12
Molecular dynamics simulation of uid ow passing through a nanochannel: Effects of geometric shape of roughnesses Pedram Alipour a , Davood Toghraie a, , Arash Karimipour b , Mehdi Hajian c a Department of Mechanical Engineering, Khomeinishahr Branch, Islamic Azad University, Khomeinishahr, Iran b Department of Mechanical Engineering, Najafabad Branch, Islamic Azad University, Najafabad, Iran c Department of Mechanical Engineering, Isfahan University of Technology, Isfahan 84156-83111, Iran abstract article info Article history: Received 23 June 2018 Received in revised form 17 August 2018 Accepted 14 November 2018 Available online 20 November 2018 In this paper, we investigate the effects of geometric shape of various roughnesses on the uid ow passing through a nanochannel by using of molecular dynamics simulation. The results of simulations are presented for the modeled structures (the ve models dened) as number density, velocity, and system temperature pro- les for various conditions. By applying roughness to the inner surface of the ideal nano-channel at a thrust force of 0.002 eV/Å, the amplitude of number density of the uid particles near the walls decreased, while the mean and maximum velocities increased by 6.5% and 2.5% in the presence of square cuboid and hemispheroid rough- ness, respectively. Furthermore, the dimensionless slip velocity and slip length were, respectively, increased by a maximum of 41.1% and 21.5% in the presence of square cuboid roughness and by a minimum of 0.9% and 0.5% in the presence of hemispheroid roughness. The temperature of the particles at the center of the nano-channel was increased by a maximum of 9.1% and a minimum of 2.8% in the presence of square cuboid and hemispheroid roughness, respectively. Calculation of the Argon-Argon radial distribution function indicated that the maximum of this function decreased by a maximum of 11.8% and a minimum of 8.5% in the presence of rectangular cuboid and ellipsoid roughness, respectively, compared to the ideal nano-channel. © 2018 Elsevier B.V. All rights reserved. Keywords: Geometric shape Nanochannel Molecular dynamics simulation Radial distribution function Roughness 1. Introduction The molecular dynamics method has been effective in treating nanouidic environments. Such atomistic simulations have been em- bedded in numerous studies of nanochannel ows over the last years. Kim et al. [1] present a new interactive thermal wall model for sim- ulate the ow and heat transfer in nanoscale channels. They found that when the interaction potential is weak or the wall has a large crystal bonding stiffness, temperature jumps on the wall/uid interface are ob- served. Das et al. [2] studied the effect of interface wettability on the transport of water (assuming it to be a simple Lennard Jones uid) in a nanochannel by using of molecular dynamics simulation. They showed that in channels with alternate hydrophobic and hydrophilic patches, the velocity variation across a hydrophilic patch shows a strong backow at the near-wall locations. Rudyak et al. [3] simulated microows in plane channels by the molecular dynamics (MD) method. The effects of uid density, accommodation coefcients, and channel di- mensions on ow properties are investigated. Sofos et al. [4] performed non-equilibrium molecular dynamics investigation of parameters affecting planar nanochannel ows. They found that uid velocity can be successfully tted by parabolas as system temperature, the magni- tude of the external force and channel width increase or when the average uid density decreases. Toghraie and Azimian [5] simulated Poiseuille ow in a narrow channel using the molecular dynamics sim- ulation (MDS). Their results show that velocity prole, slip length and slip velocity depends on the driving force. Liu et al. [6] studied electroos- motic ows in rough nanochannels by using of MDS. Their results show that roughness reduces the electroosmotic ow rate dramatically even though the roughness is very small compared to the channel width. Xu et al. [7] investigated the effect of wall roughness on uid transport resistance in nanopores. They showed that as the wavelength of roughness decreases or the magnitude of roughness increases, both the effective shear stress and the nominal viscosity increase. Kamali and Kharazmi [8] studied roughness effects on nanoscale ows by using of molecular dynamics simulation. They found that the wall- uid interaction and surface roughness are important and should be considered simultaneously in determining the nanostructures and pro- les of monatomic uid ow in a nanochannel. Barisik and Beskok [9] performed equilibrium molecular dynamics studies on nanoscale- conned uids. They concluded that the particleparticle virial increases with increased density, whereas the surfaceparticle virial develops because of the surface-force eld effects. Kannam et al. [10] Journal of Molecular Liquids 275 (2019) 192203 Corresponding author at: Department of Mechanical Engineering, Islamic Azad University, Khomeinishahr Branch, Khomeinishahr 84175-119, Iran. E-mail address: [email protected] (D. Toghraie). https://doi.org/10.1016/j.molliq.2018.11.057 0167-7322/© 2018 Elsevier B.V. All rights reserved. Contents lists available at ScienceDirect Journal of Molecular Liquids journal homepage: www.elsevier.com/locate/molliq

Journal of Molecular Liquidsresearch.iaun.ac.ir/pd/karimipour/pdfs/PaperM_6514.pdf · 2018. 12. 24. · performed equilibrium molecular dynamics studies on nanoscale-confined fluids

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Journal of Molecular Liquidsresearch.iaun.ac.ir/pd/karimipour/pdfs/PaperM_6514.pdf · 2018. 12. 24. · performed equilibrium molecular dynamics studies on nanoscale-confined fluids

Journal of Molecular Liquids 275 (2019) 192–203

Contents lists available at ScienceDirect

Journal of Molecular Liquids

j ourna l homepage: www.e lsev ie r .com/ locate /mol l iq

Molecular dynamics simulation of fluid flow passing through ananochannel: Effects of geometric shape of roughnesses

Pedram Alipour a, Davood Toghraie a,⁎, Arash Karimipour b, Mehdi Hajian c

a Department of Mechanical Engineering, Khomeinishahr Branch, Islamic Azad University, Khomeinishahr, Iranb Department of Mechanical Engineering, Najafabad Branch, Islamic Azad University, Najafabad, Iranc Department of Mechanical Engineering, Isfahan University of Technology, Isfahan 84156-83111, Iran

⁎ Corresponding author at: Department of MechaniUniversity, Khomeinishahr Branch, Khomeinishahr 84175

E-mail address: [email protected] (D. Toghraie).

https://doi.org/10.1016/j.molliq.2018.11.0570167-7322/© 2018 Elsevier B.V. All rights reserved.

a b s t r a c t

a r t i c l e i n f o

Article history:Received 23 June 2018Received in revised form 17 August 2018Accepted 14 November 2018Available online 20 November 2018

In this paper, we investigate the effects of geometric shape of various roughnesses on the fluid flow passingthrough a nanochannel by using of molecular dynamics simulation. The results of simulations are presentedfor the modeled structures (the five models defined) as number density, velocity, and system temperature pro-files for various conditions. By applying roughness to the inner surface of the ideal nano-channel at a thrust forceof 0.002 eV/Å, the amplitude of number density of the fluid particles near the walls decreased, while the meanand maximum velocities increased by 6.5% and 2.5% in the presence of square cuboid and hemispheroid rough-ness, respectively. Furthermore, the dimensionless slip velocity and slip length were, respectively, increased by amaximum of 41.1% and 21.5% in the presence of square cuboid roughness and by a minimum of 0.9% and 0.5% inthe presence of hemispheroid roughness. The temperature of the particles at the center of the nano-channel wasincreased by a maximum of 9.1% and a minimum of 2.8% in the presence of square cuboid and hemispheroidroughness, respectively. Calculation of the Argon-Argon radial distribution function indicated that themaximumof this function decreased by a maximum of 11.8% and a minimum of 8.5% in the presence of rectangular cuboidand ellipsoid roughness, respectively, compared to the ideal nano-channel.

© 2018 Elsevier B.V. All rights reserved.

Keywords:Geometric shapeNanochannelMolecular dynamics simulationRadial distribution functionRoughness

1. Introduction

The molecular dynamics method has been effective in treatingnanofluidic environments. Such atomistic simulations have been em-bedded in numerous studies of nanochannel flows over the last years.

Kim et al. [1] present a new interactive thermal wall model for sim-ulate the flow and heat transfer in nanoscale channels. They found thatwhen the interaction potential is weak or the wall has a large crystalbonding stiffness, temperature jumps on thewall/fluid interface are ob-served. Das et al. [2] studied the effect of interface wettability on thetransport of water (assuming it to be a simple Lennard Jones fluid) ina nanochannel by using of molecular dynamics simulation. Theyshowed that in channels with alternate hydrophobic and hydrophilicpatches, the velocity variation across a hydrophilic patch shows a strongbackflow at the near-wall locations. Rudyak et al. [3] simulatedmicroflows in plane channels by themolecular dynamics (MD)method.The effects offluid density, accommodation coefficients, and channel di-mensions on flow properties are investigated. Sofos et al. [4] performednon-equilibrium molecular dynamics investigation of parameters

cal Engineering, Islamic Azad-119, Iran.

affecting planar nanochannel flows. They found that fluid velocity canbe successfully fitted by parabolas as system temperature, the magni-tude of the external force and channel width increase or when theaverage fluid density decreases. Toghraie and Azimian [5] simulatedPoiseuille flow in a narrow channel using the molecular dynamics sim-ulation (MDS). Their results show that velocity profile, slip length andslip velocity depends on the driving force. Liu et al. [6] studied electroos-motic flows in rough nanochannels by using of MDS. Their results showthat roughness reduces the electroosmotic flow rate dramatically eventhough the roughness is very small compared to the channel width.Xu et al. [7] investigated the effect of wall roughness on fluid transportresistance in nanopores. They showed that as the wavelength ofroughness decreases or the magnitude of roughness increases, boththe effective shear stress and the nominal viscosity increase. Kamaliand Kharazmi [8] studied roughness effects on nanoscale flows byusing of molecular dynamics simulation. They found that the wall-fluid interaction and surface roughness are important and should beconsidered simultaneously in determining the nanostructures and pro-files of monatomic fluid flow in a nanochannel. Barisik and Beskok [9]performed equilibrium molecular dynamics studies on nanoscale-confined fluids. They concluded that the particle–particle virialincreases with increased density, whereas the surface– particle virialdevelops because of the surface-force field effects. Kannam et al. [10]

Page 2: Journal of Molecular Liquidsresearch.iaun.ac.ir/pd/karimipour/pdfs/PaperM_6514.pdf · 2018. 12. 24. · performed equilibrium molecular dynamics studies on nanoscale-confined fluids

193P. Alipour et al. / Journal of Molecular Liquids 275 (2019) 192–203

calculated the flow rate and the slip length from the streaming velocityprofiles obtained using non-equilibrium molecular dynamics. Theyfound that the slip length sensitive to the curvature of the velocityprofile and it possesses large statistical errors. Toghraie et al. [11]performed Molecular dynamic simulation of Copper and Platinumnanoparticles Poiseuille flow in a nanochannel. They found that the ag-glutination time in nanochannel with Copper particles is faster that inPlatinum nanoparticles.

In this paper, we investigate the effects of geometric shape of variousroughnesses on the fluid flow passing through a nanochannel by usingof molecular dynamics simulation.

Investigating the effects of geometric shapes of internal wall rough-ness on the flow of Argon fluid in a Platinum nano channel, one of theinnovations of this research, which may be said to have not beenreported in any of the previous work. This is one of the most importantissues that the previous researchers have simply eliminated. The resultsof simulations are presented for the modeled structures (the fivemodels defined) as number density, velocity, and system temperatureprofiles for various conditions.

2. Theoretical analysis

2.1. Equations of motion and phase-space trajectory

For N particles (N is of the order of 100), a molecular dynamics sim-ulator generates 6N values for the coordinates and the momenta of theparticles; in other words, it is safe to say that it has 6N dimensions.Sometimes, the simulator advances thousands or millions of timesteps, thus facing a plethora of numbers. Therefore, if computer simula-tion is used as the numerical equivalent of experiments, onemust knowwhat type of averages are to be calculated. One way of estimating thesenumbers is phase space, which will be discussed in the following. Gibbs[12] was the first to propose using ensemble averages instead of timeaverages due to the complexity of time transformations. An ensembleis defined as a group of particular points in the phase space, this way,statistical mechanics relates microscopic to macroscopic properties. InNewtonian dynamics, the translation of a spherical particle (molecule,atom, etc.) is a result of the force F

!i which is exerted on the particle

by an external factor. Newton's second law establishes a direct relation-ship (Eq. (1)) between the motion and the applied force,

F!

i ¼ m €r!i ð1Þ

where m is the mass of the particle. Furthermore, the acceleration iscalculated using Eq. (2),

€r!i ¼d2 r!i

dt2ð2Þ

r!i is the vector specifying the particle's location with respect to a fixedreference framework. For N spherical particles, Newton's second law(Eq. (1)) transforms into 3N second-order ordinary differential equation.Themain goal inmolecular dynamics simulation is to specify the classicaltrajectory of themolecules. In a systemofNmolecules, everymolecule in-teracts with one another. Therefore, its coordinates change based onNewton's second law. As a single molecule follows along a trajectory, itsposition is then a function of time and can be denoted by ri(t). pi(t). Ifthe instantaneous coordinates and the linear momenta of N moleculesare plotted in a 6N-dimensional superspace, the space may be dividedinto two parts:

1. The 3N-dimensional configuration space ri(t) in which the coordinateaxes are the components of the position vectors of the molecules.

2. The 3N-dimensional linear momentum space pi(t) in which thecoordinate axes are the components of the linear momenta of themolecules.

From the perspective of theorists, the importance of moleculardynamics is that it provides accurate semi-experimental results for afully-specified model [13]. In general, there are four common, constantstatistical ensembles, namely the canonical ensemble (NVT), themicrocanonical ensemble (NVE), the isothermal, isobaric ensemble(NpT), and the grand canonical ensemble (μVT). The aforesaid thermo-dynamic variables are constant for each ensemble and the othervariablesmust be calculated by averaging. Obtaining a physical propertyof each of these ensembleswith the help ofmolecular dynamics simula-tion involves the following steps:

1. Specifying the phase-space trajectory ð r!N; p!NÞ

2. Calculating the function Að r!N; p!NÞ for every property

3. Averaging over an interval (Eq. (4)).4. Assuming ⟨A⟩time = Am

Physical properties are functions of the positions and momenta ofevery particle (N particles), therefore, the instantaneous magnitude ofa general physical quantity (A) at the moment t may be defined asEq. (3),

A tð Þ ¼ f r!1 tð Þ; r!2 tð Þ;…; r!N tð Þ; p!1 tð Þ; p!2 tð Þ;…; p!N tð Þ� �

¼ A r!Ntð Þ; p!N

tð Þ� �

ð3Þ

Moreover, the average may be calculated using Eq. (4),

Ah itime ¼1M

XMi¼1

Ai tð Þ ð4Þ

where M is the number of time steps and the subscript i denotes thesimulation time steps. However, in practice, the time average may becalculated in two ways:

1. During simulation, the software calculates A(t) in every step, addsit to the sum

Xi

AiðtÞ, and divides the final average by the number

of stages.2. The position andmomentumof the particles are saved in a file at reg-

ular intervals during simulation. Then, at the end of the simulation, aprogram is developed to analyze the file containing phase spacetrajectories and, ultimately, to calculate the needed variables.

In molecular dynamics simulation, each particle interacts withseveral other particles and the particles' trajectories are not straightlines between collisions. Therefore, due to this complex nature, whichmakes the potential function more complex, there is no analytical solu-tion for the equations of motion and they must be solved numerically.The Finite Difference Method (FDM) is to be used to numerically inte-

grate the Newtonian function of motion ð F!ðtÞ

m ¼ d2 r!ðtÞdt Þ. The FDM is

based on solving equations by approximation with the help of Taylor'sseries. Therefore, Newton's equations of motion are solved simulta-neously and step-by-step in short time steps and stored in the trajectoryfile (which shows the particles' trajectories at regular intervals through-out the simulation). Then, at the end of the simulation, a program isdeveloped to analyze the file containing the trajectories and, ultimately,to calculate the needed variables. To be able to integrate the second-degree motion function, two initial conditions are required foreach particle. The two initial conditions are the initial locations andvelocities of the particles. However, it must be noted that when themicrocanonical ensemble (NVE) is used, the particles' initial velocitiesmay be assumed zero since the algorithm corrects initial velocitiesvery quickly [14]. The molecular dynamics simulation method is oftenapplicable to isolated systems containing afixed number (N) of particlesin afixed volume (V) andwithfixed total energy (the sumof kinetic and

Page 3: Journal of Molecular Liquidsresearch.iaun.ac.ir/pd/karimipour/pdfs/PaperM_6514.pdf · 2018. 12. 24. · performed equilibrium molecular dynamics studies on nanoscale-confined fluids

Fig. 1. The Poiseuille flow of the fluid (Ar) between solid Pt plates.

194 P. Alipour et al. / Journal of Molecular Liquids 275 (2019) 192–203

potential energies of all particles, E). Macroscopic properties arealways equivalent to the overall average of a statistical ensemble andpresent a mean property for molecular systems (equilibrium or non-equilibrium),

Aobs ¼ Ah iens ð5Þ

where A is any measurable property. Two important conclusions shallbe made from this in molecular modeling:

1. The required information about an isolated system is not sufficient(even if it has theminimumpotential energy). Therefore, to calculatethe macroscopic properties, it is necessary to generate a set of speci-mens at a certain temperature;

2. Atomic details, such as the position, momentum, etc. are delineatedin molecular simulations. However, given that such details belongto macroscopic properties, another approach is required to explainthe interactions and to average the properties more simply.

In molecular dynamics simulation, the particles' positions ( r!N) are

found by solving the equations of motion as in Eq. (6),

F!

i tð Þ ¼ −∂U r!N� �∂ r!i

¼ m €r!i tð Þ ¼Xi

Xi≠ j

F!

ij þXi

Xi≠ j

F!

ij;w þ F!

ext ð6Þ

where F!

iðtÞ is the total force exerted on the particle ith by the other

N − 1 particles (Xi

Xi≠ j

F!

ij), from the solid walls (Xi

Xi≠ j

F!

ij;w) and

also the externally applied forces present in the problem ( F!

ext). F!

ext

denotes the external force resulting from an external field that is ap-plied to the system to create the required non-equilibrium conditionsso the positions of the particles are determined in non-equilibriummolecular dynamics simulations based on the system's response tothis force. Every thermodynamic property that can be calculated, such

as Að r!NðtÞ; p!NðtÞÞ, is a function of the locations and momenta of all

constituent particles of the system. ð r!NðtÞ; p!NðtÞÞ, and consequently A

ð r!NðtÞ; p!NðtÞÞ, change over time. It is obvious that the experimental

macroscopic property A, is actually the time average of Að r!NðtÞ; p!NðtÞÞin a long time interval [15,16],

Aobs ¼ Ah itime ¼ Að r!Ntð Þ; p!N

tð ÞÞD E

time¼ lim

t→∞

1t

Z t0þt

t0A r!N

τð Þ; p!Nτð Þ

� �dτ ð7Þ

In Eq. (7),Aobs is the value of themeasured property,Að r!NðτÞ; p!NðτÞÞis a function of location anddisplacement of all constituent particles of thesystem, t is the total time considered for computing A and ⟨A⟩time by thetime-averaged A. The relations determining the time transformation areNewton's equations of motion that are ordinary differential equationsand can be solved for approximately 103 particles using a computer. Todo so, the user will need to solve the equations of motion step-by-step,showing the importance and the role of molecular dynamics simulation.With the help of Eq. (7), the kinetic energy and temperature may be ob-tained using Eq. (8) based on the phase trajectory [16],

Ekh i ¼ limt→∞

1t

Z t0þt

t0E p!N� �

dt ¼ 32NkBT ð8Þ

where T is the system temperature, kB is the Boltzmann's constant, N is

the number of molecules in the system, p!Nis the momenta of the

constituent particles of the system, and ⟨Ek⟩ is the average kinetic energy.To calculate pressure, the effects of the two phrases are taken into ac-count. Pm is related to the momentum and is transported by the atomspassing a hypothetical plane in the time interval dt, the other one is Pf

which results from the momentum of the forces between the atoms onthe two sides of the plane [16],

P ¼ Pm þ P f ¼2N3V

Ekh i þ 13V

XNi¼1

XNib j

Fij�!

:rij!

* +→ PV ¼ NkBT þ Wh i ð9Þ

where V is the volume of the simulation cell and F!

ij is the force

exerted on the atom i by the atom j, and r!ij shows the distancebetween the two atoms. Virial is defined as the expected value of thesum of the product of the particles' coordinates and the force exerted

on them ½W ¼ − 13 hPN

i¼1PN

ib j Fij�!

:rij!i�. Ultimately, the total energy is

obtained from Eq. (10) [16],

E ¼ Ek p!N� �

þXNi¼1

XNib j

U r!ij

� �ð10Þ

where Uð r!ijÞ is the potential energy between the atoms i and j.

2.2. Simulating the Poiseuille flow in a nanochannel

Fig. 1 shows a nanochannel limited by parallel Platinum plates con-taining an argon fluid.

In this study, the channel was assumed to be 20 Å long (L), 7.5 Åhigh (H), and 10 Å wide (W). The wall thickness was also assumed1 Å (except for the top wall which is assumed to be 1.5 Å thick),and the hydraulic diameter (DH = 4 RH = 2 W) 20 Å. Where A = WHis the wetted area and P = 2(H + W) is the wetted perimeter of thenanochannel cross-section (H ≫ W). Each wall is composed ofthree layers of Pt atoms that are simulated in a Face-CenteredCubic (FCC) lattice with a constant of 3.92 Å. Moreover, at the startof the simulation, the atoms of the fluid phase (Ar) are assumed tobe in a Simple Cubic (SC) configuration with a lattice constant of3.4 Å. Periodic boundary conditions were assumed in the flow direction(x and z axes) to eliminate surface effects and to maintain a constantnumber of particles in the system by having a particle exit the systemfrom one side for every one that enters the system from the otherside. A set of general rules were assumed for modeling the structures,considering the following items in all models (different nanochannelexamples):

a) The Embedded AtomMethod (EAM) potential was used for investi-gating and simulating the interactions between Pt atoms in thenanochannel.

b) The Lennard-Jones (LJ) potential was used for investigating andsimulating the interactions between Ar atoms within the fluid.

Page 4: Journal of Molecular Liquidsresearch.iaun.ac.ir/pd/karimipour/pdfs/PaperM_6514.pdf · 2018. 12. 24. · performed equilibrium molecular dynamics studies on nanoscale-confined fluids

195P. Alipour et al. / Journal of Molecular Liquids 275 (2019) 192–203

c) The Lennard-Jones potential was used for investigating and sim-ulating the interactions between Pt and Ar atoms and describesthe wall-fluid interactions at the nanochannel-fluid interface.

d) The nanochannel walls were assumed to be rigid and fixed.e) In all investigations, the fluid particles were assumed tomove along

the x-axis.

The Lennard-Jones potential 6–12 is presented in Eq. (11),

U rij� � ¼ 4ε

σrij

� �12

−σrij

� �6" #

rij≤rc ð11Þ

where σ is the particle's diameter, or the length scale of the problem,and the distance at which the potential is zero. ε shows the intensityof the interactions between the particles and shares the same dimen-sion with energy. This parameter shows the magnitude of the interac-tions and the depth of the potential well. The parameters σ and εdepend on the type of themolecules of the fluid and thewall. The cutoffradius (rc) is the distance beyond which the interactions betweenparticles are ignored (for rij N rc, we have U(rij) = 0). In other words,beyond the cutoff distance, the particle pairs do not see each other andonly particles within a sphere of this radius are able to see each other.σ and ε are often specified by experimental fitting or through accuratequantum chemistry calculations. rc is different when distinguishingbetween vapor-vapor, vapor-liquid, liquid-liquid, vapor-solid, andliquid-solid particle pairs. The Lennard-Jones potential was also used forAr atoms and the constants σAr−Ar, εAr−Ar, and rij were assumed 3.4 Å,0.0104 Å, and 3.4 Å, respectively in keeping with previous studies [17].The interactions between the fluid and solid-surface particles play animportant role in determining near-surface properties such as slipvelocity, the shear stress, etc. Whenever the Lennard-Jones potential isused for fluid-fluid interactions, it is best to use the same potential,although with some modifications, for fluid-solid interactions too. Theextended Lennard-Jones potential is used to model this interaction byreplacing σ and εwith σsf and εsf. The overall form of this potential func-tion is presented in Eq. (12) [18],

Uwall rð Þ ¼ 4αεsfσ sf

r

� �12

−βσ sf

r

� �6" #( )

ð12Þ

where r is the distance between the fluid particles and the solid wall.α is a potential coefficient showing how hydrophilic the surface is,while, β is another potential coefficient showing the how hydrophobicthe surface is. Different laws were used to calculate σsf and εsf [19,20],including the Waldman-Hagler combining rule [21] which is stated inEq. (13) [18],

εsf ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiεsσ s

6ε fσ f6

qσ sf

6 σ sf ¼σ s

6 þ σ f6

2

!16

ð13Þ

Furthermore, the interactions between Ar and Pt atoms are definedby the Lennard-Jones potential with the constants σAr−Pt, εAr−Pt, and rijfound to be 3.085 Å, 0.0056 Å, and 3.4 Å, respectively according to theWaldman-Hagler law of mixture. The EAM potential calculates binaryinteractions between metals and metallic alloys and is expressed asEq. (14),

Ei ¼ FαXj≠i

ρβ rij� �2

435þ 1

2

Xj≠i

ϕαβ rij� � ð14Þ

where the latent energy F is a function of the electronic density of theatoms ρ. ϕ is the binary interaction potential and αβ are the differenttypes of i and j atoms. The multi-particle property of the EAM potentialis due to the latent energy term. Both summations in Eq. (12) are ap-plied to all the neighbors of i and jwithin the cutoff distance. The cutoff

distance is a function ofϕ and ρ. The EAMpotential was used formodel-ing the potential between the Pt atoms in the wall. The constants of thepotential function were defined using the pair_coeff command and thePt_u3.eam file. In all simulation cases in this study, to simulate thenanochannel and fluid system, the Pt nanochannel and the said latticewere initially modeled with an FCC lattice and four walls (on top, bot-tom, left and right) in the simulation toolbox. Moreover, the fluidphase (Ar) was assumed to have an SC configuration initially. However,as expected, after a few time steps pass the structure loses its initialorder due to the Lennard-Jones potential between Ar atoms and assumea stochastic configuration. On the other hand, the nanochannel main-tains its initial configuration, the FCC lattice, whichwas also anticipated.This proves that a correct potential function was selected for thenanochannel, in other words, the EAM potential function is a goodchoice for this nanochannel. Combination models in which pairs of cer-tain atoms interact with different binary potentials are also formulatedusing the hybridmethod. The atomicmasses of Pt and Ar were assumed195.08 and 39.95 g/mol and are consistent with previously reportedresults [17]. Themolecular dynamics simulation programstarts by orga-nizing the particles in the simulation domain and assigning every parti-cle in the system a location and an initial velocity. After calculating theintermolecular forces based on the potential function and applying thesimulation conditions, the equations of motion were integrated andnew locations and velocities stored in every step. This loop continuesuntil the system reaches equilibriumwhen the average thermodynamicvariables and the simulation outputs are investigated, ultimately. Inother words, the molecular dynamics simulation starts by integratingNewton's second law; then, the dynamic parameters such as thepositions and velocities of the particles are determined by properintegration methods such as the Verlet method, predicting the systemtrajectory in the phase space. Therefore, the overall trajectory of theparticles in the simulated structure within the phase space can be plot-ted using these outputs. Furthermore, all thermodynamic properties ofthe simulated system can be calculated using the trajectory and themacroscopic physical properties, such as the pressure, average velocity,temperature, the number density of particles, etc. are determined bystatistical mechanics analysis. The initial velocities of the particles aredefined randomly. The initial fluid region velocity is generated at atemperature that is calculated by the Compute command. Moreover,the initial velocitymust be specified by selecting the initial temperaturewhich was assumed to be the room temperature (300 K). Themicrocanonical ensemble, NVE is used to maintain a fixed number ofparticles, volume, and energy in the simulated system. In this model,the thermostat Temp/rescale is applied to each row of atoms simulatingthe wall and the vibrating wall atoms are connected by springs. Thisway, the temperature jumps due to the accumulation of error areprevented during simulation. The combination of the vibrations andthe thermostats generates the required temperature (200K) accurately.The applied thermostat works by velocity scaling. In this method, thetemperature is fixed at the required level by limiting the kinetic energyof the system. After specifying the initial positions of the atoms, thevelocities of the different parts of the atoms must be also specified tofully define the initial conditions of the system. Given the rigidity ofthe simulated nanochannel, the Pt atoms on the top, bottom, left, andright walls were assumed to be at zero velocity with the external forceadjusted along the positive direction of the x axis using the Fix addforcecommand. Ultimately, by applying the said conditions, which is toassign the Pt and Ar atoms their initial positions and velocities, the sys-tem becomes fully defined, enabling us to investigate the output results.In the next step, to print and display the thermodynamic outputs indifferent parts of the nanochannel the atoms in the simulation wereshelved along the width (y) dividing the channel into 150 bins. Theseshelves are used to report the thermodynamic properties in differentparts of the channel and the situation of Ar atoms and to comparethem. Depending on the lattice structure and the density, and in idealconditions (with no roughness), the fluid contains 2016 particles and

Page 5: Journal of Molecular Liquidsresearch.iaun.ac.ir/pd/karimipour/pdfs/PaperM_6514.pdf · 2018. 12. 24. · performed equilibrium molecular dynamics studies on nanoscale-confined fluids

196 P. Alipour et al. / Journal of Molecular Liquids 275 (2019) 192–203

the walls are composed of 4920 particles in total. In the case of squareroughness, the fluid contains 1752 particles, while the walls and theroughness are composed of 4920 and 62 particles in total, respectively.In the case of rectangular roughness, the fluid contains 1812 particles,while the walls and the roughness are composed of 4920 and 38particles in total, respectively. In the case of ellipsoid roughness,the fluid contains 1872 articles, while the walls and the roughnessare composed of 4920 and 24 particles in total, respectively. In thecase of hemispherical roughness, the fluid contains 1872 articles,while the walls and the roughness are composed of 4920 and 35particles in total, respectively.

Fig. 2 shows a schematic of a simulated model for the flow of argonat various Platinum nanoscale sections.

Fig. 2. A schematic of a simulated model for the flow

3. Results and discussion

Ultimately, the thermodynamic outputs are printed at every 100time-step using the Thermo command and plotted in the diagrams,assuming 0.001 ps time steps to match the time interval defined inLAMMPS. The final command in the input code was the Run commandto start the simulation. To ensure the stability of the system, 1,000,000time-steps were used in this simulation. It must be noted that thesystem was optimized in terms of the geometry and energy beforedisplaying the outputs to minimize the error of the thermodynamicresults with respect to the experimental and anticipated results. Afteroptimizing the simulated fluid and Pt nanochannel systems in termsof geometry and energy, the systemwas equalized at room temperature

of argon at various Platinum nanoscale sections.

Page 6: Journal of Molecular Liquidsresearch.iaun.ac.ir/pd/karimipour/pdfs/PaperM_6514.pdf · 2018. 12. 24. · performed equilibrium molecular dynamics studies on nanoscale-confined fluids

Fig. 5. The number density profile of Ar in a Platinum nanochannel with rectangularroughness.

Fig. 3. The number density profile of Ar in an ideal Platinum nanochannel.

197P. Alipour et al. / Journal of Molecular Liquids 275 (2019) 192–203

(300 K) and the thermodynamic outputs were generated, in the end. Inotherwords, in anymodel, the system reaches equilibriumafter a certainnumber of time-steps. Then, after equilibrium, the non-equilibriummo-lecular dynamics simulation continues until the 1,000,000th time-step,modeling and generating the density, velocity, and temperature profilesof the system under various conditions. Generally, the procedure forgenerating thermodynamic outputs starts with the organization of theparticles in different parts of the nanochannel and the fluid accordingto the said lattices at the start of the simulation. Then, the equations ofmotion are integrated using the velocity Verlet algorithm. At this stage,it is necessary to calculate inter-atomic forces to obtain the particles'positions and velocities in every time-step. The forces exerted on theatoms in the fluid and the rigid Pt walls were calculated separatelyusing the fluid-fluid, wall-wall, fluid-wall interaction potentials. In thefollowing, the results of molecular dynamics simulations are presentedfor the modeled structures (the fivemodels defined) as number density,velocity, and system temperature profiles for various conditions. Fluiddensity in the nanochannel is an important variable that can help

Fig. 4. The number density profile of Ar in a Platinumnanochannelwith square roughness.

evaluate the performance of the nanochannel in conducting the Arfluid affected by the roughness. Itmust be noted that the number densityof the fluid is stated as the number of atoms in each shelf in a specifictime-step. The number of atoms is averaged over the total simulationtime resulting in a time-averaged value for each section. Fig. 3 showsthe density profile along the width of an ideal nanochannel at the1,000,000th time-step, for an external force of 0.002 eV/Å. Consideringthat the open-source software, LAMMPS was used in this study for sim-ulation, the results were compared with the previous studies to evaluatethe accuracy of the numerical code. As evident from Fig. 3, the structureof the fluid molecules between the walls is constantly changing and thefluctuations are much larger in the vicinity of the walls compared withthe middle of the nanochannel. This was also reported in moleculardynamics simulation of the of Ar atoms in nanochannels [22,23]. Thenumber density of the fluctuations at the center of the nanochannel isaround 0.018 cm−3, reaching approximately 0.045 cm−3 near thewalls. This may be attributed to the confrontation between the fluid

Fig. 6. The number density profile of Ar in a Platinum nanochannel with ellipsoidroughness.

Page 7: Journal of Molecular Liquidsresearch.iaun.ac.ir/pd/karimipour/pdfs/PaperM_6514.pdf · 2018. 12. 24. · performed equilibrium molecular dynamics studies on nanoscale-confined fluids

Fig. 9. The velocity profile of Ar in a Platinum nanochannel with square roughness.Fig. 7. The number density profile of Ar in a Platinum nanochannel with hemisphericalroughness.

198 P. Alipour et al. / Journal of Molecular Liquids 275 (2019) 192–203

particles and the wall. In the vicinity of the walls, the fluid particles havemore interaction with the walls and the force exerted by the wall affectsthedistribution of the fluid properties. The impact of the force exerted bythe wall on the distribution of the fluid properties is smaller in thefurther parts of the nanochannel walls being minimized at the centerof the nanochannel.

Figs. 4, 5, 6, and 7 show the density profile along the width of thenanochannels with square, rectangular, ellipsoid, and hemisphericalroughness, respectively, at the 1,000,000th time-step for an externalforce of 0.002 eV/Å. As evident from the figures, the structure of thefluid molecules between the walls is constantly changing and thefluctuations are much larger in the vicinity of the walls compared withthe middle of the nanochannel. The number density of the fluctuationsat the center of the nanochannels is around 0.015 cm−3, while reachingto approximately 0.040 cm−3 near thewalls, which is smaller comparedwith the number densities at the center and near the walls of thenanochannel in the case of smooth walls. This seems plausible

Fig. 8. The velocity profile of Ar in an ideal Platinum nanochannel.

considering that the number density of the fluid is stated as the numberof atoms in each shelf in a specific time-step. This was also reported inmolecular dynamics simulation of the Poiseuille flow of Ar atoms inrough nanochannels [23–25].

To evaluate the behavior of the particles in the nanochannel at thedifferent sections of the nanochannel, the velocity profile was obtainedin the simulated nanochannels using the shelving process. In thisregard, the nanochannel was shelved along its width (y axis). To obtainthe velocity profiles, the average velocity was calculated in everybin and in every time-step, and the results were all averaged. Theseprofiles are obtained by statistical averaging of the velocities of everysingle particle. Fig. 8 shows the velocity profile along the width ofan ideal nanochannel at the 1,000,000th time-step, for an externalforce of 0.002 eV/Å. As illustrated in Fig. 8, the Poiseuille flow has a par-abolic velocity profile. This was also reported in molecular dynamicssimulation of the Poiseuille flow of Ar atoms in smooth nanochannels[17,22,26].

Fig. 10. The velocity profile of Ar in a Platinum nanochannel with rectangular roughness.

Page 8: Journal of Molecular Liquidsresearch.iaun.ac.ir/pd/karimipour/pdfs/PaperM_6514.pdf · 2018. 12. 24. · performed equilibrium molecular dynamics studies on nanoscale-confined fluids

Fig. 13. Slip velocity and length. [22].

Fig. 11. The velocity profile of Ar in a Platinum nanochannel with ellipsoid roughness.

199P. Alipour et al. / Journal of Molecular Liquids 275 (2019) 192–203

Figs. 9, 10, 11, and 12 show the velocity profile along thewidth of thenanochannels with square, rectangular, ellipsoid, and hemisphericalroughness, respectively, at the 1,000,000th time-step for an externalforce of 0.002 eV/Å. As evident from the figures, all velocity profilesare parabolic. Moreover, it is obvious that the velocity is not zero onthe walls. This shows that the fluid behaves differently at nanoscalecompared with the macroscale and a velocity jump occurs, makingNavier-Stokes equations with the no-slip condition inapplicable. Thevelocity profile near the walls is affected by the force exerted by themand the fluid velocity tends to that of the wall (vwall = 0). In otherwords, the wall's force field attracts the fluid particles and keeps themnear the wall for a longer time, so the fluid velocity approaches that ofthe wall. This was also reported in molecular dynamics simulation ofthe Poiseuille flow of Ar atoms in rough nanochannels [26,27].

As illustrated in Fig. 13, fluid in this study, the slip velocity (vs) isdefined as the difference between the wall velocity and the velocity ofits adjacent.

Fig. 12. The velocity profile of Ar in a Platinum nanochannel with hemisphericalroughness.

The slip length is defined based on Eq. (15) [22]:

γ ¼ vsls→ls ¼ vs

γð15Þ

where γ is the shear velocity. Furthermore, Eq. (16) is used to calculatethe dimensionless slip velocity [22],

vsUc

¼ 1þ Dz

4ls

� �−1

ð16Þ

whereUc is thefluid velocity at the center of the channel.Dz is illustratedin Fig.14. σl represents the particle's diameter, which is the length scalein this problem.

Figs. 15 and 16 show the slip length and the dimensionless velocityin the ideal nanochannel and the nanochannels with square, rectangu-lar, ellipsoid, and hemispherical roughness at the 1,000,000th time-stepfor various external forces (0.0014, 0.0017, 0.0018, and 0.002 eV/Å).Evidently, a larger driving force increases the slip length and velocityand the variations of the slip length and velocity can be representedby a quadratic function. This was also reported in molecular dynamicssimulation of the Poiseuille flow of Ar atoms in nanochannels [17,22].The slip velocity is proportional to the transverse velocity gradient

ð∂Vx∂y jwallÞ and the Knudsen number. The slip length is proportional to the

shear stress. The fluid slips over the hydrophobic surfaces, leading to adiscontinuity between the fluid and wall velocities.

The shelving process was used to obtain the temperature profile inthe simulated nanochannels to evaluate the behavior of the particlesin the nanochannel at the different sections of the nanochannel. In

Fig. 14. Calculating the dimensionless slip velocity [22].

Page 9: Journal of Molecular Liquidsresearch.iaun.ac.ir/pd/karimipour/pdfs/PaperM_6514.pdf · 2018. 12. 24. · performed equilibrium molecular dynamics studies on nanoscale-confined fluids

Fig. 17. The temperature profile of Ar in an ideal Platinum nanochannel.Fig. 15. The slip lengths in Poiseuille flow for different extreme driving forces.

200 P. Alipour et al. / Journal of Molecular Liquids 275 (2019) 192–203

this regard, the nanochannel was shelved along its width (y axis). Toobtain the temperature profiles, the average temperature is calculatedin every bin and in every time-step, and the results are all averaged.These profiles are obtained by statistical averaging of the temperaturesof every single particle. Fig. 17 shows the temperature profile along thewidth of an ideal nanochannel at the 1,000,000th time-step, for an ex-ternal force of 0.002 eV/Å. As evident from Fig. 17, the Poiseuille flowhas a quadratic temperature profile. This was also reported inmoleculardynamics simulation of the Poiseuille flow of Ar atoms in smoothnanochannels [28].

Figs. 18, 19, 20, and 21 show the fluid temperature profile along thewidth of the nanochannels with square, rectangular, ellipsoid, andhemispherical roughness, respectively, at the 1,000,000th time-stepfor an external force of 0.002 eV/Å. As evident from the figures, alltemperature profiles are parabolic. This was also reported in moleculardynamics simulation of the Poiseuille flow of Ar atoms in roughnanochannels [26].

Fig. 16. The slip velocities in Poiseuille flow for various driving forces.

The radial distribution function is one of the most important factorsin identifying the phases of the studied materials. The function showstheprobability offinding a central atomat a certain distance fromanad-jacent atom comparedwith the ideal gas distribution and is abbreviatedg(r), where r represents the atomic spacing. In other words, the radialdistribution function shows the local density of the atoms located atthe distance r from the central atom in equilibrium. The radial distribu-tion function depends on the density and temperature. By knowing theradial distribution and assuming binary interactions between particles,thermodynamic properties such as pressure, internal energy, andchemical potential can be calculated. On theother hand, the informationprovided by the radial distribution function makes it possible to inte-grate the direct correlation function, which has an essential role in thefluctuation theory. In addition, the radial distribution function is thebest characteristic of the structural nature of materials, especially fluids,and a really good guide to identifying phases inmaterials and simulatedsystems. The radial distribution function shows the arrangementof atoms around one atom. Specifically, the binary distribution function

Fig. 18. The temperature profile of Ar in a Platinum nanochannel with square roughness.

Page 10: Journal of Molecular Liquidsresearch.iaun.ac.ir/pd/karimipour/pdfs/PaperM_6514.pdf · 2018. 12. 24. · performed equilibrium molecular dynamics studies on nanoscale-confined fluids

Fig. 21. The temperature profile of Ar in a Platinum nanochannel with hemisphericalroughness.

Fig. 19. The temperature profile of Ar in a Platinum nanochannel with rectangularroughness.

201P. Alipour et al. / Journal of Molecular Liquids 275 (2019) 192–203

g(r) states the probability of finding an atom at a distance r fromanother atom compared with the case of an ideal gas distribution.Therefore, g(r) is dimensionless. In this paper, radial distribution functionrefers to the same binary distribution function [29]. The locations ofsingle atoms are considered functions of time in molecular dynamicscalculations; therefore, the radial distribution function can be easilyobtained. In this regard, the atoms adjacent to each atom, are recordedin a histogram at certain time intervals and the corresponding radialdistribution function is calculated using Eq. (17) [29].

g rð Þ ¼ N r;Δrð Þh i1

2

� �NρV r;Δrð Þ ð17Þ

whereN(r,Δr) is the number of atoms in a spherical shell of radius r andthicknessΔr, ρ is the number density (ρ ¼ N

V),N is the total number of

atoms, and V(r,Δr) is the volume of the spherical shell. Fig. 22 illustrates

Fig. 20. The temperature profile of Ar in a Platinumnanochannel with ellipsoid roughness.

the radial distribution function of the system in an ideal nanochannelcontaining anAr fluid. An investigation of the Ar atomdistribution func-tion in the ideal Platinum nanochannel based on the distance suggests amaximum at 3.3337 Å where g(r) = 6.0441. This shows the configura-tion of atoms with respect to one another andmeans the molecules are6.0441 times more likely to locate at this distance from each othercompared to the case of an ideal gas. It must be noted that the diagramcan be plotted after enough time steps to describe the final, stableconfiguration offluid atoms. At small distances (smaller than the atomicdiameter), g(r) is zero due to strong repulsion. After 3.3337 Å, the radialdistribution function drops, reaching aminimumat 4.6269Å. Comparedto the case of an ideal gas, it is less likely to find two molecules at thisdistance. At large distances, g(r) tends to the ideal gas value, suggestingthe absence of long-range forces at such distances. As illustrated inFig. 22, considering what was said about calculating the radial distribu-tion function, the distance increases with the spherical shell radius,

Fig. 22. The Ar-Ar radial distribution in the smooth nanochannel containing fluid.

Page 11: Journal of Molecular Liquidsresearch.iaun.ac.ir/pd/karimipour/pdfs/PaperM_6514.pdf · 2018. 12. 24. · performed equilibrium molecular dynamics studies on nanoscale-confined fluids

Fig. 25. The Ar-Ar radial distribution function of the system in the nanochannel withellipsoid roughness containing fluid.

Fig. 23. The Ar-Ar radial distribution function of the system in the nanochannel withsquare roughness containing fluid.

202 P. Alipour et al. / Journal of Molecular Liquids 275 (2019) 192–203

therefore, the number of atoms inside the shell divided by the shellvolume equals material density and g(r) converges to 1.

Fig. 23 shows the radial distribution function of the system in thenanochannel with square roughness containing an Ar fluid. With amaximumat 3.3337Åwhere g(r)=5.10549, the radial distribution func-tion, in this case, is similar to that of the ideal nanochannel containing Ar.This shows the configuration of atoms with respect to one another andmeans the molecules are 5.10549 times more likely to locate at thisdistance from each other compared to the case of an ideal gas. At smalldistances (smaller than the atomic diameter), g(r) is zero due to strongrepulsion. After 3.3337 Å, the radial distribution function drops, reachinga minimum at 4.7011 Å. Compared to the case of an ideal gas, it is lesslikely to find two molecules at this distance. As illustrated in Fig. 23, atlarge distances, g(r) tends to the ideal gas value, converging to 1. Thisshows the absence of long-range forces at such distances.

Fig. 24. The Ar-Ar radial distribution function of the system in the nanochannel withrectangular roughness containing fluid.

Fig. 24 shows the radial distribution function of the system in thenanochannel with rectangular roughness containing an Ar fluid. Witha maximum at 3.4397 Å where g(r) = 5.33143, the radial distributionfunction, in this case, is only slightly different from that of the idealnanochannel and the nanochannel with square roughness containingAr. This shows the configuration of atoms with respect to one anotherand means the molecules are 5.33143 times more likely to locate atthis distance from each other compared to the case of an ideal gas. Atsmall distances (smaller than the atomic diameter), g(r) is zero due tostrong repulsion. After 3.4397 Å, the radial distribution function drops,reaching a minimum at 4.7753 Å. Compared to the case of an idealgas, it is less likely to find two molecules at this distance. As illustratedin Fig. 24, at large distances, g(r) tends to the ideal gas value, convergingto 1. This shows the absence of long-range forces at such distances.

Fig. 26. The Ar-Ar radial distribution function of the system in the nanochannelhemispherical roughness containing fluid.

Page 12: Journal of Molecular Liquidsresearch.iaun.ac.ir/pd/karimipour/pdfs/PaperM_6514.pdf · 2018. 12. 24. · performed equilibrium molecular dynamics studies on nanoscale-confined fluids

203P. Alipour et al. / Journal of Molecular Liquids 275 (2019) 192–203

Fig. 25 shows the radial distribution function of the system in thenanochannel with ellipsoid roughness containing an Ar fluid. In thiscase, the radial distribution function has a maximum at 3.3973 Åwhere g(r) = 5.52784. This shows the configuration of atoms with re-spect to one another and means the molecules are 5.52784 timesmore likely to locate at this distance from each other compared tothe case of an ideal gas. At small distances (smaller than the atomicdiameter), g(r) is zero due to strong repulsion. After 3.3973 Å, the radialdistribution function drops, reaching aminimumat 4.7435 Å. Comparedto the case of an ideal gas, it is less likely to find two molecules at thisdistance. As illustrated in Fig. 25, at large distances, g(r) tends to theideal gas value, converging to 1. This shows the absence of long-rangeforces at such distances.

Fig. 26 shows the radial distribution function of the system in thenanochannel with hemispherical roughness containing an Ar fluid. Witha maximum at 3.3761 Å where g(r) = 5.47964, the radial distributionfunction, in this case, is similar to that of the ideal nanochannel containingAr. This shows the configuration of atomswith respect to one another andmeans the molecules are 5.47964 times more likely to locate at this dis-tance from each other compared to the case of an ideal gas. At small dis-tances (smaller than the atomic diameter), g(r) is zero due to strongrepulsion. After 3.3761 Å, the radial distribution function drops, reachinga minimum at 4.8177 Å. Compared to the case of an ideal gas, it is lesslikely to find two molecules at this distance. As illustrated in Fig. 26, atlarge distances, g(r) tends to the ideal gas value, converging to 1. Thisshows the absence of long-range forces at such distances.

4. Conclusion

In this paper, we investigate the effects of geometric shape of variousroughnesses on the fluid flow passing through a nanochannel by usingofmolecular dynamics simulation. The following results can be deducedfrom our investigation:

• By applying roughness to the inner surface of the ideal nano-channelat a driving force of 0.002 eV/Å, the amplitude of number density ofthe fluid particles near the walls decreased.

• The mean and maximum velocities increased by 6.5% and 2.5% in thepresence of square cuboid and hemispheroid roughness, respectively.

• The dimensionless slip velocity and slip length were, respectively, in-creased by a maximum of 41.1% and 21.5% in the presence of squarecuboid roughness, and by aminimumof 0.9% and 0.5% in the presenceof hemispheroid roughness.

• The temperature of theparticles at the center of thenano-channelwasincreased by a maximum of 9.1% and a minimum of 2.8% in the pres-ence of square cuboid and hemispheroid roughness, respectively.

• Calculation of the Argon-Argon radial distribution function indicatedthat the maximum of this function decreased by a maximum of 11.8%and aminimum of 8.5% in the presence of rectangular cuboid and ellip-soid roughness, respectively, compared to the ideal nano-channel.

The extension of this paper for MD and nano scale simulation ac-cording previous works [30–38] affords engineers a good option forsimulations.

References

[1] Sisir Kumar Das, Siddhartha Das, Suman Chakraborty, Molecular Dynamics Simula-tion of Water Flow Through Nanochannels, 19th National & 8th ISHMT-ASME Heatand Mass Transfer Conference January 3–5JNTU Hyderabad, India, 2008.

[2] Bo Hung Kim, Ali Beskok, Tahir Cagin, Thermal interactions in nanoscale fluid flow:molecular dynamics simulations with solid–liquid interfaces, Microfluid. Nanofluid.5 (2008) 551–559.

[3] Valery Ya Rudyak, Aleksandr A. Belkin, Veniamin V. Egorov, Denis A. Ivanov, Molec-ular Dynamics Simulation of Plane Poiseuille Flow in Nanochannels2nd Micro andNano Flows Conference West London, UK, 1–2 September, 2009.

[4] F. Sofos, T.E. Karakasidis, A. Liakopoulos, Non-equilibrium molecular dynamics in-vestigation of parameters affecting planar nanochannel flows, Contemp. Eng. Sci. 2(6) (2009) 283–298.

[5] D. Toghraie, A.R. Azimian, Non-equilibruim molecular dynamics simulation ofPoiseuille flow in a nanochannel, Iran. J. Mech. Eng. 11 (1) (March 2010).

[6] Jin Liu, MoranWang, Shiyi Chen, Mark O. Robbins, Molecular simulations of electro-osmotic flows in rough nanochannels, J. Comput. Phys. 229 (2010) 7834–7847.

[7] Baoxing Xu, Yibing Li, Taehyo Park, Xi Chen, Effect of wall roughness on fluid trans-port resistance in nanopores, J. Chem. Phys. 135 (2011), 144703.

[8] Reza Kamali, Ali Kharazmi, Molecular dynamics simulation of surface roughness ef-fects on nanoscale flows, Int. J. Therm. Sci. 50 (2011) 226–232.

[9] Murat Barisik, Ali Beskok, Equilibrium molecular dynamics studies on nanoscale-confined fluids, Microfluid. Nanofluid. 11 (2011) 269–282.

[10] Sridhar Kumar Kannam, B.D. Todd, J.S. Hansen, Peter J. Daivis, Slip length of water ongraphene: limitations of non-equilibrium molecular dynamics simulations, J. Chem.Phys. 136 (2012), 024705.

[11] Davood Toghraie, MajidMokhtari, Masoud Afrand, Molecular dynamic simulation ofcopper and platinum nanoparticles Poiseuille flow in a nanochannels, Phys. E. 84(2016) 152–161.

[12] J.W. Gibbs, Elementary Principles in Statistical Mechanics, Courier Corporation, 2014.[13] A. Jafari, V. Saljooghinezhad, Discrete element method to study the impact of different

parameters on the screen performance, Modares Mech. Eng. 14 (9) (2014) 27–34.[14] G.A. Mansoori, Principles of Nanotechnology: Molecular-based Study of Condensed

Matter in Small Systems, World Scientific Publishing Co Inc., 2005[15] J. Haile, I. Johnston, A.J. Mallinckrodt, S. Mckay, Molecular dynamics simulation: el-

ementary methods, Comput. Phys. 7 (6) (1993) 625.[16] M. Yousefi, F. Feyzi, Introduction to calculate the thermodynamic properties of nano

systems to molecular dynamics method, Iran. Chem. Eng. J. 9 (46) (2010) 13–23.[17] H. Haibao, B. Luyao, Molecular Dynamics Simulations of Nanoscale Poiseuille Flow,

2014.[18] M.P. Allen, D.J. Tildesley, Computer Simulation of Liquids, OxfordUniversity Press, 1989.[19] E. Van den Akker, A. Frijns, A. Van Steenhoven, P. Hilbers, Thermodynamic analysis

of molecular dynamics simulations of evaporation and condensation, Molecules1001 (2008) 1–8.

[20] G. Sutmann, Classical Molecular Dynamics and Parallel Computing, FZJ-ZAM, 2002.[21] M. Waldman, A.T. Hagler, New combining rules for rare gas van der Waals parame-

ters, J. Comput. Chem. 14 (9) (1993) 1077–1084.[22] D.T. Semiromi, A. Azimian, Nanoscale Poiseuille flow and effects of modified

Lennard–Jones potential function, Heat Mass Transf. 46 (7) (2010) 791–801.[23] P. Alipour, D. Toghraie, A. Karimipour, M. Hajian, Modeling different structures

in perturbed Poiseuille flow in a nanochannel by using of molecular dynamicssimulation: Study the equilibrium, Physica A 515 (2019) 13–30.

[24] H. Noorian, D. Toghraie, A.R. Azimian, Molecular dynamics simulation of Poiseuilleflow in a rough nano channel with checker surface roughnesses geometry, HeatMass Transf. 50 (1) (2014) 105–113.

[25] Y. Bakhshan, J.A. Shadloo, Molecular dynamics simulation of surface specifics effectson the nanoscale fluid flow, Modares Mech. Eng. 15 (5) (2015) 176–184.

[26] D. Toghraie, A.R. Azimian, Molecular dynamics simulation of nanodroplets with themodified Lennard-Jones potential function, Heat Mass Transf. 47 (5) (2011) 579–588.

[27] F. Faraji, A. Rajabpour, F. Kowsary, Temperature profile for nanoscale Poiseuille flow:a multiscale study, J. Mech. Sci. Technol. 30 (2) (2016) 803–807.

[28] R. Ranjith, J.H. Park, R.S. Myong, Temperature Distribution in the Force-drivenPoiseuille Gas Flow by Molecular Dynamics, UCL, London, UK, 2014 1–8.

[29] S.M. Fatemi Sh, M. Foroutan, Study of some properties of polymer-based nanocom-posites using molecular dynamic simulation, Besparesh 5 (1) (2015) 4–26.

[30] M.M.A. Pourshirazi, D. Toghraie, A.R. Azimian, Study effect of deformationnanochannel wall roughness on the water-copper nano-fluids Poiseuille flow be-havior, J. Simul. Anal. Nov. Technol. Mech. Eng. 8 (1) (2015) 29–40.

[31] M. Rezaei, A.R. Azimian, D. Toghraie, Molecular dynamic simulation of copper and plat-inum nanoparticles Poiseuille flow in a nanochannels, Physica A 84 (2016) 152–161.

[32] M. Rezaei, A.R. Azimian, D. Toghraie, The surface charge density effect on theelectro-osmotic flow in a nanochannel: a molecular dynamics study, Heat MassTransf. 51 (5) (2015) 661–670.

[33] Masoumeh Tohidi, Davood Toghraie, The effect of geometrical parameters, rough-ness and the number of nanoparticles on the self-diffusion coefficient in Couetteflow in a nanochannel by using of molecular dynamics simulation, Physica B 518(2017) 20–32.

[34] H. Noorian, D. Toghraie, A.R. Azimian, The effects of surface roughness geometry offlow undergoing Poiseuille flow by molecular dynamics simulation, Heat MassTransf. 50 (2014) 95–104.

[35] D. Toghraie, A.R. Azimian, Molecular dynamics simulation of annular flow boilingwith the modified Lennard-Jones potential function, Heat Mass Transf. 48 (1)(2012) 141–152.

[36] D. Toghraie, A.R. Azimian, Molecular dynamics simulation of liquid–vapor phaseequilibrium by using the modified Lennard-Jones potential function, Heat MassTransf. 46 (3) (2010) 287–294.

[37] M.R. Faridzadeh, D. Toghraie, A. Niroomand, Analysis of laminar mixed convectionin an inclined square lid-driven cavity with a nanofluid by using an artificial neuralnetwork, Heat Transf. Res. 45 (4) (2014).

[38] M.R. Shamsi, O.A. Akbari, A. Marzban, D. Toghraie, R. Mashayekhi, Increasing heattransfer of non-Newtonian nanofluid in rectangular microchannel with triangularribs, Phys. E Low-Dimensional Syst. Nanostruct. 93 (2017) 167–178.