Sie sind auf Seite 1von 19

PHYSICAL REVIEW E 85, 056705 (2012)

Control-volume representation of molecular dynamics


E. R. Smith, D. M. Heyes, D. Dini and T. A. Zaki
Department of Mechanical Engineering, Imperial College London,
Exhibition Road, London SW7 2AZ, United Kingdom∗
(Dated: Received 13 October 2011; revised manuscript received 2 March 2012; published 22 May 2012)
A Molecular Dynamics (MD) parallel to the Control Volume (CV) formulation of fluid mechanics is devel-
oped by integrating the formulas of Irving and Kirkwood, J. Chem. Phys. 18, 817 (1950) over a finite cubic
volume of molecular dimensions. The Lagrangian molecular system is expressed in terms of an Eulerian CV,
which yields an equivalent to Reynolds’ Transport Theorem for the discrete system. This approach casts the dy-
namics of the molecular system into a form that can be readily compared to the continuum equations. The MD
equations of motion are reinterpreted in terms of a Lagrangian-to-Control-Volume (LCV) conversion function
ϑi , for each molecule i. The LCV function and its spatial derivatives are used to express fluxes and relevant
arXiv:1203.2453v2 [math-ph] 24 May 2012

forces across the control surfaces. The relationship between the local pressures computed using the Volume
Average (VA, Lutsko, J. Appl. Phys 64, 1152 (1988) ) techniques and the Method of Planes (MOP , Todd et al,
Phys. Rev. E 52, 1627 (1995) ) emerges naturally from the treatment. Numerical experiments using the MD CV
method are reported for equilibrium and non-equilibrium (start-up Couette flow) model liquids, which demon-
strate the advantages of the formulation. The CV formulation of the MD is shown to be exactly conservative,
and is therefore ideally suited to obtain macroscopic properties from a discrete system.

DOI: 10.1103/PhysRevE.85.056705 PACS number(s): 05.20.y, 47.11.Mn, 31.15.xv

I. INTRODUCTION tion to the differential form of continuum fluid mechanics.


However, the resulting equations at a point were expressed
The macroscopic and microscopic descriptions of mechan- in terms of the Dirac δ function — a form which is difficult
ics have traditionally been studied independently. The former to manipulate and cannot be applied directly in a molecular
invokes a continuum assumption, and aims to reproduce the simulation. Furthermore, a Taylor series expansion of the
large-scale behaviour of solids and fluids, without the need to Dirac δ functions was required to express the pressure ten-
resolve the micro-scale details. On the other hand, molecu- sor. The final expression for pressure tensor is neither easy
lar simulation predicts the evolution of individual, but inter- to interpret nor to compute [9]. As a result, there have been
acting, molecules, which has application in nano and micro- numerous attempts to develop an expression for the pressure
scale systems. Bridging these scales requires a mesoscopic tensor for use in MD simulation [9–21]. Some of these ex-
description, which represents the evolution of the average of pressions have been shown to be equivalent in the appropriate
many microscopic trajectories through phase space. It is ad- limit. For example, Heyes et al. [22]) demonstrated equiva-
vantageous to cast the fluid dynamics equations in a consis- lence between Method of Planes (MOP Todd et al. [13]) and
tent form for both the molecular, mesoscale and continuum Volume Average (VA Lutsko [16]) at a surface.
approaches. The current works seeks to achieve this objec- In order to avoid use of the Dirac δ function, the current
tive by introducing a Control Volume (CV) formulation for work adopts a Control Volume representation of the MD sys-
the molecular system. tem, written in terms of fluxes and surface stresses. This ap-
The Control Volume approach is widely adopted in con- proach is in part motivated by the success of the control vol-
tinuum fluid mechanics, where Reynolds Transport Theorem ume formulation in continuum fluid mechanics. At a molecu-
[1] relates Newton’s laws of motion for macroscopic fluid lar scale, control volume analyses of NEMD simulations can
parcels to fluxes through a CV. In this form, fluid mechanics facilitate evaluation of local fluid properties. Furthermore,
has had great success in simulating both fundamental [2, 3] the CV method also lends itself to coupling schemes between
and practical [4–6] flows. However, when the continuum as- the continuum and molecular descriptions [23–34].
sumption fails, or when macroscopic constitutive equations The equations of continuum fluid mechanics are presented
are lacking, a molecular-scale description is required. Exam- in Section II A, followed by a review of the Irving and Kirk-
ples include nano-flows, moving contact lines, solid-liquid wood [8] procedure for linking continuum and mesoscopic
boundaries, non-equilibrium fluids, and evaluation of trans- properties in Section II B. In section III, a Lagrangian to Con-
port properties such as viscosity and heat conductivity [7]. trol Volume (LCV) conversion function is used to express the
Molecular Dynamics (MD) involves solving Newton’s mesoscopic equations for mass and momentum fluxes. Sec-
equations of motion for an assembly of interacting discrete tion III C focuses on the stress tensor, and relates the cur-
molecules. Averaging is required in order to compute proper- rent formulation to established definitions within the litera-
ties of interest, e.g. temperature, density, pressure and stress, ture [13, 16, 17]. In Section IV, the CV equations are derived
which can vary on a local scale especially out of equilib- for a single microscopic system, and subsequently integrated
rium [7]. A rigorous link between mesoscopic and continuum in time in order to obtain a form which can be applied in MD
properties was established in the seminal work of Irving and simulations. The conservation properties of the CV formula-
Kirkwood [8], who related the mesoscopic Liouville equa- tion are demonstrated in NEMD simulations of Couette flow
in Section IV C.

∗ edward.smith05@imperial.ac.uk; d.heyes@imperial.ac.uk; d.dini@imperial.ac.uk; t.zaki@imperial.ac.uk;

1539-3755/2012/85(5)/056705(19) 056705-1
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

II. BACKGROUND and ri , respectively, and N is the number of molecules in a


single system. The momentum density at a point in space is
This section summarizes the theoretical background. First, similarly defined by,
the macroscopic continuum equations are introduced, fol-
lowed by the mesoscopic equations which describe the evolu- N  
X
tion of an ensemble average of systems of discrete molecules. ρ(r, t)u(r, t) ≡ pi δ(ri − r); f , (8)
The link between the two descriptions is subsequently dis- i=1
cussed.
where the molecular momentum, pi = mi ṙi . Note that pi is
A. Macroscopic Continuum Equations
the momentum in the laboratory frame, and not the peculiar
The continuum conservation of mass and momentum bal- value pi which excludes the macroscopic streaming term at
ance can be derived in an Eulerian frame by considering the the location of molecule i, u(ri ), [7],
fluxes through a Control Volume (CV). The mass continuity  
equation can be expressed as, pi
pi ≡ mi − u(ri ) . (9)
Z I mi

ρdV = − ρu · dS, (1)
∂t V S The present treatment uses pi in the lab frame. A discussion
of translating CV and its relationship to the peculiar momen-
where ρ is the mass density and u is the fluid velocity. The
tum is given in Appendix A.
rate of change of momentum is determined by the balance of
Finally, the energy density at a point in space is defined by
forces on the CV,
Z I N 
∂ X 
ρudV = − ρuu · dS + Fsurface + Fbody . (2) ρ(r, t)E(r, t) ≡ ei δ(ri − r); f , (10)
∂t V S
i=1
The forces are split into ones which act on the bounding sur-
faces, Fsurface , and body forces, Fbody . Surface forces are where the energy of the ith molecule is defined as the sum of
expressed in terms the pressure tensor, Π, on the CV sur- the kinetic energy and the inter-molecular interaction poten-
faces, tial φij ,
I
Fsurface = − Π · dS. (3) p2i 1X
N
S ei ≡ + φij (11)
2mi 2
The rate of change of energy in a CV is expressed in terms of j6=i
fluxes, the pressure tensor and a heat flux vector q,
Z I It is implicit in this definition that the potential energy of an
∂ interatomic interaction, φij , is divided equally between the
ρEdV = − [ρEu + Π · u + q] · dS, (4)
∂t V S two interacting molecules, i and j.
As phase space is bounded, the evolution of a property, α,
here the energy change due to body forces is not included. in time is governed by the equation,
The divergence theorem relates surface fluxes to the diver-
gence within the volume, for a variable A,   X N  
∂ ∂α pi ∂α
I Z α; f = Fi · + · ;f , (12)
A · dS = ∇ · AdV (5) ∂t ∂pi mi ∂ri
i=1
S V

In addition, the differential form of the flow equations can be where Fi is the force on molecule i, and α = α(ri (t), pi (t))
recovered in the limit of an infinitesimal control volume [35], is an implicit function of time. Using Eq. (12), Irving and
I Kirkwood [8] derived the time evolution of the mass (from
1 Eq. 7), momentum density (from Eq. 8) and energy density
∇ · A = lim A · dS. (6)
V →0 V S (from Eq. 10) for a mesoscopic system. A comparison of the
resulting equations to the continuum counterpart provided a
B. Relationship Between the Continuum and the Mesoscopic term-by-term equivalence. Both the mesoscopic and contin-
Descriptions uum equations were valid at a point; the former expressed in
A mesoscopic description is a temporal and spatial average terms of Dirac δ and the latter in differential form. In the
of the molecular trajectories, expressed in terms of a proba- current work, the mass and momentum densities are recast
bility function, f. Irving and Kirkwood [8] established the within the CV framework which avoids use of the Dirac δ
link between the mesoscopic and continuum descriptions us- functions directly, and attendant problems with their practi-
ing the Dirac δ function to define the macroscopic density at cal implementation.
a point r in space,
III. THE CONTROL VOLUME FORMULATION
N 
X 
ρ(r, t) ≡ mi δ(ri − r); f . (7) In order to cast the governing equations for a discrete
i=1
system in CV form, a ‘selection function’ ϑi is introduced,
which isolates those molecules within the region of interest.
The angled brackets hα; f i denote the inner product of α with This function is obtained by integrating the Dirac δ func-
f, which gives the expectation of α for an ensemble of sys- tion, δ(ri − r), over a cuboid in space, centered at r and
tems. The mass and position of a molecule i are denoted mi of side length ∆r as illustrated in figure 1(a) [37]. Using

056705-2
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

FIG. 1. (Color online) The CV function and its derivative applied to a system of molecules. The figures were generated using the VMD
visualization package, [36]. From left to right, (a) Schematic of ϑi which selects only the molecules within a cube, (b) Location of cube
center r and labels for cube surfaces, (c) Schematic of ∂ϑi /∂x which selects only molecules crossing the x+ and x− surface planes.


δ(ri − r) = δ(xi − x)δ(yi − y)δ(zi − z), the resulting triple and similarly for the left surface, dSxi , the total flux Eq. (14)
integral is, in any direction r is then,
∂ϑi −
+
Zx+ Zy Zz + = dS+ i − dSi ≡ dSi . (17)
∂r
ϑi ≡ δ(xi − x)δ(yi − y)δ(zi − z)dxdydz The LCV function is key to the derivation of a molecular-
x− y − z − level equivalent of the continuum CV equations, and it will
 x+ y+ z + be used extensively in the following sections. The approach
= H(xi − x)H(yi − y)H(zi − z) in sections III A, III B and III D shares some similarities with
x− y − z − the work of Serrano and Español [38] which considers the
  time evolution of Voronoi characteristic functions. However
= H(x+ − xi ) − H(x− − xi )
  the LCV function has precisely defined extents which allows
× H(y + − yi ) − H(y − − yi ) the development of conservation equations for a microscopic
  system. In the following treatment, the CV is fixed in space
× H(z + − zi ) − H(z − − zi ) ,
(13) (i.e., r is not a function of time). The extension of this treat-
ment to an advecting CV is made in Appendix A.
where H is the Heaviside function, and the limits of integra- A. Mass Conservation for a Molecular CV
tion are defined as, r− ≡ r − ∆r + ∆r
2 and r ≡ r + 2 , for each In this section, a mesoscopic expression for the mass in a
direction (see Fig. 1(b)). Note that ϑi can be interpreted as cuboidal CV is derived. The time evolution of mass within
a Lagrangian-to-Control-Volume conversion function (LCV) a CV is shown to be equal to the net mass flux of molecules
f̃or molecule i. It is unity when molecule i is inside the across its surfaces.
cuboid, and equal to zero otherwise, as illustrated in Fig. The mass inside an arbitrary CV at the molecular scale can
1(a). Using L’Hôpital’s rule and defining, ∆V ≡ ∆x∆y∆z, be expressed in terms of the LCV as follows,
the LCV function for molecule i reduces to the Dirac δ func- Z Z X N  
tion in the limit of zero volume,
ρ(r, t)dV = mi δ(ri − r); f dV
V V i=1
ϑi
δ(r − ri ) = lim . +
∆V →0 ∆V x+Z
N Z y Zz + 
X
The spatial derivative in the x direction of the LCV function = mi δ(ri − r); f dxdydz
for molecule i is, i=1 − − −
x y z
N  
∂ϑi ∂ϑ   X
= − i = δ(x+ − xi ) − δ(x− − xi ) Sxi , (14) = mi ϑi ; f . (18)
∂x ∂xi
i=1
where Sxi is Taking the time derivative of Eq. (18) and using Eq. (12),
  Z N  
Sxi ≡ H(y + − yi ) − H(y − − yi ) ∂ ∂ X
  ρ(r, t)dV = mi ϑi ; f
H(z + − zi ) − H(z − − zi ) . (15) ∂t V ∂t
i=1
N  
Eq. (14) isolates molecules on a 2D rectangular patch in the
X p i ∂ ∂
= · mi ϑi + Fi · mi ϑi ; f . (19)
yz plane. The derivative ∂ϑi /∂x is only non-zero when mi ∂ri ∂pi
i=1
molecule i is crossing the surfaces marked in Fig. 1(c), nor-
mal to the x direction. The contribution of the ith molecule The term ∂mi ϑi /∂pi = 0, as ϑi is not a function of pi .
to the net rate of mass flux through the control surface is ex- Therefore,
pressed in the form, pi · dSi . Defining for the right x surface, Z N  
∂ X ∂ϑi
ρdV = − pi · ;f , (20)
+
dSxi ≡ δ(x+ − xi )Sxi , (16) ∂t V ∂r
i=1

056705-3
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

where the equality, ∂ϑi /∂ri = −∂ϑi /∂r has been used. where the continuum expression {ρuux }+ is the average flux
From the continuum mass conservation given in Eq. (1), the through a flat region in space with area ∆A+x = ∆y∆z. This
macroscopic and mesoscopic fluxes over the surfaces can be kinetic component of the pressure tensor is discussed further
equated, in Section III C.
The configurational term of Eq. (24) is,
6 Z
X N 
X 
ρu · dSf = pi · dSi ; f . (21) N 
X ∂
 XN  
f aces Sf i=1 CT = Fi · pi ϑi ; f = Fi ϑi ; f , (27)
∂pi
i=1 i=1
The mesoscopic equation for evolution of mass in a control
volume is given by, where the total force Fi on particle i is the sum of pairwise-
additive interactions with potential φij , and from an external
N   N   potential ψi .
∂ X X
mi ϑi ; f = − pi · dSi ; f . (22)
∂t 
N

i=1 i=1 ∂ X
ϑi Fi = −ϑi φij + ψi  .
Appendix B shows that the surface mass flux yields the Irving ∂ri
j6=i
and Kirkwood [8] expression for divergence as the CV tends
to a point (i.e. V → 0), in analogy to Eq. (6). It is commonly assumed that the potential energy of an inter-
atomic interaction, φij , can be divided equally between the
B. Momentum Balance for a Molecular CV
two interacting molecules, i and j, such that,
In this section, a mesoscopic expression for time evolution
of momentum within a CV is derived. The starting point is to N N  
integrate the momentum at a point, given in Eq. (8), over the
X ∂φij 1X ∂φij ∂φji
ϑi = ϑi + ϑj , (28)
CV, ∂ri 2 ∂ri ∂rj
i,j i,j
Z N   P PN PN
where the notation N
X
ρ(r, t)u(r, t)dV = pi ϑi ; f . (23) i,j = i=1 j6=i has been introduced
V i=1
for conciseness. Therefore, the configurational term can be
expressed as,
Following a similar procedure to that in section III A, the for-
mula (12) is used to obtain the time evolution of the momen- N   XN  
1X
tum within the CV, CT = fij ϑij ; f + fiext ϑi ; f , (29)
2
i,j i=1
Z N  
∂ ∂ X
ρ(r, t)u(r, t)dV = pi ϑi ; f where fij = −∂φij /∂ri = ∂φji /∂rj and fiext =
∂t V ∂t −∂ψi /∂ri . The notation, ϑij ≡ ϑi − ϑj , is introduced, which
i=1
N   is non-zero only when the force acts over the surface of the
X pi ∂ ∂ CV, as illustrated in Fig. 2.
= · pi ϑi + Fi · pi ϑi ; f , (24)
m ∂r ∂pi
i=1 | i {z i } | {z }
KT CT

where the terms KT and CT are the kinetic and configura-


tional components, respectively. The kinetic part is,
N   XN  
X p i ∂ pi pi ∂ϑi
KT = · pi ϑi ; f = · ;f ,
mi ∂ri mi ∂ri
i=1 i=1
(25)

where pi pi is the dyadic product. For any surface of the CV,


here x+ , the molecular flux can be equated to the continuum
convection and pressure on that surface,
Z
ρ(x+ , y, z, t)u(x+ , y, z, t)ux (x+ , y, z, t)dydz
+
Sx
Z N  
X pi pix +
+ K+ dydz = dSxi ; f ,
+ x
Sx mi
i=1

where K+x is the kinetic part of the pressure tensor due to


molecular transgressions across the x+ CV surface. The av-
erage molecular flux across the surface is then,
FIG. 2. (Color online) A section through the CV to illustrate the
N   role of ϑij in selecting only the i and j interactions that cross the
1 X pi pix + bounding surface of the control volume. Due to the limited range of
{ρuux}+ + K+
x = dSxi ; f , (26)
∆A+x i=1
mi interactions, only the forces between the internal (red) molecules i
and external (blue) molecules j near the surfaces are included.

056705-4
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

Substituting the kinetic (KT ) and configurational (CT ) engineering literature, the stress and pressure tensors have
terms, from Eqs. (25) and (29) into Eq. (24), the time evo- opposite signs,
lution of momentum within the CV at the mesoscopic scale
is, Π = κ − σ. (33)

N   N   The separation into kinetic and configurational parts is made


∂ X X pi pi to accommodate the debate concerning the inclusion of ki-
pi ϑi ; f = − · dSi ; f
∂t mi netic terms in the molecular stress [9, 39, 40].
i=1 i=1
N   XN   In order to avoid confusion, the stress, σ, is herein de-
1 X
fined to be due to the forces only (surface tractions). This,
+ fij ϑij ; f + fiext ϑi ; f . (30)
2 combined with the kinetic pressure term κ, yields the total
i,j i=1
pressure tensor Π first introduced in Eq. (3).
Equations (22) and (30) describe the evolution of mass and
momentum respectively within a CV averaged over an en- 1. Irving Kirkwood Pressure Tensor
semble of representative molecular systems. As proposed by
Evans and Morriss [7], it is possible to develop microscopic The virial expression for the stress cannot be applied lo-
evolution equations that do not require ensemble averaging. cally as it is only valid for a homogeneous system, [12]. The
Hence, the equivalents of Eqs. (22) and (30) are derived for Irving and Kirkwood [8] technique for evaluating the non-
a single trajectory through phase space in section IV A, inte- equilibrium, locally-defined stress resolves this issue, and is
grated in time in section IV B and tested numerically using herein extended to a CV. To obtain the stress, σ, the inter-
molecular dynamics simulation in section IV C. molecular force term of Eq. (31) is defined to be equal to the
The link between the macroscopic and mesoscopic treat- divergence of stress,
ments is given by equating their respective momentum Eqs. Z N  
(2) and (30), ∂ 1X
· σdV ≡ fij ϑij ; f
I V ∂r 2
i,j
− ρuu · dS + Fsurface + Fbody
S N Z  
1 X  
N   = fij δ(ri − r) − δ(rj − r) ; f dV. (34)
X pi pi 2 V
=− · dSi ; f i,j
mi
i=1
Irving and Kirkwood [8] used a Taylor expansion of the Dirac
N   XN  
1X δ functions to express the pair force contribution in the form
+ fij ϑij ; f + fiext ϑi ; f . (31) of a divergence,
2
i,j i=1
  ∂
As can be seen, each term in the continuum evolution of mo- fij δ(ri − r) − δ(rj − r) = − · fij rij Oij δ(ri − r),
∂r
mentum has an equivalent term in the mesoscopic formula-
tion. where rij = ri − rj , and Oij is an operator which acts on the
The continuum momentum Eq. (2) can be expressed in Dirac δ function,
terms of the divergence of the pressure tensor, Π, in the con-   !
trol volume from, 1 ∂ 1 ∂ n−1
Oij ≡ 1 − rij + ...− r + ... .
Z I 2 ∂ri n! ij ∂ri

ρudV = − [ρuu + Π] · dS + Fbody (32a) (35)
∂t V S
Z
∂ Equation (34) can therefore be rewritten,
=− · [ρuu + Π] dV + Fbody . (32b)
V ∂r Z N Z 
∂ 1X ∂
In the following subsection, the right hand side of Eq. (31) · σdV = − · fij rij
V ∂r 2 V ∂r
is recast first in divergence form as in Eq. (32b), and then in i,j

terms of surface pressures as in Eq. (32a).
Oij δ(ri − r); f dV. (36)
C. The Pressure Tensor

The average molecular pressure tensor ascribed to a con- The Taylor expansion in Dirac δ functions is not straightfor-
trol volume is conveniently expressed in terms of the LCV ward to evaluate. This operation can be bypassed by integrat-
function. This is shown inter alia to lead to a number of ing the position of the molecule i over phase space [11], or by
literature definitions of the local stress tensor. In the first replacing the Dirac δ with a similar but finite-valued function
part of this section, the techniques of Irving and Kirkwood of compact support [15, 18, 19, 21]. In the current treatment,
[8] are used to express the divergence of the stress (as with the LCV function, ϑ, is used, which is advantageous because
the right hand side of Eq. (32b)) in terms of intermolecu- it explicitly defines both the extent of the CV and its surface
lar force. Secondly, the CV pressure tensor is related to the fluxes. The pressure tensor can be written in terms of the
Volume Average (VA) formula ([16, 17]) and, by considera- LCV function by exploiting the following identities (see Ap-
tion of the interactions across the surfaces, to the Method Of pendix of Ref. [8]),
Planes (MOP) [13, 14]. Finally, the molecular CV Eq. (30) Z1
is written in analogous form to the macroscopic Eq. (32a).
The pressure tensor, Π, can be decomposed into a kinetic Oij δ(ri − r) = δ(r − ri + srij )ds, (37)
κ term, and a configurational stress σ. In keeping with the 0

056705-5
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

Equation (36) can therefore be written as,


Z Z N 
∂ 1X ∂
· σdV = − · fij rij
V ∂r V 2 ∂r
i,j
Z1 
× δ(r − ri + srij )ds; f dV. (38)
0

Equation Eq. (38) leads to the VA and MOP definitions of


the pressure tensor.

2. VA Pressure Tensor

definition of the stress tensor of Lutsko [16] and Cormier


et al. [17] can be obtained by rewriting Eq. (38) as,
Z Z N 
∂ ∂ 1X
· σdV = − · fij rij
∂r V ∂r V 2
i,j
Z1 
× δ(r − ri + srij )ds; f dV. (39)
FIG. 3. (Color online) A plot of the interaction length given by the
0 integral of the selecting function ϑs defined in Eq. (41) along the
Equating the expressions inside the divergence on both sides line between ri and rj . The cases shown are for two molecules
which are a) both inside the volume (lij = 1) and b) both outside
of Eq. (39), [41], and assuming the stress is constant within
the volume with an interaction crossing the volume, where lij is the
an arbitrary local volume, ∆V , gives an expression for the fraction of the total length between i and j inside the volume. The
VA stress, line is thin (blue) outside and thicker (red) inside the volume.
N 
Z X Z1 
VA 1 applied to obtain the volume averaged kinetic component of
σ =− fij rij δ(r − ri + srij )ds; f dV.
2∆V V i,j the pressure tensor, KT , in Eq. (25),
0
(40) VA VA
{ρuu} + κ
N 
z
N 
}| {
Swapping the order of integration and evaluating the integral X pi pi

∂ X pi pi
of the Dirac δ function over ∆V gives a different form of the · dSi ; f = · ϑi ; f .
mi ∂r mi
LCV function, ϑs , i=1 i=1
Z Note that the expression inside the divergence includes both
ϑs ≡ δ(r − ri + srij )dV = VA
 V
 the advection, {ρuu}, and kinetic components of the pres-
H(x+ − xi + sxij ) − H(x− − xi + sxij ) sure tensor. The VA form [17] is obtained by combining the
  VA
× H(y + − yi + syij ) − H(y − − yi + syij ) above expression with the configurational stress σ ,
 
× H(z + − zi + szij ) − H(z − − zi + szij ) , (41) VA VA VA VA VA
{ρuu} + κ − σ = {ρuu} + Π
which is non-zero if a point on the line between the two N  N 
molecules, ri − srij , is inside the cubic region (c.f. ri with 1 X pi pi 1X
= ϑi + fij rij lij ; f . (43)
ϑi ). Substituting the definition, ϑs (Eq. 41), into Eq. (40) ∆V mi 2
i=1 i,j
gives,
In contrast to the work of Cormier et al. [17], the advection
N  
VA 1 X term in the above expression is explicitly identified, in order
σ =− fij rij lij ; f , (42) to be compatible with the right hand side of Eq. (32b) and
2∆V
i,j definition of the pressure tensor, Π.
where lij is the integral from ri (s = 0) to rj (s = 1) of the 3. MOP Pressure Tensor
ϑs function,
The stress in the CV can also be related to the tractions
Z 1
over each surface. In analogy to prior use of the molecular
lij ≡ ϑs ds. LCV function, ϑi , to evaluate the flux, the stress LCV func-
0
tion, ϑs , can be differentiated to give the tractions over each
Therefore, lij is the fraction of interaction length between i surface. These surface tractions are the ones used in the for-
and j which lies within the CV, as illustrated in Fig. 3. The mal definition of the continuum Cauchy stress tensor. The
definition of the configurational stress in Eq. (42) is the same surface traction (i.e., force per unit area) and the kinetic pres-
as in the work of Lutsko [16] and Cormier et al. [17]. The sure on a surface combined give the MOP expression for the
microscopic divergence theorem given in Appendix A can be pressure tensor [13].

056705-6
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

In the context of the CV, the forces and fluxes on the six
bounding surfaces are required to obtain the pressure inside
the CV. It is herein shown that each face takes the form of
the Han and Lee [14] localization of the MOP pressure com-
ponents. The divergence theorem is used to express the left
hand side of Eq. (38) in terms of stress across the six faces
of the cube. The mesoscopic right hand side of Eq. (38) can
also be expressed as surface stresses by starting with the LCV
function ϑs ,
6 Z N  Z1 
X 1X ∂ϑs
σ · dSf = − fij rij · ds; f .
2 ∂r
f aces Sf i,j 0
The procedure for taking the derivative of ϑs with respect to
r and integrating over the volume is given in Appendix C.
The result is an expression for the force on the CV rewritten
as the force over each surface of the CV. For the x+ face, for
example, this is,
Z N 
1X 
σ · dSS + = − fij sgn(x+ − xj )
Sx+ x 4
i,j

 + FIG. 4. (Color online) Representation of those molecules selected
− sgn(x+ − xi ) Sxij ;f .
through dSxij in Eq. (46) with molecules i on the side of the surface
+
inside the CV (red) and molecules j on the outside (blue). The CV
The combination of the signum functions and the Sxij term is the inner square on the figure.
specifies when the point of intersection of the line between i
and j is located on the x+ surface of the cube (see Appendix
C). Corresponding expressions for the y and z faces are de- selects the force contributions across the two opposite faces;
±
fined by Sαij when α = {y, z} respectively. similar notation to the surface molecular flux, dSij = dS+ij −
The full expression for the MOP pressure tensor, which dS−ij (c.f. Eq. (17)), is used. The case of the two x planes
includes the kinetic part given by Eq. (26), is obtained by located on opposite sides of the cube is illustrated in Fig. 4.
assuming a uniform pressure over the x+ surface, Taking all surfaces of the cube into account yields the final
Z form,
Π · dS+ +
x = [κ − σ] · nx ∆Ax
+
+
Sx 6 Z N  3 
  X 1X X
≡ K+
x − Tx+ ∆A+
x = P+ +
x ∆Ax , (44) σ · dSf = − fij dSαij ; f
2
f aces Sf i,j α=1
where n+
x is a unit vector aligned along the x coordinate axis,
n+
x = [+1, 0, 0]; Tx+ is the configurational stress (traction) 1
N 
X 
+
and Px the total pressure tensor acting on a plane. Hence, =− fij ñ · dSij ; f
2
i,j
N   N  
1 X pi pix + 1X
P+
x = δ(xi − x+ )Sxi ;f = ςij · dSij ; f . (46)
∆A+x i=1
mi 2
i,j
N
X 
1   +
+ fij sgn(x+ −xj ) − sgn(x+ −xi ) Sxij ;f , The vector ñ, obtained in Appendix C, is unity in each direc-
4∆A+x i,j tion. The tensor ςij is defined, for notational convenience, to
(45) be the outer product of the intermolecular forces with ñ,
where the peculiar momentum, pi has been used as in Todd  
et al. [13]. If the x+ surface area covers the entire domain   fxij fxij fxij
+
(Sxij = 1 in Eq. (45)), the MOP formulation of the pressure ςij ≡ −fij ñ = −fij 1 1 1 = − fyij fyij fyij  .
is recovered [13]. fzij fzij fzij
+
The extent of the surface is defined through Sxij , in Eq.
(45) which is the localized form of the pressure tensor con- In this form, the ϑij function for all interactions over the
sidered by Han and Lee [14] applied to the six cubic faces. cube’s surface is expressed as the sum of six selection func-
P
For a cube in space, each face has three components of stress, tions for each of the six faces, i.e. ϑij = − 3α=1 dSαij .
which results in 18 independent components over the total
control surface. The quantity,
1  + 4. Relationship to the continuum
+ +
dSαij ≡ sgn(rα − rαj ) − sgn(rα − rαi ) Sαij
2 The forces per unit area, or ’tractions’, acting over each
1 − −  − face of the CV, are used in the definition of the Cauchy stress
− sgn(rα − rαj ) − sgn(rα − rαi ) Sαij ,
2 tensor at the continuum level. For the x+ surface, the traction

056705-7
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

vector is the sum of all forces acting over the surface, surfaces — an important requirement for accurate unsteady
coupled simulations as outlined in the finite volume coupling
N 
1 X  of Delgado-Buscalioni and Coveney [45]. For solid coupling
T+
x =− + fij sgn(x+ − xj ) schemes, [30], the principle of virtual work can be used with
4∆Ax i,j
 tractions on the element corners (the MD CV) to give the
+  + state of stress in the element [48],
− sgn(x − xi ) Sxij ; f , (47)
Z I
which satisfies the definition, σ · ∇Na dV = Na TdS, (50)
V S

x =σ · n±
x,

of the Cauchy traction [42]. A similar relationship can be where Na is a linear shape function which allows stress to
written for both the kinetic and total pressures, be defined as a continuous function of position. It will be
demonstrated numerically in the next section, IV, that the CV
Kx± = κ · n±
x, formulation is exactly conservative: the surface tractions and
fluxes entirely define the stress within the volume. The trac-
P± ±
x = Π · nx , tions and stress in Eq. (50) are connected by the weak formu-
lation and the form of the stress tensor results from the choice
where n± ±
x is a unit vector, nx = [±1 0 0] .
T
of shape function Na .
The time evolution of the molecular momentum within a
CV ( Eq. (30)), can be expressed in a similar form to the
Navier-Stokes equations of continuum fluid mechanics. Di-
viding both sides of Eq. (30) by the volume, the following D. Energy Balance for a Molecular CV
form can be obtained; note that this step requires Eqs. Eq. In this section, a mesoscopic expression for time evolution
(26), Eq. (45) and Eq. (47): of energy within a CV is derived. As for mass and momen-
tum, the starting point is to integrate the energy at a point,
N  
{ρuα uβ }+ − {ρuα uβ }− given in Eq. (10), over the CV,
1 ∂ X
pαi ϑi ; f + =
∆V ∂t ∆rβ
i=1 Z N  
+ − + −
X
Kαβ − Kαβ Tαβ − Tαβ N  
1 X ρ(r, t)E(r, t)dV = ei ϑi ; f . (51)
− + + fαiext ϑi ; f , V i=1
∆rβ ∆rβ ∆V
i=1
(49)
The time evolution within the CV is given using formula (12),
where index notation has been used (e.g. T± ±
x = Tαx ) with
the Einstein summation convention. Z N  
In the limit of zero volume, each expression would be simi- ∂ ∂ X
ρ(r, t)E(r, t)dV = ei ϑi ; f
lar to a term in the differential continuum equations (although ∂t V ∂t
i=1
the pressure term would be the divergence of a tensor and not N  
the gradient of a scalar field as is common in fluid mechan- X p i ∂ ∂
= · ei ϑi + Fi · ei ϑi ; f . (52)
ics). The Cauchy stress tensor, σ, is defined in the limit that mi ∂ri ∂pi
i=1
the cube’s volume tends to zero, so that T+ and T− are re-
lated by an infinitesimal difference. This is used in contin-
uum mechanics to define the unique nine component Cauchy Evaluating the derivatives of the energy and LCV function
stress tensor, dσ/dx ≡ lim∆x→0 [T+ + T− ]/∆x. This limit results in,
is shown in Appendix B to yield the Irving and Kirkwood [8]
stress in terms of the Taylor expansion in Dirac δ functions. N   N   
Rather than defining the stress at a point, the tractions ∂ X 1X pi pj
ei ϑi ; f = − · fij + · fji ϑi ; f
can be compared to their continuum counterparts in a fluid ∂t 2 mi mi
i=1 i,j
mechanics control volume or a solid mechanics Finite Ele-
N  
ments (FE) method. Computational Fluid Dynamics (CFD) X pi p
− ei · dSi − Fi · i ϑi ; f .
is commonly formulated using CV and in discrete simula- mi mi
i=1
tions, Finite Volume [4]. Surface forces are ideal for coupling
schemes between MD and CFD. Building on the pioneering
work of O’Connell and Thompson [23], there are many MD Using the definition of Fi , Newton’s 3rd law and relabelling
to CFD coupling schemes – see the review paper by Mo- indices, the intermolecular force terms can be expressed in
hamed and Mohamad [43]. More recent developments for terms of the interactions over the CV surface, ϑij ,
coupling to fluctuating hydrodynamics are covered in a re-
view by Delgado-Buscalioni [44]. A discussion of coupling
N   N  
schemes is outside the scope of this work, however finite vol- ∂ X X pi
ei ϑi ; f = − ei · dSi ; f
ume algorithms have been used extensively in coupling meth- ∂t mi
i=1 i=1
ods [31, 32, 45–47] together with equivalent control volumes
N   XN  
defined in the molecular region. An advantage of the herein 1 X pi pi
proposed molecular CV approach is that it ensures conser- + · fij ϑij ; f + · fiext ϑi ; f .
2 mi mi
vation laws are satisfied when exchanging fluxes over cell i,j i=1

056705-8
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

The right hand side of this equation is equated to the right where the total force on molecule i has been decomposed into
hand side of the continuum energy Eq. 4, surface and ‘external’ or body terms. The time evolution of
energy flux pressure heating
energy in a molecular control volume is obtained by evaluat-
heat flux
z I }| { Iz }| { zI }| { ing,
− ρEu · dS − q · dS − Π · u · dS
S S S N N  
∂ X X ∂ϑ ∂e
N   ei ϑi = ei i + i ϑi
X pi ∂t ∂t ∂t
=− ei · dSi ; f i=1 i=1
mi N N
i=1 X pi X ṗ · p
i i
N   =− ei · dSi + ϑi
1 X pi mi mi
+ · ςij · dSij ; f , (53) i=1 i=1
2 mi N  
i,j 1 X pi pj
− · fij + · fji ϑi
where the energy due to the external (body) forces is ne- 2 mi mj
i,j
glected. The fij ϑij has been re-expressed in terms of surface
tractions, ςij · dSij , using the analysis of the previous sec-
using, dpi /dt = Fi and the decomposition of forces. The
tion. In its current form, the microscopic equation does not
manipulation proceeds as in the mesoscopic system to yield,
delineate the contribution due to energy flux, heat flux and
pressure heating. To achieve this division, the notion of the N N
peculiar momentum at the molecular location, u(ri ) is used ∂ X X p
ei ϑi = − ei i · dSi
together with the velocity at the CV surfaces u(r± ), follow- ∂t mi
i=1 i=1
ing a similar process to Evans and Morriss [7].
N N
1 X pi X pi
IV. IMPLEMENTATION + · fij ϑij + · fiext ϑi , (56)
2 mi mi
i,j i=1
In this section, the CV equation for mass, momentum and
energy balance, Eqs. (22), (30) and (53), will be proved to ap-
The average of many such trajectories defined through Eqs.
ply and demonstrated numerically for a microscopic system
(54), (55) and (56) gives the mesoscopic expressions in Eqs.
undergoing a single trajectory through phase space.
(22), (30) and (53), respectively. In the next subsection, the
A. The Microscopic System time integral of the single trajectory is considered.
Consider a single trajectory of a set of molecules through
phase space, defined in terms of their time dependent coor- B. Time integration of the microscopic CV equations
dinates ri and momentum pi . The LCV function depends on Integration of Eqs. (54), (55) and (56) over the time inter-
molecular coordinates, the location of the center of the cube, val [0, τ ] enables these equations to be usable in a molecular
r, and its side length, ∆r, i.e., ϑi ≡ ϑi (ri (t), r, ∆r). The simulation. For the conservation of mass term,
time evolution of the mass within the molecular control vol-
ume is given by, N
X Zτ X
N

N N
mi [ϑi (τ ) − ϑi (0)] = − pi · dSi dt. (57)
d X X ∂ϑ
mi ϑi (ri (t), r, ∆r) = mi i i=1 0 i=1
dt ∂t
i=1 i=1
N N The surface crossing term, dSi , defined in Eq. (16), involves
X dr ∂ϑi X a Dirac δ function and therefore cannot be evaluated directly.
= mi i · =− pi · dSi , (54)
dt ∂ri Over the time interval [0, τ ], molecule i passes through a
i=1 i=1
given x position at times, txi,k , where k = 1, 2, ..., Ntx [49]
using, pi = mi dri /dt. The time evolution of momentum in . The positional Dirac δ can be expressed as,
the molecular control volume is,
N Ntx
∂ X X δ(t − txi,k )
pi (t)ϑi (ri (t), r, ∆r) δ(xi (t) − x) = , (58)
∂t |ẋi (txi,k )|
i=1 k=1
N  
X ∂ϑ dp where |ẋi (txi,k )| is the magnitude of the velocity in the x
= pi i + i ϑi
∂t dt direction at time txi,k . Equation Eq. (58) is used to rewrite
i=1
N   dSi in Eq. (57) in the form,
X dri ∂ϑi dpi
= pi · + ϑ . h i
dt ∂ri dt i dSαi,k ≡ sgn(t+ − τ ) − sgn(t +
−0) +
Sαi,k (t+
i=1 αi,k αi,k αi,k )
h i
As, dpi /dt = Fi , then, − sgn(t− − − −
αi,k −τ ) − sgn(tαi,k −0) Sαi,k (tαi,k ),
N N   (59)
∂ X X pi pi
pi ϑi = − · dSi + Fi ϑi
∂t mi
i=1 i=1 where α = {x, y, z}, and the fluxes are evaluated at times,

N
X pi pi 1 XN N
X t+
αi,k and tαi,k for the right and left surfaces of the cube, re-
=− · dSi + fij ϑij + fiext ϑi , (55) spectively. Using the above expression, the time integral in
mi 2
i=1 i,j i=1 Eq. (57) can be expressed as the sum of all molecule cross-

056705-9
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

ings, Nt = Ntx + Nty + Ntz over the cube’s faces, The Eq. (62) demonstrates that the time average of the fluxes,
stresses and body forces on a CV during the interval 0 to τ ,
Accumulation
z }| { completely determines the change in momentum within the
N Nt
N X 3 CV for a single trajectory of the system through phase space
X X X pαi
mi [ϑi (τ ) − ϑi (0)] = − mi dS . (i.e. an MD simulation). The time evolution of the micro-
|pαi | αi,k scopic system, Eq. (62), can also be obtained directly by
i=1 i=1 k=1 α=1
| {z } evaluating the derivatives of the mesoscopic expression
Advection
(49)
(60) and invoking the ergodic hypothesis, hence replacing α; f
R
with τ1 0τ αdt. The use of the ergodic hypothesis is justified
In other words, the mass in a CV at time t = τ minus its provided that the time interval, τ , is sufficient to ensure phase
initial value at t = 0 is the sum of all molecules that cross its space is adequately sampled.
surfaces during the time interval. Finally, there are no new techniques required to integrate
The momentum balance equation Eq. (55), can also be the energy Eq. 56,
written in time-integrated form,
N
X
N
X [ei (τ )ϑi (τ ) − ei (0)ϑi (0)]
[pi (τ )ϑi (τ ) − pi (0)ϑi (0)] =
i=1
i=1  
  Zτ XN N
Zτ N N N p 1 X pi
X p p
i i 1 X X = −  ei i · dSi − · fij ϑij  dt (63)
−  · dSi − fij ϑij − fiext ϑi  dt, mi 2 mi
mi 2 0 i=1 i,j
0 i=1 i,j i=1

and using identity (59), which gives the final form, written without external forcing,

Accumulation Advection Accumulation Advection


z }| { z }| { z }| { z }| {
N Nt
N X 3 N N Nt 3
X X X pαi X X X X pαi
[pi (τ )ϑi (τ )−pi (0)ϑi (0)] + pi dS [ei (τ )ϑi (τ )−ei (0)ϑi (0)]+ ei dS
|pαi | αi,k |pαi | αi,k
i=1 i=1 k=1 α=1 i=1 i=1 k=1 α=1
N Zτ N Zτ N Zτ
X X 1 X pi (t)
= fij (t)ϑij (t)dt + fiext (t)ϑi (t)dt . = · fij (t)ϑij (t)dt .
2 mi
i,j 0 i=1 0 i,j 0
| {z } | {z }
Forcing Forcing
(61) (64)
The integral of the forcing term can be rewritten as the sum, As in the momentum balance equation, the integral of the
Zτ Nτ forcing term can be approximated by the sum,
X
fij (t)ϑij (t)dt ≈ ∆t fij (tn ) ϑij (tn ) , Zτ
n=1 pi (t)
0 · fij (t)ϑij (t)dt
mi
where Nτ is the number time steps. Equation (61) can be 0
rearranged as follows, Nτ
X pi (tn )
N ≈ ∆t · fij (tn ) ϑij (tn ) ,
X pαi (τ )ϑi (τ ) − pαi (0)ϑi (0) mi
n=1
τ ∆V
i=1 where Nτ is the number time steps.
+ −
{ρuα uβ }+ − {ρuα uβ }− K αβ − K αβ In the next section, the elements, Accumulation, Advec-
+ =− tion and Forcing in the above equations are computed indi-
∆rβ ∆rβ
vidually in an MD simulation to confirm Eqs. (60), (61) and
+ − Nτ
N X
T αβ − T αβ 1 X (64) numerically.
+ + fαiext (tn )ϑi (tn ), (62)
∆rβ Nτ ∆V C. Results and Discussion
i=1 n=1
Molecular Dynamics (MD) simulations in 3D are used in
where the overbar denotes the time average. The time-
this section to validate numerically, and explore the statisti-
averaged traction in (62) is given by,
cal convergence of, the CV formalism for three test cases.
N Nτ The first investigation was to confirm numerically the con-
± 1 1 XX ±
T αβ = − fαij (tn )dSβij (tn ), servation properties of an arbitrary control volume. The sec-
Nτ 4∆Aβ ond simulation compares the value of the scalar pressure ob-
i,j n=1
tained from the molecular CV formulation with that of the
The time-averaged kinetic surface pressure in (62) is, virial expression for an equilibrium system in a periodic do-
N Nt main. The final test is a Non Equilibrium Molecular Dy-
± 1 1 XX pαi (tk )pβi (tk ) ± namics (NEMD) simulation of the start-up of Couette flow
K αβ = dSβi,k (tk )
τ 2∆Aβ |pβi (tk )| initiated by translating the top wall in a slit channel geom-
i=1 k=1
etry. The NEMD system is analyzed using the CV expres-
−{ρuα uβ }± . sions Eqs. (60), (61) and (64), and the shear pressure was

056705-10
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

computed by the VA and CV routes. Newton’s equations of


motion were integrated using the half-step leap-frog Verlet 0.3
algorithm, [50]. The repulsive Lennard-Jones (LJ) or Weeks- (a)
Chandler-Anderson (WCA) potential [51],
"  0.2
 6 #
ℓ 12 ℓ
Φ(rij ) = 4ǫ − + ǫ, rij ≤ rc , (65)
rij rij 0.1

Momentum
was used for the molecular interactions, which is the
Lennard-Jones potential shifted upwards by ǫ and truncated 0
at the minimum in the potential, rij = rc ≡ 21/6 ℓ. The
potential is zero for rij > rc . The energy scale is set by ǫ, −0.1
the length scale by ℓ and molecular mass by m. The results 1
0.5
reported here are given in terms of ℓ, ǫ and m. A timestep of
0
0.005 was used for all simulations. The domain size in the −0.2
−0.5
first two simulations was 13.68, which contained N = 2048 −1
molecules, the density was ρ = 0.8 and the reduced tem- −0.3
perature was set to an initial value of T = 1.0. Test cases 0 0.2 0.4 0.6
1 and 2 described below are for equilibrium systems, and Time
therefore did not require thermostatting. Case 3 is for a non-
equilibrium system and required removal of generated heat,
which was achieved by thermostatting the wall atoms only. 0.3
(b)
1. Case 1
0.2
In case 1, the periodic domain simulates a constant energy
ensemble. The separate terms of the integrated mass, mo-
mentum and energy equations given in (60), (61) and (64) 0.1
were evaluated numerically for several sizes of CV. The mass
Energy

conservation can readily be shown to be satisfied as it simply


0
requires tracking the number of molecules in the CV. The
momentum and energy balance equations are conveniently
checked for compliance at all times by evaluating the resid- −0.1
2
ual quantity, 1
0
Residual = Accumulation − Forcing + Advection, (66) −0.2
−1
−2
which must be equal to zero at all times for the CV equations
to be satisfied. This was demonstrated to be the case, as may −0.3
0 0.2 0.4 0.6
be seen in Figs. 5(a) and 5(b), for a cubic CV of side length Time
1.52 in the absence of body forces. The evolution of momen-
tum inside the CV is shown numerically to be exactly equal
to the integral of the surface forces until a molecule crosses FIG. 5. The various components in Eq. 66, ‘Accumulation’ (—),
the CV boundary. Such events give rise to a momentum flux the time integral of the surface force, ‘Forcing’ (×), and momen-
tum flux term, ‘Advection’ (- - -) are shown. ‘Forcing’ symbols are
contribution which appears as a spike in the Advection and
shown every 4th timestep for clarity and the insert shows the full
Accumulation terms, as is evident in Fig. 5(a). The residual ordinate scale over the same time interval on the abscissa. From
nonetheless remains identically zero (to machine precision) top to bottom, (a) Momentum Control Volume, (b) Energy Control
at all times. The energy conservation is also displayed in Volume.
Fig. 5(b). The average error over the period of the simulation
(100 MD timeunits) was less than 1%, where the average er-
arises from the intermolecular interactions across the periodic
ror is defined as the ratio of the mean |Residual| to the mean
boundaries [12]. The CV formula for the scalar pressure is,
|Accumulation| over the simulation. The error is attributed
to the use of the leapfrog integration scheme, a conclusion 1 + − + − + −
supported by the linear decrease in error as timestep ∆t → 0. ΠCV = Pxx +Pxx +Pyy +Pyy +Pzz +Pzz , (68)
6
2. Case 2 where the Pαα ± normal pressure is defined in Eq. (45) and
As in case 1, the same periodic domain is used in case 2 includes both the kinetic and configurational components
to simulate a constant energy ensemble. The objective of this on each surface. Both routes involve the pair forces, fij .
exercise is to show that the average of the virial formula for However, the CV expression which uses MOP counts only
the scalar pressure, Πvir , applicable to an equilibrium peri- those pair forces which cross a plane while VA (Virial) sums
odic system, fij rij over the whole volume. It is therefore expected that
there would be differences between the two methods at short
N  N 
1 X pi · pi 1X times, converging at long times. A control volume the same
Πvir = + fij · rij ; f , (67) size as the periodic box was taken. The time averaged control
3V mi 2
i=1 i6=j volume, (ΠCV ) and virial (Πvir ) pressure values are shown

056705-11
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

2
6 10

5
0

Percentage Discrepancy
10
4
Pressure

Control Volume −2
3 10
virial

2
−4
10
1 Kinetic
Configurational
−6
Total
0 10 1 2 3 4 5
0 0.5 1 1.5 2 2.5 10 10 10 10 10
Time Time

FIG. 6. Πvir and ΠCV from Eqs. (67) and (68) respectively. The FIG. 7. The percentage relative difference between the virial and
configurational and kinetic pressures are separated with configura- control volume time-accumulated scalar pressures (PD defined in
tional values typically having greater magnitudes (∼ 4.0) than ki- the text). Values for the kinetic, configurational and total PD are
netic (∼ 0.6). Continuous lines are control volume pressures and shown.
dotted lines are virial pressure.
system out of equilibrium.
in Fig. 6 to converge towards the same value with increasing 3. Case 3
time. The simulation is started from an FCC lattice with a
short range potential (WCA) so the initial configurational In this simulation study, Couette flow was simulated by
stress is zero. It is the evolution of the pressure from this entraining a model liquid between two solid walls. The top
initial state that is compared in Fig. 6. The virial kinetic wall was set in translational motion parallel to the bottom
pressure makes use of the instantaneous values of the domain (stationary) wall and the evolution of the velocity profile to-
molecule’s velocities at every time step. In contrast, the wards the steady-state Couette flow limit was followed. The
CV kinetic part of the pressure is due to molecular surface velocity profile, and the derived CV and VA shear stresses are
crossings only, which may explain its slower convergence compared with the analytical solution of the unsteady diffu-
to the limiting value than the kinetic part of the virial sion equation. Four layers of tethered molecules were used
expression. To quantify this difference in convergence for to represent each wall, with the top wall given a sliding ve-
the two measures of the pressure, the standard deviation, locity of, U0 = 1.0 at the start of the simulation, time t = 0.
SD(x), is evaluated, ensuring decorrelation [47] using block The temperature of both walls was controlled by applying the
averaging [51]. For the kinetic virial, SD(κvir ) = 0.0056, Nosé-Hoover (NH) thermostat to the wall atoms [52]. The
and configurational, SD(σvir ) = 0.0619. For the kinetic two walls were thermostatted separately, and the equations
CV pressure SD(κCV ) = 0.4549 and configurational of motion of the wall atoms were,
SD(σCV ) = 0.2901. The CV pressure, which makes pi
use of the MOP formula, would therefore require more ṙi = + U0 n +
x, (69a)
samples to converge to a steady state value. However, the mi
MOP pressures are generally more efficient to calculate ṗi = Fi + fiext − ξpi , (69b)
than the VA. More usefully, from an evaluation of only the  
fiext = ri0 4k4 ri2 + 6k6 ri4 , (69c)
interactions over the outer CV surface, the pressure in a 0 0
volume of arbitrary size can be determined. " N #
˙ξ = 1
X p ·p
Figure 7 is a log-log plot of the Percentage Discrepancy (PD) n n
− 3T0 , (69d)
between the two (P D = [100 × |ΠCV − Πvir |/Πvir ]). Qξ mn
n=1
After 10 million timesteps or a reduced time of 5 × 104 ,
the percentage discrepancy in the configurational part has where n+ x is a unit vector in the x−direction, mn ≡ m, and
decreased to 0.01%, and the kinetic part of the pressure fiext is the tethered atom force, using the formula of Petravic
matches the virial (and kinetic theory) to within 0.1%. The and Harrowell [53] (k4 = 5 × 103 and k6 = 5 × 106 ). The
total pressure value agrees to within 0.1% at the end of this vector, ri0 = ri −r0 , is the displacement of the tethered atom,
averaging period. The simulation average temperature was i, from its lattice site coordinate, r0 . The Nosé-Hoover ther-
0.65, and the kinetic part of the CV pressure was statis- mostat dynamical variable is denoted by ξ, T0 = 1.0 is the
tically the same as the kinetic theory formula prediction, target temperature of the wall, and the effective time constant
κCV = ρkB T = 0.52 [51]. The VA formula for the pressure or damping coefficient, in Eq. (69d) was given the value,
in a volume the size of the domain is by definition formally Qξ = N ∆t. The simulation was carried out for a cubic do-
the same as that of the virial pressure. The next test case main of sidelength 27.40, of which the fluid region extent
compares the CV and VA formulas for the shear stress in a was 20.52 in the y−direction. Periodic boundaries were used

056705-12
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

Analytical
1
MD simulation

0.8

0.6

0
U /U
x
0.4

0.2

FIG. 8. (Color online) Schematic diagram of the NEMD simulation 0 0.2 0.4 0.6 0.8 1
geometry consisting of a sliding top wall and stationary bottom wall, y/L
both composed of tethered atoms. The simulation domain contained
a lattice of contiguous CV used for pressure averaging (shown by the FIG. 9. The y− dependence of the streaming velocity profile at
small boxes) while the thicker line denotes a single CV containing times t = 2n for n = 0, 2, 3, 4, 5, 6 from right to left. The squares
the entire liquid region. are the NEMD CV data values and the analytical solution to the
continuum equations of Eq. (72) is given at the same six times as
continuous curves.
in the streamwise (x) and spanwise (z) directions. The re-
sults presented are the average of eight simulation trajecto-
ries starting with a different set of initial atom velocities. The where the bottom and top wall-liquid boundaries are at y = 0
lattice contained 16384 molecules and was at a density of and y = L, respectively. The Fourier series solution of these
ρ = 0.8. The molecular simulation domain was sub-divided equations with inhomogeneous boundary conditions [55] is,
into 4096 (163 ) control volumes, and the average velocity 
and shear stress was determined in each of them. A larger 
 U0 y=L

X∞  nπy 
single CV encompassing all of the liquid region of the do-
ux (y, t) = un (t)sin 0<y<L (72)
main, shown bounded by the thick line in Fig. 8, was also  L
n=1


considered.
0 y=0
The continuum solution for this configuration is consid-
ered now. Between two plates, there are no body forces and
where λn = (nπ/L)2 and un (t) is given by,
the flow eventually becomes fully developed, [54] so that Eq.
(2) can be simplified and after applying the divergence theo- 2U0 (−1)n
 
λn µt
 
rem from Eq. (5) it becomes, un (t) = exp − −1 .
nπ ρ
Z Z

ρudV = − ∇ · ΠdV, The velocity profile resolved at the control volume level
∂t V V is compared with the continuum solution in Fig. 9. There
which is valid for any arbitrary volume in the domain and were 16 cubic NEMD CV of side length 1.72 spanning the
must be valid at any point for a continuum. The shear pres- system in the y direction, with each data point on the figure
sure in the fluid, Πxy (y), drives the time evolution, being derived from a local time average of 0.5 time units.
The analytic continuum solution was evaluated numerically
∂ρux ∂Πxy from Eq. (72) with n = 1000 and µ = 1.6, the latter a
=− . literature value for the WCA fluid shear viscosity at ρ = 0.8
∂t ∂y
and T = 1.0, [56]. There is mostly very good agreement
For a Newtonian liquid with viscosity, µ, [54], between the analytic and NEMD velocity profiles at all
times, although some effect of the stacking of molecules
∂ux near the two walls can be seen in a slight blunting of the
Πxy = −µ , (70) fluid velocity profile very close to the tethered walls (located
∂y
by the horizontal two squares on the far left and right of the
this gives the 1D diffusion equation, figure) which is an aspect of the molecular system that the
continuum treatment is not capable of reproducing.
∂ux µ ∂ 2 ux
= , (71)
∂t ρ ∂y 2 The VA and CV shear pressure, given by Eqs. (43)
and (45), are compared at time t = 10 in Fig. 10. The
assuming the liquid to be incompressible. This can be solved comparison is for a single simulation trajectory resolved
for the boundary conditions, into 16 cubic volumes of size 1.72 in the y−direction, with
averaging in the x and z directions and over 0.5 in reduced
ux (0, t) = 0 ux (L, t) = U0 ux (y, 0) = 0, time. The figure shows the shear pressure on the faces of the

056705-13
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

0.5 0.5
VA VA
Top Surface CV Top Surface CV
0.4 0.4
Bottom Surface CV Bottom Surface CV
Inside CV Inside CV
0.3 0.3 Analytical
Shear Pressure

Shear Pressure
0.2 0.2

0.1 0.1

0 0

−0.1 −0.1

−0.2 −0.2
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
y/L y/L

FIG. 10. The y−dependence of the shear pressure at t = 10, aver- FIG. 11. As Fig. 10, except that the NEMD results are averaged over
aged over 100 timesteps and for a single simulation trajectory. The a set of eight independent simulations of 1, 000 timesteps (5 reduced
VA value from Eq. (43) are the squares. The CV surface traction time units) each. The simulation-derived VA and CV shear pres-
from Eq. (45) is indicated by × and ◦ for the top and bottom sur- sures are compared with the continuum analytical solution given in
faces, respectively. The solid gray line displays the resulting pres- Eq. (73) (solid black line). The jump in the profile on the right of
sure field using Eq. (50) with linear shape functions. the figure is due to the presence of the tethered wall.

CV. Inside the CV, the pressure was assumed to vary linearly, arise from interactions with the wall atoms only. The mo-
and the value at the midpoint is shown to be comparable mentum equation, Eq. (55), is written as,
to the VA-determined value. Figure 10 shows that there
is good agreement between the VA and CV approaches. 1
N N z }| { X
3
N z }| {
Note that the CV pressure is effectively the MOP formula ∂ X X pi pi
applied to the faces of the cube, and hence this case study pi ϑi = − · dSi + fiext ϑi .
∂t mi
demonstrates a consistency between MOP and VA. We have i=1 i=1 i=1
shown previously that this is true for the special case of an N
1 X  
infinitely thin bin or the limit of the pressure at a plane [22]. − fij dSxij + fij dSyij + fij dSzij ,
Practically, the extent of agreement in this exercise is limited 2 | {z } |{z} | {z }
i,j
by the inherent assumptions and spatial resolution of the two 2 4 2
methods; a single average over a volume is required for VA,
but a linear pressure relationship is assumed for CV to obtain which can be simplified as follows. For term, 1 in the
the pressure tensor value corresponding to the center of the above equation, the fluxes across the CV boundaries in the
CV. streamwise and spanwise directions cancel due to the peri-
odic boundary conditions. Fluxes across the xz boundary
The continuum analytical xy pressure tensor component surface are zero as the tethered wall atoms prevent such cross-
can be derived analytically using the same Fourier series ap- ings. The force term, , 2 also vanishes because across the
+ −
proach for ∂ux /∂y,[55], periodic boundary, fij dSxij = −fij dSxij , (similarly for z).
,
The external force term, 3 is zero because all the forces
" #
µU0 X∞ λn µt  nπy  in the system result from interatomic interactions. The sum
n − ρ of the fyij force components across the horizontal bound-
Πxy (y, t) = − 1 + 2 (−1) e cos ,
L L aries will be equal and opposite, and by symmetry the two
n=1
(73) fzij terms in 4 will be zero on average. The above equation
therefore reduces to,
which is valid for the entire domain 0 ≤ y ≤ L.
A statistically meaningful comparison between the CV, N N
∂ X 1 Xh + −
i
VA and continuum analytic shear pressure profiles requires pi ϑi = − fxij dSyij − fxij dSyij . (74)
more averaging of the simulation data than for the streaming ∂t 2
i=1 i,j
velocity, [57], and eight independent simulation trajectories
over 5 reduced time units were used. Figure 11 shows As the simulation approaches steady state, the rate of change
that the three methods exhibit good agreement within the of momentum in the control volume tends to zero because
simulation statistical uncertainty. the difference between the shear stresses acting across the
top and bottom walls vanishes. The forces on the xz plane
As a final demonstration of the use of the CV equations, boundary and momentum inside the CV are plotted in Fig.
the control volume is now chosen to encompass the entire 12 to confirm Eq. (74) numerically. The time evolution of
liquid domain (see Fig. 8), and therefore the external forces these molecular momenta and surface stresses are compared

056705-14
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

boundaries. The average response nevertheless agrees well


with the analytic solution, bearing in mind the element of
1 ∆M uncertainty in the matching state parameter values. This ex-
Mol Force Top ample demonstrates the potential of the CV approach applied
Mol Force Bottom on the molecular scale, as it can be seen that computation of
0.5 the forces across the CV boundaries determines completely
the average molecular microhydrodynamic response of the
system contained in the CV. In fact, the force on only one of
Magnitude

the surfaces is all that was required, as the force terms for
0 the opposite surface could have been obtained from Eq. (74).

−0.5
Residual
Analy ∆ M V. CONCLUSIONS
Analy Force Top
In analogy to continuum fluid mechanics, the evolution
−1 Analy Force Bottom
equations for a molecular systems has been expressed within
0 20 40 60 a Control Volume (CV) in terms of fluxes and stresses across
Time the surfaces. A key ingredient is the definition and manipula-
tion of a Lagrangian to Control Volume conversion function,
FIG. 12. The evolution of surface forces and momentum change ϑ, which identifies molecules within the CV. The final ap-
for a molecular CV from Eq. (74), (points) and analytical solution pearance of the equations has the same form as Reynolds’
for the continuum (Eqs. (77), (78) and (76)), presented as lines on Transport Theorem applied to a discrete system. The equa-
the figure. The Residual, defined in Eq. (66), is also given. Each tions presented follow directly from Newton’s equation of
point represents the average over an ensemble of eight independent motion for a system of discrete particles, requiring no ad-
systems and 40 timesteps.
ditional assumptions and therefore sharing the same range of
validity.
to the analytical continuum solution for the CV, Using the LCV function, the relationship between Volume
  Average (VA) [16, 17] and Method Of Planes (MOP) pres-
Z Z Z
∂ sure [13, 14] has been established, without Fourier transfor-
ρux dV = −  Πxy dSf+ − Πxy dSf−  . mation. The two definitions of pressure are shown numer-
∂t V S+ S−
f f ically to give equivalent results away from equilibrium and,
(75) for homogeneous systems, shown to equal the virial pressure.
A Navier–Stokes-like equation was derived for the evo-
The normal components of the pressure tensor are non-zero lution of momentum within the control volume, expressed
in the continuum, but exactly balance across opposite CV in terms of surface fluxes and stresses. This pro-
faces, i.e. Π+ −
xx = Πxx . By appropriate choice of the gauge vides an exact mathematical relationship between molecular
pressure, Πxx does not appear in the governing Eq. (75). The fluxes/pressures and the evolution of momentum and energy
left hand side of the above equation is evaluated from the an- in a CV. Numerical evaluations of the terms in the conserva-
alytic expression for ux , tion of mass, momentum and energy equations demonstrated
Z ∞ λ µt consistency with theoretical predictions.
∂ µU0 X − nρ
ρux dV = 2∆x∆z [1 − (−1)n ] e . The CV formulation is general, and can be applied to de-
∂t V L rive conservation equations for any fluid dynamical property
n=1
(76) localised to a region in space. It can also facilitate the deriva-
tion of conservative numerical schemes for MD, and the eval-
The right hand side is obtained from the analytic continuum uation of the accuracy of numerical schemes. Finally, it al-
expression for the shear stress, for the bottom surface at y = lows for accurate evaluation of macroscopic flow properties,
0, in a manner consistent with the continuum conservation laws.
Z ∞
µU0 X − λnρµt
Πxy dSf+ = −2∆x∆z e , (77)
S+ L
f n=1
Appendix A: Discrete form of Reynolds’ Transport Theorem
and for the top y = L, and the Divergence Theorem

Z ∞ λ µt
In this appendix, both Reynolds’ Transport Theorem and
µU0 X − n
Πxy dSf− = −2∆x∆z (−1)n e ρ . (78) the Divergence Theorem for a discrete system are derived.
S − L The relationship between an advecting and fixed control vol-
f n=1
ume is shown using the concept of peculiar momentum.
In Fig 12, the momentum evolution on the left hand side of The microscopic form of the continuous Reynolds’ Trans-
Eq. (74) is compared to Eq. (76). Equations (77) and (78) are port Theorem [1] is derived for a property χ = χ(ri , pi , t)
also given for the shear stresses acting across the top and bot- which could be mass, momentum or the pressure tensor. The
tom of the molecular control volume (right hand side of Eq. LCV function, ϑi , is dependent on the molecule’s coordinate;
(74)). The scatter seen in the MD data reflects the thermal the location of the cube center, r, and side length, ∆r, which
fluctuations in the forces and molecular crossings of the CV are all a function of time. The time evolution of the CV is

056705-15
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

therefore, The vector derivative of the Dirac δ followed by the integral


N
over volume results in,
d X Z
χ(t)ϑi (ri (t), r(t), ∆r(t)) ∂
dt δ(xi − x)δ(yi − y)δ(zi − z)dV
i=1 ∂r
N V 
X dχ dr ∂ϑi [δ(xi − x)H(yi − y)H(zi − z)]V
= ϑ +χ i ·
dt i dt ∂ri = [H(xi − x)δ(yi − y)H(zi − z)]V 
i=1 [H(xi − x)H(yi − y)δ(zi − z)]V

dr ∂ϑi d∆r ∂ϑi  + −
 
+χ · +χ · . δ(xi − x+) − δ(xi − x− ) Sxi
dt ∂r dt ∂∆r
=  δ(yi − y ) − δ(yi − y ) Syi  = dSi ,
The velocity of the moving volume is defined as ũ = dr/dt, δ(zi − z + ) − δ(zi − z − ) Szi
which can be different to the macroscopic velocity u. Sur-
face translation or deformation of the cube, ∂ϑi /∂∆r, can be where the limits of the cuboidal volume are, r+ = r+ ∆r 2 and
included in the expression for velocity ũ. The above analysis r− = r − ∆r 2 . The mesoscopic equivalent of the continuum
is for a microscopic system, although a similar process for a divergence theorem (Eq. (5)) is therefore,
mesoscopic system can be applied and includes terms for CV
Z N N
movement in Eq. (12). ∂ X X
Hence Reynolds treatment of a continuous medium [1] is · χδ(ri − r)dV = χ · dSi .
V ∂r
extended here to a discrete molecular system, i=1 i=1

N
d X Appendix B: Relation between Control Volume and
χ(t)ϑi (ri (t), r(t), ∆r(t)) Description at a Point
dt
i=1
N     This Appendix proves that the Irving and Kirkwood [8]
X dχ pi expression for the flux at a point is the zero volume limit of
= ϑi + χ ũ − · dSi . (A1)
dt mi the CV formulation. As in the continuum, the control volume
i=1
equations at a point are obtained using the gradient operator
The conservation equation for the mass, χ = mi , in a moving in Eq. (6). the flux at a point can be shown by taking the zero
reference frame is, volume limit of the gradient operator of Eq. (6). Assuming
N N     the three side lengths of the control volume, ∆x, ∆y and ∆z,
d X X p tend to zero and hence the volume, ∆V , tends to zero,
mi ϑi = mi ũ − i · dSi . (A2)
dt mi
i=1 i=1 1
∇ · ρu = lim lim lim
In a Lagrangian reference frame, the translational velocity of ∆x→0 ∆y→0 ∆z→0 ∆x∆y∆z
CV surface must be equal to the molecular streaming veloc- N  
X ∂ϑ ∂ϑ ∂ϑ
ity, i.e., ũ(r± ) = u(ri ), so that, × pix i + piy i + piz i ; f . (B1)
∂x ∂y ∂z
N     N i=1
X p X
mi u − i · dSi = − pi · dSi . from Eq. (21). For illustration, consider the x component
mi
i=1 i=1 above, where
The evolution of the peculiar momentum, χ = pi , in a mov- xf ace
ing reference frame is, ∂ϑi
z }| {
= δ(x+ − xi ) − δ(x− − xi ) Sxi . (B2)
N N     ∂x
d X X pi
pi ϑi = Fi ϑi + pi u − · dSi Using the definition of the Dirac δ function as the limit of two
dt mi
i=1 i=1 slightly displaced Heaviside functions,
N      
X pi pi
= Fi ϑi − · dSi . H ξ + ∆ξ 2 − H ξ − ∆ξ
2
mi δ(ξ) = lim ,
i=1
∆ξ→0 ∆ξ
Here an inertial reference frame has been assumed so that
dpi /dt = dpi /dt = Fi . For a simple case (e.g. one dimen- the limit of the Sxi term is,
sional flow) it is possible to utilize a Lagrangian description lim lim Sxi = δ(yi − y)δ(zi − z)
by ensuring, ũ(r± ) = u(ri ), throughout the time evolution. ∆y→0 ∆z→0
In more complicated cases, this is not always possible and
the Eulerian description is generally adopted. The ∆x → 0 limit for xf ace (defined in Eq. (B2)) can be
Next, a microscopic analogue to the macroscopic diver- evaluated using L’Hôpital’s rule, combined with the property
gence theorem is derived for the generalized function, χ, of the δ function,
   
Z XN   ∂ ∆ξ 1 ∂ ∆ξ
∂ δ ξ− =− δ ξ− ,
· χ(ri , pi , t)δ(ri − r) dV ∂(∆ξ) 2 2 ∂ξ 2
V ∂r
i=1
so that,
N
Z X

= χ(ri , pi , t) · δ(ri − r)dV. ∂
V i=1 ∂r lim xf ace = δ (x − xi ) .
∆x→0 ∂x

056705-16
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

Therefore, the limit of ∂ϑi /∂x as the volume approaches where r+ → r and r− → r. The ̟βκγ function is the
zero is, integral between two molecules introduced in Eq. (37),
∂ϑi ∂ Z1  
lim lim lim = δ (ri − r) , 1 1
∆x→0 ∆y→0 ∆z→0 ∂x ∂x δ(r − ri + srij )ds = sgn
rxij
|rxij |
Taking the limits for the x, y and z terms in Eq. (B1) yields 0
 
the expected Irving and Kirkwood [8] definition of the diver- × H(rx − rxj ) − H(rx − rxi )
gence at a point,  
ryij
×δ ry − ryi − (rx − rxi )
N   rxij
X ∂  
∇ · ρu = · pi δ(ri − r); f . rzij
∂r ×δ rz − rzi − (rx − rxi ) .
i=1 rxij
This zero volume limit of the CV surface fluxes shows that where the sifting property of the Dirac δ function in the rx
the divergence of a Dirac δ function represents the flow of direction has been used to express the integral between two
molecules over a point in space. The advection and kinetic molecules in terms of the ̟xyz function. Hence,
pressure at a point is, from Eq. (25),
Z1
N   ̟xyz
X ∂ pi pi δ(r − ri + srij )ds = .
∇ · [ρuu + κ] = · δ(ri − r); f . rxij
∂r mi 0
i=1

The same limit of zero volume for the surface tractions de- As the choice of shifting direction is arbitrary, use of ry or
fines the Cauchy stress. Using Eq. (6) and taking the limit of rz in the above treatment would result in ̟yzx and ̟zxy , re-
Eq. (46), written in terms of tractions, spectively. Therefore, Eq. (38), without the volume integral,
can be expressed as,
6 Z
1 X N  Z1 
∇ · σ = lim σ · dSf = lim lim lim 1X ∂
∆V →0 ∆V ∆rx →0 ∆ry →0 ∆rz →0 fαij rβij δ(r − ri + srij )ds; f
f aces Sf 2 ∂rβ
" # i,j 0
− T+ − −
T+x − Tx y − Ty T+z − Tz N 
× + + . 1 X 
∂̟xyz ∂̟yxz ∂̟zxy
 
∆rx ∆ry ∆rz = fijα + + ;f .
2 ∂rx ∂ry ∂rz
i,j
For the rx+ surface, and taking the limits of ∆ry and ∆rz
using L’Hôpital’s rule, As Eq. (38) is equivalent to the Irving and Kirkwood [8]
stress of Eq. (36), the Irving Kirkwood stress is recovered in
N  
T+
x 1 X +
the limit that the CV tends to zero volume.
lim = − lim fαij ̟xyz ;f . This Appendix has proved therefore that in the limit of zero
∆V →0 ∆rx ∆rx →0 2∆rx
i,j control volume, the molecular CV Eqs. (22) and (49) recover
where ̟ is the description at a point in the same limit that the contin-
h i uum CV Eqs. (1) and (2) tend to the differential continuum
† † † equations. This demonstrates that the molecular CV equa-
̟βκγ ≡ H(rβ − rβj ) − H(rβ − rβi )
  tions presented here are the molecular scale equivalent of the
rκij  † continuum CV equations.
×δ rκ − rκi − r − rβi
rβij β
  Appendix C: Relationship between Volume Average and
rγij  † Method Of Planes Stress
×δ rγ − rγi − r − rβi . (B3)
rβij β
This Appendix gives further details of the derivation of the
The indices β, κ and γ can be x, y or z and † denotes the top Method Of Planes form of stress from the Volume Average
surface (+ superscript), bottom surface (− superscript) or CV form. Starting from Eq. (38) written in terms of the CV func-
center (no superscript). The ̟ selecting function includes tion for an integrated volume,
only the contribution to the stress when the line of interaction
6 Z N  Z1 
between i and j passes through the point r† in space. The X 1X ∂ϑs
difference between T+ − − σ · dSf = fij rij · ds; f
x and Tx tends to zero on taking the 2 ∂r
limit ∆rx → 0, so that L’Hôpital’s rule can be applied. Using f aces Sf i,j 0
the property, N  Z1   
1 X ∂ϑs ∂ϑs ∂ϑs


1
 
1
 = fij xij + yij + zij ds; f .
δ ξ − ∆ξ H ξ − ∆ξ 2 ∂x ∂y ∂z
i,j
∂(∆ξ) 2 2 0
    (C1)
1 ∂ 1 1
=− δ ξ − ∆ξ H ξ − ∆ξ , Taking only the x derivative above,
2 ∂ξ 2 2
then, x+
f ace
N  ∂ϑs z }| {
= xij δ(x+ − xi + sxij )
− 
T+
x − Tx 1X ∂̟xyz xij
lim =− fαij ;f . ∂x
∆V →0 ∆rx 2 ∂rx 
−δ(x− − xi + sxij ) G(s) (C2)
i,j

056705-17
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

where G(s) is,


  Z1
G(s) ≡ H(y + − yi + syij ) − H(y − − yi + syij ) Z
1X
N  
+
  − σ · dSS + = fij xij xf ace G(s)ds; f
× H(z + − zi + szij ) − H(z − − zi + szij ) . +
Sx x 2
i,j 0
1 δ(x) xij x+ N  
As δ(ax) = |a|
the f ace G(s) term in Eq. (C2) can 1 X  + +  +
= fij sgn(x − xj ) − sgn(x − xi ) Sxij ; f
be expressed as, 4
i,j
 
xij x+ − xi
xij x+
f ace G(s) = |x | δ + s G(s). (C3) Repeating the same process for the other faces allows Eq.
ij xij (C1) to be expressed as,
The integral can be evaluated using the sifting property of the
Dirac δ function [58] as follows, 6 Z N  Z1 
X 1X ∂ϑs
σ · dSf = − fij rij · ds; f
Z1 Z1  + 2 ∂r
f aces Sf
 i,j
xij x − xi 0
xij x+
f ace G(s)ds = δ + s G(s)ds
|xij | xij N  3 h i 
0 0
1 X X
+ −
=− fij ñα dSαij − dSαij ;f ,
  +
x − xj
  +
x − xi
 4
+ i,j α=1
= sgn(xij ) H −H Sxij .
xij xij ±  
where dSαij ± − r ) − sgn(r± − r ) S ±
≡ 21 sgn(rα
+
 αj α αi αij
where the signum function, sgn(xij ) ≡ xij /|xij |. The Sxij 1
and ñα ≡ sgn(rαij )sgn r = [1 1 1]. This is the force
term is thevalue of s on the
 cube surface, αij
+ x+ −x over the CV surfaces,Eq. (46), in section III C.
Sxij =G s= − x i which is, +
To verify the interpretation of Sxij used in this work,
ij
   consider the vector equation for the point of intersection of
+ + yij +  a line and a plane in space. The equation for a vector a
Sxij ≡ H y − yi − x − xi r
xij between ri and rj is defined as a = ri − s |rij | . The
  ij
yij  plane containing the positive face of a cube is defined by
− H y − − yi − +
x − xi 
xij r+ − p · n where p is any point on the plane and n is nor-
 
zij 
 malto that plane. Bysetting a = p and upon rearrangement
× H z + − zi − +
x − xi r
xij of r+ − ri + s |rij | · n, the value of s at the point of inter-
  ij
zij  section with the plane is,
− H z − − zi − x+ − xi . (C4)
xij 
r+ − ri · n
+ s=− rij ,
The definition Sxij (analogous to Sxi in Eq. (15)) has been ·n
|rij |
introduced as it filters out those ij terms where the point
of intersection of line rij and plane x+ has y and z com- The point on line a located on the plane is,
ponents between the limits of the cube surfaces. The cor-
± "  #
responding terms, Sijα , are defined for α = {y, z}. Tak- r+ − ri · n
+
ap ≡ ri + rij .
ing H(0) = 12 , the Heaviside function can be rewritten as rij · n
H(ax) = 21 (sgn(a)sgn(x) − 1), and,
Taking n as the normal to the x surface, i.e.
   
x+ − xj x+ − xi n → nx = [1, 0, 0], then,
H −H
xij xij  
 + x+
1
 
1   xxp y ij 
+
sgn(x+ − xj ) − sgn(x+ − xi ) ,
 
= sgn x+  +  = yi + xij x − xi 
αp = xyp
2 xij  z 
x+zp zi + xij x+ − xi
ij
so the expression, xij x+
f ace G(s) in Eq. (C2) becomes,
written using index notation with α = {x, y, z}. The vector
Z1   x+ +
p is the point of intersection of line a with the x plane. A
1 1 +
xij x+
f ace G(s)ds = sgn(xij )sgn function to check if the point xp on the plane is located on the
2 xij region between y ± and z ± , would use Heaviside functions
0
  + and is similar to the form of Eq. (15),
× sgn(x+ − xj ) −sgn(x+ − xi ) Sxij .
+   − + 
  Sxij = H y + − x+ yp − H y − xyp
1   − + 
The signum function, sgn xij , cancels the one obtained × H z + − x+ zp − H z − xzp ,
from integration along s, sgn(xij ). The expression for the which is the form obtained in the text by direct integration of
x+ face is therefore, the expression for stress, i.e. Eq. (C4).

056705-18
E. R. SMITH, D. M. HEYES, D. DINI, AND T. A. ZAKI PHYSICAL REVIEW E 85, 056705 (2012)

[1] O. Reynolds, Papers on Mechanical and Physical Subjects - [33] W. Ren, J. of Comp. Phys. 227, 1353 (2007).
Volume 3, 1st ed. (Cambridge University Press, Cambridge, [34] M. K. Borg, G. B. Macpherson, and J. M. Reese, Molec. Sims.
1903). 36, 745 (2010).
[2] T. A. Zaki and P. A. Durbin, J. Fluid Mech. 531, 85 (2005). [35] A. I. Borisenko and I. E. Tarapov, Vector and Tensor Analysis
[3] T. A. Zaki and P. A. Durbin, J. Fluid Mech. 563, 357 (2006). with applications, 2nd ed. (Dover Publications Inc, New York,
[4] C. Hirsch, Numerical Computation of Internal and External 1979).
Flows, 2nd ed. (Elsevier, Oxford, 2007). [36] W. Humphrey, A. Dalke, and K. Schulten, J. Molec. Grap.
[5] M. Rosenfeld, D. Kwak, and M. Vinokur, J. Comput. Phys 94, 14.1, 33 (1996).
102 (1991). [37] The cuboid is chosen as the most commonly used shape in
[6] T. A. Zaki, J. G. Wissink, W. Rodi, and P. A. Durbin, J. Fluid continuum mechanic simulations on structured grids, although
Mech. 665, 57 (2010). the process could be applied to any arbitrary shape.
[7] D. J. Evans and G. P. Morriss, Statistical Mechanics of Non- [38] M. Serrano and P. Español, Phys. Rev. E 64, 046115 (2001).
Equilibrium Liquids, 2nd ed. (Australian National University [39] A. K. Subramaniyan and C. T. Sun, J. Elast. 88, 113 (2007).
Press, Canberra, 2007). [40] W. G. Hoover, C. Hoover, and J. Lutsko, Phys. Rev. E 79,
[8] J. H. Irving and J. G. Kirkwood, J. Chem.. Phys. 18, 817 036709 (2009).
(1950). [41] The resulting equality satisfies Eq. (39) and both sides are
[9] M. Zhou, Proc. R. Soc. Lond. 459, 2347 (2003). equal to within an arbitrary constant (related to choosing the
[10] E. N. Parker, Phys. Rev. 96, 1686 (1954). gauge).
[11] W. Noll, Phys. Rev. 96, 1686 (1954). [42] S. Nemat-Nasser, Plasticity: A Treatise on the Finite Deforma-
[12] D. H. Tsai, J. Chem. Phys. 70, 1375 (1978). tion of Heterogeneous Inelastic Materials, 1st ed. (Cambridge
[13] B. D. Todd, D. J. Evans, and P. J. Daivis, Phys. Rev. E 52, University Press, Cambridge, 2004).
1627 (1995). [43] K. M. Mohamed and A. A. Mohamad, Microfluidics and
[14] M. Han and J. Lee, Phys. Rev. E 70, 061205 (2004). Nanofluidics 8, 283 (2009).
[15] R. J. Hardy, J. Chem. Phys 76, 622 (1982). [44] R. Delgado-Buscalioni, Lecture Notes in Computational Sci-
[16] J. F. Lutsko, J. Appl. Phys 64, 1152 (1988). ence and Engineering 82, 145 (2012).
[17] J. Cormier, J. Rickman, and T. Delph, J. Appl. Phys 89, 99 [45] R. Delgado-Buscalioni and P. Coveney, Phil. Trans. R. Soc.
(2001). Lond. 362, 1639 (2004).
[18] A. I. Murdoch, J. Elast. 88, 113 (2007). [46] G. D. Fabritiis, R. Delgado-Buscalioni, and P. Coveney, Phys.
[19] A. I. Murdoch, J. Elast 100, 33 (2010). Rev. Lett. 97, 134501 (2006).
[20] P. Schofield and J. R. Henderson, Proc. R. Soc. Lond. A 379, [47] R. Delgado-Buscalioni and G. D. Fabritiis, Phys. Rev. E 76,
231 (1982). 036709 (2007).
[21] N. C. Admal and E. B. Tadmor, J. Elast. 100, 63 (2010). [48] O. Zienkiewicz, The Finite Element Method: Its Basis and
[22] D. M. Heyes, E. R. Smith, D. Dini, and T. A. Zaki, J. Chem. Fundamentals, 6th ed. (Elsevier Butterworth-Heinemann, Ox-
Phys 135, 024512 (2011). ford, 2005).
[23] S. T. O’Connell and P. A. Thompson, Phys. Rev. E 52, R5792 [49] P. J. Daivis, K. P. Travis, and B. D. Todd, J. Chem. Phys 104,
(1995). 9651 (1996).
[24] N. G. Hadjiconstantinou, Hybrid AtomisticContinuum Formu- [50] M. P. Allen and D. J. Tildesley, Computer Simulation of Liq-
lations and the Moving Contact-Line Problem, Ph.D. thesis, uids, 1st ed. (Clarendon Press, Oxford, 1987).
MIT(U.S.) (1998). [51] D. C. Rapaport, The Art of Molecular Dynamics Simulation,
[25] J. Li, D. Liao, and S. Yip, Phys. Rev. E 57, 7259 (1997). 2nd ed. (Cambridge University Press, Cambridge, 2004).
[26] N. G. Hadjiconstantinou, J. Comp. Phys. 154, 245 (1999). [52] W. G. Hoover, Computational Statistical Mechanics, 1st ed.
[27] E. G. Flekkøy, G. Wagner, and J. Feder, Europhys. Lett. 52, (Elsevier Science, Oxford, 1991).
271 (2000). [53] J. Petravic and P. Harrowell, J. Chem. Phys. 124, 014103
[28] G. Wagner, E. Flekkøy, J. Feder, and T. Jossang, Comp. Phys. (2006).
Comms. 147, 670 (2002). [54] M. C. Potter and D. C. Wiggert, Mechanics of Fluids, 3rd ed.
[29] R. Delgado-Buscalioni and P. Coveney, Phys. Rev. E 67, (Brooks/Cole, California, 2002).
046704 (2003). [55] W. A. Strauss, Partial Differential Equations, 1st ed. (John Wi-
[30] W. A. Curtin and R. E. Miller, Modelling Simul. Mater. Sci. ley and Sons, New Jersey, 1992).
Eng. 11, R33 (2003). [56] F. D. C. Silva, L. A. F. Coelho, F. W. Tavares, and M. J. E. M.
[31] X. B. Nie, S. Chen, W. N. E, and M. Robbins, J. of Fluid Cardoso, J. Quantum Chem. 95, 79 (2003).
Mech. 500, 55 (2004). [57] N. G. Hadjiconstantinou, A. L. Garcia, M. Z. Bazant, and
[32] T. Werder, J. H. Walther, and P. Koumoutsakos, J. of Comp. G. He, J. Comp. Phys. 187, 274 (2003).
Phys. 205, 373 (2005). [58] V. Thankoppan, Quantum Mechanics, 1st ed. (New Age pub,
New Delhi, 1985).

056705-19

Das könnte Ihnen auch gefallen