02 colsua powell

Upload: nevena-celic

Post on 06-Apr-2018

219 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/3/2019 02 Colsua Powell

    1/11

    Colloids and Surfaces

    A: Physicochemical and Engineering Aspects 206 (2002) 241251

    Wetting of nanoparticles and nanoparticle arrays

    C. Powell, N. Fenwick, F. Bresme, N. Quirke *

    Department of Chemistry, Imperial College of Science, Technology and Medicine, South Kensington SW7 2AY, UK

    Abstract

    We report investigations of nanoparticulate wetting carried out using molecular dynamics simulations. Despite theirsmall size, model Lennard Jones nanoparticles in simple Lennard Jones solvents exhibit well defined contact

    angles, which for high surface tension interfaces have been shown to obey Youngs equation with surprising accuracy.

    In this paper we present new results for fully molecular models of nanoparticles at a water surface, where again

    well-defined contact angles are evident. Pressure area curves obtained in Langmuir trough experiments on (nano)

    particulates have been used in the past to determine contact angles. We have recently demonstrated by molecular

    dynamics and theory that, contrary to expectations, the collapse pressure measured in this experiment should be

    independent of contact angle and that the initial collapse mode is by surface buckling. We present new molecular

    dynamics results for arrays of nanoparticles with a contact angle of 72 at a liquidvapour interface which confirm

    our earlier work and offer more information on the details of the structure of collapsed nanoparticle arrays. 2002

    Elsevier Science B.V. All rights reserved.

    Keywords: Surface tension; Nanoparticles; Wetting

    www.elsevier.com/locate/colsurfa

    1. Introduction

    Three phases in contact exhibit a range of

    interesting behaviour. For example in the case of

    a solid in contact with liquid and vapour it is

    possible to distinguish two cases: one where the

    liquid forms a uniform (macroscopic) film cover-

    ing the solid surface, the complete wetting state

    and a second where the liquid forms a droplet onthe surface, partially covering the surface, the

    partial wetting state. For a solid particle in con-

    tact with two bulk fluid phases, the completely

    wet (or dry) phase corresponds to the particle

    being completely immersed in one phase whilstthe partially wet state has the particle sitting inthe interface between the two fluid phases. Suchstates and the transitions between wetting states

    are of both scientific and industrial interest, forexample in determining the order of wetting tran-sitions or in tailoring wetting properties for highspeed printing or emulsion stabilisation. Currentlythere is very strong interest in the description ofwetting of nanoscale materials and of nanostruc-tured/nanopatterned materials. Examples includethe behaviour of nanoparticles at fluid interfacesand its connection to self-assembly [1], the wettingof micron-sized channels [2] or nanopores for theformation of nanowires [3], and the wetting ofnanoscale patterned interfaces with application tonanolithography.

    * Corresponding author.

    E-mail address: [email protected] (N. Quirke).

    0927-7757/02/$ - see front matter 2002 Elsevier Science B.V. All rights reserved.

    PII: S 0 9 2 7 - 7 7 5 7 ( 0 2 ) 0 0 0 7 9 - 1

    mailto:[email protected]:[email protected]
  • 8/3/2019 02 Colsua Powell

    2/11

    C. Powell et al. /Colloids and Surfaces A: Physicochem. Eng. Aspects 206 (2002) 241 251242

    In this paper we report investigations of

    nanoparticulate wetting carried out using molecu-

    lar dynamics simulations. In previous work [46]

    we have shown that despite their small size, model

    LennardJones nanoparticles in simple Len-

    nardJones solvents (at both liquidvapour, LV

    and liquid liquid interfaces) exhibit well defined

    contact angles, which for high surface tensioninterfaces (in particular the liquid liquid case)

    have been shown to obey Youngs equation with

    surprising accuracy.

    Our initial results [46] were for a LV interface

    with a surface tension of 3 mN m1 (using LJs

    units for argon), which is low compared with the

    water/air surface tension, 72 mN m1 but larger

    than say a microemulsion oil interface 0.1

    mN m1. By measuring the free energies of the

    simulated fluid and solid interfaces we were able

    to show that Youngs equation is accurate for thisinterface for nanoparticles of diameter E3 nm.

    For smaller nanoparticles Youngs equation was

    less accurate and the simulations showed a wet-

    ting transition at 1 nm not predicted by

    Youngs equation. Youngs equation disregards

    one of the four interfaces in the system: the fourth

    interface being the line that separates the three

    phases. The free energy associated with this line,

    the line tension (~), may influence the contact

    angle. We estimated [46] ~ from our simulations

    and found it to be of the order of 1012 N.

    The effect of the line tension on the wetting

    properties of the nanoparticle, such as the contact

    angle, may be considered through a corrected

    Youngs equation, that reduces to the Youngs

    equation when ~=0 or 1/R0 i.e. the curvature

    of the three phase line is negligible. Our data

    [46] show that the corrected Youngs equation is

    more accurate, but still breaks down for the

    smallest nanoparticles (1.5 nm).

    The situation may be different however for

    interface with a high surface tension. To investi-

    gate this question we simulated [4 6] nanoparti-

    cles at a liquid liquid interface for which the two

    liquid phases are immiscible, analogous to the

    water/oil interface. In this case the surface tension

    of the liquid liquid interface was 14 mN m1

    (compared to 3 mN m1 for the LV case de-

    scribed above). We analysed the accuracy of

    Youngs equation in this case by varying one of

    the surface tensions, kpf, keeping the other two

    constant. Youngs equation predicts the contact

    angle is a linear function of the surface tension.

    We found Youngs equation to be very accurate

    in predicting the wetting behaviour of the

    nanoparticle at this interface (k=14 mN m1)

    even for diameters approaching 1 nm. This con-trasts with the results obtained for the LV inter-

    face. It might be that the line tension in these

    systems is negligible compared with the LV case.

    However our calculations [4 6] show that this is

    not so, indeed it is larger by one order of magni-

    tude at 1011 N. This can be understood by

    looking at the corrected Youngs equation, which

    predicts that the contact angle is a strong function

    of the line tension only when the surface tension

    of the fluidfluid interface is low. When the sur-

    face tension is high the contact angle is insensitiveto line tension. From our studies [46] of model

    LJs nanoparticles at LJs interfaces we conclude

    that it is possible to describe the wetting of

    nanoscale interfaces (down to 1 nm) using

    Youngs equation when (a) the fluidfluid interfa-

    cial tensions are large and (b) the three phase line

    has a small curvature.

    These general results have been obtained using

    idealised structureless model nanoparticulates. In

    nature however nanoparticles have internal and

    surface structures, are not necessarily rigid and

    the solvents may themselves be complex. For ex-

    ample metal nanoparticles have received consider-

    able attention during recent years due to the wide

    range of potential technological applications.

    These range from the use of dextran coated iron

    clusters [8] in target specific magnetic resonance

    imaging [9], to the use of lattices of organothiol

    stabilised gold clusters as chemical vapour sensors

    [10] and lithographic masks [11]. Particular inter-

    est has been shown in colloidal gold as a catalyst,

    especially since supported gold nanoparticles have

    been used to catalyse the oxidation of carbon

    monoxide [12]. In this paper we extend our work

    by describing a realistic molecular model of a

    thiol passivated gold nanoparticle at a water in-

    terface and reporting preliminary molecular dy-

    namics simulations of its wetting properties. In

    the final section we move from individual particles

  • 8/3/2019 02 Colsua Powell

    3/11

    C. Powell et al. /Colloids and Surfaces A: Physicochem. Eng. Aspects 206 (2002) 241 251 243

    to nanoparticle arrays and consider the pressure

    area curves obtained in Langmuir trough (LT)

    experiments on such arrays. Since its original

    conception by Pockels [13] in the late 19th century

    and subsequent refinement by Langmuir [14] in

    the 1920s the LT technique has become an impor-

    tant experimental tool in fields, ranging from

    biology to electronics. Surface pressure (P)area(A) isotherms obtained from LT experiments can

    provide important data on the geometry and in-

    termolecular forces [15 18] of a wide range of

    materials. This includes the characterisation of

    nanoparticles used in processes relevant to oil

    recovery, flotation, anti-foaming, gas sensors,

    biosensors pyro, piezo and ferroelectric dielectrics

    and has often involved the determination of con-

    tact angles and line tensions [1921]. Such charac-

    terisations depend on the correct interpretation of

    the pressurearea isotherm. We consider our re-cent molecular dynamics simulations and theoreti-

    cal work [22] that, contrary to expectations,

    shows the collapse pressure measured in this ex-

    periment should be independent of contact angle

    and that the initial collapse mode is by surface

    buckling. We present new molecular dynamics

    results for arrays of nanoparticles with a contact

    angle of 72 at a LV interface which confirm our

    earlier work and discuss the details of the struc-

    ture of the collapsed nanoparticle arrays.

    2. Results

    2.1. LennardJones nanoparticles at infinite

    dilution at fluid interfaces

    We consider the use of molecular simulation

    [23] to probe the wetting of nanoparticles at infi-

    nite dilution in a fluid interface. The situation of

    interest is displayed in Fig. 1. We consider partic-

    ulates at fluid interfaces with diameters in the

    range 1 nm upwards.

    In order to obtain general results we have

    defined idealised nanoparticles using both Len-

    nardJones spline (LJ/S) and Lennard Jones po-

    tentials as shown in Fig. 2. In the first case the

    particulate interacts with the LennardJones

    atoms forming the fluid phases through a LJ/S

    potential whose finite range and depth is indepen-

    dent of the particulate diameter (Fig. 2 and Eq.

    (1)). Thus even for an infinite diameter, where the

    particulate represents a planar structureless wall,

    the interaction with the fluid is the same. This

    produces a clear separation between geometric

    and energetic effects. The spline potential has the

    form

    Uij=

    4mij |f

    rsij

    12 |f

    rsij

    6n0BrsijBrs

    mij[a(rsijrc)2+b(rsijrc)

    3] rsBrsijBrc

    0 rcBr

    (1)

    where r is the separation between particles of

    species i and j, and |f is the diameter of the fluidparticles. The other variables for the LJ/S poten-

    tial are defined as follows:

    rs=(26/7)1/6, rc=(67/48)

    r, a=(24192/3211)/r s2,

    b=(387072/61009)/r s3,

    sff=0 and sfp=(|p|f)/2.

    In the second case the fluid fluid, and fluid

    nanoparticle interactions were modelled using

    Lennard Jones potentials, with a cutoff at 2.5|f.

    Note that now the fluid nanoparticle interaction

    is stronger than the fluidfluid interaction.Using the model described above we performed

    molecular dynamics simulations of particulates at

    LV interfaces. For the present work a key quan-

    tity is the contact angle the particulate makes with

    Fig. 1. The system of interest, a particulate (labelled 1) diame-

    ter | at a fluid interface, distance d from the fluid interface,

    with contact angle q.

  • 8/3/2019 02 Colsua Powell

    4/11

    C. Powell et al. /Colloids and Surfaces A: Physicochem. Eng. Aspects 206 (2002) 241 251244

    Fig. 2. (a) The LJ/S potential, (b) the Lennard Jones potential for fluidfluid, fluid colloidal nanoparticle, colloidal nanoparticle

    colloidal nanoparticle interactions.

    the liquid phase. At any instant of time the sur-

    face is discharged on the atomic scale; in addition

    the particulate moves in and out of the (instanta-

    neous) fluid interface during the simulation [24].

    Our concern in the present work is however, not

    with the fluctuation properties of the particulate/

    fluid interface system (interesting as they are), but

    with the thermodynamic properties which arise

    from averaging the system. Thus we define a

    contact angle using the average height d of the

    particulate as a function of time, with respect to

    the (average) LV equi-molar dividing surface. The

    time average of the quantity d is then related to

    the contact angle q (Fig. 1) using cos q=2d/|p. The contact angle defined in this way is consis-

    tent with that estimated from Youngs equation

    as discussed in Section 1. That a (time averaged)

    contact angle exists on the nanoscale can be seen

    from density profiles taken across the interface

    with respect to the particulate centre of mass (Fig.

    3). The profiles were calculated by considering a

    cylinder of radius r and length x centered on the

    particulate:

    z(x,r)=n(r, x)/6here n(r, x) is the average number of solventmolecules in a shell of thickness dx=0.2|f, with

    dr=0.15|f, with volume 6=2prdxdr. The figure

    shows a particulate with |p=10|f (mfp/mff=1.5) in

    a LJ/S LV interface at T*=0.75 for which k*=

    0.17 (stars refer to the usual Lennard Jones

    units). The contact angle is 90. The variation

    of contact angle with particle size and interaction

    strength has been described in detail for this

    system elsewhere [47].

    In Section 3 we will also discuss LJ/S particu-lates with |p=8|f, mfp/mff=1.25, at T*=0.5

    (z*l=0.83, z*6=0.002, k*=0.58) for which the

    contact angle is 10193. For the Lennard Jones

    system (Fig. 2(b)) with |p=7|f, mfp/mff=2, molec-

    ular simulations at T*=0.75 of a particulate in

    the LV interface (z*l =0.76, r*6=0.012, P*=

    0.008, k*=0.49) produced a contact angle of 75.

  • 8/3/2019 02 Colsua Powell

    5/11

    C. Powell et al. /Colloids and Surfaces A: Physicochem. Eng. Aspects 206 (2002) 241 251 245

    2.2. Molecular models of passi6ated gold

    nanoparticles at infinite dilution at a water/air

    interface

    In this section we report preliminary results for

    a molecular model of a gold nanoparticle at a

    water interface. Thiol passivated gold nanoparti-

    cles are described by a truncated octahedral (OT)motif and can be specified by indices (n,m) wheren is the number of atoms along any edge which

    joins two (1 1 1) faces and m the number along

    any edge joining (1 1 1) and (1 0 0) faces [25]. Our

    cluster consists of a face centred cubic lattice of

    140 atoms, adopting the OT morphology with

    n=4 and m=2.

    In our calculations, we use the semi-empirical

    many body SuttonChen [26] potential parame-

    terised for gold [27].

    Ui(r)=ma

    r

    nC

    2

    pinwhere pi=%

    j arijm

    (2)

    the first term represents the repulsion between

    atomic cores and the second term, pi, is the local

    density of atoms. The parameters m, a and C are

    determined by equilibrium lattice parameters and

    lattice energies. The exponent pairs, n and m are

    fitted to elastic constants. For gold, m=9.383265

    kJ, a=4.080A, n=10.0, m=8.0 and C=34.408

    [28]. In order to maintain the OT structure in the

    passivated cluster the value of m was increased bya factor of 5 (see below).

    2.2.1. The surface coating of the gold particle

    The potentials used to model the interactions

    between the gold, sulphur and alkyl chains can be

    divided into bonded and non-bonded interactions.

    The thiol chains were modelled using united-atom

    potentials [29] the potential functions and

    parameters used for describing the bonded inter-

    actions [3032] are shown in Table 1.

    For the molecular dynamics simulations of thepassivated gold cluster we used the code DLPOLY,

    [33]. For computational convenience in using

    DLPOLY, the bonded interaction between the gold

    and the sulphur [34] was approximated as a non-

    bonded Morse potential [35].

    U(r)=E0

    (nm)

    mr0

    r

    nnr0

    r

    mn(3)

    with E0=38.594 kJ, n=8, m=4, r0=2.9 A, .

    Standard non-bonded interactions were mod-

    elled with a Lennard Jones 12-6 potential [36].The appropriate parameters are shown in Table 2.

    Parameters for unlike interactions were ob-

    tained using LorentzBerthelot combining rules.

    2.2.2. Cluster preparation

    The thiol passivated gold cluster was prepared

    using the method of Luedtke and Landman: [35].

    Fig. 3. Density profil e o f a L J/S nanoparticle at the LV

    interface. The figure shows a particulate with |p=10|f (mfp/mff=1.5) in a LJ/S LV interface at T*=0.75 for which k*=

    0.17 (stars refer to the usual Lennard Jones units). The

    contact angle is 90. Only the surface region of the liquid

    phase immediately surrounding the nanoparticle is shown, the

    greyscale variations show local density oscillations around the

    nanoparticle extending down into the bulk liquid. As we move

    from bottom to top across the LV interface (on the right and

    left hand sides of the nanoparticle) the density falls until we

    reach the vapour phase.

  • 8/3/2019 02 Colsua Powell

    6/11

    C. Powell et al. /Colloids and Surfaces A: Physicochem. Eng. Aspects 206 (2002) 241 251246

    Table 1

    Potentials and parameters for describing carbon chain bonded interactions

    Atom typesInteraction ParametersPotential functions

    None, the bonds were constrained to a single, rigid lengthConstrained CH2CH3 r=1.54 A,bond

    CH2CH2 r=1.54 A,

    CH2SH r=1.82 A,

    CH2CH2CH3Harmonic angle Kq=519.73 kJUbend(qi)=0.5k(uiu0)2

    q0=114.4

    CH2CH2SH Kq=519.73 kJq0=114.4

    Dihedral angle Utorsion(I)=12 a1(1+cos )+

    12 a2(1cos(2i))+

    12 a3(1+cos(3i)) CH2CH2CH2SH a1=5.9046

    CHxCH2CH2CH2 a2=1.134a3=13.1608

    The bare, frozen, gold cluster was placed in a

    solution of butanethiol molecules and the systemwas allowed to equilibrate by molecular dynamics

    (1 500 000 timesteps, 1.5 ns) at low temperature

    (200 K), allowing an excess of butanethiol to

    absorb onto the gold surface. The temperature was

    then raised to 500 K in steps of 50 K (all 25 000

    timesteps, 0.025 ns) allowing desorption of excess

    butanethiols and exploration of absorption sites, so

    that the thiols were absorbed on the gold surface

    in an HCP structure with 62 molecules absorbed in

    total, in good agreement with other work [37]. The

    butane of the butanethiol was then replaced bydodecane and equilibrated (all 500 000 timesteps,

    1.25 ns). The gold atoms were kept frozen through-

    out, resulting in a surface annealed cluster without

    the structure of the gold nanoparticle being

    changed. However on unfreezing the gold the

    passivated cluster melted (the bare cluster was

    microcrystalline at all temperatures considered

    here). Due to the non-local nature of our AuSbond potential the sulphur atoms enter the gold

    cluster. By increasing the Au Au interactions

    (scaling mby a factor of 5) we forced the Au clusterto remain crystalline and the sulphurs bonded at

    the surface. The modified passivated Au nanopar-

    ticle was then equilibrated.

    2.2.3. The water liquid/6apour interface

    The water molecules in the water interface were

    modelled using the extended simple point charge

    potential [38]. The interface was constructed by

    using a starting configuration comprising a liquidslab of 500 (and also 1000) water molecules in a

    simulation cell with vacuum on either side and

    equilibrating for 50 000 timesteps (or 0.125 ns) at

    298 K. The surface tension was determined [23] by

    integrating the difference between the normal and

    tangential pressures across the interface giving a

    value of 6495 mN m1, 10% smaller than the

    experimental value [39] of 72.75 mN m1 at 298 K.

    2.2.4. Introducing the particle to the interface

    The passivated gold cluster was placed justabove the water interface and its trajectory fol-

    lowed at 298 K for 325 000 timesteps (0.8125 ns).

    Fig. 4 plots d, the position of the gold centre of

    mass with respect to the average interface, the

    particulate settles down to a position approxi-

    mately 1.5 nm above the mean interface. The

    contact angle is given by cos q=d/R, so that

    given R, it can be estimated from the figure (see

    below).

    Table 2LennardJones parameters for non-bonded interactions

    | (A, ) m (kJ)

    CH3 3.930 0.9478

    0.3908CH2 3.930

    4.450 1.6629SH

    2.737Au 3.2288

  • 8/3/2019 02 Colsua Powell

    7/11

    C. Powell et al. /Colloids and Surfaces A: Physicochem. Eng. Aspects 206 (2002) 241 251 247

    Fig. 4. Height of the centre of mass of the gold core of the nanoparticle from the water interface (equimolar surface) as a function

    of time at 298 K.

    For a flexible nanoparticle the value of the

    radius R fluctuates as the thiol molecules move in

    response to thermal fluctuations. Fig. 5 is a den-

    sity contour plot obtained in a similar manner to

    Fig. 3, for the passivated gold nanoparticle at a

    water interface. The nanoparticle clearly sits

    above the (average) water interface with a well

    defined radius which can be estimated from thefigure as 1.75 nm. Using cos q=d/R, the

    data plotted in Fig. 4 predicts an average contact

    angle of 150. Although there are significant

    fluctuations of the contact angle on a timescale of

    tens of picoseconds, this average value is in excel-

    lent agreement with that estimated from the den-

    sity plot taken from a subset of equilibrium

    configurations, Fig. 5. The particulate as a whole

    is partially dry in the water interface.

    3. Nanoparticulate arrays

    A popular method of characterising particulates

    is to use a LT to produce surface pressure (P)

    area (A) curves. Particulates are spread at the

    fluid interface and confined to a given area by a

    moveable bar (Fig. 6(a)).

    The surface pressure on the bar is measured at

    each area producing a curve similar to that shown

    in the Fig. 7. A standard analysis [1921] of the

    shape of the curve suggests that the size and

    contact angle made by the particle with the fluid

    surface can be obtained from the curve. In partic-

    ular the analysis suggests that the collapse pres-

    sure depends on the contact angle since it occurs

    when particles are ejected from the monolayer to

    form a bilayer. Hence we can measure nanoscale

    Fig. 5. Density profile of the passivated gold nanoparticle at

    equilibrium in the water interface at 298 K. The simulation

    contained a slab of saturated liquid water comprising 1000

    water molecules in a cell of dimension 6060100 A, 3. The

    contour plot shows a half plane 25100 A, 2 (one scale point

    to 0.25 A, ). These preliminary data were averaged over 566

    equilibrium configurations.

  • 8/3/2019 02 Colsua Powell

    8/11

    C. Powell et al. /Colloids and Surfaces A: Physicochem. Eng. Aspects 206 (2002) 241 251248

    Fig. 6. The LT experiment and the analogous simulation cell.

    tive particlefluid molecule interaction and a

    short range repulsive particulateparticulate in-

    teraction. For this model we take 16 nanoparticles

    in 27 645 solvent particles with a contact angle

    with the liquid of 101. In the second case there is

    a LJs particulatefluid molecule interaction and

    an attractive particulate particulate interaction

    and we simulate 64 nanoparticles with a contactangle of 75 in a fluid interface composed of

    43 888 solvent particles. By changing the y and z

    dimensions of the simulation box (the system

    density remains constant) it is possible to measure

    surface pressures at different surface areas (from

    difference in the surface tension of the two inter-

    faces) thus obtaining a pressurearea isotherm.

    The results of two simulations with particles

    having different contact angles are plotted in Fig.

    7. The surface pressure has been normalised with

    respect to the surface tension of the LV interface.The simulation reproduces the typical shape of

    the experimental isotherms. The transition is

    marked by a knee in the curve beyond which the

    surface pressure is essentially constant. Note that

    the collapse pressure is equal to the interfacial

    surface tension of the pure interface. Indeed

    monolayers with different contact angles but in

    the same solvent collapse at the same pressure. In

    the standard interpretation based on particle pro-

    motion out of the interface at the knee, they

    should collapse at different pressures; the first sign

    then that something is wrong with this view.

    Indeed by looking at computer graphics of the

    model particles in the interface as the area is

    reduced, a very different picture of what happens

    at the knee emerges. For the LJ/S nanoparticles

    (at T*=0.5) with contact angle of 101, the LV

    interface containing the particulates, buckles [23],

    creating new area at a free energy cost which is

    lower than that required to promote a particle out

    of the interface. Such buckling has been seen

    experimentally for micron size particles [40].

    For the LJs nanoparticles (at the higher tem-

    perature of T*=0.75) rather than buckle, the LV

    interface appears to roughen. Fig. 8(a) shows a

    snapshot of a monolayer of 64 nanoparticles in

    the vicinity of the knee. A layer of solvent parti-

    cles surrounds the colloids, indicating that at the

    collapse pressure the colloids remain strongly sol-

    contact angles, (as well as line tensions) from this

    simple experiment. Such characterisations depend

    on the correct interpretation of the pressurearea

    isotherm and it is this interpretation which we

    have investigated using molecular simulation.

    Fig. 6(b) illustrates the simulation method. We

    have a film of liquid with two LV surfaces at one

    of which we place model spherical nanoparticles(here with diameter 8 times the diameter of the

    fluid molecules). We consider particulates of two

    types as discussed in the previous section. In the

    first case (Fig. 2(a)) there is a short range attrac-

    Fig. 7. Pressure area curves from molecular dynamics of

    nanoparticles at LV interfaces.

  • 8/3/2019 02 Colsua Powell

    9/11

    C. Powell et al. /Colloids and Surfaces A: Physicochem. Eng. Aspects 206 (2002) 241 251 249

    Fig. 8. (a) Snapshot of the nanoparticle array at the knee of

    the curve in Fig. 7 for a contact angle of 72 ; (b) nanoparti-

    cle solvent radial distribution function at the collapse pres-

    sure.

    ticle MFP as a function of nanoparticle separa-

    tion by considering a nanoparticle dimer at the

    LV interface.

    The results shown in Fig. 9 represent the contri-

    bution to the MFP, DF due to the solvent only

    (subtracting out the direct LJs interaction). The

    curve shows features typical of those observed in

    colloid depletion force plots, such as the maxi-mum at |f, and the oscillations at larger distances.

    The force induced by the solvent is essentially

    negligible for separations larger than 4.5|f. At

    very short separations, DF remains positive and

    repulsive, reflecting the preference of the nanopar-

    ticle to be solvated. Clearly in these conditions the

    collapse of the array under pressure involves sol-

    vated nanoparticles with an effective size larger

    than in vacuum. We expect a collapse surface area

    which is significantly larger than the close packing

    area, as indeed observed in Fig. 7.

    4. Summary

    From our studies of model LJs nanoparticles

    at LJs interfaces we conclude that it is possible to

    describe the wetting of nanoscale interfaces (down

    to 1 nm) using Youngs equation when (a) the

    fluidfluid interfacial tensions are large and (b)

    the three phase line has a small curvature. These

    general results have been obtained using idealised

    structureless model nanoparticulates. In nature

    however nanoparticles have internal and surface

    structures, are not necessarily rigid and the sol-

    vents may themselves be complex. Our prelimi-

    nary results for passivated gold nanoparticles at

    water interfaces indicate that such particles have a

    well-defined mean contact angle with however

    significant fluctuations on a timescale of tens of

    picoseconds. Our simulations of nanoparticle ar-

    rays in the LT configuration strongly suggest that

    the collapse pressure measured in LT experiments

    is independent of contact angle and that such

    experiments cannot be used to determine contact

    angles. We expect the array to collapse by surface

    roughening when the surface pressure is equal to

    the surface tension of the pure solvent interface.

    In experiments where a different collapse pressure

    is observed we suggest that contamination of the

    vated. This conclusion is reinforced by Fig. 8(b),which shows the solvent nanoparticle radial dis-

    tribution function for this monolayer. As sug-

    gested by the snapshot there is a layer of solvent,

    which is strongly adsorbed at the nanoparticle

    surface. In addition we have analysed the mean

    force potential (MFP) of the colloids due to the

    solvent particles. We have calculated the nanopar-

  • 8/3/2019 02 Colsua Powell

    10/11

    C. Powell et al. /Colloids and Surfaces A: Physicochem. Eng. Aspects 206 (2002) 241 251250

    Fig. 9. Potential of mean force for solvated nanoparticles with a contact angle of 72.

    fluid interface by surfactant or other components

    of the system is likely to be responsible.

    Acknowledgements

    We thank EPSRC for support through grant

    GR/M94427, GR/R39726.

    References

    [1] J.R. Heath, C.M. Knobler, D.V. Leff, J. Phys. Chem. B

    101 (1997) 189.

    [2] P.S. Swain, R. Lipowsky, Langmuir 14 (1998) 6772.[3] P.M. Ajayan, S. Iijima, Nature 361 (1993) 333.

    [4] F. Bresme, N. Quirke, Phys. Rev. Lett. 80 (1998) 3791.

    [5] F. Bresme, N. Quirke, J. Chem. Phys. 110 (1999) 3536.

    [6] F. Bresme, N. Quirke, Phys. Chem. Chem. Phys. 1 (1999)

    2149.

    [7] F. Bresme, N. Quirke, J. Chem. Phys. 112 (2000) 5985.

    [8] T. Shen, R. Weissleder, M. Papisov, A. Boganov, T.J.

    Brady, MRM 29 (1993) 599.

    [9] R. Weissleder, et al., AJR 152 (1989) 167.

    [10] S.D. Evans, S.R. Johnson, Y.L. Cheng, T. Shen, J.

    Mater. Chem. 10 (2000) 183.

    [11] F. Burmeister, C. Schafle, T. Matthes, M. Bohmisch, J.

    Boneberg, Pleiderer, Langmuir 13 (1997) 2983.

    [12] G.C. Bond, D.T. Thompson, Catal. Rev.-Sci. Eng. 41

    (1999) 319.[13] A. Pockels, Nature 12 (1891).

    [14] I. Langmuir, J. Am. Chem. Soc. 39 (1848) 1917.

    [15] G. Gabrielli, G.G.T. Guarni, E. Ferroni, J. Colloid Inter-

    face Sci. 54 (1976) 424.

    [16] A. Doroszkowski, R. Lambourne, J. Polymer Sci. C. 34

    (1971) 253.

    [17] R. Aveyard, J.H. Clint, D. Nees, V. Paunov, Langmuir 16

    (2000) 1969.

    [18] H.M. McConnell, Annu. Rev. Phys. Chem. 42 (1991) 171.

    [19] J.H. Clint, S.E. Taylor, Colloids Surfaces 65 (1992) 61.

    [20] J.H. Clint, N. Quirke, Colloids Surfaces 78 (1993) 277.

    [21] R. Aveyard, J.H. Clint, J. Chem. Soc. Faraday Transac-

    tions 91 (1995) 175.[22] N.I.D. Fenwick, F. Bresme, N. Quirke, J. Chem. Phys.

    114 (2001) 7274.

    [23] N Quirke, Mol. Sim. 26 (2001) 1 and references therein.

    [24] Readers may view a movie of such a simulation at our

    website available from: http://www.ch.ic.ac.uk/quirke/

    research.html.

    [25] R.L Whetten, et al., Adv. Mater. 8 (1996) 428.

    [26] A.P. Sutton, J. Chen, Philos. Mag. Lett. 61 (1990) 139.

    [27] B.D. Todd, R.M. Lynden-Bell, Surf. Sci. 281 (1993) 191.

    http://www.ch.ic.ac.uk/quirke/research.htmlhttp://www.ch.ic.ac.uk/quirke/research.htmlhttp://www.ch.ic.ac.uk/quirke/research.htmlhttp://www.ch.ic.ac.uk/quirke/research.html
  • 8/3/2019 02 Colsua Powell

    11/11

    C. Powell et al. /Colloids and Surfaces A: Physicochem. Eng. Aspects 206 (2002) 241 251 251

    [28] B.D. Todd, R.M. Lynden-Bell, Surf. Sci. 328 (1995)

    170.

    [29] J.P. Ryckaert, A. Bellemans, Faraday Discussions of the

    Chem. Soc. 66 (1978) 95.

    [30] J. Hautman, M.L. Klein, J. Chem. Phys. 91 (1989) 4994.

    [31] S. Balasubramanian, M.L. Klein, J.I. Siepmann, J. Chem.

    Phys. 103 (1995) 3184.

    [32] W.L. Jorgensen, J. Phys. Chem. 90 (1986) 6379.

    [33] DLPOLY developed by W. Smith and T.R. Forester.[34] H. Sellers, A. Ulman, Y. Shnidman, J.E. Eilers, J. Am.

    Chem. Soc. 115 (1993) 9389.

    [35] W.D. Luedtke, U. Landman, J. Phys. Chem. B 102 (1998)

    6566.

    [36] T.K. Xia, J. Ouyang, M.W. Ribarsky, U. Landman, Phys.

    Rev. Lett. 69 (1992) 1967.

    [37] W.D. Luedtke, U. Landman, J. Phys. Chem. 100 (1996)

    13323.

    [38] H.J.C. Berendsen, J.R. Grigera, T.P. Straatsma, J. Phys.

    Chem. 91 (1987) 6269.

    [39] CRC Handbook of Chemistry and Physics, 74th ed.[40] R. Aveyard, J.H. Clint, D. Nees, N. Quirke, Langmuir 16

    (2000) 8820.