epje - homepage.univie.ac.at · dynamics simulations ... langevin, dissipative particle dynamics,...

14
EPJ E Soft Matter and Biological Physics your physics journal EPJ .org Eur. Phys. J. E (2018) 41: 80 DOI 10.1140/epje/i2018-11689-4 On the effect of the thermostat in non-equilibrium molecular dynamics simulations Jos´ e Ruiz-Franco, Lorenzo Rovigatti and Emanuela Zaccarelli

Upload: others

Post on 02-Nov-2019

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

EPJ ESoft Matter and Biological Physics

your physics journal

EPJ .org

Eur. Phys. J. E (2018) 41: 80 DOI 10.1140/epje/i2018-11689-4

On the effect of the thermostat in non-equilibriummolecular dynamics simulations

Jose Ruiz-Franco, Lorenzo Rovigatti and Emanuela Zaccarelli

Page 2: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

DOI 10.1140/epje/i2018-11689-4

Regular Article

Eur. Phys. J. E (2018) 41: 80 THE EUROPEANPHYSICAL JOURNAL E

On the effect of the thermostat in non-equilibrium moleculardynamics simulations⋆

Jose Ruiz-Franco1,a, Lorenzo Rovigatti1,2, and Emanuela Zaccarelli1,2

1 Dipartimento di Fisica, Sapienza Universita di Roma, P.le A. Moro 5, 00185 Roma, Italy2 CNR Institute for Complex Systems (ISC), Uos Sapienza, Roma, Italy

Received 5 February 2018 and Received in final form 24 May 2018Published online: 2 July 2018c© EDP Sciences / Societa Italiana di Fisica / Springer-Verlag GmbH Germany, part of Springer Nature,2018

Abstract. The numerical investigation of the statics and dynamics of systems in non-equilibrium in general,and under shear flow in particular, has become more and more common. However, not all the numericalmethods developed to simulate equilibrium systems can be successfully adapted to out-of-equilibrium cases.This is especially true for thermostats. Indeed, even though thermostats developed to work under equilib-rium conditions sometimes display good agreement with rheology experiments, their performance rapidlydegrades beyond weak dissipation and small shear rates. Here we focus on gauging the relative perfor-mances of three thermostats, Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello undervarying parameters and external conditions. We compare their effectiveness by looking at different observ-ables and clearly demonstrate that choosing the right thermostat (and its parameters) requires a carefulevaluation of, at least, temperature, density and velocity profiles. We also show that small modificationsof the Langevin and DPD thermostats greatly enhance their performance in a wide range of parameters.

1 Introduction

In molecular dynamics (MD) simulations, the motion ofN individual particles in a specified volume V evolvesaccording to Newton’s laws. Hence, since the energy isconserved, from a thermodynamic point of view the re-sulting system is a microcanonical ensemble (NV E) [1,2].However, in order to mimic experimental conditions it isoften necessary to simulate systems at constant temper-ature rather than energy, obtaining a canonical ensemble(NV T ). The control of the temperature is achieved bycoupling the system to a so-called “thermostat”, whichacts as a thermal bath. During the course of the yearsmany different thermostats have been developed to notonly reduce the side effects due to the coupling, but alsoto more accurately reproduce the phenomena observed inexperiments. For instance, many thermostats that exhibita good temperature control do not correctly reproduce hy-drodynamics, which require local momentum conservationand Galilean invariance. There exist methods that explic-itly incorporate in the simulation solvent particles, suchas multi-particle collision dynamics (MPCD) [3–5], thus

⋆ Contribution to the Topical Issue “Advances in Computa-tional Methods for Soft Matter Systems” edited by LorenzoRovigatti, Flavio Romano, John Russo.

a e-mail: [email protected]

naturally building the correct hydrodynamics in the sys-tems. However, these methods lie outside the scope of thepresent work, which deals with implicit solvent treatmentonly.

In the following excursus we provide a non-compre-hensive list of well-known thermostats. We consider onlythermostats that take into account the effects of the sol-vent in an implicit way. We start with the well-knownBerendsen thermostat [6]. It consists in rescaling all par-ticle velocities after a certain number of time steps so thatthe (instantaneous) kinetic energy matches the target one.This guarantees a constant thermodynamic temperature.The main drawback is that it does not sample the NV Tensemble [7], which implies that it is dangerous to use itfor data production. Furthermore, it is not Galilean in-variant and it does not locally conserve momentum.

A very simple, yet effective thermostat is the Ander-sen thermostat [8]. Here the coupling between the systemand a “heat bath” is explicit: particles undergo randomcollisions with the (fictitious) solvent, effectively acquir-ing new momenta extracted from a Maxwell-Boltzmanndistribution corresponding to the desired temperature T .This thermostat is local, but it is not Galilean invariant.

The Nose-Hoover (NH) thermostat comes next [9,10].It introduces in the Hamiltonian of the system an addi-tional internal degree of freedom which acts as an effec-tive friction parameter and represents the thermostat cou-

Page 3: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

Page 2 of 13 Eur. Phys. J. E (2018) 41: 80

pling. Once again, this thermostat is non-Galilean invari-ant. Furthermore, the additional degree of freedom altersthe dynamics, leading to artificial hydrodynamics [11]. Acommon improvement over the original implementation isto consider Nose-Hoover chains, that is, a NH thermostatwith more than one thermostat variable [12]. This ther-mostat is global, since the instantaneous value of the tem-perature is based on a global definition, and non-Galileaninvariant.

The Langevin thermostat [13], which will be discussedin depth in what follows, guarantees ergodicity in all pos-sible cases. Dissipative and noise forces are added to theHamiltonian to include the effective behaviour of the sol-vent [1]. Its main disadvantage is that it does not repro-duce hydrodynamics [14] because it is not Galilean invari-ant and does not locally conserve momentum. In order toovercome such a limitation dissipative particle dynamics(DPD) was introduced by Hoogerbrugge and Koelman [15]and later modified by Espanol and Warren [16] to satisfythe fluctuation-dissipation theorem. This thermostat rep-resents a modification of the Langevin thermostat. In theDPD thermostat the friction and noise terms are pairwiseand act over all pairs of neighbouring particles, i.e., at thelocal level. In addition, the Galilean invariance is ensuredby the fact that the drag force acts on the relative veloc-ity. However, it has several disadvantages. The most im-portant are: i) in order to employ large time steps to effec-tively speed up the simulation, care has to be taken whenchoosing the right integration algorithm [17–21]; ii) theSchmidt number Sc = η/ρD, defined as the ratio betweenthe viscosity η and the diffusion constant D, is found tobe close to 1, while for most common liquids Sc is of theorder of 103. iii) The value of the size of DPD particles,set by rc, cannot be determined a priori [22]. In general,the value of Sc can be improved upon by incorporating athird parameter (s) which modifies the weighting functionfor the dissipative force [23,24]. As we will discuss below,Sc can be written as a function of s, providing us withan efficient way (in computational terms) to increase thisnumber.

In addition, different improvements of the DPD schemehave been developed. For instance, in order to conserveangular momentum, it was proposed to introduce an ad-ditional variable [25, 26]. Moreover, if the objective is tostudy chemical reactions where there is a temperaturegradient, it is necessary to introduce an additional en-ergy term stemming from the interaction between pairsof particles [27]. Finally, DPD can be also combined withSmoothed Particle Hydrodynamics (SPH) [28, 29] to in-clude the Navier-Stokes equations in the modelling of thesolvent particles with the so-called Smoothed DissipativeParticle Dynamics (SDPD) [30, 31]. It should be notedthat DPD can also be used to model an explicit solvent,for example by considering solute and solvent particleswith different masses and sizes [32,33].

More recently, different thermostats implemented ascombinations of the above have been proposed. For in-stance, Stoyanov and Groot [11] introduced a combina-tion of the Lowe-Andersen (LA) and NH thermostats. Afraction of particles is thermalized with the NH thermo-

stat, while the others are thermalized with the LA. Theresulting thermostat is local, Galilean invariant and sta-ble even for large time steps Δt. However, it presents twomain problems: i) for Δt → 0 it does not converge to thestandard DPD, and ii) it has been shown that it does notsample the canonical ensemble.

Finally, we mention one last thermostat: the Bussi-Donadio-Parrinello (BDP) [7] thermostat. The BDP is areformulation of the Berendsen thermostat in which themomenta rescaling factor α is computed according to astochastic evolution of the kinetic energy rather than toa fixed value. Such a change makes the BDP correctlysample the canonical ensemble. In its global, rather thanlocal, implementation, hydrodynamics is not correctly re-produced and the specific value of the rescaling frequencydoes not affect the dynamics [34].

In this work we focus on non-equilibrium moleculardynamics (NEMD) simulations, aimed in particular to de-scribe rheology experiments. As in equilibrium, a goodcontrol of the temperature is necessary. However, it is alsoimperative that NEMD simulations faithfully reproduce arealistic dynamics, without any artefacts introduced bythe thermostat. This, in turn, requires a detailed assess-ment of the effects that the thermostat could artificiallyinduce on the behaviour of the system. To answer thisquestion, we consider a representative model system, i.e.a Lennard-Jones fluid, under steady shear flow and mon-itor its behavior under the action of three different ther-mostats (Langevin, DPD and BDP) for a wide choice ofparameters. In recent years both the Langevin [35–37]and DPD approaches [38–40] have been used to performNEMD simulations. The DPD approach has sometimesbeen the preferred choice because it better reproduces hy-drodynamic effects, while the BDP thermostat allows toconsider a wider range of shear rates [41]. However, wewill show that a poor choice of the thermostat param-eters can negatively affect the dynamic response of thesystem under shear, providing a physical picture very farfrom reality. While there exist direct comparisons betweenLangevin and DPD thermostats, for instance in the caseof coarse-grained bead-spring models for polymers [42],their absolute or relative performances as a function ofthe different parameters are seldom discussed in depth.Here we aim to fill this gap, providing a reference testcase which can be used as a guidance for choosing a ther-mostat and its parameters to carry out reliable NEMDsimulations. Finally, we note that we are interested in bulksystems, and hence we will consider systems where peri-odic boundary conditions are enforced, as we will describebelow. However, NEMD simulations can also be performedwith wall boundaries that constrain the system in a well-defined geometry. The walls can then be moved to repro-duce an oscillatory or steady shear to drive the systemaway from equilibrium [43,44]. In addition, it is also pos-sible to play with the nature of the wall in order to studyslip effects [45–47]. Also in this case, different thermostat-ing strategies are available [48].

We note on passing that the integration algorithms arealso an important point to consider [21], even if we do notexplicitly focus on this aspect.

Page 4: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

Eur. Phys. J. E (2018) 41: 80 Page 3 of 13

The paper is organised as follows. In sect. 2 we in-troduce the simulation details and describe in detail thethree thermostats studied in this work. We then discussthe different observables which we have measured to studythe behaviour of the thermostats. In sect. 3 the effects ofthe different thermostats are shown under both equilib-rium and shear flow conditions. In sect. 4 we draw ourconclusions.

2 Simulations details

2.1 Equilibrium

We study a system composed of N = 2000 monodis-perse particles of mass m and size σ interacting througha Lennard-Jones potential

V (r) = 4ǫ

[

r

)12

−(σ

r

)6]

, (1)

where ǫ controls the depth of the potential. The param-eters σ and ǫ are chosen as units of length and energy,respectively. We also set kB = 1. The potential is cut atrc = 2.5σ. We fix the number density of the system toρ = 0.844 and study two different temperatures, T = 1.5and T = 0.722, corresponding to liquid-like states in thesupercritical region and close to the triple point, respec-tively. In order to reduce the numerical errors intrinsic tointegration schemes [21, 49, 50], we fix the time step toΔt = 0.002 for both equilibrium and NEMD simulations.

2.2 Steady shear

Shear flow is applied by using Lees-Edwards boundaryconditions [51], where different layers of image boxes inthe direction of the flow gradient are considered as movingwith a shear velocity vs. Hence, the shear rate is γ = vs/L,where L is the length of the simulation box. The mod-ified periodic boundary conditions thus impose a time-dependent linear shear flow in the x direction such thatthe shear gradient is parallel to z and the vorticity is alongthe y direction. The flow velocity u = γzx depends lin-early on z (planar Couette flow) and is zero in the centreof the channel (z = L/2). In this work we use γ = 0.01 andγ = 0.1. We notice that these conditions can be also inves-tigated by solving the SLLOD equations of motion [52],for which several thermostats have been adapted [53, 54].However, this algorithm has not been used in this work.

In general, the energy injected by the shear flowneeds to be dissipated by the thermostat. The mecha-nism through which this happens affects the flow pro-file. There are two classes of thermostats that allow todissipate the extra energy: profile-unbiased thermostats(PUT), which allow the velocity profile to emerge as acharacteristic response of the system and profile-biasedthermostats (PBT), which enforce a fixed streaming ve-locity profile [55]. In the following we will present and usethermostats of both kinds.

2.3 Langevin thermostat

The equations of motion for the Langevin dynamics are

miri = vi, miri =∑

j( �=i)

FCij + FR

i + FDi , (2)

where the first term is the usual conservative pairwiseforce, FD

i = −ξmvi is the dissipative (drag) force withfriction constant ξ and FR

i is the random force due tothe thermal motion of the bath particles, modelled as astochastic white noise with zero mean and variance

FRi (t) · FR

j (t′)⟩

=√

2kBTξδijδ(t − t′). (3)

The variance of the stochastic force is set by thefluctuation-dissipation theorem. The combination of dis-sipative and random forces is used to represent the ef-fect of the solvent on the system. The parameter ξ ef-fectively controls the viscosity of such fictitious solvent.Depending on its value, the system can be found eitherin the Brownian (overdamped Langevin dynamics) or inthe Langevin regime (underdamped Langevin dynamics).The integration of eq. (2) is carried out with the self-adaptive OVRVO [56] scheme: if ξΔt ≫ 1, the algorithmreduces to the Euler-Maruyama method [57], whereas forξΔt ≪ 1 the algorithm is equivalent to the Langevinmethod [58]. This algorithm is a profile-unbiased ther-mostat, which is suitable for both equilibrium and non-equilibrium dynamics.

We further implement two variants of the Langevinthermostat. The first one consists in turning off the conser-vative and dissipative forces along one or two directions.This modification has been sometimes shown to be ableto correct some spurious artefacts appearing under non-equilibrium dynamics [38]. In what follows the Langevinthermostat acting on one, two and three directions will beindicated with ξy, ξyz and ξxyz, respectively. In the secondvariant, the drag force acts on the peculiar velocity, thatis, on the difference between the absolute velocity vi, andthe streaming velocity field or shear flow u, rather thanon the absolute velocity. With this modification eq. (2)becomes

miri =∑

j( �=i)

FCij − ξ (pi − mui) + FR

i . (4)

This enforces the velocity profile to be linear (profile-biased conditions). We integrate, eq. (4) with the so-calledBAOAB [50] algorithm. The use of OVRVO and BAOABalgorithms do not present any difference in equilibrium,when u = 0. In the following we will use the symbol ξpec

to refer to this version of the Langevin thermostat, actingon the “peculiar” velocity.

2.4 Dissipative Particle Dynamics

The DPD thermostat locally conserves momentum be-cause all forces act between pairs of particles. It is alsoGalilean invariant because the dissipative force acts on

Page 5: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

Page 4 of 13 Eur. Phys. J. E (2018) 41: 80

relative velocities only. Since it does not explicitly enforcea linear velocity profile, the DPD thermostat is a PUT.The equations of motion are written as

miri = vi, miri =∑

j �=i

[

FCij + FR

ij + FDij

]

, (5)

where we use the same notation as above and the pairwiserandom and dissipative forces are given by

FRij =

2kBTξwR (rij) θij rij , (6)

FDij = −ξwD (rij) [rij · vij ] rij , (7)

where vij = vi − vj and rij = ri − rj are the relativevelocities and distances of particles i and j, respectively,and rij = rij/rij . The variable θij = θji is a Gaussiannoise term with zero mean and variance given by

〈θij(t)θkl(t′)〉 = (δikδjl + δilδjk) δ(t − t′). (8)

The functions wD(rij) and wR(rij) are weight func-tions. They are related by the stationary solution of theFokker-Planck equation where the dissipation-fluctuationtheorem is satisfied through the relation [16]

wD (rij) =[

wR (rij)]2

. (9)

Consequently, one of these two weight functions can bechosen freely. Fan et al. introduced a generalized weightingfunction for the dissipative force [23]

ωD (rij) =[

ωR (rij)]2

=

(

1 − rij

rc

)s

, rij < rc,

0, rij ≥ rc,

(10)

where s and rc are the exponent and the cutoff radius ofthe weighting function, respectively. The latter parame-ter defines the size of DPD particles [22]. Taking rc = 1and s = 2 we recover the conventional DPD algorithm, asobtained by Groot and Warren [59].

The three parameters of the thermostat, i.e. s, rc andξ determine the transport properties of the system [23,60].Indeed, in ref. [24] it is shown that both the shear viscosityη and the constant diffusion D, and thus also the Schmidtnumber, all depend on the three DPD parameters. How-ever, the value of rc greatly affects the computational cost,since it controls the number of pairs of particles that en-ters into the thermostating procedure. Therefore, changesto s or ξ present a more efficient way to obtain a more re-alistic value of Sc. Furthermore, as it will be shown below,they play an important role not only in the control of thetemperature, but also in the generic dynamic response ofthe system under shear.

The DPD equations of motion can be integrated bydifferent schemes [21], the two most famous ones beingthe Lowe-Andersen [17] and Peters [20] algorithms. Thefirst one is inspired by the Andersen thermostat, but in-stead of thermalizing the velocity of individual particles, itthermalizes the relative velocities of two neighbouring par-ticles with a probability P = ΓΔt. However, for Δt → 0 it

does not converge to the standard DPD system. This issueis resolved by adopting the Peters scheme, as done in thiswork. Two observations are in order. First of all, we havetested that the thermalization does not need to follow arandom order, as suggested in ref. [20], but can be carriedout sequentially, as also mentioned in ref. [21]. Secondly,it has been suggested that, for the sake of efficiency, Gaus-sian random numbers of zero mean and unitary variancecan be well-approximated by using a uniform distributionbetween ±

√3 [61]. As shown below (see fig. 5(b)), this is

not always the case, as there exist combinations of param-eters for which non-Gaussian random numbers yield thewrong temperature.

2.5 Bussi-Donadio-Parrinello (BDP) thermostat

The BDP thermostat is similar in spirit to the Berend-sen thermostat, because the velocities of the particles are

rescaled by a factor α [6]. In both cases α =√

Kt

K , where

Kt is the target value of the kinetic energy at time t and Kis the instantaneous kinetic energy. However, in the BDPcase Kt is not constant but evolves stochastically accord-ing to the canonical distribution of the kinetic energy [7].Reference [7] gives an explicit expression for the rescalingfactor, which reads

α2 = e−∆t/τ +K

NfK

(

1 − e−∆t/τ)

⎝R21 +

Nf∑

i+2

R2i

+2R1e−∆t/τ

K

NfK

(

1 − e−∆t/τ)

, (11)

where K = Nf/2kBT is the target temperature, Nf isthe number of degrees of freedom, Ri’s are independentrandom numbers extracted from a Gaussian distributionwith zero mean and unitary variance and the parameterτ , which is defined as τ = (2ξ)−1 (dimensions of time),determines the time scale of the thermostat. Note that,in the non-local implementation of the BDP that we usehere, the parameter τ does not truly represent the effectof the solvent [34], and thus a change in τ has basicallyno effect on the dynamics of the system.

The equations of motion are integrated with the veloc-ity Verlet method [1]. The BDP thermostat is a PBT sinceit acts on the peculiar velocities of the particles, which arerescaled by the factor α after a specified number of timesteps.

2.6 Observables

We run simulations under equilibrium and non-equilib-rium conditions. We investigate the dynamics of the sys-tem by looking at the mean squared displacement 〈r2(t)〉from which we extract the diffusion coefficient D, definedas

D = limt→∞

r2(t)⟩

6t. (12)

Page 6: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

Eur. Phys. J. E (2018) 41: 80 Page 5 of 13

10-4

10-3

10-2

10-1

100

101

102

103

104

ξ

10-4

10-3

10-2

10-1

D

Bussi (T=0.772)

Bussi (T=1.5)

Langevin (ξxyz

)

Langevin (ξyz

)

Langevin (ξy)

a)

ξ

BDP (T=0.772)

BDP (T=1.5)

Langevin (ξxyz

)

Langevin (ξyz

)

Langevin (ξy)

0 25 50 75 100 125ξ

10-4

10-3

10-2

10-1

D

0 25 50 75 100 125ξ

b)

Fig. 1. Diffusion coefficient D for the system in equilibrium at T = 1.5 and ρ = 0.844. (a) D calculated with the Langevinthermostat acting along one, two and three directions. For comparison D calculated with the BDP thermostat is also shownat both T = 1.5 and T = 0.772 (see ref. [34]). (b) D calculated with DPD thermostat for different sets of parameter. Differentsymbols correspond to different values of the exponent: s = 0.25 (circles), s = 0.50 (squares), s = 1.0 (diamonds) and s = 2.0(triangles), while different colours indicate different cut-off values: rc = 1.12 (orange), rc = 1.54 (purple) and rc = 1.88 (darkgreen).

In rheology experiments and simulations, this parameteris fundamental to describe the competition between diffu-sion and shear flow. In equilibrium, the dynamics of a fluidis governed by Brownian (diffusive) dynamics. In the pres-ence of a shear flow an anisotropic microstructure, origi-nated from the competition between diffusion and sheareffects, appears in the stationary state. While Browniandynamics tends to restore the equilibrium of the system,the shear flow tends to distort its structure. The compe-tition between the two regimes is encoded in the Pecletnumber Pe = τB/τS = σ2γ/(4D), where τB and τS arethe characteristic times of diffusion and shear flow, respec-tively [62].

We also compute the zero-shear viscosity η by usingthe Green-Kubo relation [63]

η ≡∫ ∞

0

Cσσ(t)dt =β

3V

∫ ∞

0

α<β

σαβ(t)σαβ(0)⟩

dt,

(13)where β = 1/kBT , V is the volume of the simulation box,〈· · · 〉 indicates an average over initial conditions and themicroscopic stress tensor σαβ is defined as

σαβ =

N∑

i=1

miviαviβ +

N∑

i<j

rijαrijβ

rijF (rij) , (14)

where viα is the α-th component of the velocity of parti-cle i, F (rij) is the force between particles i and j. Sincenumerical estimates of the stress autocorrelation functionCσσ(t) are, especially at high friction, very noisy, we followprevious work and fit Cσσ(t) with the phenomenologicalexpression provided in ref. [64].

Finally, we analyse how particles exchange momentumwith their surroundings through the velocity autocorrela-tion function, defined as

Z(t) =1

3〈v(t) · v(0)〉 . (15)

The decay of this function provides information on howdifferent thermostats make the system’s dynamics decor-relate over time.

When subject to shear flow, the local properties of thesystem, both static and dynamic, change along the gra-dient direction z. It is therefore useful to divide the sim-ulation volume into layers of thickness σ and area L × Lperpendicular to the gradient direction. We thus consider:i) the local temperature T (z), extracted from the kineticenergy computed from the components of the velocitywhich are perpendicular to the shear direction, ii) the lo-cal density ρ(z), defined as the number of particles in thelayer divided by the volume of the layer (σL2) and iii) theprofile of the x-component of the velocity vx(z), definedas the average of the velocity along x over all particles inthe layer. Regarding the latter observable, in equilibriumthere is no preferred direction and hence 〈vx(z)〉 ∼ 0 in alllayers. By contrast, the velocity profile of a liquid undershear flow is linear with x and 〈vx(±L/2)〉 = ±γL/2. Ifthis is not the case then the velocity profile will exhibitanomalies, i.e., the appearance of regions where the defor-mation of the system is not uniform with respect to theexternal field. As we will show below, this is related to in-homogeneities in the density profile, which might indicatea non-physical behaviour.

3 Results

3.1 Equilibrium

In this section we focus on the system at T = 1.5 inequilibrium. We start by calculating the diffusion coef-ficient D, reported in fig. 1(a). First of all, we note thatD obtained via BDP simulations is independent of τ (andhence of ξ = (2τ)−1) [34]. Since the same holds true forall other observables investigated, in what follows we use

Page 7: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

Page 6 of 13 Eur. Phys. J. E (2018) 41: 80

10-2

10-1

100

101

102

103

t

1101001000

102

103

110

10-2

10-1

100

101

102

103

t

0

10

20

30

40

Cσσ(t

)

0.111010 1

110

a) b)

10-4

10-3

10-2

10-1

100

101

102

103

104

ξ

100

101

102

103

η

time time

Fig. 2. Stress correlation function Cσσ(t) for (a) the BDP thermostat and (b) the Langevin thermostat ξxyz. Symbols representthe numerical Cσσ(t), whereas solid lines are fits [64]. Inset: the dashed line shows the BDP reference viscosity, whereas symbolsindicate the ξxyz Langevin viscosity.

10-2

10-1

100

101

102

103

t

2.0_10_1.12

2.0_10_1.88

2.0_120_1.12

2.0_120_1.88

10-2

10-1

100

101

102

103

t

0

10

20

30

40

Cσσ(t

)

0.25_10_1.12

0.25_10_1.88

0.25_120_1.12

0.25_120_1.88

rc

1.121.88

1201.121.88

120

1010

a) b)

0 25 50 75 100 125

ξ

100

101

102

η

rc

1.121.88

1201.121.88

120

1010

0 25 50 75 100 125

ξ

100

101

102

η

time time

Fig. 3. Stress correlation function Cσσ(t) for the DPD thermostat for (a) s = 0.25 and (b) s = 2.0. Symbols represent thenumerical Cσσ(t), whereas solid lines are fits [64]. Insets: the dashed line shows the BDP reference viscosity, whereas symbolsindicate the DPD viscosity.

the BDP data as reference. We have checked that thesame reference curves are obtained if, at zero shear, nothermostats are employed, that is, if the simulations areperformed in the NVE ensemble. For both Langevin andDPD, D is a monotonically decreasing function of ξ. Forthe Langevin thermostat, the fewer directions the frictionacts on, the larger is the diffusion coefficient. However,when ξ ≤ 10−2 Brownian effects are small and D alwaystends to a plateau. For the DPD thermostat, the effect ofthe two other parameters is also monotonic: D increaseswith s and decreases with rc. The dependence on s canbe explained by considering that lower values of s causehigher correlations between the drag and random forcesin the equations of motion, thereby making the systemdecorrelate faster. By contrast, the increase of D withrc is related to the fact that the larger the cut-off, themore particles are included in the thermalization proce-

dure. Similar effects have been observed with DPD-relatedthermostats [23, 65, 66]. We note that the range of valuesof D is comparable for both thermostats, indicating thatthey share the same nature. A variation of two orders ofmagnitude in D is observed in the explored range of ther-mostat parameters.

The choice of the parameters also affects the viscositywhich, for the systems investigated here, gives similar in-formation as D. For instance, the BDP thermostat showsthat Cσσ(t) is independent of τ , which translates to a con-stant viscosity, as shown in fig. 2(a). With respect to theLangevin thermostat, in the same range where D is con-stant, η is also constant (see fig. 2(b)). Finally, fig. 3(a)and (b) shows that, for the DPD thermostat, the value ofη displays a strong dependence on s. In particular, smallervalues of s induce larger correlations on the microscopicinternal stress, effectively generating larger viscosities.

Page 8: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

Eur. Phys. J. E (2018) 41: 80 Page 7 of 13

10-2

10-1

100

101

102

time

-0.2

0

0.2

0.4

0.6

0.8

1Z

(t)/

Z(0

)

ξ=1

10-2

10-1

100

101

102

time

ξ=10

10-2

10-1

100

101

102

time

ξ=102a) b) c)

Fig. 4. Velocity-autocorrelation function Z(t) (normalized by its zero-time value) for the system in equilibrium at T = 1.5 andρ = 0.844 with the Langevin thermostat acting in three (red circles), two (blue squares) and one directions (green diamonds)and three values of the friction constant: (a) ξ = 1, (b) ξ = 10 and (c) ξ = 102. The black (reference) curve is obtained withthe BDP thermostat.

10-2

10-1

100

101

time

0

0.5

1

Z(t

)/Z

(0)

Bussi (NVE)

0.50_25/1.12/vac

0.50_25/1.54/vac

0.50_25/1.88/vac

0.50_120/1.12/vac

0.50_120/1.54/vac

0.50_120/1.88/vac

1.0_25/1.12/vac

1.0_25/1.54/vac

1.0_25/1.88/vac

1.0_120/1.12/vac

1.0_120/1.54/vac

1.0_120/1.88/vac

s rc

0.500.500.500.500.500.501.01.01.01.01.01.0

252525

252525

120120120

120120120

1.12

1.12

1.12

1.12

1.54

1.54

1.54

1.54

1.88

1.88

1.88

1.88

-6 -4 -2 0 2 4 6x

1.46

1.48

1.5

1.52

1.54

T

ξ

= 25

= 75

= 120

-6 -4 -2 0 2 4 6xz

a) b)

Fig. 5. (a) Velocity-autocorrelation function Z(t) (normalized by its zero-time value) for the system in equilibrium at T = 1.5and ρ = 0.844 with the DPD thermostat for several combinations of the parameters. The black (reference) curve is obtainedwith the BDP thermostat. (b) Temperature profile obtained with the DPD thermostat using s = 0.25, rc = 1.12 and variousvalues of ξ extracting random numbers from a uniform distribution [61]. Clearly, the temperature decreases as ξ increases.

Figure 4 shows Z(t) (normalized by its zero-time value)for the Langevin thermostat acting on three, two and onedirections for three different values of ξ. At low friction(ξ = 1), the three curves fall on top of the reference (BDP)one. However, upon increasing ξ we see that the correla-tions are removed earlier for the Langevin thermostat thatacts on all three spatial directions. By removing a direc-tion of thermalization, the overall friction coefficient is ef-fectively reduced. This is in agreement with the behaviourobserved for D (see fig. 1(a)).

The velocity autocorrelation functions for the DPDthermostat are shown in fig. 5(a). The dependence of itsdecay on s and rc is directly linked to the effect that thesetwo parameters have on D (see fig. 1(b)), and the effect ofξ is the same as in the Langevin thermostat. Overall, thecombined effect of the three DPD thermostat parametersis to tune the viscosity of the system in a finer way withrespect to Langevin dynamics.

A useful tip concerns the use of random numbers ex-tracted from uniform distributions, which allows to speedup the computation [61]. We find that this might, in prin-ciple, produce the wrong temperature profile, dependingon the parameters used. Indeed, fig. 5(b) shows an ex-ample that was generated with a DPD thermostat withs = 0.25 and rc = 1.12. For this choice of parameters, thetemperature clearly decreases with increasing ξ, a prob-lem that is avoided by using a Gaussian distribution forthe random numbers.

We conclude by noting that, at short times (t → 0),the slope of Z(t) should tend to zero, as seen for the ref-erence curve. However, when ξ increases the slope of theautocorrelation function at t = 0 becomes more and morenegative for both DPD and Langevin thermostats. This isdue to the local nature of the thermostating scheme, asdiscussed in refs. [67, 68].

Page 9: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

Page 8 of 13 Eur. Phys. J. E (2018) 41: 80

-6 -4 -2 0 2 4 6x

0.6

0.8

1

1.2

ρ

-6 -4 -2 0 2 4 6x

-6 -4 -2 0 2 4 6x

-6 -4 -2 0 2 4 6x

0.6

0.8

1

1.2

ρ

-8

-4

0

4

8

<V

x(z

)>/γ

-8

-4

0

4

8

<V

x(z

)>/γ

0.64

0.68

0.72

0.76

0.8

0.84

0.88

T

1.4

1.44

1.48

1.52

1.56

1.6

T

T = 0.722 T = 1.5= 0.01 = 0.01= 0.1 = 0.1

ξxyz

ξyz

ξy

1 102

a.1)

a.2)

a.3)

b.1)

b.2)

b.3)

c.1)

c.2)

c.3)

d.1)

d.2)

d.3)

ξpec

z z z z

<V

x (z)>/γ

Fig. 6. Temperature (top), velocity (middle) and density (bottom) profiles computed with the Langevin thermostat undershear flow, for two values of T and γ. Here ξxyz, ξyz, ξy indicate thermalization acting on three, two, one direction, respectively,while ξpec refers to the thermostat acting on the peculiar velocities. The black lines are the reference curves.

3.2 Steady shear with Langevin dynamics

We now focus on the system under steady shear using aLangevin thermostat. Figure 6 shows T (z), vx(z) and ρ(z)for both studied values of T and γ and for different valuesof the friction constant. For all values of ξ the temperatureis constant at low shear rate (see fig. 6(a.1) and (c.1)),regardless of the number of spatial directions to whichthermalization is applied. However, upon increasing γ thetemperature profile remains independent of z only for thethree directions Langevin thermostat and for large enoughvalues of ξ. The density profile (see fig. 6(a.3)–(d.3)) ex-hibits similar trends. By contrast, the velocity profile vx(z)has a more complex behaviour, as it is linear only in somecases. From the plots it is clear that a flat T (z) does notimply a linear velocity profile and vice versa. Indeed, aconstant T (z) can be obtained by increasing ξ, but theconsequent increase in the fluid viscosity can cause theappearance of a non-linear velocity profile and unrealis-tic inhomogeneities in the density profile. These inhomo-geneities are stronger close to the border of the simulationbox, where the flow velocity is higher (±γ L

2 ). The inho-mogeneities become more pronounced upon increasing γ,indicating that the Langevin dynamics cannot cope withthe increased stress, or upon decreasing T , when more sta-ble particle aggregates tend to form. These results clearly

show that a correct control of the temperature does notguarantee that other fundamental observables, such as thevelocity and density profiles, are correctly reproduced.

Figure 6 also shows that this problem can be only par-tially mitigated by turning off the thermostat along one(the flow) or two (the flow and gradient) directions: themodified thermostats work at low γ only. However, wedo not recommend using this technique as it introducesadditional spatial inhomogeneities in the system [49, 52].Better performances can be obtained instead by couplingthe system to a Langevin thermostat that acts on the pe-culiar velocity. In this case, indicated in fig. 6 as ξpec, alinear velocity profile is, by construction, always recov-ered. Furthermore, the density profile shows a perfect ho-mogeneity for any value of ξ. However, in even more ex-treme cases this modification might still not be enoughto ensure a correct thermalisation, and hence a flat tem-perature profile, as for too large values of the shear rate,the thermostat might not be able to dissipate the extrakinetic energy. However, with the peculiar Langevin ther-mostat the linear velocity profile is an enforced ratherthan an emergent property, meaning that, for instance,shear banding would not be observed even in cases whereit should be present [38, 40, 69]. For the same reason, thepeculiar Langevin thermostat could hide short-time dy-namic heterogeneities which appear in glasses [40].

Page 10: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

Eur. Phys. J. E (2018) 41: 80 Page 9 of 13

-6 -4 -2 0 2 4 6x

0.6

0.8

1

1.2

ρ

-6 -4 -2 0 2 4 6x

0.64

0.68

0.72

0.76

0.8

0.84

T

-6 -4 -2 0 2 4 6x

-6 -4 -2 0 2 4 6x

0.6

0.8

1

1.2

ρ

-8

-4

0

4

8

<V

x(z

)>/γ

-8

-4

0

4

8

<V

x(z

)>/γ

ρT

<V

x (z)>/γ

0.64

0.68

0.72

0.76

0.8

0.84

T

0.50_25/1.12/Temp1/average_Temp

0.50_25/1.88/Temp1/average_Temp

0.50_75/1.12/Temp1/average_Temp

0.50_75/1.88/Temp1/average_Temp

0.50_120/1.12/Temp1/average_Temp

0.50_120/1.88/Temp1/average_Temp

1.12 1.88

25

75120

rc

a.1)

a.2)

a.3)

b.1)

b.2)

b.3)

c.1)

c.2)

c.3)

d.1)

d.2)

d.3)

s = 0.5 s = 1.0 s = 0.5 s = 1.0

z z z z

Fig. 7. Temperature (top), velocity (middle) and density (bottom) profiles computed with the DPD thermostat under shearflow for T = 0.722; (a)–(b) γ = 0.01 and (c)–(d) γ = 0.1. The black lines are the reference curves.

3.3 Steady shear with DPD thermostat

Figure 7 shows the numerical results under shear flow forsystems thermostated with DPD at T = 0.722 for sev-eral choices of the thermostat parameters. For γ = 0.01(fig. 7(a)–(b)) the temperature profile is always constant,regardless of the specific values of the parameters. How-ever, the other two observables show that, once again, aflat T (z) does not guarantee a linear velocity profile. Infact, we observe that the velocity profile is almost butnot quite linear only for the lowest values of ξ and rc.Moreover, even though T (z) is constant for all possibleparameter combinations, its fluctuations increase at smallrc. To make things worse, the density profile often exhibitsunrealistic inhomogeneities, as with the original Langevindynamics. Upon increasing s the homogeneity of the ve-locity and density profiles improves.

When we increase the shear rate to γ = 0.1(fig. 7(c)–(d)), the situation worsens. Indeed, the temper-ature profile remains constant only at high values of ξ andrc, i.e. when the thermostat is more coupled (or, equiva-lently, when the viscosity is higher). In contrast with theLangevin thermostat, here the variation of ξ has a lessdramatic impact, as the behaviour of the velocity, temper-ature and density profiles is fairly insensitive to changesin ξ.

The data at T = 1.5 display qualitatively similartrends and are not shown here.

3.4 Steady shear with the modified DPD

We have seen that increasing the dissipation through achange of the DPD parameters helps in controlling thetemperature but negatively affects the velocity profile.The origin of this issue, which has been observed and de-scribed in recent work [49,70], can be ascribed to a wronghandling of the periodic boundary conditions along thegradient direction z. Indeed, when the z-component of aparticle i position is close to ±L/2, some of its neighboursare bound to be on the other side of the simulation boxboundary. All these will have a streaming velocity thatis opposite to the one of i, breaking the assumption oftranslational invariance and of homogeneity. Indeed, asnoted in refs. [52] and [49], particles should not be ableto sense when they cross box boundaries, so as to avoidsurface effects. The problem can be mitigated by turn-ing off the thermostat for all those pairs of particles thatare on different sides of the boundary [70–72]. However,there exists a less invasive approach that gets rid of theissue altogether. The idea is to correct the relative velocitybetween boundary-separated pairs of particles by adding

Page 11: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

Page 10 of 13 Eur. Phys. J. E (2018) 41: 80

-6 -4 -2 0 2 4 6x

0.64

0.68

0.72

0.76

0.8

0.84

T

γ=0.01

γ=0.1

γ=0.01 (Modified)

γ=0.1 (Modified)

γ=0.01

γ=0.1

γ=0.01 (Modified)

γ=0.1 (Modified)

-6 -4 -2 0 2 4 6x

-8

-4

0

4

8

<V

x(z

)>/γ

-6 -4 -2 0 2 4 6x

0.6

0.8

1

1.2

ρ

z z z

a) b) c)

Fig. 8. (a) Temperature, (b) velocity and (c) density profiles for a DPD thermostat with s = 0.25, ξ = 120 and rc = 1.88 withand without the improved treatment of the periodic boundary conditions (see text for details). The black lines are the referencecurves.

-6 -4 -2 0 2 4 6x

0.6

0.8

1

1.2ρ

LJ_1.5/shear0.01/tau001/Temp1/average_Part

LJ_1.5/shear0.01/tau10/Temp1/average_Part

LJ_1.5/shear0.1/tau001/Temp1/average_Part

LJ_1.5/shear0.1/tau10/Temp1/average_Part

-6 -4 -2 0 2 4 60.6

0.8

1

1.2

ρ

LJ_0.722/shear0.01/tau001/Temp1/average_Part

LJ_0.722/shear0.01/tau10/Temp1/average_Part

LJ_0.722/shear0.1/tau001/Temp1/average_Part

LJ_0.722/shear0.1/tau10/Temp1/average_Part

-6 -4 -2 0 2 4 6x

-8

-4

0

4

8

<V

x(z

)>/γ

LJ_1.5/shear0.01/tau001/Vel1/average_Vel

LJ_1.5/shear0.01/tau10/Vel1/average_Vel

LJ_1.5/shear0.1/tau001/Vel1/average_Vel

LJ_1.5/shear0.1/tau10/Vel1/average_Vel

-8

-4

0

4

8

<V

x(z

)>/γ

LJ_0.722/shear0.01/tau001/Vel1/average_Vel

LJ_0.722/shear0.01/tau10/Vel1/average_Vel

LJ_0.722/shear0.1/tau001/Vel1/average_Vel

LJ_0.722/shear0.1/tau10/Vel1/average_Vel

0.64

0.68

0.72

0.76

0.8

0.84

T

LJ_0.722/shear0.01/tau001_CM/Temp1/average_Temp

LJ_0.722/shear0.01/tau10_CM/Temp1/average_Temp

LJ_0.722/shear0.1/tau001_CM/Temp1/average_Temp

LJ_0.722/shear0.1/tau10_CM/Temp1/average_Temp

-6 -4 -2 0 2 4 6x

1.4

1.44

1.48

1.52

1.56

1.6

T

0.01 0.1

10

0.01

a.1)

a.2)

b.1)

b.2)

c.1)

c.2)

z z z

Fig. 9. (a) Temperature, (b) velocity and (c) density profiles for (a.1), (b.1) and (c.1) γ = 0.01 and (a.2), (b.2) and (c.2) γ = 0.1computed using the BDP thermostat for (top) T = 0.722 and (bottom) T = 1.5. The black lines are the reference curves.

a term − round(zij/L)γL that takes into account the ve-locity difference between the top and the bottom of thebox [49]. Figure 8 demonstrates that this approach resultsin a much better control over the temperature, velocityand density profiles.

3.5 Steady shear with BDP thermostat

While the original implementation of the BDP thermostatworks flawlessly in equilibrium [7], it needs to be adaptedto simulations performed under shear flow. First of all,the instantaneous kinetic energy K should not take intoaccount the velocity component along the flow direction.In addition, during the thermalization step, in a fashionsimilar to what has been done for the peculiar version ofthe Langevin thermostat, the shear velocity ux(z) = γvz issubtracted from the vx of each particle i, the thermostat isapplied and then the flow velocity is added back to the newvx [73]. With this change, the BDP thermostat becomes

a PBT.

Figure 9 shows the performance of the shear-flow ver-sion of the BDP thermostat for the two investigated tem-peratures and for two different values of τ . We have al-ready seen that τ has no effect on the dynamics of thesystem in equilibrium (see fig. 1). Similarly, here we seethat the BDP thermostat always works well for the com-binations of T and γ considered here, regardless of thespecific value of τ . However, we recall that the BDP ther-mostat we consider has a global nature and hence does notreproduce hydrodynamics. This thermostat is thus onlyrecommended to generate initial configurations or refer-ence data.

4 Conclusions

We have shown that, when studying out-of-equilibriumsystems, having a constant temperature profile is a neces-sary but not sufficient condition to ensure the correctnessof the simulation. Indeed, one has to be sure that alsothe velocity and density profiles exhibit sound behaviours.

Page 12: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

Eur. Phys. J. E (2018) 41: 80 Page 11 of 13

Table 1. Assessment of the Langevin dynamics (a) with and(b) without the subtraction of the peculiar velocity for differ-ent values of ξ and γ. The 5 and 9 symbols tell whether thecombination of parameters reproduce the correct behaviour forthat specific observable or not, respectively.

(a)

Parameters Observables

ξ γ T = 0.722 T = 1.5 vx(z) ρ

1 0.01 5 5 9 5

1 0.1 9 9 9 5

102 0.01 5 5 9 9

102 0.1 5 5 9 9

(b)

Parameters Observables

ξ γ T = 0.722 T = 1.5 vx(z) ρ

1 0.01 5 5 5 5

1 0.1 9 9 5 5

102 0.01 5 5 5 5

102 0.1 5 5 5 5

Here we have tested several thermostats under differentphysical and profile conditions. In general, we observethat, under PUT conditions, a fine tuning of the ther-mostat parameters is necessary to avoid non-physical be-haviours. When γ increases more energy is pumped intothe system and hence the thermostat has to be moretightly coupled to the system in order to maintain the de-sired temperature. If the thermostat acts on the absolutevelocities, a too-high friction might cancel the effect of theimposed shear flow on the particles that are far from thebox boundaries, which is where the flow velocity acquiresthe highest values. This, in turn, affect the density profile,which becomes inhomogeneous.

By contrast, if the thermostat is applied on the pecu-liar velocities, that is, if we use a PBT, correct velocityand density profiles will be assured. If we need to ex-plore high values of γ it is recommended to work witha PBT. However, doing so will make it impossible to ob-serve some real physical phenomena such as shear band-ing. For the Langevin thermostat, PBT conditions canbe implemented by removing the peculiar (flow) velocityin the friction term (see eq. (4)). A similar remark holdstrue for the Bussi-Donadio-Parrinello thermostat, whichturned out to be the most stable and reliable thermostat.However, the BDP thermostat is also the one exhibitingthe less realistic dynamics. Tables 1 and 2 summarise someof the results reported.

The best choice in terms of stability, realism and com-putational efficiency is probably the DPD thermostat,modified so as to take into account the relative differ-ence between the streaming velocities at the boundary ofthe simulation box [49]. However, this thermostat also de-

Table 2. Assessment of DPD thermostat (without the modifi-cation) considering the set parameter (s, ξ, rc) values. We onlyconsider the most common values for ξ and s. The 5 and 9symbols tell whether the combination of parameters reproducethe correct behaviour for that specific observable or not, re-spectively.

Parameters Observables

s rc ξ γ T = 0.722 T = 1.5 vx(z) ρ

0.50 1.12 25 0.01 5 5 5 5

0.50 1.12 25 0.1 9 9 5 5

1.0 1.12 25 0.01 5 5 5 5

1.0 1.12 25 0.1 9 9 5 5

0.50 1.88 25 0.01 5 5 9 5

0.50 1.88 25 0.1 5 5 9 9

1.0 1.88 25 0.01 5 5 9 5

1.0 1.88 25 0.1 9 5 9 9

pends on the largest number of parameters: the exponents, the cut-off rc and the friction constant ξ. Accordingto our results, the common choice of setting s = 0.5 or1 gives overall good results. One of the main difficultiesis to choose an optimal value for rc. There is not an apriori physical motivation to choose a certain value, andthe optimal choice of rc is in general independent on theother parameters. In general, making a good choice re-quires a knowledge of the structure of the system understudy. For instance, if rc is much smaller than the aver-age particle-particle distance, which might happen in di-lute systems, the thermostat is essentially decoupled fromthe system and might not be able to dissipate the extraenergy. Our results show that the cut-off radius shouldbe between the first maximum and the first minimum ofthe g(r), and that too-large values might originate non-physical behaviour such as anomalies and inhomogeneitiesin the velocity and density profiles. A value smaller thanbut close to the first minimum of the radial distributionfunction of the system is a good starting point.

Throughout the paper we have not analysed the com-putational efficiency of the different thermostats, whichgreatly depends on the computational (e.g., serial vs. par-allel codes) and model (e.g., short- vs. long-ranged poten-tials) details. However, for the systems studied here theBDP thermostat turned out to be the most efficient ther-mostat, providing an additional reason to use it to produceexploratory data and to prepare initial configurations forthe production runs. Second comes the Langevin thermo-stat. The DPD thermostat, which features pair interac-tions, comes last, even though its efficiency can be im-proved by using, e.g., Verlet neighbouring lists [74]. Giventhe great number of random numbers required by the DPDthermostat, it might be tempting to extract them fromuniform instead of Gaussian distributions [61]. However,we advise against it, as doing so yields the wrong tem-perature profile for some combinations of the thermostatparameters.

Page 13: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

Page 12 of 13 Eur. Phys. J. E (2018) 41: 80

JRF and EZ acknowledge support from ETN-COLLDENSE(H2020-MCSA-ITN-2014, Grant No. 642774). LR, EZ ac-knowledge support from the ERC Consolidator Grant 681597MIMIC. We thank N. Gnan for useful discussions.

Author contribution statement

JRF, LR and EZ performed simulations, analysed resultsand wrote the paper.

References

1. M.P. Allen, D.J. Tildesley, Computer Simulation of Liquids

(Oxford University Press, 2017).2. D. Frenkel, B. Smit, Understanding Molecular Simulation:

From Algorithms to Applications, Vol. 1 (Academic Press,2001).

3. M. Ripoll, K. Mussawisade, R. Winkler, G. Gompper,Phys. Rev. E 72, 016701 (2005).

4. R. Kapral, Adv. Chem. Phys. 140, 89 (2008).5. G. Gompper, T. Ihle, D. Kroll, R. Winkler, in Advanced

Computer Simulation Approaches for Soft Matter Sciences

III (Springer, 2009) pp. 1–87.6. H.J. Berendsen, J.v. Postma, W.F. van Gunsteren, A. Di-

Nola, J. Haak, J. Chem. Phys. 81, 3684 (1984).7. G. Bussi, D. Donadio, M. Parrinello, J. Chem. Phys. 126,

014101 (2007).8. H.C. Andersen, J. Chem. Phys. 72, 2384 (1980).9. S. Nose, J. Chem. Phys. 81, 511 (1984).

10. W.G. Hoover, Phys. Rev. A 31, 1695 (1985).11. S.D. Stoyanov, R.D. Groot, J. Chem. Phys. 122, 114112

(2005).12. G.J. Martyna, M.L. Klein, M. Tuckerman, J. Chem. Phys.

97, 2635 (1992).13. T. Schneider, E. Stoll, Phys. Rev. B 17, 1302 (1978).14. B. Dunweg, J. Chem. Phys. 99, 6977 (1993).15. P. Hoogerbrugge, J. Koelman, Europhys. Lett. 19, 155

(1992).16. P. Espanol, P. Warren, Europhys. Lett. 30, 191 (1995).17. C. Lowe, Europhys. Lett. 47, 145 (1999).18. G. Besold, I. Vattulainen, M. Karttunen, J.M. Polson,

Phys. Rev. E 62, R7611 (2000).19. T. Shardlow, SIAM J. Sci. Comput. 24, 1267 (2003).20. E. Peters, Europhys. Lett. 66, 311 (2004).21. B. Leimkuhler, X. Shang, J. Comput. Phys. 280, 72

(2015).22. M. Ellero, P. Espanol, Appl. Math. Mech. 39, 103 (2018).23. X. Fan, N. Phan-Thien, S. Chen, X. Wu, T. Yong Ng,

Phys. Fluids 18, 063102 (2006).24. N. Phan-Thien, N. Mai-Duy, Understanding Viscoelastic-

ity: An Introduction to Rheology (Springer, 2017).25. P. Espanol, Europhys. Lett. 39, 605 (1997).26. P. Espanol, Phys. Rev. E 57, 2930 (1998).27. J.B. Avalos, A. Mackie, Europhys. Lett. 40, 141 (1997).28. M. Liu, G. Liu, Archiv. Comput. Methods Eng. 17, 25

(2010).29. Z.-B. Wang, R. Chen, H. Wang, Q. Liao, X. Zhu, S.-Z. Li,

Appl. Math. Model. 40, 9625 (2016).30. P. Espanol, M. Revenga, Phys. Rev. E 67, 026705 (2003).31. A. Vazquez-Quesada, M. Ellero, J. Non-Newton. Fluid

Mech. 233, 37 (2016).

32. M. Whittle, K.P. Travis, J. Chem. Phys. 132, 124906(2010).

33. A. Boromand, S. Jamali, J.M. Maia, Soft Matter 13, 458(2017).

34. G. Bussi, M. Parrinello, Comput. Phys. Commun. 179, 26(2008).

35. B. Lander, U. Seifert, T. Speck, J. Chem. Phys. 138,224907 (2013).

36. T. Mohoric, J. Dobnikar, J. Horbach, Soft Matter 12, 3142(2016).

37. M. Kohl, M. Schmiedeberg, Eur. Phys. J. E 40, 71 (2017).38. T. Soddemann, B. Dunweg, K. Kremer, Phys. Rev. E 68,

046702 (2003).39. J. Zausch, J. Horbach, M. Laurati, S.U. Egelhaaf, J.M.

Brader, T. Voigtmann, M. Fuchs, J. Phys.: Condens. Mat-ter 20, 404210 (2008).

40. G.P. Shrivastav, P. Chaudhuri, J. Horbach, Phys. Rev. E94, 042605 (2016).

41. J. Zausch, J. Horbach, P. Virnau, K. Binder, J. Phys.:Condens. Matter 22, 104120 (2010).

42. C. Pastorino, T. Kreer, M. Muller, K. Binder, Phys. Rev.E 76, 026706 (2007).

43. P.S. Doyle, E.S. Shaqfeh, A.P. Gast, Phys. Rev. Lett. 78,1182 (1997).

44. X. Yong, L.T. Zhang, Phys. Rev. E 82, 056313 (2010).45. J.-L. Barrat, L. Bocquet, Phys. Rev. Lett. 82, 4671 (1999).46. N.V. Priezjev, S.M. Troian, J. Fluid Mech. 554, 25 (2006).47. A. Niavarani, N.V. Priezjev, Phys. Rev. E 81, 011606

(2010).48. X. Yong, L.T. Zhang, J. Chem. Phys. 138, 084503 (2013).49. B. Leimkuhler, X. Shang, J. Comput. Phys. 324, 174

(2016).50. X. Shang, M. Kroger, B. Leimkuhler, Soft Matter 13, 8565

(2017).51. A. Lees, S. Edwards, J. Phys. C: Solid State Phys. 5, 1921

(1972).52. G.P. Morriss, D.J. Evans, Statistical Mechanics of

Nonequilbrium Liquids (ANU Press, 2013).53. F. Zhang, D.J. Searles, D.J. Evans, J.S. den Toom Hansen,

D.J. Isbister, J. Chem. Phys. 111, 18 (1999).54. G. Pan, J.F. Ely, C. McCabe, D.J. Isbister, J. Chem. Phys.

122, 094114 (2005).55. B. Todd, P.J. Daivis, Mol. Simul. 33, 189 (2007).56. D.A. Sivak, J.D. Chodera, G.E. Crooks, J. Phys. Chem. B

118, 6466 (2014).57. D.L. Ermak, J. McCammon, J. Chem. Phys. 69, 1352

(1978).58. A. Brunger, C.L. Brooks, M. Karplus, Chem. Phys. Lett.

105, 495 (1984).59. R.D. Groot, P.B. Warren, J. Chem. Phys. 107, 4423

(1997).60. A. Moshfegh, A. Jabbarzadeh, Soft Mater. 13, 106 (2015).61. B. Dunweg, W. Paul, Int. J. Mod. Phys. C 2, 817 (1991).62. M. Cloitre, High Solid Dispersions, Vol. 236 (Springer,

2010).63. J.-P. Hansen, I.R. McDonald, Theory of Simple Liquids

(Elsevier, 1990).64. A.M. Puertas, C. De Michele, F. Sciortino, P. Tartaglia,

E. Zaccarelli, J. Chem. Phys. 127, 144906 (2007).65. C. Junghans, M. Praprotnik, K. Kremer, Soft Matter 4,

156 (2008).66. H.-J. Qian, C.C. Liew, F. Muller-Plathe, Phys. Chem.

Chem. Phys. 11, 1962 (2009).67. E. Koopman, C. Lowe, J. Chem. Phys. 124, 204103 (2006).

Page 14: EPJE - homepage.univie.ac.at · dynamics simulations ... Langevin, dissipative particle dynamics, and Bussi-Donadio-Parrinello under varying parameters and external conditions. We

Eur. Phys. J. E (2018) 41: 80 Page 13 of 13

68. G. Bussi, T. Zykova-Timan, M. Parrinello, J. Chem. Phys.130, 074101 (2009).

69. K. Kang, H. Kriegs, J. Dhont, O. Danko, J. Marakis, D.Vlassopoulos,Phys. Rev. Fluids 2, 043301 (2017).

70. A. Chatterjee, Mol. Simul. 33, 1233 (2007).71. J.-L. Barrat, L. Berthier, Phys. Rev. E 63, 012503 (2000).

72. F. Varnik, J. Chem. Phys. 125, 164514 (2006).73. J. Zausch, Dynamics, Rheology and Critical Properties of

Colloidal Fluid Mixtures: Molecular Dynamics Studies in

Equilibrium and Under, PhD Thesis, Universitat Mainz(2009).

74. L. Verlet, Phys. Rev. 159, 98 (1967).