Green Kubo Assessments of Thermal Transport in Nanocolloids Based Oninterfacial Effects

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Materials Today Communications 20 (2019) 100533

Contents lists available at ScienceDirect

Materials Today Communications


journal homepage: www.elsevier.com/locate/mtcomm

Green–Kubo assessments of thermal transport in nanocolloids based on T


interfacial effects

Tolga Akinera, , Emir Kocera, Jeremy K. Masonb, Hakan Erturka
a
Department of Mechanical Engineering, Bogazici University, Istanbul, Turkey
b
Department of Materials Science and Engineering, University of California, Davis, CA, USA

A B S T R A C T

Thermal transport in a water–Cu nanocolloid system was investigated using equilibrium molecular dynamics. A systematic analysis of the Green–Kubo calculations is
presented to clarify the effect of simulation parameters. Several sources of error were identified and quantified for the thermal conductivity estimations, and the
effect of the base fluid potential was investigated. Simulations were carried out with a single copper particle for different diameters and water potentials, and thermal
enhancements exceeding both theoretical and experimental results were observed in parallel with some other studies in the literature. The anomalous Green–Kubo
thermal enhancement results could be explained by the interfacial dynamics and the neglect of calibrating the interaction potential to satisfy the physically-observed
energy flow at the interface.

1. Introduction [9–11], methane hydrate [12], Ar–Cu nanofluids [13–15], and a wa-
ter–platinum nanofluid [16] have been estimated with EMD, though
Colloidal suspensions have recently been the focus of numerous most of these studies do not include a detailed error analysis. This is
scientific studies due to their potential engineering applications [1]. significant because the fluctuation–dissipation theorem is based on the
Studies of particulate systems can involve rheology, interface science chaotic movements of particles or atoms, which introduces statistical
and colloidal chemistry [2], and could lead to advances in foams, gels, noise into the calculation of the autocorrelation function (ACF). Porter
paints, coatings and wetting-dispersing agents [3]. One recently-de- and Yip [17] and Chen et al. [18] proposed to truncate the integration
veloped colloidal system is known as a nanofluid, and is a suspension of time of the heat current autocorrelation function (HACF) to minimize
nanoparticles designed for enhanced heat transfer [4]. The physics of the statistical noise. Other approaches include curve fitting [19], block
thermal transport in these systems requires further study, and simula- averaging [20], random walk modelling [21], and spectral analysis of
tions are frequently used for this purpose because of experimental time series [22]. These studies provide fundamental insights about the
limitations at the relevant length scale. statistical nature of the EMD method, and should be carefully con-
Molecular dynamics (MD) is a powerful approach to investigate sidered to estimate the errors associated with Green–Kubo (GK) calcu-
nanoscale dynamics where interatomic forces govern the system be- lations.
haviour. There are two different schemes that can be used to measure In most nanofluid studies, Maxwell's relation [23] is used as a re-
the heat transport properties in MD simulations: non-equilibrium mo- ference for the effective thermal conductivity of a simulated system.
lecular dynamics (NEMD) [5] and equilibrium molecular dynamics The Maxwell model neglects Brownian motion, nanolayering and ag-
(EMD) [6]. A heat flux is imposed on the system in NEMD and the glomeration [24] as possible heat transfer enhancement mechanisms
thermal conductivity is calculated based on the resulting temperature for nanofluids, but provides a lower limit by only considering the
gradients. However, finite size effects are severe as a result of very high thermal conductivity of the mediums. Anomalous thermal enhancement
temperature gradients, and careful consideration is required to obtain values with respect to the Maxwell limit have been reported in some
reliable temperature profiles [7]. Equilibrium Molecular Dynamics in- nanofluid studies that use the GK method, where the contribution of
stead calculates thermal conductivity from the time decay of heat flux various mechanisms to this enhancement has been investigated by
fluctuations based on the fluctuation–dissipation theorem. While this comparing simulations and theoretical calculations [15,16,25–27].
requires more computational power, the method does not suffer from However, inconsistent results and recent developments [19,18,22]
the drawbacks of NEMD and is widely used in the literature. concerning GK calculations in the literature indicate that there is no
The thermal conductivities of Lennard–Jones liquids [8], water generally accepted GK algorithm, and that the effect of the GK


Corresponding author.
E-mail address: takiner@ncsu.edu (T. Akiner).

https://doi.org/10.1016/j.mtcomm.2019.05.009
Received 6 May 2019; Accepted 16 May 2019
Available online 26 May 2019
2352-4928/ © 2019 Elsevier Ltd. All rights reserved.
T. Akiner, et al. Materials Today Communications 20 (2019) 100533

parameters on the thermal transport has not yet been fully understood.
The anomalous thermal conductivity of nanofluids as found by GK
calculations has been hypothesized to be the effect of Brownian motion
[15] or nanolayering [26]. Recent studies have found an insignificant
contribution of Brownian motion of the nanoparticles to the heat
transfer enhancement of nanocolloids [28,29]. As for the nanolayering
effect, both experimental and theoretical investigations have tried to
quantify the thermal conductivity of the adsorbed liquid layer [30–32].
However, investigating nanolayering with MD simulations can be dif-
ficult since the interface potential is expected to have a considerable
effect on the thermal transport at the solid-liquid surface. Most EMD
studies use the Lorentz–Berthelot rules at the interface for lack of an
alternative without adequately studying the effect of the surface para-
meters, even though the Lorentz–Berthelot rule has no physical justi-
fication. It has been recently shown that Lorentz–Berthelot can lead to
an overestimation of the interfacial thermal resistance for a hBN–water
system [33], and an optimization process carried out by fitting potential
parameters is required to accurately represent the interface.
This study presents a comprehensive evalution of the Green-Kubo
Fig. 1. Cross-section of water–Cu model with 5% volume fraction, 2.8 nm side
approach by considering the thermal transport in a nanocolloid system,
length and 1.3 nm particle diameter.
where several inconsistencies and anomalous thermal conductivity re-
sults have been reported in the literature. Using a standard force-field as
the interatomic potential and a common modeling scheme for interface The enthalpy exclusion in the last term of Eq. (3) for multi-phase
interactions, possible sources of the observed anomalities are revealed. systems was introduced by Babaei et al. [35] who studied a methane-Cu
A statistical assessment of several sources of error is presented and the colloidal system using EMD, and was not common in earlier GK studies
effect of different water potentials on the thermal transport calculations [15,16,27]. This term is subtracted from the heat flux because it re-
is tested. The effect of the interface parameters on the thermal con- presents the energy carried by the mediums but not transported be-
ductivity calculations is investigated. Finally, thermal resistances at the tween them. The species enthalpy is defined as:
solid-liquid interfaces are quantified to further investigate interfacial Nα Nα
effects. We believe that the presented results clarify the contributions of 1 ⎡ 1⎛ 1 ⎞⎤
hα =

∑ ⎢ei + 3⎜
mi v¯i2 +
2
∑ r¯ij·f¯ij ⎟ ⎥
surface phenomena and the associated interfacial thermal resistance to i=1 ⎢ ⎝ j=1 ⎠⎥ (4)
⎣ ⎦
the anomalous thermal enhancement found with Green–Kubo calcula-
tions. where Nα is the number of atoms of species α.

2. Methodology 2.2. Problem and simulation details

2.1. Green–Kubo relations Pure water and water–Cu models with a single copper nanoparticle
(diameter 1.3 or 1.8 nm) in the middle of cubic simulation domains
The Green-Kubo method relies on the fluctuation–dissipation the- with a side length of 2.8 nm and 4 nm were generated as shown in
orem and linear-response theory [6], and describes the system behavior Fig. 1. Volume fraction calculations are not trivial at the nanoscale,
using time autocorrelation functions (ACF). Integrating an ACF in time considering that the clearance at the solid–liquid interface is not well-
allows transport properties in the equilibrium state to be extracted [34]. defined. We calculated the volume fraction to be 5% using mass frac-
The thermal conductivity is given in the GK formalism as: tions and the bulk densities of the phases, and 4.5% using the sum of
volumes of the Voronoi polyhedra. This was kept fixed for all nanofluid
1 ∞
k=
3kB VT 2
∫0 〈J¯ (0)·J¯ (τ ) 〉 dτ
(1)
systems, and is reported as 5% throughout the text. TIP3P (flexible)
[36], TIP4P/2005 (rigid) [37] and SPC/E (rigid) [38] potentials were
where kB is Boltzmann's constant, T is the absolute temperature, V is the used to define the water interactions. Pure water systems of the same
system volume, J̄ is the heat current vector, and the integrand is the dimensions were used to estimate the thermal conductivity of the base
heat autocorrelation function (HACF). The HCAF is an ensemble fluid. The SHAKE algorithm [39] was used to enforce the bond and
average, which relates the thermal conductivity to the heat current in angle constraints in the rigid water models, and long-range Coulombic
an equilibrated system. Eq. (1) involves integrals in the continuous interactions were solved using particle–particle–particle–mesh (PPPM)
case, but since time is discrete in MD simulations, the integral is re- summations [40]. Atomic interactions within a copper nanoparticle
placed by a summation in practice. The heat current vector is defined were defined using a Lennard–Jones 6–12 potential with
as: ϵCu−Cu = 9.4353 kcal/mole and σCu−Cu = 0.2338 nm [41]. The Lor-
entz–Berthelot rule was used for interactions between oxygen atoms
1⎧ 1 ⎫ and copper atoms, and the timestep was set to 1 fs. Periodic boundary
J¯ = ∑ ei v¯i + 2
V⎨ i
∑ [f¯ij ·(v¯i + v¯j)] r¯ij − ∑ hα ∑ v¯i ⎬
i<j α i (2) conditions were applied in all directions.
⎩ ⎭
Each model was first equilibrated for 100 ps in the isobaric-iso-
where v̄i is the velocity of atom i, r̄ij and f¯ij are distance and force vectors thermal (NPT) ensemble, followed by a further 2 ns in the micro-
between atoms i and j, α is the atomic species, and hα is the enthalpy of canonical ensemble (NVE) at 300 K. The GK calculations were per-
that species. The total energy ei is the sum of the kinetic and potential formed in 5–8 ns production runs (depending on the simulation) with
energies which can be expressed as: HACFs calculated in 20 ps intervals, and the reported results are the
1 1 average of 20 independent simulations with different initial velocity
ei = mi v¯i2 + ∑ Φ(¯rij)
2 2 seedings. All simulations are carried out using the large-scale atomic/
j (3)
molecular massively parallel simulator (LAMMPS) [42], and Fig. 1 was
where Φ(¯rij) is the interatomic potential energy. generated using visual molecular dynamics (VMD) [43].

2
T. Akiner, et al. Materials Today Communications 20 (2019) 100533

3. Results and discussion

3.1. Green–Kubo error analysis

The confidence interval usually reported in the literature is for a


single ACF, and corresponds to the fluctuations in the thermal con-
ductivity within a single MD simulation. We denote this type of error as
‘short-time error’ in this study. Another source of error is denoted as
‘long-time error’, and can be observed in the scatter of calculated
thermal conductivities for different velocity seedings. The heat flux
fluctuations are governed by the local properties of the potential energy
surface (PES), and a single MD simulation explores a relatively small
part of this surface over a period of ns. This effectively introduces a
random error that depends more on the initialization, or where the
system begins on the PES, than on the fluctuations of a single ACF.
While in principle a single very long MD simulation would explore a
sufficiently large region of configuration space, in practice the com-
putational requirements are reduced by changing the initial velocity
distribution of the atoms to obtain different trajectories and molecule
orientations. This procedure is also followed in several other studies
that average the thermal conductivities of independent simulations
[44,45,25], yet is not the standard procedure in the literature.
The short-time errors occur because of the statistical noise in an
ACF, and can be minimized by following the decay of the ACF. The time
evaluation of the normalized HACF for pure water with the SPC/E
potential (2.8 nm simulation cell in one dimension) is presented in
Fig. 2a as a function of the correlation time τ, which represents the time
interval within which the ACFs are calculated and integrated, and of the
simulation time t, which is the point in the simulation at which the
integration begins. A single HACF converges to zero over a 20 ps in-
terval [11], and smoother overall HACF behavior is achieved by be-
ginning the calculation after a short interval of simulation time
(200 ps). The effect of the HACF decay can be also observed in Fig. 2b,
where the estimated thermal conductivity for a total simulation time of
8 ns is presented. Each data point is obtained by integrating the HACF Fig. 2. (a) Surface of HACF behavior for correlation time (τ) and for total si-
over a 20 ps interval. The convergence of the integration can be ob- mulation time (t) of a pure SPC/E water model. (b) Thermal conductivity of a
served by comparing the two inset figures at 200 ps and 5 ns, and the pure SPC/E water model for a total simulation time of 8 ns. Inset figures are
short-time error is reported as the standard error of the mean for the representing the thermal conductivities at 200 ps and 5 ns, respectively. They
converged region of Fig. 2a between 5-8 ns. The value of this quantity is are produced by integration of individual ACFs within a correlation time in-
0.002 W/mK, which is insignificant compared to the long-time error terval τ.
discussed below. The total production run is therefore set to 5 ns to save
computation time.
The long-time error is calculated by sampling some number of re-
latively short GK results with different velocity initializations instead of
performing one long simulation which would require more computa-
tion time. Green–Kubo results of the thermal conductivity of in-
dependent simulations with different initial velocity seedings for the
SPC/E water model (2.8 nm simulation cell in one dimension) are
shown in Fig. 3, where the thermal conductivity is seen to change
significantly with different velocity initializations. The standard error
of the mean σ k̄ for the long-time error is 0.03 W/mK, significantly larger
than the standard error of the mean for short-time-error given above
(0.002 W/mK). Therefore, the confidence interval of the thermal con-
ductivity results reported below is constructed based on the long-time
error.

3.2. Thermal conductivity calculations


Fig. 3. Thermal conductivity results of 10 different initial velocity seeding for
The thermal conductivities of pure water and water–Cu nanocolloid pure SPC/E water.
systems are calculated following the procedure described in Section 2.1.
The results for different water potentials, particle diameters and (Simulations 5 and 6) are in agreement with the findings of Mur-
number of particles are presented in Table 1. The thermal con- aleedharen et al. [25] who used the rigid SPC/E water model. It is
ductivities of pure water (Simulations 1–3) are in agreement with the significant that the same trend is not observed with the flexible TIP3P
reported results in the literature [46]. The thermal conductivity en- model (Simulation 4).
hancements (thermal conductivity of the nanofluid divided by that of The anomalous thermal enhancement observed with rigid water
pure water) of a water–Cu nanofluid with a rigid water model

3
T. Akiner, et al. Materials Today Communications 20 (2019) 100533

Table 1 thermal conductivity. It has been showed that the thermal conductivity
Green–Kubo estimations of the thermal conductivity of water–Cu nanosuspen- is increased with smaller size of nanoparticles due to the effect of en-
sions with a single nanoparticle. k̄ is the mean and σ k̄ is the standard error of hanced Brownian motion [48,49]. Considering the larger particle dia-
the mean for different water models (Model), volume fractions (ϕ), number of meter in the experiment, our effective thermal conductivity result with
particles (Np), and particle diameter (Dp). Simulations 1–3 are pure water.
TIP3P model is quite reasonable. However, switching the water po-
Standard error of the means are calculated based on the long-time errors as
tential to SPC/E or TIP4P results in anomalous thermal conductivity
mentioned above. σr in the last column is the ratio of σ k̄ to k̄ .
ratios exceeding both experiment and theory, implying that the para-
Simulation Model ϕ (%) Np Dp (nm) k̄ (W/mK) σ k̄ σr metrization of the system dynamics is responsible for the over-
estimation rather than any physical mechanism. When the results in
1 TIP3P – – – 0.86 0.03 0.03
2 SPC/E – – – 1.02 0.03 0.03
Tables 1 and 2 are considered together, our conclusion is that the
3 TIP4P/2005 – – – 1.05 0.03 0.03 surface interactions of solid–liquid interfaces in nanocolloids need to be
4 TIP3P 5 1 1.3 1.12 0.04 0.04 accurately represented to avoid any unphysical interfacial thermal
5 SPC/E 5 1 1.3 3.73 0.12 0.03 transport.
6 TIP4P/2005 5 1 1.3 3.29 0.11 0.03
7 TIP3P 5 3 1.3 1.19 0.04 0.03
8 SPC/E 5 3 1.3 5.31 0.2 0.04 3.3. Interfacial effects
9 TIP3P 5 1 1.8 1.31 0.04 0.03
10 SPC/E 5 1 1.8 7.78 0.22 0.03
The effect of surface interaction parameters on the effective thermal
conductivity of solid–liquid systems has been investigated before in
models exceeds both theoretical and experimental results, and has been several NEMD studies [33,50], but the effect of these parameters on GK
attributed to an artificial particle correlation effect arising from the use calculations has not been studied to our knowledge. Our earlier NEMD
of a single particle with periodic boundary conditions [25]. The sug- study [33] showed up to a 12% change in thermal conductivity using an
gested solution is to use multiple nanoparticles to break the symmetry interface coupling tuned to satisfy the experimentally observed contact
of the system. This was tested for the rigid SPC/E water model using a angle as compared to Lorentz–Berthelot. Surface interactions could
three-nanoparticle system with the same volume fraction and particle have an even larger effect for EMD calculations since GK considers the
diameter. Contrary to the hypothesis that artificial particle correlations energy flow associated with the motion of each atom at each time step
are responsible for the anomalous enhancement, a further increase in by means of Eq. (2), whereas NEMD uses an average temperature cal-
the thermal conductivity was observed for the rigid SPC/E water model culated by means of kinetic theory.
(Simulation 8). Moreover, an insignificant difference was found be- A Lennard–Jones 6–12 function is employed for the interfacial in-
tween the one- and three-nanoparticles cases with the flexible TIP3P teractions:
model (Simulations 4 and 7), suggesting that there may be some other 12 6
⎡ σ σ ⎤
effects responsible for the observed anomalous increase. The difference Φ(rij) = 4ϵ ⎢ ⎜⎛ ⎟⎞ − ⎜⎛ ⎟⎞ ⎥
r r
between water models persists and is even exacerbated for larger ⎣ ⎝ ij ⎠ ⎝ ij ⎠ ⎦ (5)
system sizes at a fixed volume fraction, with the larger single particle
system having a slightly higher thermal conductivity for the TIP3P where ϵ is an energy scale and σ is a characteristic length. Ideally, ϵ and
model (Simulations 4 and 9) and more than double the thermal con- σ should be derived from experimental findings or first principle cal-
ductivity for the SPC/E model (Simulations 5 and 10). The abnormal culations; however, in most of the nanofluid studies they are calculated
thermal conductivity values associated with SPC/E models tend to in- based on Lorentz–Berthelot mixing rules. Our intention is to identify
crease with increasing particle surface area within the system as addi- whether this could be the source of the anomalous thermal conductivity
tional particles are introduced or particle size is increased (Simulations rather than some other simulation input. Existing investigations con-
8 and 10). Given the reasonable thermal conductivities of the pure sider the effects of varying ϵ [50–52] and the same procedure is fol-
water models, these findings suggest that the interfacial interactions lowed here. The effect on the thermal conductivity of varying ϵ at the
that obey the Lorentz–Berthelot rules and that are different for each water–Cu interface is reported in Table 3, indicates that the interfacial
system are the cause of observed anomalous thermal conductivities. energy parameter would easily be able to account for the over-
This argument is supported by the results presented in Table 2, estimation of the thermal conductivity in Table 1. Specifically, chan-
where the estimated thermal conductivities for the three different water ging ϵ from the 1.21 kcal/mole given by the Lorentz–Berthelot rule for
models are compared with the Maxwell limit [23] and experimental the SPC/E potential to the 0.98 kcal/mole for the TIP3P potential (Si-
results of Xuan and Li [47] at the same volume fraction. The significant mulations 1 and 3) significantly decreased the thermal conductivity.
difference between our simulations and experiment regarding the par- Although these values of ϵ do not have any particular physical basis,
ticle size could be interpreted based on the findings of the experimental they do establish that the thermal conductivity is very sensitive to
studies that investigate the effect of particle size to the nanofluid surface ϵ, and that using the values predicted by the Lorentz–Berthelot
rule is likely a significant source of error. This reinforces the necessity
of calibrating the interface potential based on experimental findings
Table 2 before using the Green–Kubo method to quantify thermal transport in
Comparison of the Green–Kubo thermal conductivity estimations with the
theoretical Maxwell limit and experimental result for water–Cu nanosuspension Table 3
at the same volume fraction [47]. Np, Dp and ϕ are the number of particles, The thermal conductivities (k̄ ) of water–Cu nanosuspensions with a single na-
particle diameter and volume fraction, respectively. keff is the thermal con- noparticle for different water models. ϵ is the interfacial energy parameter
ductivity ratio of the nanosuspension, which is the ratio of nanofluid thermal between oxygen and copper. Dp is the nanoparticle diameter.
conductivity to base fluid thermal conductivity.
Simulation Model ϵ (kcal/mole) Dp (nm) k̄ (W/mK)
ϕ (%) Np Dp (nm) keff
1 SPC/E 1.21 1.3 3.73
Maxwell [23] 5 – – 1.15 2 SPC/E 1.21 1.8 7.78
Xuan and Li [47] 5 1 100 1.55 3 SPC/E 0.98 1.3 0.91
This study (TIP3P) 5 1 1.3 1.3 4 SPC/E 0.98 1.8 0.99
This study (SPC/E) 5 1 1.3 3.66 5 TIP3P 0.98 1.3 1.11
This study (TIP4P/2005) 5 1 1.3 3.13 6 TIP3P 0.98 1.8 1.31

4
T. Akiner, et al. Materials Today Communications 20 (2019) 100533

4. Conclusions

The thermal conductivity of a water–Cu nanocolloid system was


studied using the Green–Kubo method, with the intention of identifying
the source of the anomalously high nanofluid thermal enhancements in
the literature. This is crucial for the validity of the conclusions drawn
on the basis of such simulation results. A detailed error analysis iden-
tified different sources of statistical errors, denoted as short-time errors
and long-time errors. The magnitude of the error was estimated for
both, and the long-time error associated with differences in velocity
seeding was found to be larger than the short-time error associated with
the fluctuations of a single autocorrelation function.
The anomalous thermal enhancement in the literature was re-
produced using rigid SPC/E and TIP4P/2005 water models, but was not
observed for the flexible TIP3P water model. Although the rigidity of
the water model has appeared to be a possible reason for the observed
anomalies, further simulations revealed that the discrepancy is related
Fig. 4. Two water blocks have been created in between three copper blocks to to the difference in the interfacial potential, pointing to the unsuit-
apply NEMD.
ability of the Lorentz–Berthelot rules when calculating thermal con-
ductivity with the Green–Kubo formulation. We conclude that the in-
Table 4 terfacial potential parameters should be carefully optimized to correctly
The thermal resistivities (R) at water–Cu interface with for different water simulate heat transport for solid-liquid systems when using the Green-
potentials. σR is the associated error in the measurement. Kubo method.
Model R [×10−9 Km2/W] σR
References
TIP3P 2.73 0.06
SPC/E 1.94 0.06
[1] W.B. Russel, D.A. Saville, W.R. Schowalter, Colloidal Dispersions, Cambridge
TIP4P/2005 1.68 0.05
University Press, 1989.
[2] A. Nguyen, H.J. Schulze, Colloidal Science of Flotation vol. 118, CRC Press, 2003.
[3] T.F. Tadros, Handbook of Colloid and Interface Science: Industrial Applications-
nanocolloids. Agrochemicals, Paints, Coatings and Food Systems, Walter de Gruyter GmbH & Co
KG, 2017.
Another observation is that the anomalous thermal conductivity [4] R. Prasher, P. Bhattacharya, P.E. Phelan, Thermal conductivity of nanoscale col-
values observed with both rigid models in Tables 1 and 2 were not loidal solutions (nanofluids), Phys. Rev. Lett. 94 (2) (2005) 025901.
observed with the flexible TIP3P model. While this could appear as a [5] F. Müller-Plathe, A simple nonequilibrium molecular dynamics method for calcu-
lating the thermal conductivity, J. Chem. Phys. 106 (14) (1997) 6082–6085.
possible reason for the abnormally high thermal conductivites, this
[6] R. Kubo, The fluctuation–dissipation theorem, Rep. Prog. Phys. 29 (1) (1966) 255.
conclusion is not supported by the reasonable thermal conductivities of [7] P.K. Schelling, S.R. Phillpot, P. Keblinski, Comparison of atomic-level simulation
the base fluids in Table 1. Moreover, Table 3 provides direct evidence methods for computing thermal conductivity, Phys. Rev. B 65 (14) (2002) 144306.
that an inaccurate interfacial potential could be responsible for the [8] R. Vogelsang, C. Hoheisel, G. Ciccotti, Thermal conductivity of the Lennard–Jones
liquid by molecular dynamics calculations, J. Chem. Phys. 86 (11) (1987)
anomalous thermal conductivities. 6371–6375.
The significance of the interface to the observed dependence of [9] F. Bresme, J.W. Biddle, J.V. Sengers, M.A. Anisimov, Communication: minimum in
thermal transport on the water model could also be expressed in terms the thermal conductivity of supercooled water: a computer simulation study, J.
Chem. Phys. 140 (16) (2014) 161104.
of interfacial thermal resistance. Based on prior work [53], this is [10] T.W. Sirk, S. Moore, E.F. Brown, Characteristics of thermal conductivity in classical
measured using the NEMD method. Three different water-Cu systems water models, J. Chem. Phys. 138 (6) (2013) 064505.
were generated with the geometry shown in Fig. 4, and the Müller- [11] N.J. English, J.S. Tse, Thermal conductivity of supercooled water: an equilibrium
molecular dynamics exploration, J. Phys. Chem. Lett. 5 (21) (2014) 3819–3824.
Plathe algorithm [5] was applied to create a heat flux in the system in [12] E.J. Rosenbaum, N.J. English, K.J. Johnson, D.W. Shaw, R.P. Warzinski, Thermal
the z-direction. The temperature was calculated using a methodology conductivity of methane hydrate from experiment and molecular simulation, J.
presented elsewhere [53]. The interfacial thermal resistance was cal- Phys. Chem. B 111 (46) (2007) 13194–13205.
[13] L. Li, Y. Zhang, H. Ma, M. Yang, Molecular dynamics simulation of effect of liquid
culated as R = ΔT/q″ where R is the thermal resistance, ΔT is the
layering around the nanoparticle on the enhanced thermal conductivity of nano-
temperature difference at the water–Cu interface, and q″ is the heat fluids, J. Nanopart. Res. 12 (3) (2010) 811–821.
flux. Each system was equilibrated for 100 ps in the NPT ensemble, [14] K.-L. Teng, P.-Y. Hsiao, S.-W. Hung, C.-C. Chieng, M.-S. Liu, M.-C. Lu, Enhanced
thermal conductivity of nanofluids diagnosis by molecular dynamics simulations, J.
followed by a 1 ns production run in the NVE ensemble. The thermal
Nanosci. Nanotechnol. 8 (7) (2008) 3710–3718.
resistance results were obtained by averaging the results of 100 in- [15] S. Sarkar, R.P. Selvam, Molecular dynamics simulation of effective thermal con-
dependent runs, and are presented in Table 4. A higher thermal re- ductivity and study of enhanced thermal transport mechanism in nanofluids, J.
sistance was found for the TIP3P model compared to the rigid models, Appl. Phys. 102 (7) (2007) 074302.
[16] N. Sankar, N. Mathew, C.B. Sobhan, Molecular dynamics modeling of thermal
and is consistent with our thermal conductivity results. The discrepancy conductivity enhancement in metal nanoparticle suspensions, Int. Commun. Heat
between the relative difference in the thermal resistances (around 60%) Mass Transfer 35 (7) (2008) 867–872.
and the thermal conductivities (more than 200%) could be due to the [17] J. Li, L. Porter, S. Yip, Atomistic modeling of finite-temperature properties of
crystalline β-SiC: II. Thermal conductivity and effects of point defects, J. Nucl.
fact that the thermal resistances were calculated using the NEMD Mater. 255 (2-3) (1998) 139–152.
Müller-Plathe algorithm, whereas the thermal conductivities were cal- [18] J. Chen, G. Zhang, B. Li, How to improve the accuracy of equilibrium molecular
culated using the EMD Green–Kubo algorithm. A second contributing dynamics for computation of thermal conductivity? Phys. Lett. A 374 (23) (2010)
2392–2396.
factor could be the difference in geometry: whereas the nanocolloid [19] A.J.H. McGaughey, M. Kaviany, Phonon transport in molecular dynamics simula-
contains highly curved surfaces, the thermal resistances are calculated tions: formulation and thermal conductivity prediction, Adv. Heat Transfer 390
for flat surfaces. (2006) 169–255.
[20] R.E. Jones, K.K. Mandadapu, Adaptive Green–Kubo estimates of transport coeffi-
cients from molecular dynamics based on robust error analysis, J. Chem. Phys. 136
(15) (2012) 154102.
[21] L.S. Oliveira, P.A. Greaney, Method to manage integration error in the Green–Kubo
method, Phys. Rev. E 95 (2) (2017) 023308.

5
T. Akiner, et al. Materials Today Communications 20 (2019) 100533

[22] L. Ercole, A. Marcolongo, S. Baroni, Accurate thermal conductivities from optimally [37] J.L.F. Abascal, C. Vega, A general purpose model for the condensed phases of water:
short molecular dynamics simulations, Sci. Rep. 7 (1) (2017) 15835. TIP4P/2005, J. Chem. Phys. 123 (23) (2005) 234505.
[23] J.C. Maxwell, W. Garnett, P. Pesic, An Elementary Treatise on Electricity, Courier [38] H.J.C. Berendsen, J.R. Grigera, T.P. Straatsma, The missing term in effective pair
Corporation, 2005. potentials, J. Phys. Chem. 91 (24) (1987) 6269–6271.
[24] P. Keblinski, S.R. Phillpot, S.U.S. Choi, J.A. Eastman, Mechanisms of heat flow in [39] J.-P. Ryckaert, G. Ciccotti, H.J.C. Berendsen, Numerical integration of the Cartesian
suspensions of nano-sized particles (nanofluids), Int. J. Heat Mass Transfer 45 (4) equations of motion of a system with constraints: molecular dynamics of n-alkanes,
(2002) 855–863. J. Comput. Phys. 23 (3) (1977) 327–341.
[25] M.G. Muraleedharan, D.S. Sundaram, A. Henry, V. Yang, Thermal conductivity [40] P. Mark, L. Nilsson, Structure and dynamics of the TIP3P, SPC, and SPC/E water
calculation of nano-suspensions using Green–Kubo relations with reduced artificial models at 298 K, J. Phys. Chem. A 105 (43) (2001) 9954–9960.
correlations, J. Phys.: Condens. Matter 29 (15) (2017) 155302. [41] D.E. Sanders, A.E. DePristo, Metal/metal homo-epitaxy on fcc (001) surfaces: is
[26] P. Sachdeva, R. Kumar, Effect of hydration layer and surface wettability in en- there transient mobility of adsorbed atoms? Surf. Sci. 254 (1-3) (1991) 341–353.
hancing thermal conductivity of nanofluids, Appl. Phys. Lett. 95 (22) (2009) [42] S. Plimpton, Fast parallel algorithms for short-range molecular dynamics, J.
223105. Comput. Phys. 117 (1) (1995) 1–19.
[27] H. Kang, Y. Zhang, M. Yang, Molecular dynamics simulation of thermal con- [43] W. Humphrey, A. Dalke, K. Schulten, VMD: visual molecular dynamics, J. Mol.
ductivity of Cu–Ar nanofluid using EAM potential for Cu–Cu interactions, Appl. Graph. 14 (1) (1996) 33–38.
Phys. A 103 (4) (2011) 1001. [44] M.H. Khadem, A.P. Wemhoff, Comparison of Green–Kubo and NEMD heat flux
[28] H. Babaei, P. Keblinski, J.M. Khodadadi, A proof for insignificant effect of Brownian formulations for thermal conductivity prediction using the Tersoff potential,
motion-induced micro-convection on thermal conductivity of nanofluids by uti- Comput. Mater. Sci. 690 (2013) 428–434.
lizing molecular dynamics simulations, J. Appl. Phys. 113 (8) (2013) 084302. [45] J.E. Turney, E.S. Landry, A.J.H. McGaughey, C.H. Amon, Predicting phonon prop-
[29] T. Akıner, H. Ertürk, K. Atalık, Prediction of thermal conductivity and shear visc- erties and thermal conductivity from anharmonic lattice dynamics calculations and
osity of water–Cu nanofluids using equilibrium molecular dynamics, ASME 2013 molecular dynamics simulations, Phys. Rev. B 79 (6) (2009) 064301.
International Mechanical Engineering Congress and Exposition, American Society [46] Y. Mao, Y. Zhang, Thermal conductivity, shear viscosity and specific heat of rigid
of Mechanical Engineers, 2013 p. V08CT09A012. water models, Chem. Phys. Lett. 5420 (2012) 37–41.
[30] W. Yu, S.U.S. Choi, The role of interfacial layers in the enhanced thermal con- [47] Y. Xuan, L. Qiang, Heat transfer enhancement of nanofluids, Int. J. Heat Fluid Flow
ductivity of nanofluids: a renovated Hamilton–Crosser model, J. Nanopart. Res. 6 21 (1) (2000) 58–64.
(4) (2004) 355–361. [48] S.P. Jang, S.U.S. Choi, Effects of various parameters on nanofluid thermal con-
[31] Y. Feng, B. Yu, P. Xu, M. Zou, The effective thermal conductivity of nanofluids ductivity, J. Heat Transfer 129 (5) (2007) 617–623.
based on the nanolayer and the aggregation of nanoparticles, J. Phys. D: Appl. Phys. [49] C.H. Chon, K.D. Kihm, S.P. Lee, S.U.S. Choi, Empirical correlation finding the role of
40 (10) (2007) 3164. temperature and particle size for nanofluid (Al2O3) thermal conductivity en-
[32] C.-J. Yu, A.G. Richter, A. Datta, M.K. Durbin, P. Dutta, Molecular layering in a hancement, Appl. Phys. Lett. 87 (15) (2005) 153107.
liquid on a solid substrate: an X-ray reflectivity study, Phys. B: Condens. Matter 283 [50] S. Maruyama, T. Kurashige, S. Matsumoto, Y. Yamaguchi, T. Kimura, Liquid droplet
(1–3) (2000) 27–31. in contact with a solid surface, Microscale Thermophys. Eng. 2 (1) (1998) 49–62.
[33] T. Akıner, J.K. Mason, H. Ertürk, Nanolayering around and thermal resistivity of the [51] F. Leroy, F. Müller-Plathe, Solid-liquid surface free energy of Lennard–Jones liquid
water–hexagonal boron nitride interface, J. Chem. Phys. 147 (4) (2017) 044709. on smooth and rough surfaces computed by molecular dynamics using the
[34] D.A. McQuarrie, Statistical Mechanics. 2000 120 University Science Books, phantom-wall method, J. Chem. Phys. 133 (4) (2010) 044110.
Sausalito, CA, 2004, p. 641. [52] F. Leroy, S. Liu, J. Zhang, Parametrizing nonbonded interactions from wetting ex-
[35] H. Babaei, P. Keblinski, J.M. Khodadadi, Equilibrium molecular dynamics de- periments via the work of adhesion: example of water on graphene surfaces, J. Phys.
termination of thermal conductivity for multi-component systems, J. Appl. Phys. Chem. C 119 (51) (2015) 28470–28481.
112 (5) (2012) 054310. [53] T. Akıner, J. Mason, H. Ertürk, Thermal characterization assessment of rigid and
[36] W.L. Jorgensen, J. Chandrasekhar, J.D. Madura, R.W. Impey, M.L. Klein, flexible water models in a nanogap using molecular dynamics, Chem. Phys. Lett.
Comparison of simple potential functions for simulating liquid water, J. Chem. 687 (2017) 270–275.
Phys. 790 (1983) 926.

You might also like