future studies. Taken

    together, these data are consistent with a hypothesis

    wherein LapA, alone or in association with LapE, serves

    to promote stable adhesion of P. fluorescensWCS365 to

    a surface early in biofilm development.

    A role for ABC transporters in cell-to-surface and/or

    cell-to-cell interactions and biofilm development has been

    proposed in other organisms. An ABC transporter is

    important for attachment and virulence of Agrobacterium

    tumefacienson carrot cells (Matthysse et al., 1996). In a

  • 8/12/2019 ABC Pseudomonas

    11/14

    P. fluorescens biofilm formation 915

    2003 Blackwell Publishing Ltd, Molecular Microbiology, 49, 905918

    recent report Sauer and Camper (2001) showed that in

    another soil pseudomonad, P. putida, the potB gene,

    which codes for an ABC transporter component, is

    upregulated in the early stages of biofilm formation on

    silicone. This observation suggests a possible role for the

    PotB ABC transporter in early biofilm development, but

    this has not been demonstrated experimentally. In the

    oral microbe Streptococcus gordonii, an ABC transporter

    was shown to be important for self co-aggregation (cell-

    to-cell interactions) in vitro (Kolenbrander et al., 1994).

    Taken together, these data indicate that a role for ABC

    transporters in biofilm development may be conserved

    across Gram-positive and Gram-negative organisms,

    perhaps for the purpose of secreting cell surface

    adhesins.

    All of the lapmutants of P. fluorescensWCS365 and

    the lapAmutant of P. putidawere shown to be defective

    for attachment to a number of plastics (polyvinylchloride,

    polypropylene, polycarbonate and polystyrene) as well as

    borosilicate glass (OToole and Kolter, 1998b; Espinosa-

    Urgel et al., 2000). Here we report that the lapmutantsare also defective for attaching to quartz sand, a sub-

    strate they are very likely to encounter in the environ-

    ment. The P. putida mus-24 (lapA) mutant was isolated

    as defective for adhesion to corn seeds, and is also

    impaired in biofilm formation and competitive root coloni-

    zation (M. Espinosa-Urgel, unpubl. obs.). These broad

    phenotypic effects on attachment to biotic and abiotic

    surfaces caused by mutations in LapA could be

    explained in two ways. The interactions mediated by this

    protein might be somewhat non-specific, therefore LapA

    may act as a general purpose adhesin. Alternatively, the

    multiple domains may in fact be different bindingdomains, each promoting adherence to a set of sub-

    strates. The repeats in LapA (Domains 2 and 3) are rem-

    iniscent of those found in adhesion proteins of Gram-

    positive bacteria involved in biofilm formation or in cell

    cell interactions, whereas Domain 4 shows similarities

    with calcium-binding proteins and haemolysins, that play

    a role in cellhost interactions during pathogenesis.

    The presence of the lap genes in three different soil

    isolates, and their absence from two pathogenic

    Pseudomonasstrains, suggests that these genes may be

    specifically necessary for biofilm development by only a

    subset of pseudomonads. Current studies are exploring

    the extent of conservation of the lap genes and their

    organization in a wide variety of soil pseudomonads. To

    date, no ABC transporter required for biofilm formation by

    P. aeruginosahas been identified. This finding, along with

    the data presented here, suggest that although both

    pathogenic and non-pathogenic pseudomonads make

    biofilms, they may utilize distinct mechanisms to transition

    from reversible to irreversible attachment during early bio-

    film development.

    Experimental procedures

    Bacterial strains, plasmids, and culture conditions

    Pseudomonas fluorescenswas grown in LuriaBertani (LB)

    or in minimal media, as specified, at 30C. The minimal salts

    medium used was M63 (Pardee et al., 1959) supplemented

    with MgSO4 (10 mM) and either glucose (0.2%) or citrate

    (0.4%), or AB10 media without trace minerals (Tolker-Nielsen

    et al., 2000). Pseudomonas putida was grown in LB andAB10 at 30C and E. coliwas grown in LB at 37C. Antibiotics

    were added at the following concentrations: (i) E. coli: gen-

    tamycin (Gm), 10 mg ml-1; chloramphenicol (Cm), 30 mg ml-1;

    (ii) P. fluorescens: Gm, 50 mg ml-1; kanamycin (Kn),

    250 mg ml-1. Generalized transductions were performed as

    described (Jensen et al., 1998). Plasmid pSMC21 is derived

    from pSMC2 (Bloemberg et al., 1997), expresses the green

    fluorescent protein (GFP) under a constitutive promoter, and

    carries both Ap and Kn resistance markers. Plasmids pSU21

    (Martinez et al., 1988), pME6000 (Itoh et al., 1988), pSMC32

    (OToole et al., 2000b), pUCP18 (Schweizer, 1991) and

    pGEM (Promega) were utilized for cloning experiments.

    Molecular techniques

    Sequence of the DNA flanking transposon insertions was

    determined by arbitrary primed PCR (OToole et al., 1999).

    Selected transposon insertions were cloned to determine

    additional DNA sequence flanking the element. Chromo-

    somal DNA was prepared as described (Pitcher et al., 1989),

    digested with EcoRI and ligated into pSU21 previously

    digested with EcoRI. The ligations were electroporated into

    E. coliJM109 electrocompetent cells, plated on LB supple-

    mented with Cm, then replica printed onto LB supplemented

    with Cm and Gm. The CmrGmrcolonies were purified, plas-

    mid DNA was prepared, and the plasmids were sequenced

    with the Tn5Ext primer (OToole et al., 1999). Polymerase

    chain reaction using primers to sequence derived from P.fluorescens WCS365 was performed to confirm the gene

    order inferred from sequence analysis (Fig. 1A).

    Biofilm assays

    Initial biofilm formation was measured using the microtiter

    dish assay system performed as described previously

    (OToole and Kolter, 1998a,b; OToole et al., 1999) using min-

    imal M63 medium with glucose (0.2%) or citrate (0.4%) as

    the growth substrate. The visualization of bacterial cells

    attached to PVC was performed as previously reported

    (Bloemberg et al., 1997) except cultures were incubated at

    room temperature for up to 5 h before analysis. The once-

    flow through continuous culture flow cell system was assem-

    bled as described (Christensen et al., 1999) and modified

    AB10, as described above, was utilized as the growth

    medium.

    To address whether secondary mutations were responsible

    for allowing the lapstrains to form microcolonies at later time-

    periods, the attached lapAmutant cells were tested for the

    ability to make biofilms after re-culture under planktonic con-

    ditions. The lap mutant bacteria were allowed to attach to

    PVC tabs for 24 h, after which the tabs were rinsed and

    placed into an eppendorf tube containing LB. The bacteria

  • 8/12/2019 ABC Pseudomonas

    12/14

    916 S. M. Hinsa, M. Espinosa-Urgel, J. L. Ramos and G. A. OToole

    2003 Blackwell Publishing Ltd, Molecular Microbiology, 49, 905918

    adhered to the tab were removed by two alternating series

    of vortexing (10 s) and sonication (10 s, Tabletop Ultrasonic

    Cleaner, FS-60, Fischer Scientific) followed by a final 10 s

    vortex (as described Gardener and de Bruijn, 1998). The LB

    and any bacteria removed from the tab were used to inocu-

    late an LB culture that was outgrown for 24 h, and then a

    standard biofilm assay was performed using these cultures

    as an inoculum.

    Sand attachment

    Bacteria were grown overnight in LB and subcultured into

    M63 plus citrate (0.4%) at a 1:5 dilution. Sand was placed in

    the bottom of the well in a 24-well plate and covered with the

    bacterial suspension. The plates were placed on a shaker at

    room temperature for four hours. A sample of sand was

    removed from each well, placed in an eppendorf tube, and

    washed five times with 500 ml of M63. A sand sample could

    be removed at this point for visualization of bacter ia attached

    to sand. Epifluorescent microscopy indicated that wild-type

    bacteria formed a monolayer of cells on the sand particles at

    this time under the growth conditions described (see Fig. 5).

    Quantification of bacteria attached to sand was performed asfollows: 50 ml of M63 was added to the sand containing tube,

    and the bacteria were removed from the sand by two alter-

    nating series of vortexing (10 s) and sonication (10 s, Table-

    top Ultrasonic Cleaner, FS-60, Fischer Scientific) followed by

    a final 10 s vortex (as described by Gardener and de Bruijn,

    1998). Ten microlitres of suspension was removed and used

    for dilution plating. Bacterial counts are normalized to grams

    of sand assayed.

    To determine the efficacy of the vortex/sonication regimen,

    the percentage of bacteria removed from the sand was deter-

    mined. One sample of sand was treated as described above,

    whereas a second sample did not receive the vortexing and

    sonication treatment. Both treated and untreated samples

    were washed an additional three times with 500 ml of M63.Fifty microlitres of M63 was added to the sand post treatment,

    the samples were incubated for 2 h and dilution plating per-

    formed to determine the number of bacteria present (e.g.

    those bacteria shed from the sand and growing planktoni-

    cally). These control experiments demonstrated >99.9% of

    the bacteria attached to the sand are removed during the

    vortexing and sonication steps.

    Imaging

    Epifluorescent and phase-contrast microscopy were per-

    formed with a Model DM IRBE microscope (Leica Microsys-

    tems) equipped with an Orca Model C4742-5 CCD camera

    (Hamamatsu). Images were acquired and processed on aMacintosh G4 loaded with OpenLab 3.1 software (Improvi-

    sion). COMSTATanalysis was performed as described (Hey-

    dorn et al., 2000a, b).

    Protein localization and Western analysis

    Samples for Western analysis were prepared as follows.

    Twenty millilitres of minimal M63 medium supplemented with

    citrate and MgSO4 were grown shaking at 30C for ~16 h.

    The cultures were centrifuged for 10 min at 6000 gand 1 ml

    of the supernatant was removed and TCA precipitated as

    described (Kunitz, 1952). The precipitated protein pellet was

    resuspended in 100 ml of resuspension buffer (Tris-HCl,

    20 mM, pH 8 plus 10 mM MgCl2) this sample is designated

    S1. The bacterial pellet from the 20 ml culture was resus-

    pended in 400 ml of resuspension buffer and vortexed for 5 s.

    The samples were centrifuged for 5 min at 12 200 gand the

    supernatant was collected (the supernatant of this sample is

    designated S2). The inner and outer membranes were sep-

    arated as previously described (Lohia et al., 1984) with some

    modifications. The cells were grown as described above, then

    centrifuged for 10 min at 6000 gand the pellet resuspended

    in 1.5 ml of PBS. A Thermo Spectronic French press mini-

    cell was used to lyse the cells by processing twice at 20 000

    p.s.i. Next, the samples were spun at 12 200 g to pellet

    unbroken cells. The supernatant was removed and spun at

    100 000 gfor 60 min at 4C. The pellet was resuspended in

    100 ml of Hepes buffer (10 mM, pH 5.7) and incubated with

    100 mg ml-1 DNAse and RNAse for 20 min at 25C. Nine

    hundred microlitres of urea (4 M) was added to the samples

    to solubilize the inner membrane. The samples were spun at

    100 000 g for 60 min at 4C. The supernatant fraction con-

    taining the inner membrane was removed and the pellet(outer membrane) was washed with cold H2O and centrifuged

    for another 60 min at 4C at 100 000 g. SDS loading buffer

    was mixed with each sample, followed by heat denaturation

    at 75C for 10 min. The samples were resolved on a gradient

    polyacrylamide gel (415%) at 20 mA. The protein was trans-

    ferred in a to a nitrocellulose membrane in transblot buffer as

    described (Towbin et al., 1979). Western blots were devel-

    oped with ECL Western detection reagents (Amersham).

    Acknowledgements

    We thank Christian Weinel for providing us with P. putida

    sequence data prior to publication. We thank referee #3 forsuggesting the suppressor analysis experiment. This work

    was supported by grants from the NSF (CAREER 9984521)

    and The Pew Charitable Trusts to G.A.O. G.A.O. is a Pew

    Scholar in the Biomedical Sciences. We also acknowledge

    grant BMC2001-0576 from the Plan Nacional de I +D +I to

    M.E.U. M.E.U. is the recipient of a grant from the Ramn y

    Cajal Program (MCYT).

    Supplementary material

    The following material is available from

    http://www.blackwellpublishing.com/products/journals/

    suppmat/mmi/mmi3615/mmi3615sm.htm

    Time-lapse movies of biofilm-grown bacteria were made

    from images captured every 3 s over a period of 5 min. In

    this experiment, the biofilm was 8 h old and the flow direc-

    tion was from the bottom to the top of the image. Most of

    the wild-type bacteria are attached to the surface along the

    long length of the cell and do not move during the course of

    the 5 min movie. In contrast, most of the lapCmutant bacte-

    ria are attached by one pole and can be seen spinning

    rapidly or moving with the flow of the medium through the

    flow cell.

    http://www.blackwellpublishing.com/products/journals/http://www.blackwellpublishing.com/products/journals/
  • 8/12/2019 ABC Pseudomonas

    13/14

    P. fluorescens biofilm formation 917

    2003 Blackwell Publishing Ltd, Molecular Microbiology, 49, 905918

    References

    Alexandrov, N.N., Nussinov, R., and Zimmer, R.M. (1995)

    Fast Protein Fold Recognition Via Sequence to Structure

    Alignment and Contact Capacity Potentials.Pacific Sym-

    posium on Biocomputing 96. L. Hunter and T. E. Klein.

    Singapore: World Scientific Publishing, pp. 5372.

    Baumann, U., Wu, S., Flaherty, K.M., and McKay, D.B.

    (1993) Three-dimensional structure of the alkaline pro-

    tease of Pseudomonas aeruginosa: a two-domain proteinwith a calcium binding parallel beta roll motif. EMBO J12:

    33573364.

    Bloemberg, G.V., OToole, G.A., Lugtenberg, B.J.J., and

    Kolter, R. (1997) Green fluorescent protein as a marker for

    Pseudomonasspp. Appl Environ Microbiol63: 45434551.

    Buell, C.R., and Anderson, A.J. (1992) Genetic analysis of

    the aggAlocus involved in agglutination and adherence of

    Pseudomonas putida, a beneficial fluorescent

    pseudomonad. Mol Plant-Microb Interact5: 154162.

    Camacho-Carbajal, M.M. (2001) Molecular characterization

    of type 4 pili, NDHI and PyrR in rhizosphere colonization

    of Pseudomonas fluorescensWCS365. PhD Thesis, Insti-

    tute of Molecular Plant Sciences, Leiden University:

    Leiden, The Netherlands.Christensen, B.B., Sternberg, C., Andersen, J.B., Palmer,

    R.J. Jr, Nielsen, A.T., Givskov, M., and Molin, S. (1999)

    Molecular tools for study of biofilm physiology. Methods

    Enzymol310: 2042.

    Costerton, J.W., Lewandowski, Z., Caldwell, D.E., Korber,

    D.R., and Lappin-Scott, H.M. (1995) Microbial Biofilms. In

    Annual Review Microbiology. Ornston, L.N., Balows, A.,

    and Greenberg, E.P. (eds). Palo Alto: Annual Reviews, pp.

    711745.

    Cucarella, C., Solano, C., Valle, J., Amorena, B., Lasa, I., and

    Penades, J.R. (2001) Bap, a Staphylococcus aureussur-

    face protein involved in biofilm formation. J Bacteriol183:

    28882896.

    Dassa, E., and Bouige, P. (2001) The ABC of ABCS: aphylogenetic and functional classification of ABC systems

    in living organisms. Res Microbiol152: 211229.

    Davey, M.E., and OToole, G.O. (2000) Microbial biofilms:

    from ecology to molecular genetics. Microbiol Mol Biol

    Revs64: 847867.

    Dekkers, L.C., Bloemendaal, C.J.P., de Weger, L.A., Wijffel-

    man, C.A., Spaink, H.P., and Lugtenberg, B.J.J. (1998a)

    A two-component system plays an important role in the

    root-colonizing ability of Pseudomonas fluorescensstrain

    WCS365. Mol Plant-Microbe Interact11: 4556.

    Dekkers, L.C., Phoelich, C.C., van der Fits, L., and Lugten-

    berg, B.J.J. (1998b) A site-specific recombinase is required

    for competitive root colonization by Pseudomonas fluore-

    scensWCS365. Proc Natl Acad Sci USA95: 70517056.Dekkers, L.C., van der Bij, A.J., Mulders, I.H.M., Phoelich,

    C.C., Wentwoord, R.A.R., Glandorf, D.C.M., et al.(1998c)

    Role of the O-antigen of lipopolysaccharide, and possible

    roles of growth rate and of NADH: ubiquinone oxidoreduc-

    tase (nuo) in competetive tomato root-tip colonization by

    Pseudomonas fluorescens WCS365. Mol Plant-Microbe

    Interact11: 763771.

    Duong, F., Lazdunski, A., and Murgier, M. (1996) Protein

    secretion by heterologous bacterial ABC-transporters: the

    C-terminus secretion signal of the secreted protein confers

    high recognition specificity. Mol Microbiol21: 459470.

    Economou, A., Hamilton, W.D.O., Johnston, A.W.B., and

    Downie, J.A. (1990) The Rhizobiumnodulation gene nodO

    encodes a Ca2+-binding protein that is exported without N-

    terminal cleavage and is homologous to haemolysin and

    related proteins. EMBO J9: 349354.

    Espinosa-Urgel, M., Salido, A., and Ramos, J.L. (2000)

    Genetic analysis of functions involved in adhesion of

    Pseudomonas putida to seeds. J Bacteriol 182: 2363

    2369.

    Fath, M.J., and Kolter, R. (1993) ABC transporters: bacterial

    exporters. Microbiol Rev57: 9951017.

    Fletcher, M. (1996) Bacterial attachment in aquatic environ-

    ments: a diversity of surfaces and adhesion strategies. In

    Bacterial Adhesion: Molecular and Ecological Diversity.

    Fletcher, M. (ed.). New York: John Wiley & Sons, pp. 124.

    Gardener, B.B.M., and de Bruijn, F.J. (1998) Detection and

    isolation of novel rhizopine-catabolizing bacteria from the

    environment. Appl Environ Microbiol64: 49444949.

    Geels, F.P., and Schippers, B. (1983) Reduction of yield

    depressions in high frequency potato cropping soil after

    seed tuber treatments with antagonistic fluorescent

    Pseudomonasspp. Phytopathol Z108: 207214.Ghigo, J.M., and Wandersman, C. (1994) A carboxyl-

    terminal four-amino acid motif is required for secretion of

    the metalloprotease PrtG through the Erwinia chrysan-

    themi protease secretion pathway. J Biol Chem 269:

    89798985.

    Heydorn, A., Nielsen, A.T., Hentzer, M., Sternberg, C.,

    Givskov, M., Ersboll, B.K., and Molin, S. (2000a) Quantifi-

    cation of biofilm structures by the novel computer program

    COMSTAT. Microbiology146: 23952407.

    Heydorn, A., Ersboll, B.K., Hentzer, M., Parsek, M.R.,

    Givskov, M., and Molin, S. (2000b) Experimental reproduc-

    ibility in flow-chamber biofilms. Microbiology 146: 2409

    2415.

    Itoh, Y., Soldati, L., Leisinger, T., and Haas, D. (1988) Low-and intermediate-copy-number cloning vectors based on

    the Pseudomonasplasmid pVS1. Antonie Van Leeuwen-

    hoek54: 567573.

    Jensen, E.T., Kharazmi, A., Hoiby, N., and Costerton, J.W.

    (1992) Some bacterial parameters influencing the neutro-

    phil oxidative burst response to Pseudomonas aeruginosa

    biofilms. Apmis100: 727733.

    Jensen, E.C., Schrader, H.S., Rieland, B., Thompson, T.L.,

    Lee, K.W., Nickerson, K.W., and Kokjohn, T.A. (1998)

    Prevalence of broad-host-range lytic bacteriophages of

    Sphaerotilus natans, Escherichia coli, and Pseudomonas

    aeruginosa. Appl Environ Microbiol64: 575580.

    Kolenbrander, P.E., Anderson, R.N., and Ganeshkumar, N.

    (1994) Nucleotide sequence of the Streptococcus gordoniiPK488 coaggregation adhesin gene, scaA, and ATP-

    binding cassette. Infect Immun62: 44694480.

    Korber, D.R., Lawrence, J.R., and Caldwell, D.E. (1994)

    Effect of motility on surface colonization and reproductive

    success of Pseudomonas fluorescens. dual-dilution contin-

    uous culture and batch culture systems. Appl Environ

    Microbiol60: 14211429.

    Kunitz, M.J. (1952) Crystalline inorganic pyrophosphatase

    isolated from bakers yeast. J Gen Physiol35: 423450.

  • 8/12/2019 ABC Pseudomonas

    14/14

    918 S. M. Hinsa, M. Espinosa-Urgel, J. L. Ramos and G. A. OToole

    2003 Blackwell Publishing Ltd, Molecular Microbiology, 49, 905918

    Letoffe, S., and Wandersman, C. (1992) Secretion of CyaA-

    PrtB and HlyA-PrtB fusion proteins in Escherichia coli:

    involvement of the glycine-rich repeat domain of Erwinia

    chrysanthemiprotease B. J Bacteriol174: 49204927.

    Lohia, A., Chaterjee, A.N., and Das, J. (1984) Lysis of Vibrio

    choleraecells: Direct isolation of the outer membrane from

    whole cells by treatment with urea. J Gen Microbiol130:

    20272033.

    van Loosdrecht, M.C., Lyklema, J., Norde, W., and Zehnder,

    A.J.B. (1990) Influence of interfaces on microbial activity.

    Microbiol Rev54: 7587.

    Mah, T.F., and OToole, G.A. (2001) Mechanisms of biofilm

    resistance to antimicrobial agents. Trends Microbiol9: 34

    39.

    Marshall, K.C., Stout, R., and Mitchell, R. (1971) Mechanism

    of the initial events in the sorbtion of marine bacteria to

    surfaces. J Gen Microbiol68: 337348.

    Martinez, E., Bartolome, B., and de la Cruz, F. (1988)

    pACYC184-derived cloning vectors containing the multiple

    cloning site and lacZa reporter gene of pUC8/9 and

    pUC18/19. Gene68: 159162.

    Matthysse, A.G., Yarnall, H.A., and Young, N. (1996)

    Requirement for genes with homology to ABC transport

    systems for attachment and virulence of Agrobacteriumtumefaciens. J Bacteriol178: 53025308.

    McNab, R., Forbes, H., Handley, P.S., Loach, D.M., Tannock,

    G.W., and Jekinson, H.F. (1999) Cell wall-anchored CshA

    polypeptide in Streptococcus gordoniiforms surface fibrils

    that confer hydrophobic and adhesive properties. J Bacte-

    riol181: 30873095.

    McNab, R., Jenkinson, H.F., Loach, D.M., and Tannock, G.W.

    (1994) Cell-surface-associated polypeptides CshA and

    CshB of high molecular mass are colonization determi-

    nants in the oral bacterium Streptococcus gordonii. Mol

    Microbiol14: 743754.

    Nelson, K.E., Weinel, C., Paulsen, I.T., Dodson, R.J., Hilbert,

    H., Martins dos Santos, V.A.P., et al. (2002) Complete

    genome sequence and comparative analysis of the meta-bolically versatile Pseudomonas putida KT2440. Environ

    Microbiol4: 799808.

    OToole, G.A., and Kolter, R. (1998a) Flagellar and twitching

    motility are necessary for Pseudomonas aeruginosabio-

    film development. Mol Microbiol30: 295304.

    OToole, G.A., and Kolter, R. (1998b) Initiation of biofilm

    formation in Pseudomonas fluorescensWCS365 proceeds

    via multiple, convergent signalling pathways: a genetic

    analysis. Mol Microbiol28: 449461.

    OToole, G.A., Pratt, L.A., Watnick, P.I., Newman, D.K.,

    Weaver, V.B., and Kolter, R. (1999) Genetic approaches

    to the study of biofilms. In Methods in Enzymology.Doyle,

    R.J. (ed). San Diego, CA: Academic Press, 310:91109.

    OToole, G.A., Kaplan, H., and Kolter, R. (2000a) Biofilm

    formation as microbial development. Annu Rev Microbiol

    54: 4979.

    OToole, G.A., Gibbs, K.A., Hager, P.W., Phibbs,P.V. Jr and

    Kolter, R. (2000b) The global carbon metabolism regulator

    Crc is a component of a signal transduction pathway

    required for biofilm development by Pseudomonas aerug-

    inosa. J Bacteriol182: 425431.

    Pardee, A.B., Jacob, F., and Monod, J. (1959) The genetic

    control and cytoplasmic expression of inducibility in the

    synthesis of b-galactosidase in E. coli. J Mol Biol1: 165

    178.

    Pitcher, D.G., Saunders, N.A., and Owen, R.J. (1989) Rapid

    extraction of bacterial genomic DNA with guanidium thio-

    cyanate. Lett Appl Microbiol8: 151156.

    Sauer, K., and Camper, A.K. (2001) Characterization of phe-

    notypic changes in Pseudomonas putida. response to

    surface-associated growth. J Bacteriol183: 65796589.

    Sauer, K., Camper, A.K., Ehrlich, G.D., Costerton, J.W., and

    Davies, D.G. (2002) Pseudomonas aeruginosa displays

    multiple phenotypes during development as a biofilm. J

    Bacteriol184: 11401154.Schweizer, H.P. (1991) Escherichia-Pseudomonas shuttle

    vectors derived from pUC18/19. Gene103: 109112.

    Simons, M., van der Bij, A.J., Brand, I., de Weger, L.A.,

    Wijffelman, C.A., and Lugtenberg, B.J.J. (1996) Gnotobi-

    otic system for studying rhizosphere colonization by plant

    growth-promoting Pseudomonas bacteria. Mol Plant-

    Microbe Inter9: 600607.

    Simons, M., Permentier, H.P., de Weger, L.A., Wijffelman,

    C.A., and Lugtenberg, B.J.J. (1997) Amino acid synthesis

    is necessary for tomato root colonization by Pseudomonas

    fluorescensstrain WCS365. Mol Plant-Microbe Interact10:

    102106.

    Tolker-Nielsen, T., Brinch, U.C., Ragas, P.C., Andersen, J.B.,

    Jacobsen, C.S., and Molin, S. (2000) Development anddynamics of Pseudomonas sp. biofilms. J Bacteriol 182:

    64826489.

    Towbin, H., Staehelin, T., and Gordon, J. (1979) Electro-

    phoretic transfer of proteins from polyacrylamide gels to

    nitrocellulose sheets: procedure and some applications.

    Proc Natl Acad Sci USA76: 43504354.

    Young, J., and Holland, I.B. (1999) ABC transporters: bacte-

    rial exporters-revisited five years on. Biochim Biophys Acta

    1461: 177200.

    Zobell, C.E. (1943) The effects of solid surfaces upon bacte-

    rial activity. J Bacteriol46: 3956.