University of Groningen Bacterial interaction forces in adhesion dynamics … · 2018. 2. 10. ·...

161
University of Groningen Bacterial interaction forces in adhesion dynamics Boks, Niels IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below. Document Version Publisher's PDF, also known as Version of record Publication date: 2009 Link to publication in University of Groningen/UMCG research database Citation for published version (APA): Boks, N. P. (2009). Bacterial interaction forces in adhesion dynamics Groningen: s.n. Copyright Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons). Take-down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum. Download date: 10-02-2018

Transcript of University of Groningen Bacterial interaction forces in adhesion dynamics … · 2018. 2. 10. ·...

  • University of Groningen

    Bacterial interaction forces in adhesion dynamicsBoks, Niels

    IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite fromit. Please check the document version below.

    Document VersionPublisher's PDF, also known as Version of record

    Publication date:2009

    Link to publication in University of Groningen/UMCG research database

    Citation for published version (APA):Boks, N. P. (2009). Bacterial interaction forces in adhesion dynamics Groningen: s.n.

    CopyrightOther than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of theauthor(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

    Take-down policyIf you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediatelyand investigate your claim.

    Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons thenumber of authors shown on this cover page is limited to 10 maximum.

    Download date: 10-02-2018

    https://www.rug.nl/research/portal/en/publications/bacterial-interaction-forces-in-adhesion-dynamics(27692517-0d19-469b-87e2-932130ff128a).html

  • Bacterial interaction forces in adhesion dynamics

  • Copyright © 2008 by N.P. Boks All rights reserved. No part of this publication may be reproduced or transmitted in any form or by any means without permission of the author and the publisher holding the copyright of the published articles. Printed by Gildeprint drukkerijen b.v., Enschede ISBN: 978-90-367-3638-1 (printed version) ISBN: 978-90-367-3667-1 (electronic version) Cover: Fluorescence image of staphylococci attached to a tipless AFM-cantilever. Using a LIVE/DEAD Baclight viability stain (Molecular Probes Europe BV, Leiden, The Netherlands), bacteria appear red when they are dead and green when they are still alive. Omslag: Fluorescentie opname van staphylococci die aan een tiploze AFM-cantilever zijn geplakt. Met behulp van een LIVE/DEAD Baclight viability kleuring (Molecular Probes Europe BV, Leiden, The Netherlands) kleuren dode bacteriën rood en levende bacteriën groen.

  • Bacterial interaction forces in adhesion dynamics

    Proefschrift

    ter verkrijging van het doctoraat in de Medische Wetenschappen

    aan de Rijksuniversiteit Groningen op gezag van de

    Rector Magnificus, Dr. F. Zwarts, in het openbaar te verdedigen op

    woensdag 14 januari 2009 om 16.15 uur

    door

    Niels Peter Boks geboren op 3 april 1979

    te Apeldoorn

  • Promotores: Prof. dr. ir. H.J. Busscher Prof. dr. ir. W. Norde

    Prof. dr. H.C. van der Mei Beoordelingscommissie: Prof. dr. Y. Ren Prof. dr. ir. M. Cohen-Stuart Prof. dr. ir. M. van Loosdrecht

  • Paranimfen: Anika Embrechts

    Eefje Engels

    Voor Froukje Visser - de Vries * 08-12-1906 † 24-03-1995

  • CONTENTS

    Chapter 1 General introduction 1

    Chapter 2 Forces involved in bacterial adhesion to hydrophilic

    and hydrophobic surfaces

    9

    Chapter 3 Residence time dependent desorption of Staphylococcus epidermidis from hydrophilic and hydrophobic substrata

    45

    Chapter 4 Mobile and immobile adhesion of staphylococcal strains to hydrophilic and hydrophobic surfaces

    57

    Chapter 5 Bond-strengthening in staphylococcal adhesion to hydrophilic and hydrophobic surfaces using AFM

    77

    Chapter 6 Fibronectin interactions with Staphylococcus aureus with and without fibronectin-binding proteins and their role in adhesion and desorption

    95

    Chapter 7 General discussion

    121

    Summary

    131

    Samenvatting

    137

    Dankwoord

    145

    Curriculum vitae 150

  • CHAPTER 1

    GENERAL INTRODUCTION

  • Chapter 1

    2

    Microbial adhesion

    Adhesion of micro-organisms to surfaces and their subsequent growth into a

    biofilm is a problem occurring in many fields of application. Bacterial biofilms

    on pipes and heat exchangers in industry result in high costs when equipment

    has to be cleaned or replaced [1,2]. In medicine, biofilms on biomedical

    materials (like prosthetic implants or urinary catheters) lead to infections,

    causing expensive treatments and enormous discomfort, and sometimes even

    death of a patient [3-5]. In order to prevent these problems, it is important to

    know more about microbial affinity for substratum surfaces, which governs the

    first step in biofilm formation. Microbial affinity can be expressed in terms of

    initial adhesion rate [6-8], deposition (or collision) efficiency [9] or in number

    of adhering bacterial cells after a few hours [8]. However, none of these

    parameters provide information on the adhesion strength between a micro-

    organism and substratum surface.

    The strength of microbial adhesion to substratum surfaces, is at least

    equally important if not more so than data on numbers of adhering organisms. In

    medicine bacterial desorption from one location may lead to an infection

    elsewhere in the body. An example of the importance of bacterial adsorption

    and desorption is found in the daily use of contact lenses. Wearing contact

    lenses increases the risk of microbial keratitis (i.e. inflammation of the cornea)

    [10]. Bacterial adhesion to contact lenses occurs during manual handling while

    putting the lens onto the eye, but also during storage in a lens box [11]. Once the

    contact lens is placed on the epithelium of the cornea, bacterial desorption from

    the lens surfaces occurs and bacteria may adhere to the epithelium, with the

    possibility to cause microbial keratitis.

  • General introduction

    3

    Adhesion forces

    In literature, adhesion strengths between bacteria and surfaces are calculated

    theoretically using the (extended-) DLVO theory (named after Derjaguin,

    Landau, Verwey and Overbeek) [12-14] or measured e.g. by using centrifugal

    force assays [15], laser tweezers [16-18] or total internal reflection microscopy

    related techniques [19-22]. Most frequently used however, are atomic force

    microscopy (AFM) [23-25] and fluid flow devices [26-29]. Because the flow

    profile in these devices is well controlled [30], they are suitable to determine the

    shear rate to prevent adhesion of bacteria or to detach adhering bacteria. These

    shear rates provide an averaged adhesion force for a bacterial population.

    Conversely, in AFM adhesion forces are probed directly between a substratum

    surface and an individual bacterial cell.

    As can be seen from Table 1, the magnitude of the force range estimated

    for microbial interaction forces with substratum surfaces is greatly dependent on

    the method used. For example, predicting interaction forces using the DLVO-

    theory result in the weakest forces, while AFM yields forces that are up to 105

    times stronger. It is unclear why these different techniques each yield their own

    class of force values.

    Table 1. Average force ranges for the interaction of micro-organisms with substratum

    surfaces reported in the literature and obtained with different techniques.

    Method of force measurement Force range (N) References

    Fluid flow devices 10-13 – 10-11 [26,31]

    Air bubble detachment 10-9 – 10-7 [32-34]

    AFM 10-10 – 10-9 [35-38]

    DLVO 10-14 – 10-10 [12,39,40]

  • Chapter 1

    4

    In conclusion, commonly accepted, well documented data on microbial

    adhesion forces do not exist, because different methods yield widely varying

    results and in many cases studies do not contain enough strains to warrant

    generally valid conclusions [41].

    Aim of this thesis

    The main aim of this thesis is to develop an understanding of the reason(s) why

    different techniques yield different ranges for microbial interaction forces with

    substratum surfaces. To achieve this aim, adhesion forces, together with data on

    adhesion dynamics, will be systematically obtained on hydrophobic and

    hydrophilic surfaces for a wide variety of bacterial strains and using different

    techniques.

    References 1. Videla, H.A. (2002), Prevention and control of biocorrosion, Int Biodeter Biodegr 49,

    259 - 270.

    2. Visser, J. and Jeurnink, T.J.M. (1997), Fouling of heat exchangers in the dairy

    industry, Exp Therm Fluid Sci 14, 407 - 424.

    3. Dankert, J., Hogt, A.H. and Feijen, J. (1986), Biomedical polymers - Bacterial

    adhesion, colonization, and infection, Crit Rev Biocompat 2, 219 - 301.

    4. Harris, L.G. and Richards, R.G. (2006), Staphylococci and implant surfaces: A review, Injury 37, 3 - 14.

  • General introduction

    5

    5. Schierholz, J.M. and Beuth, J. (2001), Implant infections: A haven for opportunistic bacteria, J Hosp Inf 49, 87 - 93.

    6. Gallardo-Moreno, A.M., Gonzalez-Martin, M.L., Bruque, J.M. and Perez-Giraldo,

    C. (2004), The adhesion strength of Candida parapsilosis to glass and silicone as a

    function of hydrophobicity, roughness and cell morphology, Colloids and Surface A

    249, 99 - 103.

    7. Van Merode, A.E.J., Duval, J.F.L., Van der Mei, H.C., Busscher, H.J. and Krom,

    B.P. (2008), Increased adhesion of Enterococcus faecalis strains with bimodal

    electrophoretic mobility distributions, Colloids Surface B 64, 302 - 306.

    8. Roosjen, A., Kaper, H.J., Van der Mei, H.C., Norde, W. and Busscher, H.J. (2003),

    Inhibition of adhesion of yeasts and bacteria by poly(ethylene oxide)-brushes on glass

    in a parallel plate flow chamber, Microbiol-Sgm 149, 3239 - 3246.

    9. Cail, T.L. and Hochella, M.F. (2005), the effects of solution chemistry on the sticking

    efficiencies of viable enterococcus faecalis: an atomic force microscopy and modeling

    study, Geochim Cosmochim Ac 69, 2959 - 2969.

    10. Bourcier, T., Thomas, F., Borderie, V., Chaumeil, C. and Laroche, L. (2003),

    Bacterial keratitis: Predisposing factors, clinical and microbiological review of 300

    cases, Brit J Ophthalmol 87, 834 - 838.

    11. Vermeltfoort, P.B.J., van Kooten, T.G., Bruinsma, G.M., Hooymans, A.M.M., Van

    der Mei, H.C. and Busscher, H.J. (2005), Bacterial transmission from contact lenses

    to porcine corneas: An ex vivo study, Invest Ophth Vis Sci 46, 2042 - 2046.

    12. Sharma, P.K. and Rao, K.H. (2003), Adhesion of Paenibacillus polymyxa on chalcopyrite and pyrite: Surface thermodynamics and extended DLVO theory, Colloids

    Surface B 29, 21 - 38.

    13. Vijayalakshmi, S.P. and Raichur, A.M. (2003), The utility of Bacillus subtilis as a

    bioflocculant for fine coal, Colloids Surface B 29, 265 - 275.

  • Chapter 1

    6

    14. Walker, S.L., Redman, J.A. and Elimelech, M. (2004), Role of cell surface lipopolysaccharides in Escherichia coli K12 adhesion and transport, Langmuir 20,

    7736 - 7746.

    15. Prakobphol, A., Burdsal, C.A. and Fisher, S.J. (1995), Quantifying the strength of

    bacterial adhesive interactions with salivary glycoproteins, J Dent Res 74, 1212 - 1218.

    16. Fallman, E., Schedin, S., Jass, J., Andersson, M., Uhlin, B.E. and Axner, O. (2004), Optical tweezers based force measurement system for quantitating binding interactions:

    system design and application for the study of bacterial adhesion, Bios Bioelectron 19,

    1429 - 1437.

    17. Maier, B., Koomey, M. and Sheetz, M.P. (2004), A force-dependent switch reverses

    type IV pilus retraction, P Natl Acad Sci USA 101, 10961 - 10966.

    18. Liang, M.N., Smith, S.P., Metallo, S.J., Choi, I.S., Prentiss, M. and Whitesides,

    G.M. (2000), Measuring the forces involved in polyvalent adhesion of uropathogenic

    Escherichia coli to mannose-presenting surfaces, P Natl Acad Sci USA 97,

    13092 - 13096.

    19. Clapp, A.R., Ruta, A.G. and Dickinson, R.B. (1999), Three-dimensional optical

    trapping and evanescent wave light scattering for direct measurement of long range

    forces between a colloidal particle and a surface, Rev Sci Instrum 70, 2627 - 2636.

    20. Sharp, J.M., Clapp, A.R. and Dickinson, R.B. (2003), Measurement of long-range

    forces on a single yeast cell using a gradient optical trap and evanescent wave light

    scattering, Colloids Surface B 27, 355 - 364.

    21. Geggier, P. and Fuhr, G. (1999), A time-resolved total internal reflection aqueous

    fluorescence (TIRAF) microscope for the investigation of cell adhesion dynamics, Appl

    Phys A-Mater 68, 505 - 513.

    22. Prieve, D.C. (1999), Measurement of colloidal forces with TIRM, Adv Coll Int Sci 82, 93 - 125.

  • General introduction

    7

    23. Bowen, W.R., Hilal, N., Lovitt, R.W. and Wright, C.J. (1999), Characterisation of membrane surfaces: direct measurement of biological adhesion using an atomic force

    microscope, J Membrane Sci 154, 205 - 212.

    24. Van der Aa, B.C. and Dufrene, Y.F. (2002), In situ characterization of bacterial extracellular polymeric substances by AFM, Colloids Surface B 23, 173 - 182.

    25. Razatos, A., Ong, Y.L., Boulay, F., Elbert, D.L., Hubbell, J.A., Sharma, M.M. and

    Georgiou, G. (2000), Force measurements between bacteria and poly(ethylene glycol)-

    coated surfaces, Langmuir 16, 9155 - 9158.

    26. Rutter, P.R. and Vincent, B. (1988), Attachment mechanisms in the surface growth of

    microorganisms. In: Physiological models in microbiology. Bazin, M. J. and Prosser, J.

    I. (Eds.), Boca Raton, Florida:CRC Press, Inc. pp 87 - 107.

    27. Owens, N.F., Gingell, D. and Rutter, P.R. (1987), Inhibition of cell-adhesion by a

    synthetic-polymer adsorbed to glass shown under defined hydrodynamic stress, J Cell

    Sci 87, 667 - 675.

    28. Thomas, W.E., Nilsson, L.M., Forero, M., Sokurenko, E.V. and Vogel, V. (2004),

    Shear-dependent 'stick-and-roll' adhesion of type 1 fimbriated Escherichia coli, Mol

    Microbiol 53, 1545 - 1557.

    29. Duddridge, J.E., Kent, C.A. and Laws, J.F. (1982), Effect of surface shear-stress on

    the attachment of Pseudomonas fluorescens to stainless-steel under defined flow

    conditions, Biotechnol Bioeng 24, 153 - 164.

    30. Busscher, H.J. and Van der Mei, H.C. (2006), Microbial adhesion in flow

    displacement systems, Clin Microbiol Rev 19, 127 - 141.

    31. Shive, M.S., Hasan, S.M. and Anderson, J.M. (1999), Shear stress effects on bacterial adhesion, leukocyte adhesion, and leukocyte oxidative capacity on a polyetherurethane,

    J Biomed Mater Res 46, 511 - 519.

    32. Gomez-Suarez, C., Busscher, H.J. and Van der Mei, H.C. (2001), Analysis of bacterial detachment from substratum surfaces by the passage of air-liquid interfaces,

    Appl Environ Microb 67, 2531 - 2537.

  • Chapter 1

    8

    33. Roosjen, A., Busscher, H.J., Norde, W. and Van der Mei, H.C. (2006), Bacterial factors influencing adhesion of Pseudomonas aeruginosa strains to a poly(ethylene

    oxide) brush, Microbiol-Sgm 152, 2673 - 2682.

    34. Noordmans, J., Wit, P.J., Van der Mei, H.C. and Busscher, H.J. (1997), Detachment

    of polystyrene particles from collector surfaces by surface tension forces induced by air-

    bubble passage through a parallel plate flow chamber, J Adhes Sci Technol 11,

    957 - 969.

    35. Camesano, T.A. and Logan, B.E. (2000), Probing bacterial electrosteric interactions

    using atomic force microscopy, Environ Sci Technol 34, 3354 - 3362.

    36. Dufrene, Y.F., Boonaert, C.J.P., Gerin, P.A., Asther, M. and Rouxhet, P.G. (1999),

    Direct probing of the surface ultrastructure and molecular interactions of dormant and

    germinating spores of Phanerochaete chrysosporium, J Bacteriol 181, 5350 - 5354.

    37. Fang, H.H.P., Chan, K.Y. and Xu, L.C. (2000), Quantification of bacterial adhesion

    forces using atomic force microscopy (AFM), J Microbiol Meth 40, 89 - 97.

    38. Emerson, R.J. and Camesano, T.A. (2004), Nanoscale investigation of pathogenic

    microbial adhesion to a biomaterial, Appl Environ Microb 70, 6012 - 6022.

    39. Busalmen, J.P. and de Sanchez, S.R. (2001), Adhesion of Pseudomonas fluorescens

    (ATCC 17552) to nonpolarized and polarized thin films of gold, Appl Environ Microb

    67, 3188 - 3194.

    40. Jacobs, A., Lafolie, F., Herry, J.M. and Debroux, M. (2007), Kinetic adhesion of bacterial cells to sand: Cell surface properties and adhesion rate, Colloids Surface B 59,

    35 - 45.

    41. Bakker, D.P., Postmus, B.R., Busscher, H.J. and Van der Mei, H.C. (2004),

    Bacterial strains isolated from different niches can exhibit different patterns of adhesion

    to substrata, Appl Environ Microb 70, 3758 - 3760.

  • CHAPTER 2

    FORCES INVOLVED IN BACTERIAL ADHESION TO

    HYDROPHILIC AND HYDROPHOBIC SURFACES

    Parts of this chapter are reproduced with permission of the Society of General

    Microbiology from : Boks, N.P., Norde, W., Van der Mei, H.C. and Busscher, H.J. (2008),

    Microbiology 154, 3122-3133.

  • Chapter 2

    10

    Abstract

    Using a parallel plate flow chamber, the hydrodynamic shear forces to prevent

    bacterial adhesion (Fprev) and to detach adhering bacteria (Fdet) were evaluated

    for hydrophilic glass, hydrophobic, dimethyldichlorosilane (DDS)-coated glass

    and six different bacterial strains, in order to test the following three hypotheses:

    1. A strong hydrodynamic shear force to prevent adhesion relates to a strong

    hydrodynamic shear force to detach an adhering organism.

    2. A weak hydrodynamic shear force to detach adhering bacteria implies that

    more bacteria will be stimulated to detach by a passing air-liquid interface

    through the flow chamber.

    3. DLVO interactions determine the characteristic hydrodynamic shear

    forces to prevent adhesion and to detach adhering micro-organisms as

    well as the detachment induced by a passing air-liquid interface.

    Fprev varied from 0.03 to 0.70 pN, while Fdet varied between 0.31 to over 19.64

    pN, suggesting that after initial contact, strengthening of the bond occurs.

    Generally, it was more difficult to detach bacteria from DDS-coated glass than

    from hydrophilic glass, which was confirmed by air-bubble detachment studies.

    Calculated attractive forces based on the DLVO theory (FDLVO) towards the

    secondary interaction minimum were higher on glass than on DDS-coated glass.

    In general, all three hypotheses had to be rejected, showing that it is of

    importance to distinguish between forces acting parallel (hydrodynamic shear)

    and perpendicular (DLVO, air-liquid interface passages) to the substratum

    surface.

  • Forces involved in bacterial adhesion

    11

    Introduction

    Microbial adhesion and subsequent biofilm formation occur in many fields of

    industrial and medical applications, such as on ship hulls, heat exchanger plates,

    food packaging materials and biomaterials implants, including urinary catheters,

    contact lenses, and vascular grafts [1-3]. Common in most applications is the

    deposition of micro-organisms to a surface from a flowing suspension. This

    implies that a variety of forces act on depositing and already adhering

    organisms. Deposition is mainly governed by Brownian motion, sedimentation

    and hydrodynamic forces, while actual adhesion of micro-organisms to a

    substratum surface is mediated by Lifshitz-Van der Waals, electrostatic, acid-

    base and hydrophobic interaction forces [4].

    Fluid flow is an important factor in microbial deposition [5]. An increase

    in fluid flow velocity will in a first instance, yield increased microbial transport

    towards a substratum surface (convective-diffusion), but at the same time causes

    an increase in hydrodynamic detachment forces. Shear is the dominant effect of

    fluid flow and can be well controlled in experimental systems, like on rotating

    disks, at stagnation points and in parallel plate flow chambers. In principle, two

    critical shear rates can be distinguished based on current literature (see Table 1):

    a critical shear rate to prevent adhesion and a critical shear rate to stimulate

    detachment of already adhering organisms. Both critical shear rates vary from

    strain to strain and also depend on the substratum material involved. The shear

    rates and, hence, the shear forces, required to stimulate detachment are generally

    higher than the shear rates to prevent adhesion.

    Detachment can also be invoked by allowing an air-bubble to pass over

    adhering bacteria. The passage of an air-liquid interface is accompanied by a

    perpendicularly oriented force of around 10-7 N, which is much higher than the

    hydrodynamic shear forces acting parallel to a substratum surface. Yet, a

  • Cha

    pter

    2

    4Tabl

    e 1.

    Sum

    mar

    y of

    inte

    ract

    ion

    forc

    es b

    etw

    een

    bact

    eria

    and

    subs

    tratu

    m su

    rfac

    es, t

    oget

    her w

    ith th

    e m

    etho

    d ap

    plie

    d.

    Stra

    in

    Subs

    trat

    um

    Forc

    e (p

    N)

    Met

    hod

    Ref

    eren

    ce

    Esch

    eric

    hia

    coli

    prot

    ein

    coat

    ings

    0.

    2

    [6]

    Stap

    hylo

    cocc

    us e

    pide

    rmid

    is

    seve

    ral b

    iom

    ater

    ials

    1.

    2 –

    1.4

    Hyd

    rody

    nam

    ic fo

    rce

    [7-9

    ] St

    aphy

    loco

    ccus

    aur

    eus

    Col

    lage

    n 0.

    4 to

    pre

    vent

    adh

    esio

    n [1

    0]

    Pseu

    dom

    onas

    fluo

    resc

    ens

    stai

    nles

    s ste

    el

    9.2

    – 12

    .3

    [1

    1]

    Stre

    ptoc

    occu

    s san

    guis

    G

    lass

    22

    .0

    [1

    2]

    Baci

    llus c

    ereu

    s gl

    ass a

    nd si

    licon

    ized

    gla

    ss

    43.1

    – 8

    0.1

    [1

    2]

    Esch

    eric

    hia

    coli

    hydr

    opho

    bic

    subs

    trate

    s 3.

    1 –

    4.6

    [1

    3]

    Stap

    hylo

    cocc

    us e

    pide

    rmid

    is

    mod

    ified

    PV

    C

    0.1

    – 1.

    2 H

    ydro

    dyna

    mic

    forc

    e [1

    4]

    Stap

    hylo

    cocc

    us a

    ureu

    s C

    olla

    gen

    >>3.

    9 to

    det

    ach

    adhe

    ring

    [1

    5]

    Pseu

    dom

    onas

    fluo

    resc

    ens

    stai

    nles

    s ste

    el

    18.5

    ba

    cter

    ia

    [11]

    M

    ix o

    f Gra

    m p

    ositi

    ve c

    occi

    gl

    ass,

    silic

    oniz

    ed g

    lass

    and

    stee

    l 20

    .4 –

    42.

    4

    [12]

    Es

    cher

    ichi

    a co

    li Q

    uartz

    0.

    3 –

    2.4

    [1

    6]

    Stap

    hylo

    cocc

    us e

    pide

    rmid

    is

    PMM

    A

    11.1

    [17]

    Ps

    eudo

    mon

    as fl

    uore

    scen

    s G

    old

    32.1

    – 5

    7.9

    [1

    8]

    Baci

    llus c

    ereu

    s Sa

    nd

    0.03

    D

    LVO

    cal

    cula

    tion

    [19]

    Ba

    cillu

    s sub

    tilus

    C

    oal

    0.09

    [20]

    Ba

    cillu

    s sub

    tilus

    Sa

    nd

    0.03

    [19]

    Pa

    enib

    acill

    us p

    olym

    yxa

    Pyrit

    e –

    chal

    copy

    rite

    170

    – 56

    0

    [21]

    Sp

    hing

    omon

    as p

    auci

    mob

    ilis

    Gla

    ss

    0.07

    – 0

    .7

    [2

    2]

    Esch

    eric

    hia

    coli

    silic

    on su

    rfac

    es

    7400

    – 2

    2800

    [23]

    Es

    cher

    ichi

    a co

    li si

    licon

    nitr

    ide

    tip

    400

    – 21

    00

    Ato

    mic

    For

    ce

    [24]

    St

    aphy

    loco

    ccus

    epi

    derm

    idis

    si

    licon

    nitr

    ide

    tip

    2000

    M

    icro

    scop

    e [2

    5]

    Spor

    es o

    f Bac

    illus

    myo

    cide

    s hy

    drop

    hobi

    cally

    coa

    ted

    glas

    s 74

    00 –

    495

    00

    [2

    6]

    Esch

    eric

    hia

    coli

    gala

    bios

    e-fu

    nctio

    naliz

    ed b

    eads

    50

    – 1

    00

    [2

    7]

    Stap

    hylo

    cocc

    us e

    pide

    rmid

    is

    fibro

    nect

    in c

    oatin

    gs

    18

    Opt

    ical

    Tw

    eeze

    rs

    [28]

    St

    aphy

    loco

    ccus

    aur

    eus

    fibro

    nect

    in c

    oatin

    gs

    15 –

    26

    [2

    9]

    Hyd

    rody

    nam

    ic fo

    rces

    are

    cal

    cula

    ted

    usin

    g F

    = η σ

    Ap,

    in w

    hich

    η th

    e ab

    solu

    te v

    isco

    sity

    of w

    ater

    and

    Ap t

    he a

    rea

    of th

    e pa

    rticl

    e ex

    pose

    d to

    shea

    r. C

    occi

    wer

    e as

    sum

    ed to

    hav

    e a

    radi

    us o

    f 0.

    5 μm

    , whi

    le ro

    d-sh

    aped

    bac

    teria

    wer

    e ap

    prox

    imat

    ed b

    y sp

    here

    s w

    ith e

    qual

    vol

    ume,

    usi

    ng 0

    .7 μ

    m a

    s ra

    dius

    . DLV

    O-f

    orce

    s are

    take

    n as

    the

    attra

    ctiv

    e fo

    rce

    tow

    ards

    the

    pred

    icte

    d se

    cond

    ary

    min

    imum

    in th

    e to

    tal i

    nter

    actio

    n en

    ergy

    cur

    ves.

    Chapter 2

    12

  • Forces involved in bacterial adhesion

    13

    passing air-liquid interface does not cause complete bacterial detachment for all

    combinations of strains and substratum surfaces.

    Gomez-Suarez et al. [30] investigated detachment of several bacterial

    strains from hydrophilic and hydrophobic surfaces by a passing air-bubble.

    Depending on the strain involved, the presence of a conditioning film and the

    velocity of the air-bubble, detachment ranged from 0 to 90%. Although air-

    bubble induced detachment is relatively easy to measure, it only yields an

    extremely rough estimate of a detachment force threshold and it cannot be used

    to estimate the actual binding strength.

    Perpendicularly oriented interaction forces can be measured more

    directly, for instance using atomic force microscopy (AFM) or optical tweezers.

    As can be seen in Table 1, forces obtained using these techniques, differ in

    orders of magnitude. Forces measured with optical tweezers remain in the pN

    range, while AFM yields stronger forces than any other method, which are

    generally in the nN range.

    Another, often used approach for assessing adhesion strength is the

    (extended) DLVO theory (named after Derjaguin, Landau, Verwey and

    Overbeek). In the DLVO theory the binding strength between colloidal particles,

    such as micro-organisms, and substratum surfaces may be calculated on the

    basis of Lifshitz-Van der Waals, (acid-base and) electrical double layer

    interactions. Usually, also the theoretical values provide a distinct class of force

    values, that cannot be easily matched with experimental values, as reported in

    the literature.

    From Table 1, it is obvious that throughout the literature different types of

    forces may be distinguished for every strain-substratum combination.

    Furthermore, conclusions on bacterial adhesion mechanisms are often based

    on not more than two strains [31]. Comparing all reported data is further

    complicated by the fact that different suspending media are used to determine

  • Chapter 2

    14

    adhesion parameters on different substrata. It is currently unclear why different

    methods to evaluate bacterial binding forces yield distinct classes of force values

    that often differ by orders of magnitude. The aim of our research is to gain more

    insight in the relevance of the different bacterial interaction force indicators,

    including theoretically predicted interaction forces from the DLVO-theory, and

    their mutual relationships. To this end, the following hypotheses were tested:

    1. A strong hydrodynamic shear force to prevent adhesion relates to a strong

    hydrodynamic shear force to detach an adhering organism.

    2. A weak hydrodynamic shear force to detach adhering bacteria implies that

    more bacteria will be stimulated to detach by a passing air-liquid interface

    through the flow chamber.

    3. DLVO interactions determine the characteristic hydrodynamic shear

    forces to prevent adhesion and to detach adhering micro-organisms as

    well as the detachment induced by a passing air-liquid interface.

    To test these hypotheses, the critical shear force to prevent bacterial adhesion

    and to stimulate detachment of adhering bacteria are determined. Hydrophilic

    glass and hydrophobic, dimethyldichlorosilane-coated, glass are employed as

    substrata. To allow for more general conclusions to be drawn, six widely

    different bacterial strains are included. In addition, theoretical DLVO interaction

    forces, as calculated from measured zeta potentials and contact angles are

    determined. Furthermore, the detachment force threshold is evaluated for

    detachment caused by a passing air-liquid interface.

  • Forces involved in bacterial adhesion

    15

    Materials and Methods

    Bacterial strains and culture conditions. Staphylococcus epidermidis ATCC

    35983, S. epidermidis HBH2 169, Pseudomonas aeruginosa D1, P. aeruginosa

    KEI 1025 were cultured aerobically from blood agar plates in 10 ml Tryptone

    Soya Broth (OXOID, Basingstoke, England) for 24 h at 37 ºC. Raoultella

    terrigena ATCC 33527 was precultured aerobically from nutrient agar (Nutrient

    Broth, OXOID) in 10 ml nutrient broth for 24 h at 37 ºC. Streptococcus

    thermophilus ATCC 19258 was precultured from a frozen stock in 10 ml M17

    broth for 24 h at 37 ºC. After 24 h, precultures were used to inoculate 200 ml

    main cultures, which were grown for 16 h under similar conditions as the

    corresponding precultures. S. epidermidis and P. aeruginosa strains were

    harvested by centrifugation for 5 min at 6500 x g, while R. terrigena and S.

    thermophilus were harvested at 10000 x g. All strains were washed twice with

    10 mM potassium phosphate buffer at pH 7 and resuspended in the same buffer.

    To break bacterial chains or clusters, sonication at 30 W (Vibra Cell model 375,

    Sonics and Materials Inc., Danbury, CT, USA) was carried out for

    staphylococcal (3 times 10 s) and streptococcal (2 times 10 s) suspensions, while

    cooling in an ice/water bath. Subsequently, bacteria were resuspended to a

    concentration of 3 x 108 cells ml-1. In the calculations discussed below, the cocci

    were assumed to have a radius of 0.5 μm. Rod-shaped P. aeruginosa (2.5 μm x

    0.9 μm) and R. terrigena (3.2 μm x 1.4 μm) were approximated as spheres with

    equal volume, using a radius of 0.6 μm and 0.9 μm, respectively as they adhere

    in different orientations, i.e. “end-on” and “side-on”.

    Substratum surfaces. Glass slides were sonicated during 3 min in 2% RBS35

    (Omnilabo International BV, The Netherlands) followed by thorough rinsing

    with tap water, demineralised water, methanol, tap water and finally

  • Chapter 2

    16

    demineralised water again to obtain a hydrophilic surface. After washing, the

    slides were either directly used or dried for 4 h at 80 ºC prior to applying a

    hydrophobic coating. To obtain a hydrophobic surface, the dried glass slides

    were submerged during 15 min in a solution of dimethyldichlorosilane (DDS,

    Merck, Germany) in trichloroethylene (0.05 w/v %) and washed with

    trichloroethylene, methanol and ultrapure water. Prepared slides were stored for

    no longer than 3 days at room temperature and rinsed with 10 mM potassium

    phosphate buffer before use.

    Bacterial adhesion in the parallel plate flow chamber. The parallel plate flow

    chamber (PPFC) and image analysis have been described previously [32]. The

    flow chamber used in this study has a length of 175 mm, a depth of 0.75 mm

    and a width of 17 mm. Prior to use, the flow chamber was washed with 2%

    Extran (Merck, Germany) and rinsed thoroughly with tap water and

    demineralised water before mounting a clean substratum surface in the PPFC.

    Subsequently, the flow chamber was installed between two communicating

    vessels and the system was filled with 10 mM potassium phosphate buffer while

    care was taken to remove all air-bubbles. When the PPFC was positioned under

    the microscope, the vessels containing bacterial suspension were positioned at

    different heights to create a flow. The difference in fluid levels was maintained

    by a roller-pump to ensure a circulating pulse free flow throughout the duration

    of an entire experiment. Deposition of bacteria was monitored with a phase

    contrast microscope (Olympus HB-2) equipped with a 40x ultra long working

    distance objective (Olympus ULWD-CD Plan 40 PL) which was connected to a

    CCD-MXRi camera (Basler A101F, Basler AG, Germany). Images were

    obtained by summation of 15 consecutive images (time interval 0.25 s) in order

    to enhance the signal to noise ratio and eliminate moving bacteria from analysis.

    Analysis of the images was done using proprietary software based on the Matlab

    Image processing Toolkit (The MathWorks, MA, USA)).

  • Forces involved in bacterial adhesion

    17

    Shear rate dependent adhesion. The bacterial suspension was allowed to flow

    through the flow chamber during 1 h at flow rates (Q) of 1, 5, 10, 19, 57, 77, 105

    and 153 ml min-1 which corresponds to shear rates (σ) of 10, 50, 100, 200, 600,

    800, 1100 and 1600 s-1. Under these conditions the flow is laminar and bacterial

    transport occurs by convective-diffusion. Adhesion is monitored on both the top

    (negative contribution of sedimentation) and bottom (positive contribution of

    sedimentation) plate of the PPFC. For each shear rate, the number of bacteria

    adhering per unit area was recorded as a function of time. Adhesion was then

    expressed in initial deposition rates j0 (cm-2 s-1), while at the end of each

    experiment an air-bubble was passed through the flow chamber to stimulate

    detachment (only evaluated for the bottom plate).

    Initial deposition rates for the top and bottom plate were averaged and expressed

    as deposition efficiencies by normalization with respect to the Von

    Smoluchowski-Levich (SL) theoretical upper limit for deposition in the parallel

    plate flow chamber. The SL upper limit for bacterial deposition is an

    approximate solution of the convective-diffusion equation and assumes perfect

    sink conditions at the substratum surface (i.e. every particle that arrives at the

    surface actually adheres) in the absence of sedimentation. The theoretical upper

    limit for deposition is given by [33]:

    31

    *0 9

    289.0 ⎥⎦

    ⎤⎢⎣⎡ ⋅= ∞

    xbPe

    rcDj (1)

    in which D∞ is the diffusion coefficient of the particles (taken 3.1 x 10-13 m2 s-1

    for micron-sized bacteria [34]), c the concentration of bacteria in suspension, r

    the bacterial radius, x the longitudinal distance from the flow chamber entrance,

    b the half-depth of the PPFC and Pe the dimensionless Péclet number. This

    latter is defined as:

  • Chapter 2

    18

    =Dwb

    QrPe 33

    43 (2)

    in which Q is the applied flow rate and w the width of the flow chamber

    Detachment induced by a passing air-liquid interface. Following the deposi-

    tion measurement, an air-liquid interface was introduced by passing an air-

    bubble through the flow chamber, which is accompanied by a perpendicularly

    oriented detachment force equal to [35]:

    sw,bw, ΘΘ cos

    2sin2 2max ⎟

    ⎞⎜⎝

    ⎛⋅⋅= lvrF γπγ for Θw,s < 90 (3)

    sw,bw, ΘΘ cos

    2sin2 2max ⎟

    ⎞⎜⎝

    ⎛ +⋅⋅−=π

    γπγ lvrF for Θw,s > 90 (4)

    in which γlv represents the interfacial surface tension of the liquid and vapour,

    Θw,b and Θw,s denote the bacterial- and substratum-water contact angles,

    respectively.

    Shear rate dependent detachment of adhering bacteria. The flow system was

    filled and positioned as described before. Bacteria were resuspended in

    potassium phosphate buffer to a high concentration of 7.5 x 108 cells ml-1 to

    accelerate deposition and allowed to adhere to the collector surface at a shear

    rate of 25 s-1. After 20 min, flow was switched to fresh buffer without bacteria at

    25 s-1 to wash out the bacterial suspension for 30 min, after which the shear rate

    was increased to either 250, 1000, 3000, 6650 or 7320 s-1 for 30 min. The

    number of bacteria that remained adhering was enumerated after each step.

  • Forces involved in bacterial adhesion

    19

    Surface characterization. To determine the zeta potentials of the substrata,

    streaming potentials were measured in 10 mM phosphate buffer at pH 7.

    Collector surfaces were mounted in a homemade parallel plate flow chamber,

    separated by a 0.1 mm Teflon spacer. A platinum electrode was placed at either

    end of the chamber. Streaming potentials were measured at 10 different

    pressures ranging from 5 x 103 to 20 x 103 Pa. Each pressure was applied during

    10 s in both directions. Zeta potentials were deduced by linear least squares

    fitting from the pressure dependent streaming potentials [36].

    For bacterial zeta potentials, bacteria were washed with demineralised water and

    resuspended in 10 mM potassium phosphate buffer at pH 7 to a concentration of

    1 x 108 cells ml-1. The electrophoretic mobilities of these suspensions were

    measured at 150 V using a Lazer Zee Meter 501 (PenKem, USA). The

    electrophoretic mobilities were converted to apparent zeta potentials assuming

    the Helmholtz-Von Smoluchowski approximation holds, which is appropriate

    considering the high value for κ r (i.e. ≈ 150) in the systems used (N.B. κ

    denotes the reciprocal Debeye length which is directly related to the ionic

    strength [37]).

    To calculate surface free energies of the substratum and bacterial cell surfaces,

    sessile drop contact angles were measured with water, formamide, α-

    bromonaphthalene and methylene iodide. In order to measure contact angles

    with liquids on bacteria, bacterial lawns were prepared by depositing bacteria

    from suspension in demineralised water on cellulose acetate membrane filters

    (Millipore, pore diameter 0.45 μm) under negative pressure until approximately

    50 layers were stacked. Subsequently, filters were fixed on a sample holder and

    left to dry until “plateau contact angles” could be measured, i.e. water contact

    angles that remained stable over time for 30–60 min. All contact angles were

    measured in triplicate, implying separate substrata and different bacterial

  • Chapter 2

    20

    cultures. Measured contact angles were converted into surface free energies

    using:

    lv

    pluslv

    plussv

    lv

    lvsv

    lv

    LWlv

    LWsv

    γγγ

    γγγ

    γγγ 222

    1cos +++−=minusminus

    Θ (5)

    in which γLWsv is the Lifshitz-Van der Waals component of the surface free

    energy of the surface of interest (i.e. substratum surface or bacterial lawn) and γlv

    is the surface free energy of the liquid vapour interface. The acid-base

    component of the surface free energies was separated into an electron donor

    (γminussv) and electron acceptor (γplussv) parameter, according to: plussvsv

    ABs γγγ

    minus2= (6)

    Interaction forces using the extended DLVO theory. In the extended DLVO

    theory, the interaction energy is divided in a Lifshitz-Van der Waals, acid-base

    and electrostatic contribution, while accounting for their distance dependencies.

    The Lifshitz-Van der Waals contribution can be derived by first calculating the

    Lifshitz-Van der Waals component of the free energy of adhesion of a bacterium

    to a substratum surface, which reads at contact [38]:

    ( )( )LWlvLWsvLWlvLWbvLWslbG γγγγ −−−=Δ 2 (7)

    Equation 7 can be used to calculate the Hamaker constant A according to [39]: 2012 dGA

    LWslb ⋅⋅Δ= π (8)

    where d0 denotes the minimal separation distance (0.157 nm [40]) and ΔGLWslb is

    obtained from thermodynamic analysis [38]. Lifshitz-Van der Waals attractive

    interaction energies (ΔGLW) were subsequently calculated as a function of

    distance assuming a sphere-plane geometry using [39]:

    ( )( ) ⎥⎦

    ⎤⎢⎣

    ⎡⎟⎠⎞

    ⎜⎝⎛ +−

    ++

    −=Δd

    rdrddrdrAdG LW 2ln)2

    26

    )( (9)

  • Forces involved in bacterial adhesion

    21

    in which d denotes the separation distance. The acid-base component, ΔGABslb,

    can be calculated from [38]:

    ( ) ( ) ( )( ) ( ) ( )⎥⎥⎦

    ⎢⎢

    −⋅−−−

    ⋅−−−⋅−=Δ

    minusminusplusplusminusminus

    pluspluminusminusplusplus

    lvsvlvsvlvbv

    lvs

    bvsvbvsvbvABslbG

    γγγγγγ

    γγγγγγ2 (10)

    in which the subscript “s” denotes the substratum surface and “b” the bacterial

    cell surface.

    Using Equation 10, the distance dependence of the acid-base interaction

    energies (ΔGAB) can then be calculated according to [39]:

    ⎟⎠⎞

    ⎜⎝⎛ −Δ⋅⋅=Δ

    λλπ

    ddGrdG ABslb

    AB 0exp2)( (11)

    in which λ denotes the correlation length of molecules in the liquid medium

    (estimated to be 0.6 nm [39]).

    Distance dependent electrostatic interaction energies (ΔGEL) were calculated

    using [41]:

    ( ) ( )( ) ( )[ ]⎭⎬⎫

    ⎩⎨⎧

    −−+⎥⎦

    ⎤⎢⎣

    ⎡−−−+

    ++=Δ d

    ddrdG

    sb

    sbsb

    EL κκκ

    φφφφ

    φφπεε 2exp1lnexp1exp1ln

    2)( 22

    220 (12)

    in which εε0 denotes the dielectric permittivity of the medium (i.e. water), φb and

    φs the surface (zeta) potentials of the bacterial cell surface and collector surface

    and κ the reciprocal Debye length.

    Summation and differentiation with respect to distance of these three

    components lead to the total DLVO-interaction energy and interaction force,

    respectively, as a function of separation distance. All DLVO interaction forces

    reported in this chapter represent the maximal attractive force towards the

    secondary interaction minimum, which was present in all bacterium-substratum

    systems investigated.

  • Chapter 2

    22

    Results

    Shear rates to prevent bacterial adhesion. Figure 1 presents an example of

    bacterial deposition to the bottom and top plate of the parallel plate flow

    chamber as a function of shear rate. Deposition is higher to the bottom plate than

    to the top plate, especially at lower shear rates. Moreover, at low shear rates an

    initial increase in deposition to the bottom plate can be seen with increasing

    shear rate up to 200 s-1 due to increased mass transport, above which deposition

    decreases with increasing shear due to detachment. A similar effect is observed

    on the top plate.

    Figure 1. Initial deposition rates (j0) for S. epidermidis ATCC 35983 on the bottom (●) and top plate (○) in a parallel plate flow chamber as a function of the shear rate (σ) applied on glass.

  • Forces involved in bacterial adhesion

    23

    The influence of sedimentation on mass transport can be eliminated by

    averaging bottom and top plate depositions. Figure 2 shows the deposition

    efficiencies (α) in the absence of sedimentation, as calculated from averaged

    initial deposition rates and the theoretical upper limit for deposition (Eq. 1) as a

    function of shear rate. From Figure 2, critical shear rates to prevent adhesion

    (σprev) were deduced using

    ⎪⎭

    ⎪⎬⎫

    ⎪⎩

    ⎪⎨⎧−⋅=

    prevσσαα exp0 (13)

    with α0 the extrapolated deposition efficiency in absence of shear. Subsequently,

    values for σprev were expressed in shear forces using

    ipi AF ση ⋅⋅= (14)

    in which η is the absolute viscosity of the buffer (1 x 10-3 Pa s) and Ap is the area

    of the adhering bacterium subject to shear flow. Furthermore, subscript “i”

    denotes the type of hydrodynamic force calculated: prev for the hydrodynamic

    force to prevent adhesion and det for the hydrodynamic force to detach adhering

    micro-organisms. Hydrodynamic shear forces to prevent adhesion (Fprev) are

    listed in Table 2. All values for Fprev remain in the low pN range and are

    influenced by the substratum surface, although not consistently higher on any of

    the two surfaces. Depending on the strain used, the difference between Fprev on

    glass and DDS-coated glass can be as large as a factor 6.

  • Chapter 2

    24

    Figure 2. Bacterial deposition efficiency (α) in the absence of a mass transport contribution

    from sedimentation as a function of the shear rate (σ) applied on glass (●) and DDS-coated

    glass (○) for the six different bacterial strains included. Black and grey lines represent the

    fits of equation 13 to the datapoints on glass and DDS-coated glass, respectively.

  • Forces involved in bacterial adhesion

    25

    Table 2. Critical shear forces to prevent (Fprev) bacterial adhesion and to detach (Fdet) adhering bacteria from a hydrophilic (glass) and hydrophobic (DDS-coated) substratum, as derived for different bacterial strains, together with the theoretically calculated DLVO interaction forces. Reported uncertainties are based on the standard error of the predicted fitting curve. Fprev (pN) Fdet (pN) FDLVO (pN)

    Bacterial

    strain

    Glass DDS Glass DDS Glass DDS

    S. epidermidis HBH2 169

    0.13 ± 0.06 0.40 ± 0.12 0.31 ± 0.03 5.52 ± 0.09 0.08 0.05

    S. epidermidis ATCC 35983

    0.57 ± 0.22 0.10 ± 0.03 5.39 ± 0.19 >5.75* 0.05 0.03

    R. terrigena ATCC 33527

    0.10 ± 0.00 0.11 ± 0.01 > 19.64* 14.30 ± 2.45 0.10 0.06

    S. thermophilus ATCC 19258

    0.12 ± 0.34 0.03 ± 0.01 0.55 ± 0.05 0.68 ± 0.04 0.00 0.00

    P. aeruginosa D1

    0.06 ± 0.02 0.04 ± 0.01 0 3.41 ± 1.00 0.05 0.03

    P. aeruginosa KEI 1025

    0.24 ± 0.02 0.70 ± 0.28 4.53 ± 0.82 9.93 ± 0.03 0.08 0.05

    * No detachment could be stimulated within the shear rates applied and the value indicated denotes the highest shear applied.

    Shear rates to remove adhering bacteria. Figure 3 presents the detachment of

    bacteria from glass and DDS as a function of the shear rate applied, expressed as

    the fraction (f) of bacteria removed from the substratum surface. For a given

    shear rate, f is defined as the number of removed bacteria after 30 min exposure

    to that shear divided by the number of adhering bacteria before application of

    the shear. From the plots in Figure 3 critical shear rates to detach adhering

    bacteria (σdet) were derived, defined as the shear rate at which 63% of the

    adhering bacteria had detached. Subsequently, these shear rates were expressed

    in detachment forces (Fdet) using Equation 14, and their values are listed in

  • Chapter 2

    26

    Table 2. In most cases, bacteria are more readily detached from glass than from

    DDS-coated glass. All forces remain in the pN range, but are an order of

    magnitude larger than Fprev. Note that the critical detachment level could not be

    reached within the range of shear rates possible in our experimental set-up for S.

    epidermidis ATCC 35983 on DDS-coated glass and for R. terrigena ATCC

    33527 on glass.

    Air-bubble induced bacterial detachment. Table 3 summarizes the effect of

    an air-bubble passing over the adhering bacteria. At first sight, binding affinity

    on DDS-coated glass seems to be less than on hydrophilic glass, as judged from

    air-bubble-induced detachment. However, on DDS-coated glass, the force

    exerted by an air-liquid interface on adhering bacteria is calculated to be up to

    five times larger than on glass. For the two Staphylococcus epidermidis strains

    and R. terrigena, this results in higher detachment percentages from DDS-coated

    glass. In contrast, the percentages detached from glass and DDS-coated glass,

    respectively, are for the pseudomonas strains and S. thermophilus not

    significantly different. It should be noted that detachment by a passing air-

    bubble does not give any indication of the magnitude of the interaction forces.

    For example, for the staphylococcal strains and R. terrigena on glass, it cannot

    be established at which force detachment would be stimulated to a larger extent.

    Air-bubble detachment studies are indecisive here with respect to binding

    strength information. However, for the pseudomonas strains and S. thermophilus

    it is clear that, even though the exerted force on DDS-coated glass is stronger,

    detachment percentages are not higher. Results for these strains suggest stronger

    interaction forces with the hydrophobic DDS-coated glass.

  • Forces involved in bacterial adhesion

    27

    Figure 3. Shear-induced detachment, expressed as the fraction (f) of bacteria that are removed, as a function of the shear rate (σ) applied for glass (●) and DDS-coated glass (○) after 30 min of flow.

    Surface characterization and calculation of theoretical interaction forces.

    Measured contact angles, together with the surface free energy components of

    the wetting liquids used, are listed in Table 4. All bacteria have a surface

    hydrophilicity comparable to the one of glass, as judged from the water contact

    angles. DDS-coated glass is significantly more hydrophobic.

  • Chapter 2

    28

    Table 3. Number of adhering bacteria on the bottom plate of the parallel plate flow chamber after 1 h of flow (N1h, averaged over adhesion experiments at σ = 10, 50, 100 and 200 s-1; n=1 for each shear rate), detachment percentages from glass and a DDS-coating and the corresponding maximal detachment force (Fγmax) a liquid/air interface exerts. Glass DDS

    Bacterial strain N1h

    (x 106 cm-2)

    Detach-

    ment

    (%)

    Fγmax

    (nN)

    N1h

    (x 106 cm-2)

    Detach-

    ment

    (%)

    Fγmax

    (nN)

    S. epidermidis HBH2 169

    4.9 ± 0.5 9 ± 10 14 3.8 ± 0.5 92 ± 9 40

    S. epidermidis ATCC 35983

    4.0 ± 0.8 4 ± 5 20 3.5 ± 0.8 62 ± 47 39

    R. terrigena ATCC 33527

    0.8 ± 0.8 27 ± 6 16 1.4 ± 1.0 87 ± 14 72

    S .thermophilus ATCC 19258

    0.4 ± 0.3 56 ± 16 17 0.5 ± 0.5 47 ± 21 39

    P. aeruginosa D1

    0.3 ± 0.4 71 ± 40 37 0.4 ± 0.2 40 ± 14 48

    P. aeruginosa KEI 1025

    1.3 ± 2.1 53 ± 10 12 2.9 ± 0.7 51 ± 13 54

    Bacterial cell surfaces and the glass substratum surface are predominantly

    electron-donating, as evidenced by their larger γminus surface free energy

    parameter as compared with γplus. Hydrophobic DDS-coated glass is neither a

    good electron donor nor acceptor. All surfaces are negatively charged and

    whereas bacterial zeta potentials vary between -22 and -50 mV, the zeta

    potentials of glass and DDS-coated glass are similarly negative (-33 to -35 mV).

    The bacterial cell and substratum surface properties listed in Table 4 have

    been used in the extended DLVO theory, yielding interaction free energy- and

    force-distance profiles for all combinations of bacteria and substratum surfaces,

    as illustrated in Figure 4 for P. aeruginosa KEI 1025. Note the reversed force-

  • Forces involved in bacterial adhesion

    29

    axis (right hand side) in Figure 4 indicating that negative values correspond to

    attractive interaction forces according to the definition of force:

    ( ) ( )dEd

    dFδδ

    −= (15)

    Residing in the secondary minimum of the interaction energy corresponds to

    zero interaction force, resulting from compensating attractive (Van der Waals)

    and repulsive (electrostatic) forces. However, the approach toward the

    secondary minimum yields a maximum net attraction force (Figure 4) at a

    distance of about 40 nm from the surface. On glass, these interaction forces are

    generally higher than on DDS-coated glass (see also Table 2), due to larger

    Hamaker constants for glass as a substratum. Additionally, on DDS-coated glass

    a primary minimum (closer to the surface) is predicted due to acid-base

    interaction. The height of the energy barrier between the secondary and primary

    minimum varies from 229 kT for S. thermophilus to 1030 kT for R. terrigena and

    therefore it is very unlikely that a depositing micro-organism will cross the

    barrier to adhere in the primary minimum. On glass, a primary interaction

    minimum is absent.

    Discussion

    The forces that govern microbial deposition, adhesion and detachment are still

    not fully understood, and difficult to relate with each other. In a previous study

    we successfully investigated the characteristic shear force to prevent adhesion of

    microbial strains [42]. In the current research we used a more systematic

    approach by including not only the shear forces to prevent adhesion, but also

    those that stimulate detachment of adhering bacteria, as well as theoretical

    adhesion forces calculated using the extended DLVO theory. In addition, the

    effect of a passing air-liquid interface, which invokes a high, perpendicularly

  • Cha

    pter

    2

    30

    Tab

    le 4

    . Phy

    sico

    -che

    mic

    al c

    hara

    cter

    istic

    s of

    the

    bact

    eria

    l stra

    ins

    and

    colle

    ctor

    sur

    face

    s us

    ed. B

    acte

    rial c

    hara

    cter

    izat

    ions

    wer

    e ba

    sed

    on th

    ree

    sepa

    rate

    ly g

    row

    n cu

    lture

    s. Pe

    r cul

    ture

    , con

    tact

    ang

    les

    of w

    ater

    ( Θw),

    form

    amid

    e (Θ

    form

    ), α

    -bro

    mon

    apht

    hale

    ne (Θ

    br) a

    nd m

    ethy

    lene

    iodi

    de (Θ

    met)

    wer

    e m

    easu

    red

    on fo

    ur b

    acte

    rial l

    awns

    usi

    ng o

    ne d

    ropl

    et p

    er li

    quid

    per

    bac

    teria

    l law

    n. Z

    eta

    pote

    ntia

    ls (ζ

    ) wer

    e de

    term

    ined

    in tr

    iplic

    ate.

    Con

    tact

    an

    gle

    and

    stre

    amin

    g po

    tent

    ial

    mea

    sure

    men

    ts o

    n su

    bstra

    tum

    sur

    face

    s w

    ere

    perf

    orm

    ed i

    n qu

    adru

    plic

    ate.

    Fre

    e su

    rfac

    e en

    ergy

    com

    pone

    nts

    are

    deriv

    ed fr

    om c

    onta

    ct a

    ngle

    mea

    sure

    men

    ts g

    ivin

    g an

    ele

    ctro

    n-do

    natin

    g ( γ

    min

    us) a

    nd -a

    ccep

    ting

    (γpl

    us) p

    aram

    eter

    for t

    he a

    cid-

    base

    com

    pone

    nt (γ

    AB),

    the

    Lifs

    hitz

    -Van

    der

    Waa

    ls c

    ompo

    nent

    (γLW

    ) and

    the

    tota

    l sur

    face

    free

    ene

    rgy

    (γTo

    t ).

    Bac

    teri

    al st

    rain

    Θ

    w

    (º)

    Θfo

    rm

    (º)

    Θbr

    ) Θ

    met

    (º)

    γmin

    us

    (mJ

    m-2

    ) γp

    lus

    (mJ

    m-2

    ) γA

    B

    (mJ

    m-2

    ) γL

    W

    (mJ

    m-2

    ) γT

    ot

    (mJ

    m-2

    ) ζ

    (m

    V)

    S. e

    pide

    rmid

    is

    HB

    H2 1

    69

    31 ±

    4

    31 ±

    4

    34 ±

    5

    50 ±

    3

    47.8

    0.

    4 9

    40

    49

    -50

    ± 6

    S. e

    pide

    rmid

    is

    ATC

    C 3

    5983

    38

    ± 5

    40

    ± 5

    36

    ± 1

    54

    ± 4

    45

    .8

    0.5

    10

    34

    44

    -51

    ± 2

    R. te

    rrig

    ena

    ATC

    C 3

    3527

    24

    ± 3

    24

    ± 3

    40

    ± 4

    51

    ± 4

    49

    .9

    1.8

    19

    34

    53

    -49

    ± 5

    S. th

    erm

    ophi

    lus

    ATC

    C 1

    9258

    35

    ± 2

    31

    ± 4

    58

    ± 2

    77

    ± 2

    41

    .2

    5.3

    30

    22

    52

    -22

    ± 5

    P. a

    erug

    inos

    a

    D1

    44 ±

    6

    42 ±

    4

    48 ±

    8

    58 ±

    6

    38.8

    1.

    2 14

    30

    44

    -3

    0 ±

    3

    P. a

    erug

    inos

    a

    KEI

    102

    5 25

    ± 2

    31

    ± 2

    40

    ± 2

    49

    ± 4

    54

    .8

    0.8

    14

    35

    49

    -39

    ± 5

    Subs

    trat

    um su

    rfac

    e

    Gla

    ss

    28

    ± 8

    25

    ± 3

    51

    ± 2

    64

    ± 1

    45

    .8

    3.7

    26

    28

    54

    -35

    ± 5

    DD

    S-co

    atin

    g

    101

    ± 2

    85 ±

    3

    59 ±

    4

    65 ±

    4

    2.2

    0.0

    0 26

    26

    -3

    3 ±

    2

    Chapter 2

    30

  • Forces involved in bacterial adhesion

    31

    Figure 4. Example of the extended DLVO interaction energy (⎯) and –force (⋅⋅⋅⋅) as a function of distance for P. aeruginosa KEI 1025 on glass and DDS-coated glass. Arrows indicate the correct axis for both plots. Note the reversed force-axis.

  • Chapter 2

    32

    oriented detachment force on adhering bacteria, was determined. Furthermore,

    all experiments were carried out with six different bacterial strains in order to

    allow general conclusions to be drawn. As a first step in the experimental

    analysis, the gravitational force and its impact on bacterial deposition [43,44]

    and adhesion was eliminated by averaging the deposition rates on bottom-and

    top plate. At low shear rates, deposition efficiencies (α) exceed unity especially

    for the S. epidermidis strains, indicating that deposition is more favourable than

    theoretically predicted. Often such deviations are ascribed to the presence of

    surface structures [45]. With respect to possible relations between the different

    forces distinguished, we test the following hypotheses:

    1) A strong hydrodynamic shear force to prevent adhesion relates to a strong

    hydrodynamic shear force to detach an adhering organism. This hypothesis

    implies a positive correlation between attachment and detachment. Comparison

    between Fprev and Fdet (Table 2) show that regardless of the substratum involved,

    Fdet is always larger than Fprev. In the experimental set-up used, bacteria are

    adhering to the substratum surface for at least half an hour before being subject

    to high shear. Therewith, over time the bond between a bacterium and the

    substratum surface may become stronger. Supporting evidence for this is

    provided by others who have used AFM and found that the adhesion force

    increases with prolonged contact time [46,47]. Thus, even though initial

    adhesion forces are rather weak, they may be indicative for forces after a

    prolonged time, i.e. a relatively strong Fprev might be expected to correspond

    with a relatively strong Fdet. However, from Figure 5 it is clear that no

    correlation exists between Fprev and Fdet. It implies that attachment and

    detachment should be regarded as independent processes and the hypothesis of

    an unambiguous relation between attachment and detachment forces should be

    rejected.

  • Forces involved in bacterial adhesion

    33

    Figure 5. Graphical presentation of possible relations between Fprev and Fdet (A), FDLVO and Fprev (B), FDLVO and Fdet (C) and detachment percentage and Fmax (D).

    2) A weak hydrodynamic shear force to detach adhering bacteria implies that

    more bacteria will be stimulated to detach by a passing air-liquid interface

    through the flow chamber. Table 2 clearly indicates that Fdet for hydrophobic

    DDS-coated glass is larger than for hydrophilic glass, indicating stronger

    interaction forces on the hydrophobic substratum. Table 3 summarizes

    parameters involved in air-bubble-induced detachment. An air-liquid interface

    exerts forces 104 times larger than Fdet, yet it does not result in complete

    detachment. Combining the data in Tables 2 and 3, reveals the absence of a clear

    relation between shear-induced detachment and detachment by passing an air-

    bubble. Thus a weaker Fdet does not result in higher air-bubble-stimulated

    detachment and this hypothesis has to be rejected too. In this respect it must be

    realized that different mechanisms of detachment are involved in both processes.

  • Chapter 2

    34

    Hydrodynamic detachment forces are measured while the system is completely

    submerged in liquid whereas an extra phase is introduced in air-bubble-induced

    detachment. Furthermore, Fdet is a force acting parallel to the substratum surface,

    whereas the air-liquid interface acts perpendicularly to the substratum surface.

    3) DLVO interactions determine the characteristic hydrodynamic shear forces

    to prevent adhesion and to detach adhering micro-organisms as well as the

    detachment induced by a passing air-liquid interface. Further analysis revealed

    the absence of quantitative relations between FDLVO and Fprev, as well as between

    FDLVO and Fdet (Figure 5). DLVO-predictions have often been demonstrated to

    deviate from experimental observations of bacterial interaction phenomena,

    which is usually ascribed to the presence of surface appendages [48,49] or

    chemical surface heterogeneities. However, the direction of action of the

    DLVO-forces should be taken into account as well. DLVO-forces act

    perpendicularly to the substratum surface, whereas both Fprev and Fdet are

    directed parallel to the substratum surface.

    When the fluid flow is increased to high enough values, the bacterium

    most likely detaches in a rolling fashion [50]. It can be argued that in this mode

    of detachment, forces normal to the surface (i.e. DLVO- and lift forces) are

    related to forces directed parallel to the surface. However, in similar detachment

    studies it was found that lift forces are negligible and surface roughness may

    play a decisive role in determining the hydrodynamic force to remove adhering

    particles from the surface [51,52]. This feature is not accounted for in the

    DLVO-theory. Table 2 shows only slight differences between the theoretical

    FDLVO-values for the various microbial strains, but substantial differences

    between the experimentally obtained forces Fprev and Fdet. If a correlation

    between DLVO forces and shear forces would exist, an increase of these parallel

    directed forces implies an increase in normally directed forces. However, this is

  • Forces involved in bacterial adhesion

    35

    not observed in FDLVO. Hence, the parallel detachment forces do not correlate to

    the perpendicularly directed DLVO-forces.

    The DLVO theory predicts a secondary minimum of interaction at a

    distance of about 30 to 40 nm away from the surface (see Figure 4). On

    hydrophilic glass, closer approach is impossible due to strong repulsion and

    adhesion can only occur in the secondary minimum. On DDS-coated glass, also

    primary minimum interactions are predicted. However, due to the prohibitive

    high barrier of the free energy (ranging from 229 kT to 1030 kT depending on

    the strain used), it is very unlikely that adhesion in the primary minimum can

    occur. Therefore, also on the hydrophobic DDS-coated glass, only adhesion in

    the secondary minimum is expected to occur. As can be seen in Table 2, Fdet-

    values are much higher than FDLVO. Often, a transition of adhesion from the

    secondary interaction minimum towards the primary minimum is used as

    explanation [53]. However, in this study this is considered to be impossible as

    on glass a primary minimum is absent and on DDS-coated glass it is considered

    to be unreachable due to the high energy barrier. It is therefore more likely that

    the higher Fdet values are the result of attachment of surface appendages, or

    “extracellular polymeric substances” produced, capable to reach the surface.

    These structures are known to extend as much as hundreds of nm away from the

    bacterial cell wall [38], which is more than enough to bridge the distance

    between secondary minimum and the substratum surface. Unfortunately,

    although it is known for instance that some streptococci may possess surface

    fibrils, structural information about the cell surface of far most all strains studied

    in the literature are lacking, let alone detailed knowledge about the length,

    diameter and micro(nano-)scopic physico-chemical properties of these

    structures. The use of the DLVO-theory as currently done in the literature as

    well as in this chapter, can therefore only pertain to long-distance approach,

    where fine surface structures do not play a role. Up to what distance of approach

    and up to what extent this statement is valid, is hard to say. However, while the

  • Chapter 2

    36

    DLVO-theory predicts interactions for the entire micro-organism, it is likely that

    the experimentally obtained detachment forces are related to a number of

    distinct (hydrogen) bonds. When these linkages break, due to parallel directed

    forces, the bacterium can be transported away from the surface due to lift forces

    which are induced by the tangential flow [54]. In this respect, parallel directed

    hydrodynamic forces (i.e. Fprev and Fdet) can serve as useful parameters to

    indicate adhesion strength.

    When combining the detachment parameters (i.e. Fdet and the air-bubble

    detachment percentage), our results suggest that bridging between the bacterium

    and the substrate surface is more favourable for DDS-coated glass. Fdet on

    hydrophobic DDS-coated glass is always higher than on glass and even though

    one has to be cautious in interpreting air-bubble detachment percentages, the

    higher detachment force exerted by the air-bubble on DDS-coated glass does not

    necessarily lead to more detachment. The hydrophobicity of the surface likely

    enhances the possibility of bridging as removal of water from in between the

    interacting surfaces is more favourable. This matter is further complicated by the

    influence of the type of medium in which adhesion occurs. The DLVO-theory is

    based on averaged properties of the surfaces of the bacterial cell and substratum.

    However, it was found that ions in the suspending medium, especially divalent

    ions, can greatly influence the adhesion of bacteria to a surface, probably due to

    surface charge heterogeneities resulting from complexation of different ions

    with the (bacterial cell) surface(s) [55]. Since our experiments were performed

    in potassium phosphate buffer, we cannot rule out similar effects caused by the

    divalent phosphate anion.

    Even though no quantitative correlation between the DLVO theory and

    detachment behaviour could be established, and the above forwarded hypothesis

    should therefore be rejected, this theory does help to provide a better insight in

    the mechanism of bacterial adhesion to a substratum surface.

  • Forces involved in bacterial adhesion

    37

    Conclusions

    The hydrodynamic force to prevent adhesion (Fprev) is lower than the

    hydrodynamic force to stimulate detachment (Fdet) showing that the bond

    between a substratum surface and a bacterium becomes stronger after initial

    adhesion. Consequently, Fprev and Fdet should be considered as independent

    parameters.

    There is no unambiguous relation between the hydrodynamic forces (Fprev

    and Fdet) directed parallel to the substratum surface and perpendicularly oriented

    parameters (FDLVO, air-liquid interface detachment), because these forces act in

    different directions. DLVO forces maybe wrongfully estimated because of local

    charge heterogeneities and bridging between cell appendages and/or exudates on

    the one hand and substrate surface on the other. Furthermore, air-liquid interface

    induced detachment relies on a three-phase system, whereas the other forces are

    obtained for a two-phase environment, complicating establishment of a possible

    correlation.

  • Chapter 2

    38

    References

    1. Costerton, J.W., Stewart, P.S. and Greenberg, E.P. (1999), Bacterial biofilms: A

    common cause of persistent infections, Science 284, 1318 - 1322.

    2. Flemming, H.C. (2002), Biofouling in water systems - Cases, causes and

    countermeasures, Appl Microbiol Biot 59, 629 - 640.

    3. Von Eiff, C., Jansen, B., Kohnen, W. and Becker, K. (2005), Infections associated

    with medical devices - Pathogenesis, management and prophylaxis, Drugs 65,

    179 - 214.

    4. Van Oss, C.J., Good, R.J. and Chaudhury, M. (1986), The role of Van der Waals

    forces and hydrogen bonds in hydrophobic interactions between biopolymers and low

    energy surfaces, J Colloid Interf Sci 111, 378 - 390.

    5. Bakker, D.P., Busscher, H.J. and Van der Mei, H.C. (2002), Bacterial deposition in a

    parallel plate and a stagnation point flow chamber: Microbial adhesion mechanisms

    depend on the mass transport conditions, Microbiol-Sgm 148, 597 - 603.

    6. Thomas, W.E., Nilsson, L.M., Forero, M., Sokurenko, E.V. and Vogel, V. (2004),

    Shear-dependent 'stick-and-roll' adhesion of type 1 fimbriated Escherichia coli,

    Molecular Microbiology 53, 1545 - 1557.

    7. Higashi, J.M., Wang, I.W., Shlaes, D.M., Anderson, J.M. and Marchant, R.E.

    (1998), Adhesion of Staphylococcus epidermidis and transposon mutant strains to

    hydrophobic polyethylene, J Biomed Mater Res 39, 341 - 350.

    8. Shive, M.S., Hasan, S.M. and Anderson, J.M. (1999), Shear stress effects on bacterial

    adhesion, leukocyte adhesion, and leukocyte oxidative capacity on a polyetherurethane,

    J Biomed Mater Res 46, 511 - 519.

    9. Wang, I.W., Anderson, J.M., Jacobs, M.R. and Marchant, R.E. (1995), Adhesion of

    Staphylococcus epidermidis to0 biomedical polymers - Contributions of surface

    thermodynamics and hemodynamic shear conditions, J Biomed Mater Res 29, 485 - 493.

  • Forces involved in bacterial adhesion

    39

    10. Mohamed, N., Rainier, T.R. and Ross, J.M. (2000), Novel experimental study of

    receptor-mediated bacterial adhesion under the influence of fluid shear, Biotechnol

    Bioeng 68, 628 - 636.

    11. Duddridge, J.E., Kent, C.A. and Laws, J.F. (1982), Effect of surface shear-stress on

    the attachment of Pseudomonas fluorescens to stainless-steel under defined flow

    conditions, Biotechnol Bioeng 24, 153 - 164.

    12. Rutter, P.R. and Vincent, B. (1988), Attachment mechanisms in the surface growth of

    microorganisms. In: Physiological models in microbiology. Bazin, M.J. and Prosser, J.I.

    (Eds.), Boca Raton, Florida:CRC Press, Inc. pp 87 - 107.

    13. Owens, N.F., Gingell, D. and Rutter, P.R. (1987), Inhibition of cell-adhesion by a

    synthetic-polymer adsorbed to glass shown under defined hydrodynamic stress, J Cell

    Sci 87, 667 - 675.

    14. Katsikogianni, M. and Missirlis, Y.F. (2004), Concise review of mechanisms of

    bacterial adhesion to biomaterials and of techniques used in estimating bacteria-material

    interactions, Eur Cell Mater 8, 37 - 57.

    15. Mohamed, N., Teeters, M.A., Patti, J.M., Hook, M. and Ross, J.M. (1999),

    Inhibition of Staphylococcus aureus adherence to collagen under dynamic conditions,

    Infect Immun 67, 589 - 594.

    16. Walker, S.L., Redman, J.A. and Elimelech, M. (2004), Role of cell surface

    lipopolysaccharides in Escherichia coli K12 adhesion and transport, Langmuir 20,

    7736 - 7746.

    17. Meinders, J.M., Van der Mei, H.C. and Busscher, H.J. (1995), Deposition efficiency

    and reversibility of bacterial adhesion under flow, J Colloid Interf Sci 176, 329 - 341.

    18. Busalmen, J.P. and de Sanchez, S.R. (2001), Adhesion of Pseudomonas fluorescens

    (ATCC 17552) to nonpolarized and polarized thin films of gold, Appl Environ Microb

    67, 3188 - 3194.

  • Chapter 2

    40

    19. Jacobs, A., Lafolie, F., Herry, J.M. and Debroux, M. (2007), Kinetic adhesion of

    bacterial cells to sand: Cell surface properties and adhesion rate, Colloid Surface B 59,

    35 - 45.

    20. Vijayalakshmi, S.P. and Raichur, A.M. (2003), The utility of Bacillus subtilis as a

    bioflocculant for fine coal, Colloid Surface B 29, 265 - 275.

    21. Sharma, P.K. and Rao, K.H. (2003), Adhesion of Paenibacillus polymyxa on

    chalcopyrite and pyrite: Surface thermodynamics and extended dlvo theory, Colloid

    Surface B 29, 21 - 38.

    22. Azeredo, J., Visser, J. and Oliveira, R. (1999), Exopolymers in bacterial adhesion:

    Interpretation in terms of DLVO and XDLVO theories, Colloid Surface B 14, 141 - 148.

    23. Cao, T., Tang, H.Y., Liang, X.M., Wang, A.F., Auner, G.W., Salley, S.O. and Ng,

    K.Y.S. (2006), Nanoscale investigation on adhesion of E. coli surface modified silicone

    using atomic force microscopy, Biotechnol Bioeng 94, 167 - 176.

    24. Abu-Lail, N.I. and Camesano, T.A. (2003), Role of lipopolysaccharides in the

    adhesion, retention, and transport of Escherichia coli JM109, Environ Sci Technol 37,

    2173 - 2183.

    25. Mendez-Vilas, A., Gallardo-Moreno, A.M. and Gonzalez-Martin, M.L. (2006),

    Nano-mechanical exploration of the surface and sub-surface of hydrated cells of

    Staphylococcus epidermidis, Anton Leeuw Int J Gen Mol Microbiol 89, 373 - 386.

    26. Bowen, W.R., Fenton, A.S., Lovitt, R.W. and Wright, C.J. (2002), The measurement

    of Bacillus mycoides spore adhesion using atomic force microscopy, simple counting

    methods, and a spinning disk technique, Biotechnol Bioeng 79, 170 - 179.

    27. Fallman, E., Schedin, S., Jass, J., Andersson, M., Uhlin, B.E. and Axner, O. (2004),

    Optical tweezers based force measurement system for quantitating binding interactions:

    System design and application for the study of bacterial adhesion, Biosens Bioelectron

    19, 1429 - 1437.

  • Forces involved in bacterial adhesion

    41

    28. Simpson, K.H., Bowden, M.G., Hook, M. and Anvari, B. (2002), Measurement of

    adhesive forces between S. epidermidis and fibronectin-coated surfaces using optical

    tweezers, Laser Surg Med 31, 45 - 52.

    29. Simpson, K.H., Bowden, A.G., Peacock, S.J., Arya, M., Hook, M. and Anvari, B.

    (2004), Adherence of Staphylococcus aureus fibronectin binding protein a mutants: An

    investigation using optical tweezers, Biomol Eng 21, 105 - 111.

    30. Gomez-Suarez, C., Busscher, H.J. and Van der Mei, H.C. (2001), analysis of

    bacterial detachment from substratum surfaces by the passage of air-liquid interfaces,

    Appl Environ Microb 67, 2531 - 2537.

    31. Bakker, D.P., Postmus, B.R., Busscher, H.J. and Van der Mei, H.C. (2004),

    Bacterial strains isolated from different niches can exhibit different patterns of adhesion

    to substrata, Appl Environ Microb 70, 3758 - 3760.

    32. Busscher, H.J. and Van der Mei, H.C. (2006), Microbial adhesion in flow

    displacement systems, Clin Microbiol Rev 19, 127 - 141.

    33. Elimelech, M. (1994), Particle deposition on ideal collectors from dilute flowing

    suspensions - Mathematical formulation, numerical-solution, and simulations, Separ

    Technol 4, 186 - 212.

    34. Van Holde, K.E. (1971), Introduction in transport processes:diffusion. In: Physical

    biochemistry. Van Holde, K.E. (Eds.), Englewood Cliffs, New Jersey:Prentice-Hall, Inc.

    pp 79 - 111.

    35. Leenaars, A.F.M. and O'Brien, S.B.G. (1989), Particle removal from silicon

    substrates using surface-tension forces, Philips J Research 44, 183 - 209.

    36. Van Wagenen, R.A. and Andrade, J.D. (1980), Flat-plate streaming potential

    investigations - Hydrodynamics and electrokinetic equivalency, J Colloid Interf Sci 76,

    305 - 314.

    37. Lyklema, J. (1991), Electrochemistry and its application to colloids and interfaces. In:

    Fundamentals of interface and colloid science. London:Academic Press Ltd. pp

  • Chapter 2

    42

    38. Bos, R., Van der Mei, H.C. and Busscher, H.J. (1999), Physico-chemistry of initial

    microbial adhesive interactions - Its mechanisms and methods for study, FEMS

    Microbiol Rev 23, 179 - 230.

    39. Van Oss, C.J. (1994), Rate of decay with distance. In: Interfacial forces in aqueous

    media. Van Oss, C.J. (Eds.), New York:Marcel Dekker. pp 75 - 88.

    40. Van Oss, C.J. (1994), Relation between the Hamaker constant and the apolar surface

    tension component. In: Interfacial forces in aqueous media. Van Oss, C.J. (Eds.), New

    York:Marcel Dekker. pp 154 - 160.

    41. Norde, W. and Lyklema, J. (1989), Protein adsorption and bacterial adhesion to solid-

    surfaces - A colloid-chemical approach, Colloid and Surface 38, 1 - 13.

    42. Roosjen, A., Boks, N.P., Van der Mei, H.C., Busscher, H.J. and Norde, W. (2005),

    Influence of shear on microbial adhesion to PEO-brushes and glass by convective-

    diffusion and sedimentation in a parallel plate flow chamber, Colloid Surface B 46,

    1 - 6.

    43. Walt, D.R., Smulow, J.B., Turesky, S.S. and Hill, R.G. (1985), The effect of gravity

    on initial microbial adhesion, J Colloid Interf Sci 107, 334 - 336.

    44. Agladze, K., Wang, X. and Romeo, T. (2005), Spatial periodicity of Escherichia coli

    K12 biofilm microstructure initiates during a reversible, polar attachment phase of

    development and requires the polysaccharide adhesin pga, J Bacteriol 187, 8237 - 8246.

    45. Triandafillu, K., Balazs, D.J., Aronsson, B.O., Descouts, P., Quoc, P.T., Van

    Delden, C., Mathieu, H.J. and Harms, H. (2003), Adhesion of Pseudomonas

    aeruginosa strains to untreated and oxygen-plasma treated poly(vinyl chloride) (PVC)

    from endotracheal intubation devices, Biomaterials 24, 1507 - 1518.

    46. Vadillo-Rodriguez, V., Busscher, H.J., Norde, W., De Vries, J. and Van der Mei,

    H.C. (2004), Atomic force microscopic corroboration of bond aging for adhesion of

    Streptococcus thermophilus to solid substrata, J Colloid Interf Sci 278, 251 - 254.

  • Forces involved in bacterial adhesion

    43

    47. Xu, L.C., Vadillo-Rodriguez, V. and Logan, B.E. (2005), Residence time, loading

    force, pH, and ionic strength affect adhesion forces between colloids and biopolymer-

    coated surfaces, Langmuir 21, 7491 - 7500.

    48. Jucker, B.A., Zehnder, A.J.B. and Harms, H. (1998), Quantification of polymer

    interactions in bacterial adhesion, Environ Sci Technol 32, 2909 - 2915.

    49. Ong, Y.L., Razatos, A., Georgiou, G. and Sharma, M.M. (1999), Adhesion forces

    between E.coli bacteria and biomaterial surfaces, Langmuir 15, 2719 - 2725.

    50. Das, S.K., Schechter, R.S. and Sharma, M.M. (1994), The role of surface-roughness

    and contact deformation on the hydrodynamic detachment of particles from surfaces, J

    Colloid Interf Sci 164, 63 - 77.

    51. Batra, A., Paria, S., Manohar, C. and Khilar, K.C. (2001), Removal of surface

    adhered particles by surfactants and fluid motions, Aiche J 47, 2557 - 2565.

    52. Yiantsios, S.G. and Karabelas, A.J. (1995), Detachment of spherical microparticles

    adhering on flat surfaces by hydrodynamic forces, J Colloid Interf Sci 176, 74 - 85.

    53. Van Loosdrecht, M.C.M. and Zehnder, A.J.B. (1990), Energetics of bacterial

    adhesion, Experientia 46, 817 - 822.

    54. Cantat, I. and Misbah, C. (1999), Lift force and dynamical unbinding of adhering

    vesicles under shear flow, Phys Rev Lett 83, 880 - 883.

    55. De Kerchove, A.J. and Elimelech, M. (2008), Calcium and magnesium cations

    enhance the adhesion of motile and nonmotile Pseudomonas aeruginosa on alginate

    films, Langmuir 24, 3392 - 3399.

  • Chapter 2

    44

  • CHAPTER 3

    RESIDENCE TIME DEPENDENT DESORPTION OF

    STAPHYLOCOCCUS EPIDERMIDIS FROM HYDROPHILIC AND

    HYDROPHOBIC SUBSTRATA

    Reproduced with permission of Elsevier b.v. from: Boks, N.P., Kaper, H.J., Norde, W.,

    Busscher, H.J. and Van der Mei, H.C. (2008), Colloids and Surfaces B: Biointerfaces 67,

    276-278.

  • Chapter 3

    46

    Abstract

    Adhesion and desorption are simultaneous events during bacterial adhesion to

    surfaces, although desorption is far less studied than adhesion. Here, desorption

    of Staphylococcus epidermidis from substratum surfaces is demonstrated to be

    residence time dependent. Initial desorption rate coefficients were similar for

    hydrophilic and hydrophobic dimethyldichlorosilane (DDS) -coated glass, likely

    because initial desorption is controlled by attractive Lifshitz-Van der Waals

    interactions, which are comparable on both substratum surfaces. However,

    significantly slower decay times of the desorption rate coefficients are found for

    hydrophilic glass than for hydrophobic DDS-coated glass. This difference is

    suggested to be due to the acid-base interactions between staphylococci and

    these surfaces, which are repulsive on hydrophilic glass and attractive on

    hydrophobic DDS-coated glass. Final desorption rate coefficients are higher on

    hydrophilic glass than on hydrophobic DDS-coated glass, due to the so called

    hydrophobic effect, facilitating a closer contact on hydrophobic DDS-coated

    glass.

  • Residence time dependent desorption

    47

    Introduction

    Microbial adhesion to substratum surfaces is generally considered to consist of

    two steps [1,2]. In the first step, adhesion is reversible and detachment may

    occur spontaneously. Gradually, as a second step, adhesive forces increase to

    cause more irreversible adhesion [3-6] with a lower desorption probability [7].

    Although residence time dependent desorption has been investigated previously,

    these studies were subject to technical limitations [8]. Analyses were based on a

    series of images taken with 12 s time interval between consecutive pictures.

    Nowadays, consecutive images can be taken with a time interval of 1 s, or even

    shorter if desired, enabling more accurate registration and analysis of adhesion

    and desorption events in microbial adhesion.

    Desorption of bacteria from surfaces can be described by a so called

    residence time dependent desorption rate coefficient (β(t-τ)). After adhesion,

    immediate s