dpd simulation

Upload: abhishek-kumar-gupta

Post on 14-Apr-2018

259 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/27/2019 DPD Simulation

    1/16

    Journal of Colloid and Interface Science 283 (2005) 251266

    www.elsevier.com/locate/jcis

    Comparison between theoretical values and simulation results of viscosityfor the dissipative particle dynamics method

    Akira Satoh a,, Tamotsu Majima b

    a Department of Machine Intelligence and System Engineering, Faculty of System Science and Technology, Akita Prefectural University, 84-4, Ebinokuchi,

    Tsuchiya-aza, Honjo 015-0055, Japanb Graduate School of Science and Technology, Chiba University, 1-33, Yayoicho, Inage-ku, Chiba 263-8522, Japan

    Received 10 October 2003; accepted 13 September 2004

    Abstract

    In order to investigate the validity of the dissipative particle dynamics method, which is a mesoscopic simulation technique, we have derived

    an expression for viscosity from the equation of motion of dissipative particles. In the concrete, we have shown the FokkerPlanck equation

    in phase space, and macroscopic conservation equations such as the equation of continuity and the equation of momentum conservation.

    The basic equations of the single-particle and pair distribution functions have been derived using the FokkerPlanck equation. The solutions

    of these distribution functions have approximately been solved by the perturbation method under the assumption of molecular chaos. The

    expressions of the viscosity due to momentum and dissipative forces have been obtained using the approximate solutions of the distribution

    functions. Also, we have conducted nonequilibrium dynamics simulations to investigate the influence of the parameters, which have appeared

    in defining the equation of motion in the dissipative particle dynamics method. The theoretical values of the viscosity due to dissipative forces

    in the HoogerbruggeKoelman theory are in good agreement with the simulation results obtained by the nonequilibrium dynamics method,

    except in the range of small number densities. There are restriction conditions for taking appropriate values of the number density, numberof particles, time interval, shear rate, etc., to obtain physically reasonable results by means of dissipative particle dynamics simulations.

    2004 Published by Elsevier Inc.

    Keywords: Dissipative particle dynamics method; Mesoscopic approach; Nonequilibrium dynamics method; Transport coefficients; Viscosity

    1. Introduction

    The hydrodynamic solution for a three-particle system

    has to be combined into a simulation method in order to take

    into account multibody hydrodynamic interactions among

    colloidal particles more precisely. However, it is highly dif-

    ficult even to solve analytically the flow field for a three-

    particle system [1,2], and, therefore, it seems to be almost

    hopeless to obtain an analytical solution for a nonspherical

    particle system such as a system composed of rodlike par-

    ticles. Thus, we have the choice to take another approach

    to develop a more precise simulation method for colloidal

    dispersions. If both colloidal particles and solvent mole-

    * Corresponding author. Fax: +81-184-27-2188.

    E-mail address: [email protected] (A. Satoh).

    cules are simulated, we can obtain the particle motion and

    the solution of the flow field simultaneously. However, if

    molecules themselves are treated in simulations, we cannot

    develop an effective simulation method, since the character-

    istic time for the motion of colloidal particles is significantly

    different from that of solvent molecules. In other words, this

    kind of molecular-dynamics-like method is unrealistic as a

    technique for simulation of a colloidal dispersion from a

    simulation time point of view. To circumvent this difficulty,

    the concept of fluid particles seems to be promising. Hooger-

    brugge and Koelman [3,4] have developed the dissipative

    particle dynamics method in terms of fluid particles. In their

    theory, a fluid is regarded as being composed of such virtual

    particles, and the flow field is solved by simulating these par-

    ticles. The fluid particles interact with each other, exchange

    momentum, and should make random motion like Brownian

    0021-9797/$ see front matter 2004 Published by Elsevier Inc.

    doi:10.1016/j.jcis.2004.09.050

    http://www.elsevier.com/locate/jcismailto:[email protected]:[email protected]://www.elsevier.com/locate/jcis
  • 7/27/2019 DPD Simulation

    2/16

    252 A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266

    particles. From now on, such fluid particles are called dissi-

    pative particles.

    For the simulation of the flow field for a colloidal dis-

    persion, the motion of colloidal particles is dependent on

    the interaction between colloidal particles themselves, the

    interaction between colloidal particles and dissipative par-

    ticles, and the interaction between colloidal particles andan applied field such as a magnetic field. Hence, in this

    simulation technique, the solution of pair or three-body hy-

    drodynamic interactions need not be combined with it, in

    order to simulate colloidal particles. This is in high con-

    trast with the ordinary simulation methods such as Stokesian

    dynamics methods [511]. Multibody hydrodynamic inter-

    actions among colloidal particles are automatically taken

    into account from the interactions of colloidal particles with

    dissipative ones. This is a significant feature in the dissi-

    pative particle dynamics method compared with ordinary

    simulation methods. It is clear from these discussions that

    the concept of the dissipative particle dynamics method isstraightforwardly applicable to a colloidal dispersion com-

    posed of nonspherical particles, and, therefore, seems to be a

    highly promising technique from a simulation point of view.

    As will be shown later, the equation of motion for the dis-

    sipative particle dynamics, which has been presented to date

    [3,4,1214], is not unique, but has considerable freedom;

    for example, the equation of motion has unknown constants

    which are parameters representing the strength of conserv-

    ative and dissipative forces. However, whether or not the

    equation of motion can yield physically reasonable solu-

    tions has not been sufficiently clarified [3,4,15], although

    there is uncertainty in the equation of motion for dissipa-

    tive particles. Hence, in order to apply the dissipative par-

    ticle dynamics method widely to the usual flow problems

    and physicsengineering problems of colloidal dispersions,

    it is very important to investigate the influence of such un-

    certainty in the equation of motion on the solution of the

    flow properties such as the flow field and transport coeffi-

    cients. This may be accomplished by comparing simulation

    results with theoretical solutions concerning transport coef-

    ficients, which can be derived analytically from the equation

    of motion.

    The present study, therefore, attempts first to derive an

    analytical expression for viscosity using the equation of mo-

    tion, secondly to obtain the numerical data in terms of dis-sipative particle dynamics simulations, and finally to com-

    pare the simulation results with the analytical solutions. In

    the concrete, we show the FokkerPlanck equation in phase

    space, and macroscopic conservation equations such as the

    equation of continuity and the equation of momentum con-

    servation. The basic equations of the single-particle and pair

    distribution functions are derived using the FokkerPlanck

    equation. These equations of the distribution functions are

    approximately solved to obtain an expression for the viscos-

    ity due to momentum and dissipative forces. Nonequilibrium

    dynamics simulations have been conducted under circum-

    stances of simple shear flow. Finally, these simulation results

    are compared with the theoretical solutions to investigate the

    influence of the model parameters, which have appeared in

    definitions of the equation of motion, number density, num-

    ber of particles, time interval, etc., on this transport coeffi-

    cient and also the internal structure of a dissipative particle

    system. The present study may stimulate the development of

    a new equation of motion for the dissipative particle dynam-ics, which gives rise to more physically reasonable simula-

    tion results.

    2. Dynamics of dissipative particles

    2.1. Kinetic equation of dissipative particles

    The equation of motion for dissipative particles has to

    obey some physical constraints in order for the solution

    obtained by the dissipative particle dynamics method to

    agree with that obtained by the NavierStokes equation. Theconservation of the total momentum of a system and the

    Galilean invariance have to be satisfied by the equation of

    motion as a minimum request [3,4,12,13].

    We concentrate our attention on particle i and consider

    the forces acting on this particle. The following three kinds

    of forces may be physically reasonable as forces acting on

    particle i: a repulsive conservative force FCij exerted by the

    other particles, a dissipative force FDij providing a viscous

    drag to the system, and a random or stochastic force FRij in-

    ducing the thermal motion of particles. With these forces, the

    equation of motion of particle i can be written as [1214]

    (1)mdvi

    dt=

    j (=i)

    FCij +

    j (=i)FDij +

    j (=i)

    FRij,

    in which m is the mass of particle i , vi is the velocity, and,

    concerning the subscripts, for example, FCij is the force act-

    ing on particle i from particle j.

    Now we have to embody specific forms of the above-

    mentioned forces. It may be reasonable to assume that the

    conservative force FCij depends only on the relative position

    rij (= ri rj), and not on the particle velocities. An ex-plicit expression for this force will be shown later. Since the

    Galilean invariance has to be satisfied, the dissipative force

    FDij and the random force FRij should not be dependent onthe position ri and velocity vi themselves, but should be

    functions of the relative position rij and relative velocity vij(= vi vj), if necessarily. Additionally, it may be reason-able to assume that the random force FRij does not depend on

    the relative velocity but the relative position rij alone. Fur-

    thermore, we have to take into account the isotropy of the

    particle motion and the decrease in the magnitude of forces

    with particleparticle separation. The following expressions

    for FDij and FRij

    satisfy these physical requirements [1214],

    (2)FDij = wD(rij)(eij vij)eij,

    (3)FRij = wR(rij)eijij,

  • 7/27/2019 DPD Simulation

    3/16

    A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266 253

    in which rij = |rij|, and eij is the unit vector denoting thedirection of particle i from particle j, expressed as eij =rij/rij. Also ij is a random variable inducing the random

    motion of particles and has to satisfy the following stochastic

    properties:

    ij(t )

    = 0,(4)

    ij(t)ij (t

    )= (ii jj + ij j i )(t t).

    This variable satisfies the characteristic of the symmetry

    ij = j i , which ensures that the total momentum of thesystem is conserved. The wD(rij) and wR(rij) are weight

    functions to reproduce the decrease in forces with particle

    particle separation, and and are constants specifying the

    magnitude of forces. These constants can be related to the

    system temperature and friction coefficient, which will be

    shown later. It is seen from the expression for FDij that this

    force acts in such a way as to relax the relative motion of

    particles i and j. On the other hand, FRij acts as if inducing

    the thermal motion of particles, but this force satisfies theactionreaction formula, so that the conservation of the total

    momentum of the system is ensured.

    The substitution of Eqs. (2) and (3) into Eq. (1) leads to

    the following equation:

    mdvi

    dt=

    j (=i)

    FCij(rij)

    j (=i) wD(rij)(eij vij)eij

    (5)+

    j (=i) wR(rij)eijij.

    If this equation is integrated with respect to time over a small

    time interval from t to t+

    t, then the finite difference equa-

    tions governing the particle motion in simulations can be

    obtained as

    (6)ri = vi t ,

    (7)

    vi = 1m

    j (=i)

    FCij(rij)

    j (=i) wD(rij)(eij vij)eij

    t

    + 1m

    j (=i)

    wR(rij)eijWij,

    in which

    (8)Wij =

    t+tt

    ij d.

    This Wij has to satisfy the following stochastic properties,

    which arise from Eq. (4):

    Wij = 0,(9)WijWij = (ii jj + ij j i )t.

    If a new stochastic variable ij is introduced from the defini-

    tion Wij = ij

    t, the third term in Eq. (7) can be written

    as

    (10)1

    m

    j (=i) wR(rij)eijij

    t ,

    in which ij has to satisfy the following stochastic proper-

    ties:

    (11)ij = 0, ijij = (ii jj + ij j i ).It is seen from Eq. (11) that the stochastic variable ij is sam-

    pled from a uniform or normal distribution with zero average

    value and unit variance [3,4,1214].

    The equation of motion in Eq. (5) satisfies the conserva-

    tion law of the total momentum of the system, but not the

    energy conservation law. The equation of motion has to be

    modified to satisfy the conservation of the system energy

    [16,17].

    The coefficients and and the weight functions

    wD(rij) and wR(rij) cannot be determined independently,

    because there are some relationships among these quanti-

    ties. The derivation of these relationships will be shown in

    the next section.

    2.2. FokkerPlanck equation

    We use the notation ri for the position vector of particle i

    and vi for the velocity vector. Also, for simplicity of expres-

    sion, the vector vN is used for describing the velocity vectors

    of all the particles (that is, vN represents v1, v2, v3, . . .), and

    similarly rN is for the position vectors of all the particles

    (that is, rN represents r1, r2, r3, . . .). If the probability that

    a particle position and velocity are found within the range

    from (rN, vN) to (rN + rN, vN + vN) is denoted byW (rN, vN, t) drN dvN, then the probability density function

    W (rN, vN, t) satisfies the following stochastic formula, i.e.,

    the ChapmanKolmogorov equation,

    W (rN, vN, t) =

    W (rN rN, vN vN, t t )

    (12)

    (rN rN, vN vN; rN, vN) d(rN) d(vN),

    in which (rNrN, vNvN; rN, vN) is the transi-tion probability for the state (rN rN, vN vN) trans-ferring to another state (rN, vN) during a time interval t

    and this does not depend on the time point t. If the veloc-

    ity vN does not change appreciably during the small time

    interval t, then can be written as

    (rN rN, vN vN; rN, vN)

    (13)

    = (rN rN, vN vN; vN)(rN vNt),

    in which is the transition probability and is independent

    of rN. The substitution of Eq. (13) into Eq. (12) leads to

    the following different form of the ChapmanKolmogorov

    equation:

    W (rN + vNt , vN, t+ t )=

    W (rN, vN vN, t)

    (14) (rN

    , v

    N

    vN

    ; vN

    ) d(v

    N

    ).

  • 7/27/2019 DPD Simulation

    4/16

    254 A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266

    The FokkerPlanck equation in phase space (or the Chan-

    drasekhar equation) can be derived by reforming Eq. (14) in

    terms of Taylor series expansions with the finite difference

    equations (6) and (7). Since the detailed derivation proce-

    dure is shown in Ref. [18], we here show the final result:

    Wt

    +

    i

    vi W ri

    +

    i

    j

    (i=j )

    FCij

    m W

    vi

    =

    i

    j

    (i=j )

    eij vi

    1

    m wD(rij)(eij vij)W

    + 12

    i

    j

    (i=j )

    1

    m22w2R(rij)

    (15) eij

    vi eij

    vi eij

    vjW.

    2.3. Final expression of equation of motion and its

    nondimensional form

    If a system composed of dissipative particles is in equi-

    librium, then the equilibrium distribution Weq becomes the

    canonical distribution for an ensemble which is specified by

    a given particle number N, volume V, and temperature T.

    With the notation U for the potential energy, Weq is written

    as

    (16)Weq =1

    Zexp

    1

    kTN

    i=1

    mv2i

    2+ U,

    in which Z is the partition function, and U is related to the

    conservative force FCij by

    (17)FCij = U

    rij.

    The equilibrium distribution Weq has to satisfy the Fokker

    Planck equation in phase space in Eq. (15). Since the left-

    hand side in Eq. (15) vanishes for the substitution of Weq,

    the right-hand side also has to become zero. This is accom-

    plished by the requirements

    (18)wD(rij) = w2

    R(rij), 2

    = 2 k T ,in which k is Boltzmanns constant. The second equation is

    the fluctuationdissipation theorem for the dissipative parti-

    cle dynamics.

    We now have to determine the explicit expressions for the

    conservative force FCij(rij) and the weight function wR(rij).

    FCij(rij) is a repulsive force preventing unphysical exces-

    sive overlaps between particles, and wR(rij) has to be set

    so that interparticle forces decrease with increasing particle

    particle separations. These requirements are satisfied by the

    expressions

    (19)FCij = wR(rij)eij,

    (20)wR(rij) =

    1 rijrc

    for rij rc,

    0 for rij > rc,

    in which is a constant representing the magnitude of the re-

    pulsive forces. By substituting these equations into Eqs. (6)

    and (7) with considering Eq. (10), the final expression for

    the equation of motion of the dissipative particle dynamicscan be written as

    (21)ri = vi t ,

    (22)

    vi =

    m

    j (=i)

    wR(rij)eijt

    m

    j (=i)

    w2R(rij)(eij vij)eijt

    + (2 k T )1/2

    m

    j (=i)

    wR(rij)eijij

    t ,

    in which the expression for wR(rij) has already been shownin Eq. (20). The stochastic variable ij has to obey the sto-

    chastic properties shown in Eq. (11) and is sampled from a

    uniform or normal distribution with zero average and unit

    variance, as already pointed out. The dissipative particle dy-

    namics method uses Eqs. (21) and (22) to simulate the mo-

    tion of dissipative particles.

    Lastly, we show an exampleof the nondimensionalization

    method used for actual simulations. To nondimensionalize

    each quantity, the following representative values are used:

    (kT/m)1/2 for velocities, rc for distances, rc(m/kT )1/2 for

    time, (1/r 3c ) for number densities, etc. With these represen-

    tative values, Eqs. (21) and (22) can be nondimensionalized

    as

    (23)ri = vi t,

    (24)

    vi =

    j (=i)wR(r

    ij)eijt

    j (=i)w2R(r

    ij)(eij vij)eijt

    + (2)1/2

    j (=i)wR(r

    ij)eijij

    t,

    in which

    (25)wR(rij) =

    1 rij for rij 1,0 for rij > 1,

    (26) = rckT

    , = rc(mkT )1/2

    .

    In these equations, the quantities with superscript * are di-

    mensionless. Equations (23) and (24) clearly show that one

    can start a dissipative particle dynamics simulation if appro-

    priate values of the nondimensional parameters and are adopted, and also the time interval t , the number den-sity n0 (= r3c N/V , V is the system volume), and the particlenumber N are properly specified. Results obtained by the

  • 7/27/2019 DPD Simulation

    5/16

    A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266 255

    simulations should not depend on the nondimensional para-

    meters which have been introduced in embodying the equa-

    tion of motion; one has to be, therefore, careful in setting

    appropriate values for these values in conducting dissipa-

    tive particle dynamics simulations. For example, since time

    is nondimensionalized by the representative value based on

    the mean velocity v ( (kT/m)1/2) and the radius rc, themagnitude of the dimensionless time interval t may haveto be taken sufficiently smaller than unity [19].

    3. Transport equations

    In the present section, the dissipative particle dynamics

    method, the equation of continuity, and the momentumequa-

    tion of the fluid are derived using the equation of motion of

    the dissipative particle dynamics.

    If an arbitrary physical quantity A(rN, vN) is not de-

    pendent on time explicitly, the time average A can beexpressed, using the probability density function W which

    satisfies Eq. (15), as

    (27)A =

    AW (rN, vN, t) drN dvN,

    in which

    (28)

    W (rN, vN, t) drN dvN = 1.

    Hence, the time variation of A can be expressed, fromEq. (15), as

    tA =

    A

    W

    tdrN dvN

    =

    A

    i

    vi W ri

    i

    j

    (i=j )

    FCij

    m W

    vi

    +

    i

    j

    (i=j )

    eij vi

    1

    m wD(rij)(eij vij)W

    + 12 i j

    (i=j )

    1

    m22w2R(rij)

    (29)

    eij

    vi

    eij

    vi eij

    vj

    W

    drN dvN.

    By applying integration by parts, this equation is written as

    tA =

    i

    vi A

    ri

    i

    j

    (i=j )

    FCij

    m A

    vi

    i j

    (i=j )

    mwD(rij)(eij vij)eij

    A

    vi

    + 12

    i

    j

    (i=j )

    1

    m22w2R(rij)

    eij

    vi

    (30)

    eij

    vi eij

    vj

    A

    .

    Next we derive the equation of continuity and the momen-

    tum equation of the fluid using Eq. (30). If A is defined by

    the equation

    (31)A =N

    i=1m(r ri ),

    then the average A, which is evaluated from Eq. (27), isequal to a local density (r). That is,

    (32)A =

    N

    i=1

    m(r ri )

    = (r).

    IfA is taken as

    (33)A =N

    i=1mvi (r ri ),

    then the average A is now

    (34)A = (r)u(r),in which u(r) is a macroscopic fluid velocity at position r.

    The substitution of Eq. (31) into Eq. (30) leads to the fol-

    lowing expression:

    (35)

    t+

    r (u) = 0.

    This is no other than the equation of continuity.

    Next the substitution of Eq. (33) into Eq. (30) leads to the

    following equation:

    t(u) =

    i

    mvi ri

    vi (r ri )

    +

    i

    j

    (i=j )

    FCij

    vi

    vi (r ri )

    i

    j

    (i=j )

    wD(rij)(eij vij)

    eij vi

    vi (r ri )

    + 12

    i

    j

    (i=j )

    1

    m2w2R(rij)

    eij

    vi

    (36) eij

    vi

    vi (r ri )

    vj

    vj(r rj)

    .

    Thus, by taking into account Eq. (2), and conducting a simi-

    lar reformation, Eq. (36) can be simplified as

  • 7/27/2019 DPD Simulation

    6/16

    256 A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266

    t(u) =

    r

    i

    mvi vi (r ri )

    (37)+

    i

    j

    (i

    =j )

    FCij + FDij

    (r ri )

    .

    By taking account of the relation FDij = FDijeij, the followingequation can be obtained:

    i

    j

    (i=j )

    FDij(r ri )

    = 12

    i

    j

    (i=j )

    FDij(r ri ) + FDj i (r rj)

    (38)= 12 i j

    (i=j )

    FDijeij

    (r ri ) (r rj)

    .

    If the following relationship concerning the Dirac delta func-

    tion is taken into account,

    (r rj) = (r ri + rij)

    (39)

    = (r ri ) + r

    eij

    rij0

    (r ri + eij) d

    ,

    then Eq. (38) can be reformed further as

    i

    j(i=j )

    FDij(r

    ri )

    (40)

    = 12

    i

    j

    (i=j )

    r

    FDijeijeij

    rij0

    (r ri + eij) d

    .

    By conducting a similar reformation concerning FCij, Eq. (37)

    can finally be written as

    (41)

    t(u) =

    r (uu) +

    r (K + U),

    in which

    K =

    i

    m(vi u)(vi u)(r ri )

    ,

    (42)

    U = 1

    2

    i

    j

    (i=j )

    FCij + FDij

    eijeij

    rij0

    (r ri + eij) d

    .

    If the equation of continuity in Eq. (35) is taken into consid-

    eration, Eq. (41) reduces to the momentum equation of the

    fluid,

    (43)u

    t +u

    u

    r=

    r (K

    +

    U),

    in which K and U are stress tensors due to the particle

    momentum and the forces acting between particles, respec-

    tively. It has now been shown that the equation of motion of

    the dissipative particle dynamics gives rise to the momen-

    tum equation of the fluid. The virial equation of state can be

    derived by evaluating directly K and U in Eq. (42).

    4. Expressions for transport coefficients

    4.1. Distribution functions

    The local number density n(r, t) at position r at time t

    can be expressed using W (rN, vN, t ) as [18]

    n(r, t) =

    Ni=1

    (r ri )

    = i

    (r ri )W (rN, vN, t ) drN dvN

    = N

    W (r, r2, . . . , rN, vN, t)

    (44) dr2 . . . drN dvN.Similarly, the pair correlation function g(r, r, t) can be writ-ten as [18]

    g(r, r, t )

    = 1n20

    N

    iN

    j(i=j )

    (r ri )(r rj)

    W (rN, vN, t ) drN dvN

    = N(N 1)n20

    (45)

    W (r, r, r3, . . . , rN, vN, t) dr3 . . . drN dvN,

    in which n0 is the mean number density, given by n0 =N/V, as defined before.

    If the distribution function f (r, v, t) is defined as [14]

    (46)f (r, v, t) = N

    i=1(r ri )(v vi )

    ,

    f (r, v, t ) can be written, using W (rN, vN, t ), as

    (47)

    f (r, v, t) = N

    W (r, r2, . . . , rN, v, v2, . . . , vN, t )

    dr2 . . . drN dv2 . . . dvN.By comparing Eq. (47) with Eq. (44), it is seen that there is

    a relationship between f (r, v, t) and n(r, t ):

    (48)n(r, t)

    = f (r, v, t) dv.

  • 7/27/2019 DPD Simulation

    7/16

    A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266 257

    Similarly,

    (49)n(r, t )u(r, t) =

    Ni

    vi (r ri )

    =

    vf (r, v, t) dv.

    Next, if we define the pair distribution f(2)(r, r, v, v, t )by the equation [14]

    f(2)(r, r, v, v, t)

    =

    Ni=1

    Nj=1

    (i=j )

    (r ri )(r rj)(v vi )(v vj)

    = N(N 1)

    (50)

    W (r, r, r3, . . . , rN, v, v, v3, . . . , vN, t)

    dr3 . . . drN dv3 . . . dvN,

    then thepair distribution can be related to the pair correlationfunction g(r, r, t) as

    (51)n20g(r, r, t) =

    f(2)(r, r, v, v, t) dv dv.

    It is clear from Eq. (30) that the average of an arbi-

    trary quantity A(rN, vN) can be evaluated using f (r, v, t )

    and f(2)(r, r, v, v, t ) without obtaining the expression forW (rN, vN, t ). Hence, we here first show the basic equations

    for f and f(2), and then solve the equations analytically to

    obtain the approximate solutions.

    If we set A =i

    (r ri )(v vi ) in Eq. (30), and re-form the equation with the formulae of vector operations, the

    following equation can be obtained:

    tf (r, v, t) + v

    rf (r, v, t)

    = m

    wD(R)RR

    : v

    (v v)f(2)(r, r + R, v, v, t)dR dv

    + 2

    2m2

    w2R(R)RR

    (52)

    :2

    vv

    f(2)(r, r

    +R, v, v, t) dR dv.

    It has been assumed in deriving this equation that there are

    no conservative forces. In Eq. (25), R is the unit vector,

    denoted by R = R/|R|. It is seen from Eq. (52) that thepair distribution function f(2)(r, r, v, v, t ) is necessary forobtaining the solution off (r, v, t), and, similarly, the multi-

    body distribution functions more than the pair distribution is

    required to get the solution of f(2). This clearly shows that

    Eq. (52) is not closed, but a hierarchy equation. To close the

    equation for f, the following assumption of molecular chaos

    is introduced [14]:

    (53)f

    (2)

    (r, r, v, v, t) f (r, v, t ) f (r, v,t).

    This assumption theoretically enables us to solve the equa-

    tion of the distribution function. The substitution of Eq. (53)

    into Eq. (52) leads to the following equation:

    tf (r, v, t) + v

    rf (r, v, t)

    = m

    wD(R)f(r + R, v, t)RR

    : v

    (v v)f (r, v, t)dR dv

    + 2

    2m2

    w2R(R)f (r + R, v, t )RR

    (54): 2

    v vf (r, v, t ) dR dv.

    The validity of solutions which are obtained from Eq. (54)

    is strongly dependent on the physical reasonability of the

    assumption of molecular chaos.

    The nondimensionalform of Eq. (54) may be more under-standable to be solved analytically. According to the nondi-

    mensionalization method, which has been shown in Sec-

    tion 2.3, Eq. (54) is nondimensionalized as

    tf(r, v, t) + v

    rf(r, v, t)

    = n0

    wD(R)f(r + R, v , t)RR

    : v

    (v v )f(r, v, t)dR dv

    + n0 wD(R)f(r + R, v , t)RR

    (55): 2

    v vf(r, v, t) dR dv ,

    in which the distribution function f has been nondimension-

    alized as f = (1/n0)(kT/m)3/2f. If we set = 1/n0and assume to be much smaller than unity, the perturba-

    tion method is applicable. Hence, f is expanded, with theperturbation parameter , as

    f(r, v, t) = f0 (r, v, t) + f1 (r, v, t)(56)+ 2f2 (r, v, t) + .

    By substituting this equation into Eq. (55) and neglecting the

    higher-order terms, we can obtain the following expression:

    tf0 + v

    rf0

    + 2

    tf1 + v

    rf1

    =

    wD(R)

    f0 (r + R, v , t)

    + f1 (r + R, v , t)

    RR

    : v

    (v v )f0 (r, v, t)+ f1 (r, v, t)

    dR dv

    + wD(R)f

    0 (r

    + R, v , t)

    + f1 (r + R, v , t)RR

  • 7/27/2019 DPD Simulation

    8/16

    258 A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266

    (57)

    : 2

    vv

    f0 (r, v, t) + f1 (r, v, t)

    dR dv .

    The equation for the zeroth order of can be written, from

    this equation, as

    wD(R)f0 (r + R, v , t)RR

    : v

    (v v )f0 (r, v, t)

    dR dv

    +

    wD(R)f0 (r

    + R, v , t)RR

    (58): 2

    vvf0 (r

    , v, t) dR dv = 0.Similarly, the equation for the first order of can be ex-

    pressed as

    tf0

    +v

    rf0

    =

    wD(R)

    f0 (r + R, v , t)RR

    :

    v(v v )f1 (r, v, t)

    + f1 (r + R, v , t)RR

    : v

    (v v )f0 (r, v, t)

    dR dv

    +

    wD(R)

    f0 (r + R, v , t)RR

    :2

    vv f1 (r, v, t)+ f1 (r + R, v , t)RR

    (59): 2

    vvf0 (r

    , v, t)

    dR dv .

    Equation (58) can straightforwardly be solved as

    (60)f0 (r, v, t) = n

    1

    2

    3/2exp

    1

    2(v u)2

    .

    By reforming Eq. (59) with the formulae of vector opera-

    tions, the following equation is obtained:

    t

    f0 + v r f0 =

    n(r + R, t)wD(R)RR

    (61):

    v

    (v u)f1 (r, v, t)

    + 2

    vvf1 (r

    , v, t)

    dR.

    If it is assumed that n(r +R, t) (= n/n0) is independentof the position and is constant, and the integration is carried

    out using the polar axis coordinates of R, then Eq. (61) can

    be simplified as

    t f0 + v

    r f0

    = [w]3

    v (v u)f1 (r, v, t)

    (62)+ v

    v

    f1 (r, v, t)

    ,

    in which [w] is defined as

    (63)[w] =1

    0

    wD dR =

    10

    wD4 R2 dR.

    The left-hand side of Eq. (62) can finally be reformed, by

    taking account of Eq. (60), as

    tf0 + v

    rf0

    (64)= 12

    f0 UU :

    ru +

    rut

    ,

    in which the notation U, defined by U

    =(v

    u), has

    been used. Also, the superscript t means a transposed ten-sor. Now we have prepared for solving Eq. (62) to get the

    solution off1 .As the first trial to find the solution of f1 , we substitute

    f0 UU into f1 in Eq. (62) to get the following expression

    for the right-hand side of Eq. (62):

    (65)1

    3[w]2UUf0 + 2If0 .

    As the second trail, we substitute f0 U2I into f1 in

    Eq. (62) to get the following equation for the right-hand side

    of Eq. (62):

    (66)13[w]2U2f0 I + 6f0 I.

    As the third trial, we substitute f0 I into f1 in Eq. (62) to

    obtain zero for the right-hand side of Eq. (62). Finally, if the

    expression for f1 is assumed to be

    (67)UUf0 1

    3U2f0 I,

    then the right-hand side of Eq. (62) becomes the following

    expression:

    (68)23[w]U

    Uf0 1

    3U2f0 I.

    After all, from these considerations, the solution of Eq. (62)

    can straightforwardlybe seen to be the following expression:

    f1 = 3

    2

    1

    n0[w]

    UUf0 1

    3U2f0 I

    (69): 12

    ru +

    rut

    .

    If the equation of continuity, / r u = 0, is taken intoaccount, it is seen that the following relationship is satisfied:

    (70)I

    :

    ru

    +

    ru

    t

    = 0.

  • 7/27/2019 DPD Simulation

    9/16

    A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266 259

    We now have obtained the zeroth and first order solutions

    off,

    (71)f = f0 + f1 ,in which f0 and f

    1 are given in Eqs. (60) and (69), respec-

    tively.

    4.2. Viscosity due to momentum

    The viscosity due to momentum, K, can be obtainedfrom Eqs. (42) and (71) as

    K = n0

    UU

    f0 + f1

    dv

    = n0I +3

    2

    1

    2

    3/2 n0[w]

    UU

    UU 13

    U2I

    exp1

    2U2

    dU : 1

    2

    ru +

    rut

    (72)= n0I +3

    2

    1

    [w]

    ru +

    rut

    ,

    in which the following relationship has been used:

    UUUU 1

    3

    U2I :D exp

    1

    2

    U2dU

    (73)

    = (2 )3/2

    2D11 D12 + D21 D13 + D31

    D12 + D21 2D22 D23 + D32D13 + D31 D23 + D32 2D33

    .

    In this equation, D is defined by

    (74)D = 12

    ru +

    rut

    .

    Hence, by comparing Eq. (72) with the general expression

    of a stress tensor for Newtonian fluids,

    (75) = PI +

    ru +

    rut

    ,

    the viscosity due to momentum, K, can be expressed as

    (76)K = 32

    1

    [w] =45

    4

    1

    .

    The dimensional expression for this expression is written as

    (77)K

    =

    3

    2

    mkT

    r

    3

    c [w].

    4.3. Viscosity due to dissipative forces

    We derive an expression for the stress tensor D. Withthe following Taylor series expansion of the Dirac delta

    function,

    (r ri + eij)= (r ri ) + eij

    r(r ri )

    (78)+ 12!

    2eijeij :

    r

    r(r ri ) + ,

    the stress tensor due to dissipative forces can be written as

    D = 1

    2

    i

    j

    (i=j )

    wD(rij)(eij vij)eijrij(r ri )

    + 14

    ri j

    (i=j )

    wD(rij)

    (79) (eij vij)eijrijrij(r ri )

    + .

    The second term on the right-hand side in this equation is

    negligible unless there is a significant nonuniformity of the

    system. Equation (79) is, therefore, rewritten as

    D = 1

    2n20

    RwD(R)

    R (v v )RR(80) f(2)(r, r + R, v, v , t) dR dv dv .

    By the assumption of molecular chaos, written in Eq. (53),with Eq. (71) for f, the expression for f(2) can be writtenas

    f(2)(r, r , v, v , t) = f0 (r, v, t)f0 (r , v , t)+ f0 (r, v, t)f1 (r , v , t)+ f0 (r , v , t)f1 (r, v, t)

    (81)+ 2f1 (r, v, t)f1 (r , v , t).By evaluating Eq. (80) with Eq. (81), the expression for Dcan be obtained, which is now shown in the following.

    We first derive the expression for the stress D0 due to thefirst term on the right-hand side in Eq. (81). From Eq. (80),

    we get the following expression:

    D0 = 12 n

    20

    RwD(R)

    R (v v )RR(82) f0 (r, v, t)f0 (r , v , t) dR dv dv .

    If we take into account the relationship

    (83)

    v v = (v u) (v u ) R r

    u(r, t),

    then Eq. (82) can be rewritten as

    (84)D0 =1

    2n20

    wD(R)R2RRRR : D dR.

    The integral on the right-hand side in this equation can be

    carried out using the polar coordinate system to give the final

  • 7/27/2019 DPD Simulation

    10/16

    260 A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266

    expression for D0 ,

    (85)D0 =n20

    15[R2w]1

    2

    ru +

    rut

    ,

    in which

    [R2w] =

    R2wD(R) dR

    (86)=1

    0

    R2wD(R)4 R2 dR.

    Next, we have to evaluate the stresses due to the second,

    third, and fourth terms on the right-hand side in Eq. (81).

    From a similar procedure, it can straightforwardly be shown

    that all these stresses become zero.

    By taking into account all these results, the expression for

    D can easily be derived as

    (87)D = D0 .Hence, the comparison of Eq. (85) with Eq. (75) leads to

    the final expression for the viscosity D due to dissipativeforces:

    (88)D = n2030

    [R2w] = 21575

    n20 .

    The dimensional form is expressed as

    (89)D = n20

    30

    r 3c R

    2w

    .

    The viscosity is finally obtained as the sum ofK and D

    as = K + D in the present case.It is clearly seen from the above discussion that the per-

    turbation method with perturbation parameter using the

    nondimensional basic equation written in Eq. (55) can be

    significantly understandable in notifying the validity range

    of the analytical expression of the viscosity, Eq. (88) or

    Eq. (89), although the result essentially agrees with that

    which was derived by Marsh et al. [14].

    4.4. Relationship with HK theory

    In order to compare the present results with the HK the-

    ory [3,4] more straightforwardly, we rewrite Eq. (89) as

    (90)D = n20

    30

    r2

    1 rrc

    2dr = n

    20

    30

    r 2wD(r)dr.

    The viscosity due to dissipative forces, DHK, in the HK

    theory is written as

    (91)DHK =mn20

    30t

    r2wHK(r) dr,

    in which wHK(r) is a weight function in the HK theory

    and this weight function corresponds to wD(r) in the present

    theory. However, it should be noted that wHK(r) does not

    satisfy the fluctuationdissipation theory [12,13], which is,

    in constant, satisfied by the present theory. If we take into

    account that the notation in the HK theory is related to the

    notations here as = t/m, it is seen that D essentiallyagrees with DHK. This clearly shows that the expression in

    the HK theory, DHK, has a similar validity range to the

    present expression D, which has derived for the cases of

    small values of.

    5. Evaluation of viscosity by means of the

    nonequilibrium dynamics method

    5.1. Evaluation by GreenKubo expression

    There are two methods for evaluating transport coef-

    ficients by means of simulations: the first method is for

    thermodynamic equilibrium and evaluates them using the

    GreenKubo expression; the second one is the nonequilib-

    rium dynamics method, shown in the following, in whichtransport coefficients are evaluated under circumstances of a

    simple shear flow.

    The theory of the nonequilibrium molecular dynamics

    method [20] for a molecular system is applicable to the

    present case, if a simple shear flow is considered as a flow

    field. In this case, the viscosity yx is evaluated from the fol-

    lowing equation in simulations [18,20],

    (92)yx = 1

    VJyx ne,

    in which

    Jyx (t) =N

    i=1

    mviy (t)vix (t ) + yi (t)Fix (t )

    (93)=N

    i=1mviy (t)vix (t ) +

    Ni=1

    Nj=1

    (i

  • 7/27/2019 DPD Simulation

    11/16

    A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266 261

    cases for a molecular system. According to the nonequi-

    librium molecular dynamics method, the finite difference

    equations of the equation of motion for the present system

    are modified as [20]

    xi = vix + yi t, yi = viy t,(96)zi = viz t,

    vix =

    Fix viy

    t, viy = Fiy t,(97)viz = Fiz t,

    in which Fi denotes the force acting on particle i.

    5.2. Parameters for simulations

    Simulations have been conducted under various condi-

    tions in order to clarify the characteristics of the equation

    of motion for dissipative particles dynamics by comparing

    simulation results with theoretical solutions. Typical cases

    such as n0 = 0.5, 3.16, and 10.0 were taken for the numberdensity, and the particle number N was 500, unless specif-

    ically notified. From sufficient numbers of simulations for

    a small system, the time interval t was taken as t =0.0001, and the total time steps were adopted as 5,000,000,

    which ensures sufficient accuracy of average values. The er-

    ror analysis [20] concerning almost all simulation results

    was conducted to show the error ranges of simulation data;

    the error ranges of K and D are denoted by Kerror andDerror, respectively, and these values will be shown withresults in Section 6. Also, the parameters and , whichhave appeared in the equation of motion for dissipative parti-

    cles, were taken as = 10, 25 and = , /10, unlessspecifically noted.

    6. Results

    6.1. Influence of on viscosity

    Figs. 1 and 2 show the influence of the values of onthe viscosity, which were obtained by the nonequilibrium

    dynamics method. Fig. 1 is for the viscosity D due to dis-sipative forces and Fig. 2 is for the viscosity K due tomomentum.

    It is seen from Fig. 1 that the results obtained by thenonequilibrium dynamics simulations are in good agreement

    with the theoretical solutions shown in Eq. (88) for both

    cases of = /10 and = . However, for the casesof small number densities such as n0 = 0.5, the simula-tion results have significant errors and, therefore, qualitative

    properties are relatively difficult to identify. That the num-

    ber density n0 is smaller than unity means that a sufficientlylarge number of particles do not exist around an arbitrary

    particle within the range of the radius rc from its center. In

    this situation, the accuracy of the viscosity data is strongly

    dependent on whether or not there are other particles which

    interact with the particle of interest. Hence such a situation

    (a)

    (b)

    Fig. 1. Influence of on viscosity due to dissipative forces obtainedby nonequilibrium dynamics method (a) for = /10 and (b) for = (Derror = (5.2, 1.5) for n20 = (10,000, 1000), respec-tively, for n0

    =10,

    (1.5, 0.32) for n20

    =(1000, 100), respectively,

    for n0 = 3.16, and (0.049, 0.022) for n20 = (25, 2.5), respectively, forn0 = 0.5, in (a); similar errors are included in (b)).

    causes significant errors of the simulation results of viscos-

    ity, which has been seen for n0 = 0.5 in Fig. 1. We mayconclude from this fact that it is desirable to take the number

    density as n0 1.0 in simulating a flow problem by meansof the dissipative particle dynamics method.

    Fig. 2 clearly shows that the simulation data are larger

    than the theoretical solutions, and this tendency becomes

    more significant for larger values of . When the influenceof dissipative forces is of the same order of the momentum,

    that is, when 1, the simulation results of K agreewell with the theoretical solutions. On the other hand, at

    1, where dissipative forces are more dominant thanthe momentum, the discrepancy between the simulation re-

    sults and the analytical solutions has a tendency to increase

    with values of . As will be shown later in Fig. 8, the ra-dial distribution function does not significantly change for

    values of, so that this discrepancy is presumed to be dueto the mechanism other than the assumption of molecular

    chaos, adopted here. Some researchers attempted to explain

    the discrepancy between simulation results and theoretical

    solutions concerning the viscosity due to the momentum, in

    terms of the pair collision theory [21].

  • 7/27/2019 DPD Simulation

    12/16

    262 A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266

    (a)

    (b)

    Fig. 2. Influence of on viscosity due to momentum obtained by non-equilibrium dynamics method (a) for = /10 and (b) for = (Kerror = (0.57, 1.47) for 1/ = (0.01, 0.1), respectively, for n0 = 10,(0.40, 0.62) for 1/ = (0.01, 0.1), respectively, for n0 = 3.16, and(0.44, 0.12) for 1/ = (0.01, 0.1), respectively, for n0 = 0.5, in (a);similar errors are included in (b)).

    6.2. Influence of and number density on viscosity

    Fig. 3 shows the influence of on the viscosity due todissipative forces. It is seen from Fig. 3 that the values of

    do not have significant influence on the viscosity for bothcases ofn0 = 3.16 and 10.0. As will be shown later in Fig. 9,the radial distribution function changes significantly for dif-

    ferent values of. Hence, this means that the viscosity dueto dissipative forces, D, is not strongly dependent on theshape of the radial distribution function.

    Fig. 4 shows the dependence of the viscosity on the num-

    ber density of particles: Fig. 4a is for the viscosity D dueto dissipative forces, and Fig. 4b is for the viscosity Kdue to momentum. Since the theoretical solutions are valid

    for n0 1.0, the simulation results are in good agree-

    ment with the analytical solutions for the values of n0 whichsatisfy the condition n0

    1.0 for a given value of .This agreement becomes worse with decreasing values ofn0smaller than unity; especially, the results for = 10 and = 1 significantly deviate from the theoretical values withlarge error ranges. As already pointed out, this is presumed

    to be due to the situation in which the ambient particles of an

    Fig. 3. Influence of on viscosity due to dissipative forces (Derror =(1.2, 3.1) for = (100, 1), respectively, in the case of n0 = 10 and = 25, and (0.17, 0.32) for = (40, 1), respectively, in the case ofn0 = 3.16 and = 10).

    (a)

    (b)

    Fig. 4. Influence of number density on viscosity: (a) for viscosity due

    to dissipative forces and (b) for viscosity due to momentum (Derror =(2.7, 0.12) for n0 = (10, 1), respectively, in the case of = 25 and = 2.5, and (0.78, 0.022) for n0 = (10, 0.5), respectively, in the caseof = 10 and = 1, in (a); Kerror = (1.1, 0.24) for n0 = (10, 1),respectively, in the case of = 25 and = 2.5, and (1.0, 0.12) forn0 = (10, 0.5), respectively, in the case of = 10, and = 1, in (b)).

    arbitrary particle come to be less for decreasing the number

    density smaller than unity, which leads to the large fluctu-

    ation in the contribution of dissipative particles to physical

  • 7/27/2019 DPD Simulation

    13/16

    A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266 263

    quantities such as transport coefficients with time. Although

    the viscosity K due to momentum shown in Fig. 4b has alarge error range, it is clearly seen that there is a significant

    discrepancy between the simulation and theoretical values,

    and this tendency does not strongly depend on the values

    ofn0. As already pointed out in Fig. 2, since the radial distri-bution function does not significantly deviate from g(r) = 1(shown later in Fig. 10a), it is presumed that this discrepancy

    for = 10 and = 1 is induced by the mechanism otherthan the invalidity of the molecular chaos assumption.

    6.3. Influence of time interval on mean velocity of particles

    and viscosity

    Fig. 5 shows the influence of the time interval h (=t), which is used in conducting simulations based on thefinite difference equations, on the mean velocity of particles

    and the viscosity. It is seen from Fig. 5a that the viscos-

    ity D does not strongly depend on the time interval forh 0.1. On the other hand, as shown in Fig. 5b, the vis-cosity due to momentum, K, is significantly dependenton the values of h. Especially, this dependence comes toappear from h 0.001, and values of K significantlyincrease with increasing values of the time interval. A sim-

    ilar tendency can be observed in Fig. 5c, in which the mean

    velocity of particles significantly increases from h 0.01with the values of h. Since the mean velocity v is nondi-mensionalized in such a way of being v = 1, this conditionhas to be satisfied in actual simulations. We may, therefore,

    conclude that the simulations have to be conducted using a

    sufficiently small time interval such as h = 0.0001, adoptedhere.

    6.4. Influence of shear rate and number of particles on

    viscosity

    Fig. 6 shows the influence of the shear rate of a simple

    shear flow and the number of particles on the viscosity. It

    is seen that the simulation results come to have significant

    errors with decreasing shear rate, since just a weak simple

    shear flow is generated for such situations. This character-

    istic is observed for both cases of D in Fig. 6a and K

    in Fig. 6b, especially in the results of K. It is seen fromFig. 6c that the mean velocity v significantly increases withthe shear rate after 1. Hence, there is an appropriaterange for the shear rate , which is sufficiently small forsatisfying the condition of v = 1, and sufficiently large forgenerating a simple shear flow with an appropriate shear

    strength. By taking account of these constraints, we here

    adopted = 0.01 in conducting simulations.Fig. 7 shows the influence of the number of particles on

    the viscosity D due to dissipative forces. It is seen fromFig. 7 that a larger system is required for a smaller num-

    ber density to remove the dependence of the results on the

    number of particles. This means that the interactions of the

    (a)

    (b)

    (c)

    Fig. 5. Influence of time interval on (a) viscosity due to dissipative forces,

    (b) viscosity due to momentum, and (c) average velocity of particles

    (Derror = (0.43, 2.7) for h = (0.01, 0.0001), respectively, in the caseof n0 = 10, = 25, and = 2.5 (Case A), and (0.051, 0.31) forh = (0.1, 0.0001), respectively, in the case of n0 = 3.16, = 10, and = 1 (Case B), in (a); Kerror = (0.34, 1.1) for h = (0.01, 0.0001),respectively, in Case A, and (1.1, 0.62) for h = (0.1, 0.0001), respec-tively, in Case B, in (b)).

    particle of interest with the ambient particles come to de-

    pend on the dimensions of the system, unless a sufficiently

    large system is used for the cases of small number densities.

    As already pointed out, the number density has to be taken

    as sufficiently large, and a large system such as N = 500,adopted here, at least has to be adopted to remove the de-

    pendence of results on the system size.

  • 7/27/2019 DPD Simulation

    14/16

    264 A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266

    (a)

    (b)

    (c)

    Fig. 6. Influence of shear rate on (a) viscosity due to dissipative forces,

    (b) viscosity due to momentum, and (c) average velocity of particles

    (Derror = (0.055, 2.7) for = (0.5, 0.01), respectively, in Case A,and (0.0064, 0.32) for = (0.5, 0.01), respectively, in Case B, in (a);Kerror = (0.028, 1.1) for = (0.5, 0.01), respectively, in Case A, and

    (0.017, 0.62) for

    =(0.5, 0.01), respectively, in Case B, in (b)).

    6.5. Influence of, , and number density on radialdistribution function

    Finally, we show the results of the radial distribution

    function [18] in Figs. 810 in order to clarify the quantitative

    dependence of the internal structure on , , and the num-ber density of particles. Figs. 8, 9, and 10 show the influence

    of, , and the number density, respectively.It is seen from Fig. 8 that the radial distribution function

    is almost independent of the values of , and the curve for

    = 10 agrees well with that for = 100 for both cases

    Fig. 7. Influence of number of particles on viscosity due to dissipative forces

    (Derror = (1.7, 6.1) for N = (1372, 108), respectively, in Case A, and(0.19, 0.68) for N= (1372, 108), respectively, in Case B).

    (a)

    (b)

    Fig. 8. Dependence of radial distribution function on for = 10 (a) forn0 = 3.16 and (b) for n0 = 10.

    ofn0 = 3.16 and 10.0. Although the theoretical solutions ofthe viscosity have been derived under circumstances of ne-

    glecting the internal structure of the system, that is, under

    the molecular chaos assumption, it is very interesting that

    the theoretical solutions agree well with the simulation re-

    sults (Fig. 1), irrespective of the system being in a gas-like

    or liquid-like internal structure, as shown in Fig. 8.

    Fig. 9 clearly shows that the radial distribution function

    strongly depends on the values of . A more significant

    repulsive force acts between particles with increasing the

  • 7/27/2019 DPD Simulation

    15/16

    A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266 265

    (a)

    (b)

    Fig. 9. Dependence of radial distribution function on (a) for n0 = 3.16and = 10, and (b) for n0 = 10 and = 25.

    values of , so that the radial distribution function shows aliquid-like structure for = 25 and = 40. That the valueof is much larger than that ofmeans that the influenceof dissipative and random forces is much smaller than that

    of repulsions due to conservative forces. This means that the

    merits of the dissipative particle dynamics method cannot

    function effectively under such circumstances. It is, there-

    fore, presumed that has to be taken as < to makethe dissipative particle dynamics method effective. It is seen

    from Fig. 9a that the molecular chaos assumption is satisfied

    to a certain degree for = 0.4. To the contrary, the resultsfor = 4 and 25 cannot sufficiently satisfy this assumption.Nevertheless, as already shown in Fig. 3, the viscosity Ddue to dissipative forces is almost independent of the values

    of . We may conclude form these results that dissipativeforces do not strongly depend on the internal structure of the

    system.

    It is seen from Fig. 10 that the radial distribution function

    is not strongly dependent on the number density. However,

    the curve for the case of n0 = 6.0, in which the particle ofinterest can interact with a sufficient number of ambient par-

    ticles, slightly deviates from the results for n0 = 0.2 and 1.0,in which there are not sufficient particles around the particle

    of interest.

    (a)

    (b)

    Fig. 10. Dependence of radial distribution function on number density of

    particles (a) for = 10 and = 1, and (b) for = 25 and = 2.5.

    Fig. 11. Results of radial distribution function obtained by Espanol et al.

    [22].

    Finally, we compare the present results of the radial distri-

    bution function with other simulation results obtained by Es-

    panol et al. [22]. They investigated the relationship between

    the dissipative particle dynamics and the smoothed particle

    dynamics methods. In the concrete, the behavior of the clus-

    ters of particles, which correspond to dissipative particles

    in the present study, was investigated by means of conduct-

    ing equilibrium molecular dynamics simulations of a system

    composed of particles with repulsive interactions. Fig. 11

  • 7/27/2019 DPD Simulation

    16/16

    266 A. Satoh, T. Majima / Journal of Colloid and Interface Science 283 (2005) 251266

    shows one of the results concerning the radial distribution

    function, which were obtained by Espanol et al. [22]. From

    the potential curves of the mean force in their paper [22], it

    is seen that R in their notation roughly corresponds to 3rin the present one, and that lower curves correspond to re-

    sults for a larger value of in the present study. We clearlysee from comparing Fig. 9 (also Fig. 10) with Fig. 11 that theshape of the curves are quite similar between the present and

    their results, in which a small increase at very short distances

    can be observed more significantly with decreasing values

    of. A system with larger values of approaches a usualsystem such as a molecular one, so that dissipative parti-

    cles cannot penetrate each other due to a high-energy barrier.

    This leads to the fact that the curves of the radial distribution

    function comes to have an ordinary shape for a liquid or gas,

    which has no small increase within a short distance range.

    7. Conclusions

    In order to investigate the validity of the dissipative par-

    ticle dynamics method, which is a mesoscopic simulation

    technique, we have derived expressions for transport coef-

    ficients such as viscosity, from the equation of motion of

    the dissipative particles. In the concrete, we have shown the

    FokkerPlanck equation in phase space, and macroscopic

    conservation equations such as the equation of continuity

    and the equation of momentum conservation. The basic

    equations of the single-particle and pair distribution func-

    tions have been derived using the FokkerPlanck equation.

    The solutions of these distribution functions have approx-

    imately been solved by the perturbation method under the

    assumption of molecular chaos. The expressions of the vis-

    cosity due to momentum and dissipative forces have been

    obtained using the approximate solutions of the distribution

    functions. Also, we have conducted nonequilibrium dynam-

    ics simulations to investigate the influence of the parame-

    ters, which have appeared in defining the equation of motion

    in the dissipative particle dynamics method. The theoreti-

    cal values of the viscosity due to dissipative forces in the

    HK theory are in good agreement with the simulation re-

    sults obtained by the nonequilibrium dynamics method, ex-

    cept in the range of small numberdensities. There are restric-

    tion conditions for taking appropriate values of the numberdensity, number of particles, time interval, shear rate, etc., to

    obtain physically reasonable results by means of dissipative

    particle dynamics simulations.

    References

    [1] P. Mazur, W. van Saarloos, Physica A 115 (1982) 21.

    [2] W. van Saarloos, P. Mazur, Physica A 120 (1983) 77.

    [3] P.J. Hoogerbrugge, J.M.V.A. Koelman, Europhys. Lett. 19 (1992) 155.

    [4] J.M.V.A. Koelman, P.J. Hoogerbrugge, Europhys. Lett. 21 (1993) 263.

    [5] G. Bossis, J.F. Brady, J. Chem. Phys. 80 (1984) 5141.

    [6] J.F. Brady, G. Bossis, J. Fluid Mech. 155 (1985) 105.

    [7] L. Durlofsky, J.F. Brady, G. Bossis, J. Fluid Mech. 180 (1987) 21.

    [8] A. Satoh, R.W. Chantrell, G.N. Coverdale, S. Kamiyama, J. Colloid

    Interface Sci. 203 (1998) 233.

    [9] A. Satoh, G.N. Coverdale, R.W.Chantrell, J. Colloid Interface Sci. 231

    (2000) 238.

    [10] A. Satoh, J. Colloid Interface Sci. 243 (2001) 342.

    [11] A. Satoh, J. Colloid Interface Sci. 255 (2002) 98.

    [12] P. Espanol, P. Warren, Europhys. Lett. 30 (1995) 191.

    [13] P. Espanol, Phys. Rev. E 52 (1995) 1734.

    [14] C.A. Marsh, G. Backx, M.H. Ernst, Phys. Rev. E 56 (1997) 1676.

    [15] E.S. Boek, P.V. Coveney, H.N.W. Lekkerkerker, P. van Schoot, Phys.

    Rev. E 55 (1997) 3124.

    [16] P. Espanol, Europhys. Lett. 40 (1997) 631.

    [17] J.B. Avalos, A.D. Mackie, Europhys. Lett. 40 (1997) 141.

    [18] A. Satoh, Studies in Interface Science: Introduction to Molecular-Microsimulation of Colloidal Dispersions, Elsevier Science, Amster-

    dam, 2003.

    [19] K.E. Novik, P.V. Coveney, J. Chem. Phys. 109 (1998) 7667.

    [20] M.P. Allen, D.J. Tildesley, Computer Simulation of Liquids, Claren-

    don, Oxford, 1987.

    [21] A.J. Masters, P.B. Waeewn, Europhys. Lett. 48 (1999) 1.

    [22] P. Espanol, M. Serrano, I. Zuniga, Int. J. Mod. Phys. C 8 (1997) 899.