You are on page 1of 9

Journal of Molecular Liquids 319 (2020) 114183

Contents lists available at ScienceDirect

Journal of Molecular Liquids

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

The molecular dynamics simulation of thermal manner of Ar/Cu


nanofluid flow: The effects of spherical barriers size
Amirhosein Mosavi a,b, Maboud Hekmatifar c, As'ad Alizadeh d, Davood Toghraie c,
Roozbeh Sabetvand e, Aliakbar Karimipour f,⁎
a
Environmental Quality, Atmospheric Science and Climate Change Research Group, Ton Duc Thang University, Ho Chi Minh City, Vietnam
b
Faculty of Environment and Labour Safety, Ton Duc Thang University, Ho Chi Minh City, Vietnam
c
Department of Mechanical Engineering, Khomeinishahr Branch, Islamic Azad University, Khomeinishahr, Iran
d
Department of Mechanical Engineering, College of Engineering, University of Zakho, Zakho, Iraq
e
Department of Energy Engineering and Physics, Faculty of Condensed Matter Physics, Amirkabir University of Technology, Tehran, Iran
f
Institute of Research and Development, Duy Tan University, Da Nang 550000, Vietnam

a r t i c l e i n f o a b s t r a c t

Article history: In this computational work, we focus on spherical barrier effects on the thermal behaviour of Ar/Cu nanofluid
Received 8 July 2020 with molecular dynamics simulation. LAMMPS software is implemented in our study with Universal Force
Received in revised form 26 August 2020 Field and Embedded Atom Model force field for various atomic structures in the simulation box. The thermal be-
Accepted 28 August 2020
haviour study of Ar/Cu nanofluid is done with physical parameters calculations such as atomic temperature, total
Available online 31 August 2020
energy, number of nanofluid atoms at gas phase, radial distribution function, and thermal conductivity of Ar/Cu
Keywords:
nanofluid. By atomic barrier adding to our simulated plates, the atomic phase transition occurs in fewer time
Molecular dynamics simulation steps. Numerically, phase transition in the simulated nanofluid occurs in 610,000-time steps by Pt spherical bar-
Spherical barrier riers simulation (with 15 Å radius). By increasing the atomic barrier size, the number of nanofluid atoms in which
Nanofluid phase transition occur in them is increased. From these simulations results, we conclude that, heat flux in Ar/Cu
Argon nanofluid increases but thermal conductivity of the foremost constant. Numerically, the thermal conductivity of
Copper Ar/Cu nanofluid reaches to 0.016 W/m.K by atomic barrier radius increasing to 15 Å.
Phase transition © 2020 Elsevier B.V. All rights reserved.
Thermal conductivity
Thermal behaviour

1. Introduction optimization of the nanoparticles rates in base-fluid [12,13]. Numeri-


cally, experimental research calculated thermal conductivity enhance-
Nanotechnology is one of the promising areas of science. The con- ments in base-fluids were in the range 10–50%. Other experiment
cepts that seeded nanotechnology were first discussed in 1959 by works showed enhancement values of the thermal conductivity coeffi-
renowned physicist Richard Feynman in his talk There's Plenty of cient in some nanofluids more than 100% [14]. To the efficient applica-
Room at the Bottom, in which he described the possibility of synthe- tion of nanofluids thermal manner, the understanding of the heat
sis via direct manipulation of atoms [1]. Today, this method used in transfer mechanism in these nanostructures was essential. Physically,
various areas such as electronic [2–6], heat transfer [7–11], etc. In ad- Nanofluids are a promising type of heat transfer fluids, containing parti-
dition to the above, one of the achievements of nanotechnology is cles with 1 nm to 100 nm atomic size, which dispersed in a common
nanofluid. Nanofluids are suspensions of nanoparticles in common base-fluids [15–18]. The nanostructures (nanoparticles), which dis-
base-liquids such as water, argon, etc. Several works in recent de- persed in base-fluid, commonly made of metal or metal oxide materials.
cades showed that the addition of an optimized rate of nanoparticles This thermal manner of nanofluids allowing for heat energy transfer in
in cooling base-fluids optimized the thermal conductivity of these various systems in industrial applications. Nanofluid structures have
structures significantly. been assumed for applications as heat transfer fluids optimization.
Further, nanofluids convective heat transfer coefficients show a sim- However, due to the variety of these nanostructures, no agreement
ilar thermal manner, and this physical parameter increases by has been reported on the rate of benefits of using nanoparticles in com-
mon base-fluids for heat transfer mechanisms. For the first time, Choi
⁎ Corresponding author.
introduced a promising class of fluids that depends on suspending
E-mail addresses: amirhosein.mosavi@tdtu.edu.vn (A. Mosavi), nanostructures [19,20]. Choi et al. [21] calculated the thermal conduc-
aliakbarkarimipour@duytan.edu.vn (A. Karimipour). tivity of carbon nanotube (CNT) structures which inserted in oil fluid,

https://doi.org/10.1016/j.molliq.2020.114183
0167-7322/© 2020 Elsevier B.V. All rights reserved.
2 A. Mosavi et al. / Journal of Molecular Liquids 319 (2020) 114183

behaviour in simulated structures, we report physical parameters


Nomenclature
such as temperature, total energy, the temperature of Ar/Cu
nanofluid atoms, radial distribution function, number of nanofluid
Fij interparticle force on particle i by particle j (eV/Å)
atoms in the gas phase, and thermal conductivity of nanofluid struc-
Vij interparticle potential (eV)
ture. We expected our MD simulation results could be used in the op-
Fext external applied force (eV/Å)
timization of heat transfer procedures in industrial usages.
m particle mass (u)
rc cut-off distance (Å)
2. Computational method
rij distance between particle i and j (Å)
t time step (ps)
In this computational work, we use molecular dynamics simulations
T temperature (K)
to research the dynamical behaviour of Ar/Cu nanofluid between non-
v velocity of particle (Å/ps)
ideal Pt plates. Non-ideal plates in our simulations made of spherical
Natom number of particles
atomic barriers with various radiuses. Theoretically, in MD simulations
time evolution of particles over time steps and all simulation time. For
Greek symbols
estimation of the displacement of the atoms in the simulated time inter-
ε energy parameter in Lennard-Jones (LJ) potential
val, Newton's second law used as below equation [30],
σ length parameter of LJ potential
ϕ interaction potential 2
d ri dv
θ0 Equilibrium angle F i ¼ ∑F ij ¼ mi ¼ mi i ð1Þ
i≠j dt 2 dt

where mi is the atomic mass, ri is the atom position, and finally, vi is the
atom velocity. Further, Gaussian formalism used to particles tempera-
this research group conclude that the thermal conductivity of the stud- ture definition, which can be calculated by below equation [30]:
ied oil increased significantly. Ahammed et al. [22] reported the thermal
conductivity of the H2O base fluid at various temperatures. They 3 1 N 1
kB T ¼ ∑ mv2i ð2Þ
showed that the thermal conductivity of nanofluid improves with in- 2 Natom i¼1 2
creasing the optimized atomic rate of graphene nanoparticles. Numeri-
cally, the thermal behaviour of base fluid improves to 37.2% for 0.15% Computationally, these computational equations calculated by the
concentration of graphene nanosheets when compared with H2O velocity-Verlet algorithm as [30]:
base-fluid. Selvam et al. [23] prepared the promising nanofluid with
1
ethylene glycol and H2O as based-fluid. They said that the thermal con- r ðt þ Δt Þ ¼ r ðt Þ þ vðt Þ Δt þ aðt Þ Δt 2 ð3Þ
2
ductivity of nanofluids enhances with inserting nanoparticles to base-
fluid significantly. Ding et al. [24] researched about the thermal manner 1
of H2O/CNT nanofluid with an experimental approach. This research vðt þ Δt Þ ¼ vðt Þ þ ½aðt Þ þ aðt þ Δt Þ Δt ð4Þ
2
group reported the optimization of the thermal behaviour approxi-
mately reach to 350% by adding CNT nanoparticles with a 0.5% atomic In Eqs. (3) and (4), the r(t + Δt), v(t + Δt), and a(t + Δt) are final
ratio. position, velocity, and acceleration of particles. Further, r(t), v(t),
Akhavan et al. [25] showed the transport coefficients of H2O base- and a(t) are the initial rate of these parameters. In the initial step
fluid increases by graphene nanoparticle, adding to base-fluid. They of our MD simulations, we use periodic boundary conditions in y
conclude the thermal conductivity of H2 O molecules can be opti- and x directions and fix one implemented in the z-direction which
mized by adding nanostructures to them. Witharana et al. and these conditions commonly used in the MD approach [30]. The MD
Chen et al. [26,27] expressed that the science of the atomic manner simulation box length set to 100 Å, 100 Å, and 130 Å in x, y, and z di-
and effect of nanofluids is a significant physical factor in planning rections. Further, the 1 fs rate chooses for the simulation time steps
their thermal behaviour for the heat transfer process. In various ex- in all MD simulation steps. Theoretically, to get correct calculations
perimental studies, nanoparticles adding to common base-fluids in MD based simulations, one should choose the atomic force-field
cause phase transition to occur in base-fluid in shorter times. So we based on the physical phenomena which studied. The force-field
can conclude by inserting nanostructures to common base-fluids, for nanofluid and metal structures (Pt non-ideal plates and Cu nano-
the heat transfer rate in these structures improved. Today, theoreti- particle) in our simulations was based on a universal force field
cal methods like Molecular Dynamics (MD) simulation method can (UFF), and Embedded Atom Model (EAM) potentials, respectively
be used to the atomic study of various substances [28–30]. This com- [44,45]. In the UFF force field, the interaction between particles was
putational approach is used in this study of nanostructures thermal, stated by the Lennard-Jones (LJ) potential as [45]:
membranes and mechanical behaviours [31–41]. Li et al. in their sim- "   6 #
ulation work which similar our research, studied the phase transi- σ 12 σ
U ðr Þ ¼ 4ε − r ≪ rc ð5Þ
tion process of argon fluid in confined non-ideal Pt plates [42]. The r ij r ij
simulation outcomes were that the molecular dynamics simulation
approach accuracy on phase transition phenomena prediction in In Eq. (5), rij is the distance between j and j atoms, rc is the cutoff ra-
base-fluid, such as argon atoms. Further, Hekmatifar et al. studied dius, and σ and ε parameters are the energy and length scales—these
the cone barriers effect on nanofluid phase transition phenomena. computational constants selected from force-field references. Further,
The results of this computational work showed the significant im- in our MD simulations, metal structures simulated by EAM potential.
pact of atomic barriers in the thermal manner of Ar/Cu nanofluid EAM potential represented by the below equation [46]:
and the nanofluid phase transition phenomena [43]. In this work, !
we simulate Ar base-fluid thermal manner in a confined space   1  
Ei ¼ F α ∑ρβ r ij þ ∑ϕαβ r ij ð6Þ
which made of Pt plates by adding Cu nanoparticles. Pt plate in the i≠j 2 i≠j
upper region of our simulation box has a non-ideal surface with 9
spherical barriers in which the atomic size of these spherical struc- where rij is the distance between j and i atoms, ϕαβ is a pair-wise poten-
tures affected to nanofluid thermal manner. For the study of thermal tial function, ραβ is the contribution to the electron charge density from
A. Mosavi et al. / Journal of Molecular Liquids 319 (2020) 114183 3

Fig. 1. Schematic of Pt plates and Ar fluid modelled in our MD simulation study.

atom j of type β at the location of atom i, and Fα is an embedding func- Fig. 1 [51]. In this structure, the temperature of heat source and heat sink
tion that represents the energy required to place atom i of type α into defined with below equations:
the electron cloud. After the force-field definition for various atomic
structures, atom's time evolution described by Newton's law. For this, T HeatSource ¼ 107:5−Δt ð7Þ
we use the Large Scale Molecular/Atomic Massively Simulator
(LAMMPS) package. This simulation package released by Sandia Na-
T HeatSink ¼ T 4 ¼ 107:5 þ Δt ð8Þ
tional Laboratories in 1995 for the first time [47–50]. To use this stimu-
lation package to describe Ar/Cu nanofluid thermal behaviour and
phase transition phenomena, Pt non-ideal plates and H2O/Cu nanofluid In these two equations, the Δt is 22.50 K. Simulation box in perspec-
structures simulated as depicted in Fig. 1. This schematic of the atomic tive view. These temperature setting in the simulated structure imple-
system visualized by the Open Visualization Tool (OVITO) is shown in mented by Nose-Hoover thermostat [30].

Fig. 2. Temperature variation of Ar/Cu nanofluid and non-ideal Pt plates as a function of simulation time steps.
4 A. Mosavi et al. / Journal of Molecular Liquids 319 (2020) 114183

3. Discussion and results

3.1. Equilibration process of simulated structures

In the first step of this computational study to validate our MD sim-


ulation method, we equilibrated the Ar/Cu nanofluid and Pt non-ideal
plates atomic structures. This atomic behaviour was similar to previous
research and showed the appropriate choice for MD simulation setting
[43]. In this section, the temperature and total energy of structures are
calculated. Our result shows that 1,000,000-time steps are sufficient
for thermodynamic equilibration of Ar/Cu nanofluid and Pt non-ideal
plates. Figs. 2 and 3 show the temperature and total energy deference
of these atomic structures. Numerically, from these figs. We conclude
the temperature of simulated structures reaches to 85 K after
1,000,000-time steps. Further, the total energy of atomic structures
shows that this physical parameter reaches to −381 eV. The negative
magnitude of total energy shows that we choose the force field and
atomic structures appropriately, and these structures have thermal sta-
bility. Computationally, the thermal stability of simulated structures
arises from the atomic position and interatomic force-field mathemati-
cally adapting. These atomic manners show the initial atomic positions
for various structures and interatomic force between them, defined
with the appropriate way. After total energy and temperature equilibra-
tion process, we report the temperature, number of nanofluid atoms in
the gas phase, Radial Distribution Function (RDF), and thermal conduc-
tivity of simulated nanofluid.

3.2. Average temperature of Ar fluid and Ar/Cu nanofluid structures

In the second step of this molecular dynamics study, we simulated


spherical barriers with the various radius as depicted in Fig. 4. After
this atomic simulation, we calculated the average temperature of simu-
lated nanofluid to predict the phase transition phenomena in Ar/Cu
nanofluid. MD simulation results show by simulation time passing the
heat flux movement in atomic structures rises and so the average tem-
perature of Aar/Cu nanofluid increases. This atomic manner arises from
the contact surface area between Ar/Cu nanofluid and Pt plate atoms in-
creasing. This atomic contact increasing cause improves the atomic in-
teraction and finally, the nanofluid particles mobility and temperature
increases, too. Physically, by Ar/Cu nanofluid temperature increasing
the phase transition in this atomic structure occur. Fig. 5 shows the
nanofluid temperature variation in the middle region of the simulation
package as a function of time. Our simulations show that the tempera-
ture of nanofluid atoms changes from 90 K to 113 K after 1,000,000-
time steps. Numerically, the phase transition phenomena detectable
Fig. 4. Atomic modelling of Pt spherical barrier with a) 5 Å, b) 7.5 Å, and c) 10 Å radius.

Fig. 3. Total energy difference of Ar/Cu nanofluid and non-ideal Pt plates as a function of MD simulation time.
A. Mosavi et al. / Journal of Molecular Liquids 319 (2020) 114183 5

Fig. 5. Temperature variation of Ar/Cu nanofluid as a function of spherical barrier radius and simulation time steps.

after 1,000,000-time steps. By increasing of spherical barriers radius 3.3. Number of nanofluid atoms in gas atomic phase
from 5 Å to 10 Å in non-ideal Pt plate, the average temperature of
nanofluid reach to 126 K. Further, the phase transition in simulated The number of nanofluid atoms in various atomic phase is an impor-
nanofluid detectable after 425,000-time steps. Time evolution of Ar/Cu tant parameter for heat transfer applications. We report the number of
nanofluid temperature and phase transition time of nanofluid atomic Ar/Cu nanofluid atoms in the gas phase after the phase transition pro-
structure reported in Table 1 as a function of barrier radius. cess in our MD simulations. Our results show that the number of
nanofluid atoms in the gas phase increases to 3136 from 2562 after
1,000,000-time steps. By radius of spherical barrier enlarging from 5 Å
to 10 Å, the number of nanofluid atoms in the gas phase reaches to
3214. Previous computational work shows a similar atomic manner
[42]. Li et al. show that atomic barriers have a direct effect on the num-
ber of atoms in the phase transition process, and by adding this atomic
Table 1 parameter to simulated structures, phase transition phenomena occur
Average temperature and phase transition time in Ar/Cu nanofluid as a function of spher- effectively. Physically, this atomic improvement arises from increasing
ical barrier radius and simulation time steps.
of atomic contact between plates and nanofluid particles which cause
Barrier radius The temperature in final time step Phase transition time the heat flux increasing in the simulated structures. By heat flux increas-
(Å) (K) (ns) ing between the plate and nanofluid particles, the acceleration of
5 113 771,000 nanofluid particles rises, and displacement of these particles get harder.
7.5 120 510,000 By this mechanical parameter changes, the atomic distance between
10 126 425,000
nanofluid particles increases and the gas phase in this structure

Fig. 6. Number of gas atoms in Ar/Cu nanofluid as a function of the molecule dynamics simulation time step and Pt barriers radius.
6 A. Mosavi et al. / Journal of Molecular Liquids 319 (2020) 114183

Table 2 3.5. Thermal conductivity of simulated nanofluid


Number of gas atoms in Ar/Cu nanofluid and needed MD simulation time step for atomic
phase transition to occur.
In industrial applications of nanostructures as a heat transfer appara-
Barrier Number of nanofluid atoms in gas Time step of maximum rate of tus, the thermal conductivity of them is an important parameter. In the
radius (Å) phase (atoms) phase transition last step of this MD simulations, we compute the thermal conductivity
5 3136 780,000 of Ar/Cu nanofluid with 2 simulation methods: equilibrium and non-
7.5 3189 670,000 equilibrium methods. In equilibrium calculations, we use the Green-
10 3214 610,000
Kubo (GK) method [53,54]. In this computational method, thermal coef-
ficient calculated by integrating time autocorrelation functions of heat
flux with the equation below [54]:
detectable. Fig. 6 shows the number of nanofluid atoms in the gas phase
as a function of spherical barrier radius and MD simulation time steps. Z∞
1
Further, our MD simulations show that by atomic barriers enlarging, λ¼ <J ðt Þ:J ð0Þ > ð10Þ
T 2 kB V
the time step of MD simulations to reach a maximum rate of atoms in 0

the gas phase decreases from 780,000 to 610,000-time steps, as re-


ported in Table 2. In this formula, V is the volume of the atomic system, J(t) is the heat
flux and T is the temperature of structure, and finally, kB is the
3.4. Phase transition process of simulated Nanofluid Boltzmann constant. GK approach predicted the thermal conductivity
of Ar/Cu nanofluid is 0.016 W/m.K. Further, the Fourier approach used
The radial distribution function (RDF) is one of the physical param- for non-equilibrium method simulation of nanofluid thermal conduc-
eter in common molecular dynamics simulation. In our study, RDF re- tivity. According to the heat conduction, that indicates that the heat
ported shows the probability of fluid atom finding at a distance r from flow per unit area in unit time is proportional to the temperature varia-
other fluid atoms. This physical parameter presented by this Eq. (9) tion at the cross-section of structures. In this method, we can write [54]:
[30]:
Q=AΔt
λ¼ ð11Þ
dnr dT=dz
g ðr Þ ¼ ð9Þ
4π  dr  ρ
In Eq. (11), Q is the heat flow of the structure, and A is the cross-
In this equation, ρ is the density, dnr is a function that calculates the section, Δt is the time difference and, dT/dz. is the temperature gradient
number of fluid atoms within a shell of thickness dr. RDF analyze result in the atomic system. By estimating the temperature and heat flow dif-
difference in this simulations show the phase transition phenomena ference in MD simulations, the thermal conductivity calculated by
and convert liquid phase to gas one in Ar/Cu nanofluid structure. Eq. (11). To thermal conductivity calculation from the non-
Based on Fig. 7, we can express that the Ar atomic structure in 0-time equilibrium approach, the simulation package is divided into 100 slabs
step has liquid phase, by MD simulation time evolution, RDF curve of with 100
D thickness, where D is the length of the simulation box. The
Ar fluid show the phase transition process after 450,000-time steps. Fur- temperature of the heat source region is fixed at 157.5 K, and the tem-
ther, this calculated RDF curve for liquid phase of Ar fluid is similar to perature of heat sink temperature is equilibrated at 57.5 K. For temper-
earlier reports, which shows the correctness of this molecular dynamic ature equilibration in our MD simulations, we use the NPT ensemble
simulations settings and validation of our simulation procedure [52]. with 0.01 and 0.1 values for pressure and temperature damping rates.
Further, the phase transition process and attraction of fluid atoms to After this equilibrium process, we implement the NVE ensemble for
up non-ideal Pt plate as a function of simulation time steps, depicted 1,000,000-time steps to reach constant heat flow in a simulated
in Fig. 8. As depicted in this Fig. By simulation time passing the number nanofluid. After this computational process, MD simulations run contin-
of fluid atoms in the gas phase increases, and we conclude the phase ued for 1,000,000-time steps later and finally, the thermal conductivity
transition process occurs in the simulated structure. of simulated nanofluid reported. This computational procedure

Fig. 7. RDF curve of Ar\


\Ar in nanofluid structure as a function of the MD simulation time step.
A. Mosavi et al. / Journal of Molecular Liquids 319 (2020) 114183 7

estimated the thermal conductivity of nanofluid equal by 0.016 W/mK.


Fig. 9 depicted the thermal conductivity changes of Ar/Cu nanofluid as a
function of MD simulation time steps. From this Fig. We conclude as MD
simulation running, the thermal conductivity of nanofluid decreases
and then converged to a constant value. By atomic barrier enlarging,
we expect the heat flux in simulated structures increases, but thermal
conductivity of them remain consistent. The maximum rate of thermal
conductivity reaches to converged rate after 355,000 and 493,000-
time steps for the smallest and most significant atomic barriers in equi-
librium MD simulations. This atomic manner shows the thermal inter-
action between nanofluid particles and Pt atoms done in shorter time
steps by atomic barrier size increasing. This MD simulation result can
be used for thermal conductance structures designing to reach the opti-
mal performance of them. Numerically, 0.018 W/m.K and 0.016 W/m.K
rates for thermal conductivity of Ar/Cu nanofluid thermal conductivity
calculated with non-equilibrium MD simulations as reported in Fig. 10
and Table 3. We can conclude that by atomic barriers increasing the
atomic interaction between nanofluid and plate atoms increases and
the amplitude of atomic fluctuations increases. These physical phenom-
ena cause mobility increasing in simulated structures, and so the ther-
mal coefficient of simulated nanofluid increases.

4. Conclusion

In our computational study, the thermal behaviour of Ar/Cu nonfluid


between the ideal and non-ideal Pt plates reported. In these MD simula-
tions, the barriers size effect on Ar/Cu nanofluid phase transition proce-
dure reported for the first time, and we expected these effects, optimize
the thermal conductance procedures in industrial applications. Our re-
sults show that the presence of the atomic barrier is having an essential
impact on the phase transition process of simulated nanofluid. Atomic
barriers in our simulation are spherical shape. Physically, by spherical
barrier's radius increasing, the total heat flux in atomic structures in-
creases and thermal conductivity of Ar/Cu nanofluid converged in a
shorter time. Numerically, we conclude that:
• Spherical barrier radius increasing cause phase transition in Ar/Cu
nanofluid occur in shorter time steps (T = 780,000 to T = 610,000).
• The average temperature of Ar/Cu nanofluid increases from 113 K to
126 K by spherical barrier radius enhancing from 5 Å to 10 Å.
• The number of Ar/Cu nanofluid atoms in the gas phase increases from
3136 to 3214, by spherical barriers radius increasing from 5 Å to 10 Å.
• Thermal conductivity of Ar/Cu nanofluid reach to 0.016 W/m.K by
using molecular equilibrium dynamics, and non-equilibrium molecu-
lar dynamics approach.

These outcomes of our simulations from LAMMPS software can be


used in difference industrial procedures like heat transfer mechanisms
designing to increase their efficiency.

CRediT authorship contribution statement

Amirhosein Mosavi: Writing - review & editing, Investigation.


Maboud Hekmatifar: Methodology, Software, Validation, Writing - re-
view & editing, Investigation. As'ad Alizadeh: Methodology, Software,
Validation, Investigation. Davood Toghraie: Methodology, Software,
Validation, Writing - original draft, Investigation. Roozbeh Sabetvand:
Methodology, Software, Validation, Writing - review & editing, Writing
- original draft. Aliakbar Karimipour: Writing - review & editing, Writ-
ing - original draft.

Fig. 8. The Ar atoms arrangement in the simulation box as a function of the molecular
dynamics simulation time step: a) T = 0, b) T = 500,000, and c) T = 1,000,000 time steps.
8 A. Mosavi et al. / Journal of Molecular Liquids 319 (2020) 114183

Fig. 9. Thermal conductivity variation of Ar/Cu nanofluid as a function of the simulation time step and atomic barrier height (equilibrium MD simulations).

Fig. 10. Thermal conductivity variation of Ar/Cu nanofluid as a function of the simulation time step and atomic barrier height (non-equilibrium MD simulations).

References

Table 3 [1] K. Eric Drexler, Engines of Creation: The Coming Era of Nanotechnology, Doubleday,
Calculated thermal conductivity of Ar/Cu nanofluid by using equilibrium and non-equilib- 1986 (ISBN 978-0-385-19973-5).
rium methods as a function of atomic barrier heights. [2] H. Karimi-Maleh, M. Sheikhshoaie, I. Sheikhshoaie, M. Ranjbar, J. Alizadeh, N.W.
Maxakato, A. Abbaspourrad, A Novel Electrochemical Epinine Sensor Using Ampli-
Barrier height Equilibrium MD method Non-equilibrium MD method fied CuO Nanoparticles and an-hexyl-3-methylimidazolium.
(Å) (W/mK) (W/mK) [3] F. Tahernejad-Javazmi, M. Shabani-Nooshabadi, H. Karimi-Maleh, 3D reduced
graphene oxide/FeNi3-ionic liquid nanocomposite modified sensor; an electrical
5 0.017 0.018
synergic effect for development of tert-butylhydroquinone and folic acid sensor,
7.5 0.016 0.016
Compos. Part B 172 (2019) 666–670.
10 0.016 0.016
[4] H. Karimi-Maleh, M. Sheikhshoaie, I. Sheikhshoaie, M. Ranjbar, J. Alizadeh, N.W.
Maxakato, A. Abbaspourrad, A novel electrochemical epinine sensor using amplified
CuO nanoparticles and an-hexyl-3-methylimidazolium hexafluorophosphate elec-
trode, New J. Chem. 43 (2019) 2362–2367.
[5] A. Khodadadi, E. Faghih-Mirzaei, H. Karimi-Maleh, A. Abbaspourrad, S. Agarwal, V.K.
Gupta, A new epirubicin biosensor based on amplifying DNA interactions with poly-
Declaration of competing interest pyrrole and nitrogen-doped reduced graphene: experimental and docking theoret-
ical investigations, Sensors Actuators B Chem. 284 (2019) 568–574.
The authors declare that they have no known competing financial [6] Z. Shamsadin-Azad, M.A. Shamsadin-Azad, S. Cheraghi, H. Karimi-Maleh, A nano-
structure voltammetric platform amplified with ionic liquid for determination of
interests or personal relationships that could have appeared to influ-
tert-butylhydroxyanisole in the presence kojic acid, Journal of Food Measurement
ence the work reported in this paper. and Characterization 13 (2019) 1781–1787.
A. Mosavi et al. / Journal of Molecular Liquids 319 (2020) 114183 9

[7] M. Zarringhalam, H. Ahmadi-Danesh-Ashtiani, D. Toghraie, R. Fazaeli, The effects of [32] E. Hosseini, W. Cha-Umpong, A. Razmjou, M. Zakertabrizi, A. HabibnejadKorayem, V.
suspending Copper nanoparticles into Argon base fluid inside a microchannel under Chen, New molecular understanding of hydrated ion trapping mechanism during
boiling flow condition by using of molecular dynamic simulation, J. Mol. Liq. 293 thermally-driven desalination by pervaporation using GO membrane, J. Membr.
(2019) 111474, https://doi.org/10.1016/j.molliq.2019.111474. Sci. Res. 598 (2020), 117687, https://doi.org/10.1016/j.memsci.2019.117687.
[8] M. Zarringhalam, H. Ahmadi-Danesh-Ashtiani, D. Toghraie, R. Fazaeli, Molecular dy- [33] M. Rezaei, A.R. Azimian, D. Toghraie, Molecular dynamics study of an electro-kinetic
namic simulation to study the effects of roughness elements with cone geometry on fluid transport in a charged nanochannel based on the role of the stern
the boiling flow inside a microchannel, Int. J. Heat Mass Transf. 141 (2019) 1–8. layerapproaches, Physica A 426 (2015) 25–34, https://doi.org/10.1016/j.physa.
[9] P. Alipour, D. Toghraie, A. Karimipour, Investigation the atomic arrangement and 2015.01.043.
stability of the fluid inside a rough nanochannel in both presence and absence of dif- [34] M. Goel, S.P. Harsha, S. Singh, A.K. Sahani, Analysis of temperature, helicity and size
ferent roughness by using of accurate nano scale simulation, Physica A: Statistical effect on the mechanical properties of carbon nanotubes using molecular dynamics
Mechanics and Its Applications 524 (2019) 639–660, https://doi.org/10.1016/j. simulation, Materials Today: Proceedings 26 (2) (2020) 897–904, https://doi.org/
physa.2019.04.243. 10.1016/j.matpr.2020.01.130.
[10] A. Ahmadi Balootaki, A. Karimipour, D. Toghraie, Nano scale lattice Boltzmann [35] D. Toghraie, M. Hekmatifar, Y. Salehipour, M. Afrand, Molecular dynamics simula-
method to simulate the mixed convection heat transfer of air in a lid-driven cavity tion of Couette and Poiseuille water-copper nanofluid flows in rough and smooth
with an endothermic obstacle inside, Physica A: Statistical Mechanics and Its Appli- nanochannels with different roughness configurations, Chem. Phys. 527 (2019),
cations 508 (2018) 681–701. 110505.
[11] D. Toghraie, M. Mokhtari, M. Afrand, Molecular dynamic simulation of copper and [36] M. Aghababaie, M. Beheshti, A. Razmjou, A. Bordbar, Covalent immobilization of
platinum nanoparticles Poiseuille flow in a nanochannels, Physica E: Low- Candida rugosa lipase on a novel functionalized Fe3O4@SiO2 dip-coated nanocom-
Dimensional Systems and Nanostructures 84 (2016) 152–161. posite membrane, Food Bioprod. Process. 100 (2016) 351–360, https://doi.org/10.
[12] B.A. Bhanvase, D.P. Barai, S.H. Sonawane, N. Kumar, S.S. Sonawane, Intensified heat 1016/j.fbp.2016.07.016.
transfer rate with the use of nanofluids, Handbook of Nanomaterials for Industrial
[37] M. Mohammad, A. Razmjou, K. Liang, M. Asadnia, V. Chen, Metal–Organic-Frame-
Applications 2018, pp. 739–750.
work-Based Enzymatic Microfluidic Biosensor via Surface Patterning and Biominer-
[13] B.A. Bhanvase, M.R. Sarode, L.A. Putterwar, A. K.A., M.P. Deosarkar, S.H. Sonawane,
alization, ACS Appl. Mater. Interfaces 11 (2) (2019) 1807–1820, https://doi.org/10.
Intensification of convective heat transfer in water/ethylene glycol-based
1021/acsami.8b16837.
nanofluids containing TiO2 nanoparticles, Chem. Eng. Process. Process Intensif. 82
[38] H. Khan, A. Razmjou, M. Ebrahimi Warkian, A. Kottapalli, M. Asadnia, Sensitive and
(2014) 123–131.
Flexible Polymeric Strain Sensor for Accurate Human Motion Monitoring, Sensors
[14] P. Gurav, S. Naik, B.A. Bhanvase, D.V. Pinjari, S.H. Sonawane, M. Ashokkumar, Heat
18 (2) (2018), 418, https://doi.org/10.3390/s18020418.
transfer intensification using polyaniline based nanofluids: preparation and applica-
tion, Chem. Eng. Process. Process Intensif. 95 (2015) 195–201. [39] Z. Changani, A. Razmjou, A. Taheri-Kafrani, M. Ebrahimi Warkiani, M. Asadnia, Sur-
[15] C. Yue, D. Han, W. Pu, W. He, Parametric analysis of a vehicle power and cooling/ face modification of polypropylene membrane for the removal of iodine using
heating cogeneration system, Energy 115 (2016) 800–810. polydopamine chemistry, Chemosphere 249 (2020), 126079, https://doi.org/10.
[16] U. Choi, D.M. France, B.D. Knodel, Impact of advanced fluids on costs of district 1016/j.chemosphere.2020.126079.
cooling systems, Argonne National Lab, 1992. [40] A. Ghanbari, F. Warchomicka, C. Sommitsch, A. Zamanian, Investigation of the oxida-
[17] U. Choi, T. Tran, Experimental studies of the effects of non-Newtonian surfactant so- tion mechanism of dopamine functionalization in an AZ31 magnesium alloy for bio-
lutions on the performance of a shell-and-tube heat exchanger, Recent Develop- medical applications, Coatings 9 (9) (2019) 584.
ments in Non-Newtonian Flows and Industrial Applications, the American Society [41] S.G. Volz, G. Chen, Molecular-dynamics simulation of thermal conductivity of silicon
of Mechanical Engineers, New York, NY, USA, 1991. crystals, Phys. Rev. B 61 (4) (2000) 2651–2656.
[18] Y. Xuan, Q. Li, Heat transfer enhancement of nanofluids, Int. J. Heat Fluid Flow 21 (1) [42] L. Li, P. Ji, Y. Zhang, Molecular dynamics simulation of condensation on the nano-
(2000) 58–64. structured surface in a confined space, Applied Physics A 122 (5) (2016).
[19] S.U.S. Choi, Jeffrey Eastman, Enhancing thermal conductivity of fluids with nanopar- [43] M. Hekmatifar, D. Toghraie, B. Mehmandoust, F. Aghadavoudi, S.A. Eftekhari, Molec-
ticles, Proceedings of the ASME International Mechanical Engineering Congress and ular dynamics simulation of the phase transition process in the atomic scale for Ar/
Exposition (1995) 66. Cu nanofluid on the platinum plates, International Communications in Heat and
[20] M. Hashemian, S. Jafarmadar, J. Nasiri, H. Sadighi Dizaji, Enhancement of heat trans- Mass Transfer 117 (2020), 104798.
fer rate with structural modification of double pipe heat exchanger by changing the [44] A.K. Rappe, C.J. Casewit, K.S. Colwell, W.A. Goddard, W.M. Skiff, UFF, a full periodic
cylindrical form of tubes into conical form, Appl. Therm. Eng. 118 (2017) 408–417. table force field for molecular mechanics and molecular dynamics simulations, J.
[21] S.U.S. Choi, Z.G. Zhang, W. Yu, F.E. Lockwood, E.A. Grulke, Anomalous thermal con- Am. Chem. Soc. 114 (25) (1992) 10024–10035.
ductivity enhancement in nanotube suspensions, Appl. Phys. Lett. 79 (14) (2001) [45] Murray S. Daw, Mike Baskes, Embedded-atom method: derivation and application
2252–2254. to impurities, surfaces, and other defects in metals, Physical Review B. American
[22] N. Ahammed, L.G. Asirvatham, J. Titus, J.R. Bose, S. Wongwises, Measurement of Physical Society 29 (12) (1984) 6443–6453 (Bibcode: 1984PhRvB..29.6443D).
thermal conductivity of graphene–water nanofluid at below and above ambient [46] J.E. Lennard-Jones, On the determination of molecular fields, Proc. R. Soc. Lond. A
temperatures, International Communications in Heat and Mass Transfer 70 (2016) 106 (738) (1924) 463–477 (Bibcode: 1924RSPSA.106..463J).
66–74. [47] S. Plimpton, Fast parallel algorithms for short-range molecular dynamics, J. Comput.
[23] C. Selvam, D.M. Lal, S. Harish, Thermal conductivity enhancement of ethylene glycol Phys. 117 (1) (1995) 1–19.
and water with graphene nanoplatelets, Thermochim. Acta 642 (2016) 32–38. [48] S.J. Plimpton, A.P. Thompson, Computational aspects of many-body potentials, MRS
[24] Y. Ding, H. Alias, D. Wen, R.A. Williams, Heat transfer of aqueous suspensions of car- Bull. 37 (05) (2012) 513–521.
bon nanotubes (CNT nanofluids), Int. J. Heat Mass Transf. 49 (1–2) (2006) 240–250.
[49] W.M. Brown, P. Wang, S.J. Plimpton, A.N. Tharrington, Implementing molecular dy-
[25] H. Akhavan-Zanjani, M. Saffar-Avval, M. Mansourkiaei, M. Ahadi, F. Sharif, Turbulent
namics on hybrid high-performance computers – short-range forces, Comput. Phys.
convective heat transfer and pressure drop of graphene–water nanofluid flowing in-
Commun. 182 (4) (2011) 898–911.
side a horizontal circular tube, J. Dispers. Sci. Technol. 35 (9) (2014) 1230–1240.
[50] W.M. Brown, A. Kohlmeyer, S.J. Plimpton, A.N. Tharrington, Implementing molecular
[26] S. Witharana, H. Chen, Y. Ding, Stability of nanofluids in quiescent and shear flow
dynamics on hybrid high-performance computers – particle–particle particle-mesh,
fields, Nanoscale Res. Lett. 6 (2011) 231, http://www.nanoscalereslett.com/con-
Comput. Phys. Commun. 183 (3) (2012) 449–459.
tent/6/1/231/.
[27] H. Chen, S. Witharana, et al., Predicting thermal conductivity of liquid suspensions of [51] A. Stukowski, Visualization and analysis of atomistic simulation data with OVITO–
nanoparticles (nanofluids) based on rheology, Particuology 7 (2) (2009) 151–157. the open visualization tool, Model. Simul. Mater. Sci. Eng. 18 (1) (2009) 015012.
[28] B.J. Alder, T.E. Wainwright, Studies in molecular dynamics. I. General method, J. [52] B.J. Yoon, M.S. Jhon, H. Eyring, The radial distribution function of liquid argon ac-
Chem. Phys. 31 (2) (1959) 459 (Bibcode: 1959JChPh..31..459A). cording to significant structure theory, Proc. Natl. Acad. Sci. 78 (11) (1981)
[29] A. Rahman, Correlations in the motion of atoms in liquid argon, Phys. Rev. 136 (2A) 6588–6591.
(19 October 1964) A405–A411 (Bibcode: 1964PhRv..136..405R). [53] Melville S. Green, Markoff random processes and the statistical mechanics of time-
[30] D.C. Rapaport, The Art of Molecular Dynamics Simulation, 1996 (ISBN 0-521- dependent phenomena. II. Irreversible processes in fluids, J. Chem. Phys. 22 (3)
44561-2). (1954) 398–413.
[31] N.A. Jolfaei, N.A. Jolfaei, M. Hekmatifar, A. Piranfar, D. Toghraie, R. Sabetvand, S. [54] Ryogo Kubo, Statistical-mechanical theory of irreversible processes. I. General the-
Rostami, Investigation of thermal properties of DNA structure with precise atomic ory and simple applications to magnetic and conduction problems, J. Phys. Soc.
arrangement via equilibrium and non-equilibrium molecular dynamics approaches, Jpn. 12 (6) (1957-06-15) 570–586.
Comput. Methods Prog. Biomed. (2019) 105169.

You might also like