Appendix A: DL_POLY_4 Dissipative Particle Dynamics

Introduction

Although in a Molecular Dynamics sense Dissipative Particle Dynamics (DPD) is regarded as a type of thermostat, on its own it is an off-lattice, discrete particle method for modelling mesoscopic systems in a fluid state. It has little in common with Lattice Gas Automata and Lattice Boltzmann dynamics methods 42, except in its application to systems of similar length and time scales, and last but not least that it captures hydrodynamics behaviour.

The DPD method inherits its methodology from Brownian/Langevin Dynamics (BD). However, it differs from BD in an important way: it is Galilean invariant and for this reason conserves hydrodynamic behaviour, while the BD method does not (its microscopic behaviour is only diffusive. Many systems in their fluid state are crucially dependent on hydrodynamic interactions and it is essential to retain this feature in their models. DPD is particularly useful for simulating coarse-grained systems on the near-molecular scale, such as polymers, biopolymers, lipids, emulsions and surfactants – systems in which large scale structure evolves on a time scale that is too long to be modelled effectively by traditional MD. The particles in DPD 28 are not regarded as molecules in a fluid but as lumps of molecules grouped to form a fluid particle in much the same spirit as the renormalisation group has been applied in polymer physics where lumps of monomers are grouped to form a bead. Hence, the beads are regarded as carriers of momentum.

It is worth noting that DPD may also be used when such systems experience shear and flow gradients.

Outline of Method

Following 39 the DPD algorithm can be summarised by the following:

  • A condensed phase system may be modelled as a system of ‘free’ particles interacting directly through soft forces. Note that DL_POLY_4 allows for the application of the DPD thermostat beyond systems of free particles only. Thus it will be valid on systems with any inratamolecular like interactions.

  • The system is coupled to a heat bath via stochastic forces, which act on the particles in a pairwise manner.

  • The particles also experience a damping or drag force, which also acts in a pairwise manner.

  • Thermodynamic equilibrium is maintained through the balance of the stochastic and drag forces, i.e. the method satisfies the fluctuation-dissipation theorem.

  • At equilibrium (or steady state) the properties of the system are calculated as averages over the individual particles, as in traditional Molecular Dynamics.

Therefore, the equation of motion are the same as these for the microcanonical ensemble (NVE) but force, \(f_{i}\), on particle \(i\) is now a sum of pair forces:

(172)\[\underline{f}_i = \sum_{j \neq i}^N \left( \underline{f}_{ij}^{C} + \underline{f}_{ij}^{D} + \underline{f}_{ij}^{R} \right)~~, \label{DPD}\]

in which \(\underline{f}_{ij}^{C}\), \(\underline{f}_{ij}^{D}\) and \(\underline{f}_{ij}^{R}\) are the conservative, drag and random (or stochastic) pair forces respectively. Each represents the force exerted on particle \(i\) due to the presence of particle \(j\).

The conservative interactions are usually soft (i.e. weakly interacting) so that the particles can pass by each other (or even through each other) relatively easily so that equilibrium is achieved quickly. A common form of interaction potential is an inverse parabola (part of VDW types of potentials, see Section Short Ranged (van der Waals) Potentials)

(173)\[\begin{split}V(r_{ij}) = \left\{ \begin{array} {l@{\quad:\quad}l} \frac{A_{ij}}{2}~r_{c}~\left(1-\frac{r_{ij}}{r_{c}}\right)^{2} & r_{ij} < r_{c} \\ 0 & r_{ij} \ge r_{c} \end{array} \right.~~, \label{DPDU}\end{split}\]

where \(r_{ij} = |\underline{r}_{j}-\underline{r}_{i}|\), \(r_{c}\) is a cutoff radius and \(A_{ij}\) is the interaction strength (that may be the same for all particle pairs or may be different for different particle types).

Equation (172) gives rise to a repulsive force of the form:

(174)\[\underline{f}_{ij}^{C} = A_{ij}~w^{C}(r_{ij}) \frac{\underline{r}_{ij}}{r_{ij}} = A_{ij} \left( 1 - \frac{r_{ij}}{r_{c}} \right) \frac{\underline{r}_{ij}}{r_{ij}}~~. \label{DPDF}\]

This is the deterministic or conservative force \(\underline{f}_{ij}^{C}\) exerted on particle \(i\) by particle \(j\). Note the switching function:

(175)\[\begin{split}w^{C}(r_{ij}) = \left\{ \begin{array} {l@{\quad:\quad}l} \left(1-\frac{r_{ij}}{r_{c}}\right) & r_{ij} < r_{c} \\ 0 & r_{ij} \ge r_{c} \end{array} \right.~~, \label{DPDS}\end{split}\]

and the force are zero when \(r_{ij} \ge r_{c}\) and thus the particles have an effective diameter of \(1\) in units of the cutoff radius \(r_{c}\). In the DL_POLY_4 context all inter- and intra-molecular forces will fall into this category of force!

The stochastic forces experienced by the particles is again pairwise in nature and takes the form:

\[\underline{f}_{ij}^{R} = \sigma_{ij} w^{R}(r_{ij}) \zeta_{ij} \Delta t^{-\frac{1}{2}} \frac{\underline{r}_{ij}}{r_{ij}}~~,\]

in which \(\Delta t\) is the time step and \(w^{R}(r_{ij})\) is a switching function which imposes a finite limit on the range of the stochastic force. \(\zeta_{ij}\) is a random number with zero mean and unit variance. The constant \(\sigma_{ij}\) is related to the temperature, as is understood from the role of the stochastic force in representing a heat bath.

Finally, the particles are subject to a drag force, which depends on the relative velocity between interacting pairs of particles:

\[\underline{f}_{ij}^{D} = -\gamma_{ij} w^{D}(r_{ij}) \left(\underline{r}_{ij} \cdot \underline{v}_{ij}\right) \frac{\underline{r}_{ij}}{r_{ij}^2}~~,\]

where \(w^{D}(r_{ij})\) is once again a switching function and \(\underline{v}_{ij} = \underline{v}_{j}-\underline{v}_{i}\) is the inter-particle relative velocity. The constant \(\gamma_{ij}\) is the drag coefficient. It follows from the fluctuation-dissipation theorem that for thermodynamic equilibrium to result from this method the following relations must hold:

(176)\[\sigma_{ij}^2 = 2~\gamma_{ij} k_{B} T \label{DPDC1}\]
(177)\[w^{D}(r_{ij}) = \left[w^{R}(r_{ij})\right]^{2}. \label{DPDC2}\]

In practice, the switching functions are defined through:

\[w^{R}(r_{ij}) = \left[w^{C}(r_{ij})\right]^{2}~~,\]

which ensures that all interactions are switched off at the range \(r_{ij} = r_{c}\).

In many DPD simulations, the stochastic and drag coefficients are often constant for all interactions, i.e. \(\sigma_{ij} \equiv \sigma\) and \(\gamma_{ij} \equiv \gamma\), although this assumption does not have to apply. In DL_POLY_4 the \(\gamma_{ij}\) coefficients may be supplied at the end of each specified vdw interaction potential as a parameter further to the last one for the particular vdw potential form. For a DPD thermostat to work correctly all possible two body interactions must be defined and all \(\gamma_{ij} \ne 0\). What DL_POLY_4 will attempt first, if a two body interaction is missing, is to derive it using mixing rules (default may be overridden by user specification). However, if any of \(\gamma_{ij} = 0\) then DL_POLY_4 will check for the existence of a global \(\gamma\) that may be optionally supplied by the user on the ensemble nvt dpdINT line and if it is non-zero a global override will occur. Otherwise, when the requirements for a DPD thermostat are not satisfied, everything else will result in a controlled termination.

Equation of state and dynamic properties

The form of the conservative force determines the equation of state for a DPD fluid, which can be derived using the virial theorem to express system pressure as follows:

\[\begin{split}\begin{aligned} {\cal P} =& \rho k_{B}T + \frac{1}{3V} \left\langle \sum_{j>i} (\underline{r}_i-\underline{r}_j) \cdot \underline{f}_{ij}^{C} \right\rangle \\ =& \rho k_{B} T + \frac{2 \pi}{3} \rho^{2} \int_{0}^{r_{c}} A \left( 1 - \frac{r}{r_{c}} \right) r^{3} g(r)~dr~~, \end{aligned}\end{split}\]

where \(g(r)\) is a radial distribution function for the soft sphere model 39 and \(\rho\) is the DPD particle density. For sufficiently large densities (\(\rho > 2\)), \(g(r)\) takes the same form and the equation of state can be well-approximated by:

\[{\cal P} = \rho k_{B}T + \alpha A \rho^{2}~~,\]

where the parameter \(\alpha \approx 0.101 \pm 0.001\) has units equivalent to \(r_{c}^{4}\). This expression permits the use of fluid compressibilities to obtain conservative force parameters for bulk fluids, e.g. for water \(A \approx 75 k_{B} T/\rho\). Alternative equations of state may be obtained by modifying the functional form of conservative interactions to include localized densities (i.e. many-body DPD) 70,118.

Transport coefficients for a DPD fluid without conservative forces can be derived using the expressions for the drag and stochastic forces:cite:groot-97a,koelman-93a,marsh-97a. The kinematic viscosity can be found to be

\[\nu \approx \frac{45 k_{B} T}{4 \pi \gamma \rho r_{c}^{3}} + \frac{2 \pi \gamma \rho r_{c}^{5}}{1575}~~,\]

while the self-diffusion coefficient is given as

\[D \approx \frac{45 k_{B} T}{2 \pi \gamma \rho r_{c}^{3}}.\]

The ratio of these two properties, the Schmidt number (\(\textnormal{Sc} = \nu / D\)), is therefore:

\[\textnormal{Sc} \approx \frac{1}{2} + \frac{(2 \pi \gamma \rho r_{c}^{4})^{2}}{70875 k_{B} T}\]

and for values of the drag coefficient and density frequently used in DPD simulations, this value is of the order of unity, which is an appropriate magnitude for gases but three orders of magnitude too small for liquids.

This property of standard DPD does not rule it out for simulations of liquid phases except when hydrodynamics are important. It may also be argued that the self-diffusion of DPD particles might not correspond to that of individual molecules and thus a Schmidt number of the order \(10^{3}\) is unnecessary for modelling liquids 72. Alternative thermostats are available in the DL_MESO 91 (http://www.ccp5.ac.uk/DL_MESO/) package, which can model systems with higher Schmidt numbers 57,106.

Derivation of Equilibrium

The derivation of the DPD algorithm is based on the Fokker-Planck equation

(178)\[\frac{\partial \rho}{\partial t} = \mathcal{L} \rho \label{FokkerPlanck}\]

where \(\rho\) is the equilibrium distribution function and \(\mathcal{L}\) is the evolution operator, which may be split into conservative and stochastic+dissipative parts:

\[\mathcal{L} = \mathcal{L}^{C} + \mathcal{L}^{R+D}\]

with

(179)\[\begin{split}\begin{aligned} \mathcal{L}^{C} =& -\sum_{i=1}^{N} \frac{\underline{p}_{i}}{m_{i}} \frac{\partial}{\partial \underline{r}_{i}} - \sum_{i \neq j}^{N} \underline{f}_{ij}^{C} \frac{\partial}{\partial \underline{p}_{i}} \\ \mathcal{L}^{R+D} =& \sum_{i=1}^{N} \hat{e}_{ij} \cdot \frac{\partial}{\partial \underline{p}_{i}} \left[ \frac{\sigma^{2}}{2} \left\{w^{R} \left(r_{ij} \right) \right\}^{2} \hat{e}_{ij} \cdot \left\{ \frac{\partial}{\partial \underline{p}_{i}} - \frac{\partial}{\partial \underline{p}_{j}} \right\} + \gamma w^{D} \left( \hat{e}_{ij} \cdot \underline{v}_{ij} \right) \right]~~, \label {DPDEvolution}\end{aligned}\end{split}\]

where \(\hat{e}_{ij} = \frac{\underline{r_{ij}}}{r_{ij}}\).

When \(\sigma = \gamma = 0\) then equation (178) becomes

\[\frac{\partial \rho}{\partial t} = \mathcal{L}^{C} \rho~~,\]

for which the equilibrium solution is evidently

\[\rho^{eq} = \frac{1}{Z} \exp \left( \frac{1}{k_{B} T} \left[ \sum_{i=1}^{N} \frac{p_{i}^{2}}{2 m_{i}} + \frac{1}{2} \sum_{j \neq i}^{N} \phi (r_{ij}) \right] \right)\]

which is, of course, the Boltzmann distribution function for an equilibrium system. Thus it is apparent that for the simulation based on equation (178) to maintain the same distribution function, the terms in the operator \(\mathcal{L}^{R+D}\) of equation (179) must sum to zero. It follows that the conditions given in equations (176) and (177) must apply.

Summary of Dissipative Particle Dynamics

DPD is a simple method that can be viewed as a novel thermostatting method for molecular dynamics. All that is required is a system of spherical particles enclosed in a periodic box undergoing time evolution as a result of the above forces. It should be noted that all computed interactions are pairwise, which means that the principle of the conservation of momentum in the system, or Galilean invariance, is preserved. The conservation of momentum is required for the preservation of hydrodynamic forces. Therefore, the DPD method is an NVT method that preserves hydrodynamics. The presence of hydrodynamics is important in annealing defects in ordered mesophases 37. Thus DPD has an intrinsic advantage over other methods such as traditional molecular dynamics, dynamic density functional theory (which are purely diffusive!) or Monte Carlo methods, in trying to evolve a system towards an ordered thermodynamic equilibrium state.