Effect of hydrophobicity on the water flow in carbon nanotube—A molecular dynamics study

Effect of hydrophobicity on the water flow in carbon nanotube—A molecular dynamics study

Theoretical & Applied Mechanics Letters 8 (2018) 284-290   Contents lists available at ScienceDirect Theoretical & Applied Mechanics Letters journal...

2MB Sizes 0 Downloads 40 Views

Theoretical & Applied Mechanics Letters 8 (2018) 284-290  

Contents lists available at ScienceDirect

Theoretical & Applied Mechanics Letters journal homepage: www.elsevier.com/locate/taml    

Letter

Effect of hydrophobicity on the water flow in carbon nanotube—A molecular dynamics study Hamed Esmaeilzadeh, Junwei Su, Majid Charmchi, Hongwei Sun*   Department of Mechanical Engineering, University of Massachusetts Lowell, Lowell, MA 01854, USA

H  I  G  H  L  I  G  H  T  S

•  The effect of hydrophobicity on the water flow in carbon nanotubes (CNTs) is studied using a molecular dynamics approach. •  The hydrophobicity was characterized by interaction-energy-coefficient (IEC). •  The static contact angles between water and carbon surface decreases with increasing the IEC. •  The hydrophobicity of CNT surface has a significant impact on the pressure drops of water flow in CNT. A  R  T  I  C  L  E      I  N  F  O

A  B  S  T  R  A  C  T

 

 

Article history: Received 28 April 2018 Received in revised form 25 June 2018 Accepted 2 July 2018 Available online 2 July 2018 *This article belongs to the Fluid Mechanics.

   

Keywords: Molecular dynamics Contact angle Interaction-energy-coefficient Water flow Carbon nanotube

This work focuses on the study of the effect of hydrophobicity on the water flow in carbon nanotubes (CNTs) using a molecular dynamics (MD) approach for a wide range of potential applications such as water purification and high efficiency of nanofluid energy absorption systems (NEAS). The hydrophobicity between liquid water and surface of CNTs was characterized by interaction-energy-coefficient (IEC)—a parameter describing the energy interaction strength between water molecules and carbon atoms. It is shown that the static contact angles between water and carbon surface decrease from 155° to 44° when the values of IEC increase from 0.042 kJ/mol to 2.196 kJ/mol. In addition, the pressure drops in CNT became independent of IEC when the IEC value was higher than 1.192 kJ/mol for a given flow rate. It was found that the hydrophobicity of CNT surface has a significant impact on the pressure drop of water flow in the CNTs and MD method provides a quantitative evaluation of the impact. ©2018 The Authors. Published by Elsevier Ltd on behalf of The Chinese Society of Theoretical and Applied Mechanics. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).  

 

  Carbon nanotube (CNT) is one of the most promising nanomaterials due to its high thermal and electrical conductivities and low thermal expansion coefficient that make them well suitable candidates for various applications such as water purification [1], biomolecular sensors [2], drug delivery systems [3], hydrogen storage medium in fuel cells [4], carbon nanopipettes [5], and high efficiency of nanofluid energy absorption systems (NEAS) [6]. For example, CNT membrane has been studied as the filter materials for water purification or desalination system [7]. Also, some researchers proposed to use CNT membrane as nanosyringes to deliver drugs in which the drug is released

 

   

* Corresponding author. E-mail address: [email protected] (H.W. Sun).

 

through cellular membrane [8]. Therefore, it is very critical to understand the water flow behavior in CNTs for a better design of nanofluidic-based systems for the stated applications. As the effect of surface forces in the nanoscale becomes significant, the continuum assumption breaks down and the water flow in nanoscale channels exhibits significantly different behavior from the flow in the macroscale and/or microscale channels [9]. The hypothesis of nonslip boundary condition, applied in large-scale analyses, may not be valid in nanoscale systems [10, 11 ]. As a result, the slip velocity at the wall could lead to higher flux of flow in comparison to nonslip continuum condition [12-15]. Considering the experimental study in nanoscales is still a great challenge, theoretical and numerical models are widely

http://dx.doi.org/10.1016/j.taml.2018.04.007 2095-0349/© 2018 The Authors. Published by Elsevier Ltd on behalf of The Chinese Society of Theoretical and Applied Mechanics. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

H. Esmaeilzadeh et al. / Theoretical & Applied Mechanics Letters 8 (2018) 284-290

used for understanding the behaviors of fluid flow in CNTs under various situations [16-19]. As an experimental computation method in simulating kinetic behavior of substance at molecular and atomic scales, molecular dynamics (MD) has been widely utilized to simulate fluid behaviors at nanoscale channels, enabling prediction of behavior of fluid flows and observation of the fluid/wall interface effects [20-22]. Briefly, MD is a deterministic simulation method, in which the atomic trajectories of fluid particles are tracked by numerically solving Newton's equation of motion based on a specific interatomic potential and certain initial and boundary conditions [23, 24]. For the simulation of water molecules in CNT, the pairwise interaction between water molecules and carbon atoms (Uwc), was generally modeled by the so-called Lennard–Jones (LJ) potential energy [25]. Hummer et al. [26] employed two different values for the interaction parameters between water molecules and carbon atoms in their MD simulation. The results showed that small changes in these parameters caused different occupancy of water molecules in CNT—two state transitions between empty and filled states. Thomas and McGaughey investigated the pressure driven water flows through CNTs with diameters ranging from 1.66 nm to 4.99 nm using MD simulation. Their results revealed that the ratio of the observed flow rate from MD simulation to that predicted from the nonslip Hagen–Poiseuille relation decreased with increasing CNT diameter and varied from 433 to 47 [27]. Joly [28] studied the effect of liquid/wall slip on dynamic of capillary tube filling in nanometric pores where it was found that for water flow in CNT, there was no shear stresses inside the liquid. Therefore, viscous term is only based on the friction coefficient of liquid at the wall and the liquid viscosity does not play a role. Nichollas et al. [29] employed the non-equilibrium molecular dynamic (NEMD) method to study the water flow in CNT to examine the length effect on the internal flow dynamics. Their results showed that under similar applied pressure difference, an increase in CNT length had no effect on the mass flow rate and fluid flow velocity. Liu et al. [30] investigated the pressuredriven water transport inside the nanotube formed by graphene bilayers through molecular dynamics simulation. Their results showed the water flow rate increases linearly as the applied pressure gradient increases. The weak viscous effect and therefore the very low friction at the channel walls is the key factor for achieving fast flow rates in nanotubes. Recently, Ganjiani and Hossein Nezhad [31] used MD simulation to study the effect of carbon/water interaction parameters on NEAS. It was found that the contact angle was reduced by increasing values of the energy and distance parameters of LJ potential. As mentioned earlier, these parameters are related to interaction’s strength and distance between water molecules and carbon atoms, respectively. Cao and Liu [32, 33] studied the working mechanism of energy absorption in NEAS under impact load using MD simulation. Their results confirmed that NEAS can be a promising tool to protect against high speed load rate. All of simulations performed so far employed different values of interaction parameters for MD simulation which heavily rely on the methods which were used to determine these interaction parameters [34]. These methods include Moller-plesset perturbation theory of the second order (MP2) [35], random phase approximation (RPA) [36], diffusion Monte Carlo (DMC) [36], density functional theory-symmetry-adapted perturbation theory (DFT-SAPT) [37],

285

density functional theory/coupled cluster (DFT/CC) [38], density functional theory with dispersion correction (DFT-D) [39], and coupled cluster treatment with single and double excitations and perturbative triples (CCSD(T)) [40]. These interactionenergy-coefficients (IECs) values leads to different surface hydrophobicities of CNT from hydrophilic [28] to hydrophobic ones [41, 42]. In addition, a review of the previous works showed that CNT can be modified to form different surface hydrophobicities [43-45] and a different surface property can be achieved for a desired application by selecting specific functional groups coated onto the CNT surface [46-48]. The focus of this research is to explore the hydrodynamic behavior of water flow in a CNT connected between two reservoirs, where water is driven by two moving pistons. First, the relationship between IEC and hydrophobicity of water molecules on carbon atoms was quantified through the change of contact angles of a water drop on carbon surface. Then the effect of the hydrophobicity on the water flow in the nanotubes was investigated in terms of pressure drop, axial density and axial velocity distributions. The MD model system consists of a CNT, two reservoirs at the two ends of CNT, two moving pistons of graphene, and water molecules confined in CNT and reservoirs, as shown in Fig. 1. The simple point charge (SPC) model for water molecules was utilized in this research since it reproduces the transport properties of water close to experimental results [49]. The integration of equations of motions were performed using Verlet algorithm in the canonical network virtual terminal (NVT) ensemble by using LAMMPS software [50]. The time step of simulation was set to be 1 fs (1 fs = 1 × 10–15 s) and the cut-off radius of LJ potential energy was 8 Å (1 Å = 1 × 10–10 m). The energy and distance scales between water molecules and carbon atoms were "wc = 0.389 kJ/mol and ¾wc = 3.283 Å, respectively [51]. The long range Columbic interaction was computed with Ewald technique to consider the electrostatic charge between water molecules [52]. The boundary conditions were periodic in all of directions. The CNT has different atomic arrangements including armchair, zigzag, and chiral [53]. The zigzag CNT was used in this study with radius of 5.45 Å and length of 25.3 Å. Based on dimensions of CNT and density of water at 298 K (1 g/cm3), the number of water molecules contained in the zigzag CNT was about 80 [51]. The water reservoirs with the dimensions of 29.82 Å × 27.69 Å × 10.2 Å contains 297 water molecules assuming the  

Piston

Water molecules

CNT

Piston

  Fig. 1.   Schematic of the water flow system in a CNT including water molecules, two moving pistons of graphene and CNT atoms.

H. Esmaeilzadeh et al. / Theoretical & Applied Mechanics Letters 8 (2018) 284-290

density of the water 1 g/cm3. Two pistons are moving downward with constant velocity of 10 m/s. The initial locations of the water molecules and carbon atoms were built using the opensource MOLTEMPLATE software, a general text-based molecule builder [54]. To validate the MD model, the simulation result performed in this study by LAMMPS software [55] was compared with the data reported by Liu et al. [51] and the comparison is presented in Fig. 2. The CNT radius of 7.85 Å was used in the simulation. It can be seen that the density profile obtained from MD simula-

n =1+

90

 

80 This study Results of Ref. [51]

Density (g/cm3)

70 60 40 30 10 0 2

 

3 4 5 Radius (Å)

6

7

8

Fig. 2.   Comparison of the simulation results of water flow in a CNT (radius = 7.85 Å) without the presense of reservoirs and pistons with the data published by Liu et al. [51] at 298 K and water density of 1 g/cm3.

As Eq. (2) shows, with increasing CNT diameter, the density of the water molecule is approaching to that of bulk water condition. Our previous results also showed that by increasing the diameter of CNT, the results of density approached to the bulk condition (1 g/cm3) [58] as predicted by Eq. (2). The effect of various IEC on the hydrophobicity was inspected as the IEC is directly related to the interaction force and energy between water molecules and carbon atoms [25] and the values of IEC were changed from 0.042 kJ/mol to 2.196 kJ/mol [45, 59]. Figure 3(a) describes the interaction force and energy between water molecules and carbon atoms versus particle distances for IEC at "wc = 2.196 kJ/mol. It can be seen that at equi-

tion is in a good agreement with Liu’s result. The MD simulation results shows that there is almost no molecule in the region near the wall of the CNT. The reason is that the carbon atoms and water molecules tend to keep a distance of approximately ¾C¡O = 3.283 Å from each other [56]. There are two density peaks appearing at radius of r of 2.5 Å and 5.0 Å. Wang et al. [57] proposed an equation to calculate the number of layers n that are formed completely in the CNT with radius of Rn as follows: a

300 250

Strong repulsive forces

200 F(r) (pN)

Force Potential energy

r<σ

150 100

3

Weak attractive force

2

r=2σ

1 0

0 −50

−1

Repulsion

0.25

0.75

b

5 4

Separation at energy minimum

r = 1.12 σ

50

6

Attraction 1.25

1.75 r/σwc

−2 2.25

2.75

−3

25 20

U(r) (kJ/mol)

 

(2)

where ½0 is the density of bulk water. The values of ½¤ and d ¤ are caluclated by: ( 1:4 g=cm3 ; d < 24:8 º A; ½¤ = 3 0:6 g=cm ; d ¸ 24:8 º A; ( º º 20:6 A ; d < 24:8 A ; (3) d¤ = º 55:0 A ; d ¸ 24:8 º A:

20

1

(1)

½ ½¤ =1¡ µ ¶2 ; ½0 d +1 d¤

50

0

R n ¡ ¾C¡O : ¾O¡O

Based on values of ¾O¡O = 3.166 Å, ¾C¡O = 3.283 Å and radius of Rn = 7.85 Å, two layers of water were formed completely which is consistent with the results shown in Fig. 2. Equation (1) indicates that with increasing the CNT diameter, the number of layers will increase and the ordered structure of water molecules in CNT will be disappearing to the bulk condition. The similar layered structure and solid-like distribution of water molecules in CNT with various diameters was described in our previous work [58]. Alexiadis and Kassinos [56] proposed a correlation between the overall density of water and CNT diameter as follows:

U(r) (kJ/mol)

286

εwc = 0.25 kJ/mol εwc = 0.527 kJ/mol εwc = 1.192 kJ/mol εwc = 2.196 kJ/mol

15 10 5 0 −5

0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6 r/σwc

  Fig. 3.   a Force vs. distance between water molecules and carbon atoms for IEC of 2.196 kJ/mol (1 pN = 1×10-12 N). b Interaction energy vs. distances for different values of IECs.

H. Esmaeilzadeh et al. / Theoretical & Applied Mechanics Letters 8 (2018) 284-290

librium distance ro = 1:12¾ the interaction force vanished and the interaction energy reached the minimum value (−"wc ). When the distances between particles is less than ro , the particles will repel each other while when the distance is greater than 1:12¾ , the interaction force between particles became attractive. Figure 3(b) shows the interaction energy between water molecules and carbon atoms versus different values of "wc . As can be seen, with increasing values of the IEC, the interaction energy Uwc increases and the equilibrium distance, where dU=drjr=r equilibrium = 0 becomes larger. The effect of IEC on the hydrophobicity was identified by measuring the static contact angle of a water droplet on carbon surface using MD model where the value of IEC was varied from 0.042 kJ/mol to 2.196 kJ/mol [60-63]. For this purpose, 2000 water molecules and a graphene surface consisting of 5408 carbon atoms were employed. The total simulation time was 0.045 ns (1 ns = 1×10–9 s) so that a stable contact angle was reached between the water drop and graphene surface. Then, using the curve fitting toolbar in MATLAB software, the droplet contact angle was determined. The change of static contact angle with IEC for a simulation duration of 0.045 ns is presented in Fig. 4. The MD results show that the contact angle as high as 155°

can be reached representing a superhydrophobic state when the value of IEC is low ("wc = 0.042 kJ/mol). With the increasing "wc , the contact angle decreased from 155° down to as low as 40° demonstrating the transition from hydrophobic to hydrophilic states. In addition, after the value of IEC reached about 1.192 kJ/mol, the further increase of IEC values could not reduce the contact angle value any further and the graphene surface displayed a hydrophilic state. Further study is underway to explore this phenomenon. The water flows in CNTs were simulated using the MD model under different IEC values to quantify the effect of hydrophobicity. Figure 5(a) shows the solid/liquid interaction energy per unit area of the CNT as function of IEC ("wc ) at the simulation time of 0.04 ns (tsimulation = 0.04 ns). The results show that the interaction energy (Uwc) increases almost linearly with the increasing of the IEC value as predicted by LJ potential energy [25]. Then the water flow in a CNT described in Fig. 1 was simulated under different IEC values and the pressure drops along the CNT are presented in Fig. 5(b). It can be seen that at the low IEC values the pressure drops along the CNT are low due to the weak interactions between the water molecules and carbon surface. As the interaction between water molecules and carbon atoms become stronger, as indicated by larger IEC values, the pressure drops increased sharply due to the increased friction forces caused by strong interactions between water molecules and carbon surface. The pressure drop reached a stable value when the IEC reached 1.192 kJ/mol, where the surface became hydrophilic. The trend of the pressure drop increase is consistent with the decrease of the contact angle presented in Fig. 4. Figure 6(a) presents the mass density distribution in the axial direction of the water molecules under different IECs. As the IEC increase, the mass density peaks of water molecules tends to move to the centerline of CNT. It is believed that the redistribution of the water molecules in CNT was due to the change of the ordered configuration of the hydrogen bond network between molecules. The hydrogen bonding is an attractive interaction between a hydrogen atom of one water molecule and a negative oxygen atom of another water molecule [64]. As a result, the water molecule can make four hydrogen bonds such that the average number of hydrogen bonds per molecule for the bulk water

 

160

Contact angle (º)

140 120 100 80 60 40 0

0.3

0.6

0.9 1.2 1.5 εwc (kJ/mol)

1.8

2.1

  Fig. 4.   The dependence of static contact angles of water droplets on a graphene sheet vs. IECs at t = 0.045 ns.  

a

b

500 450

21

Pressure drop (MPa)

Interaction energy (kJ/mol)

24

18 15 12 9 6

400 350 300 250 200 150 100

3 0

287

50 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00 2.25 εwc (kJ/mol)

0

0

0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00 2.25 εwc (kJ/mol)

  Fig. 5.   a Interaction energy between water molecules and CNT atoms. b Pressure drop along CNT vs. IECs at tsimulation = 0.04 ns for the water flow at 298 K and density of 1 g/cm3 in the presense of two reservoirs and two moving pistons with the constant velocity of 10 m/s.

a

Mass density (g/cm3)

 

H. Esmaeilzadeh et al. / Theoretical & Applied Mechanics Letters 8 (2018) 284-290

60 50 40

b εwc = 0.042 kJ/mol εwc = 0.251 kJ/mol εwc = 0.69 kJ/mol εwc = 2.196 kJ/mol

30 20

Increasing εwc

10 0

 

0 0.5 1.0 1.5 2.0 2.5 3.5 3.5 4.0 4.5 5.0 5.5 Radius (Å)

Vz (m/s)

288

3500 3000 2500 2000 1500 1000 500 0 −500 −1000 −1500 −2000

εwc = 0.042 kJ/mol εwc = 0.251 kJ/mol εwc = 0.69 kJ/mol εwc = 2.196 kJ/mol

Increasing εwc 0 0.5 1.0 1.5 2.0 2.5 3.5 3.5 4.0 4.5 5.0 5.5 Radius (Å)

Fig. 6.   a Mass density and b velocity distribution in axial direction inside CNT for different values of IECs for the water flow at 298 K and density of 1 g/cm3 in presense of two reservoirs and two moving pistons with constant velocity of 10 m/s.

varies between 2.4 to 3.9 for different water models [65, 66]. However, for the water molecules in the CNT, the average number of hydrogen bonds per molecule decreases to 0.9, where only one single chain is allowed to move due to the space limitation [67, 68 ]. In the meantime, the layered structures of water molecules in a CNT behave as a spring with a resonant frequency of vibration due to the hydrogen bonds. The transport of layered structures in confined systems proceed as a wave motion with Eigen frequencies which can be determined by the CNT size [69, 70]. As can be seen in Fig. 6(a), two mass density peaks were formed in the CNT—one at the center of the CNT and the other at the radius of 2:5¾ . With the increasing value of IEC from 0.250 kJ/mol to 2.196 kJ/mol, the values of mass densities at the two peaks increased which is consistent with the results reported by Köhler et al. [71]. Furthermore, the location of the second peak moved closer to the CNT center. It is believed this movement is due to the stronger interaction between the carbon wall and the water molecules which is less favored energetically. This is consistent with the results of pressure drop in Fig. 5(b) which realved that for larger values of IECs a higher pressure (energy) was required to push water molecules through the CNT. Therefore, the water molecules preferred to move toward the center with increasing value of IEC. Figure 6(b) presents the results for velocity distribution of water molecules in axial direction through the CNT for different values of IECs. As can be seen there, the water flow in the CNT was similar to a plug-like flow which was reported in Refs. [72-74]. The friction force between water molecules and the CNT atoms resulted in a stepwise variation of velocity profile which was consistent with those results of Hanasaki and Nakatani [72]. With the increasing IEC, the friction force became larger leading to a larger jump in the streaming velocity profile near the wall. In addition, the increasing values of IEC caused the increases of the axial velocity in the CNT. This is because with the increasing value of IEC, higher pressure (energy) was transferred on water molecules as shown in Fig. 5b leading to increases of axial velocity. In this research, the effect of wall hydrophobicity of CNTs on the hydrodynamic behavior of water flow was studied through MD simulation. The hydrophobicity of CNT was changed by selecting various values of IEC from 0.042 kJ/mol to 2.196 kJ/mol while the contact angle decreased from 155° to 40° indicating the

transition from hydrophobic to hydrophilic states. When the values of IEC were higher than 1.192 kJ/mol, the static contact angle didn’t reduce anymore and the graphene surface displays hydrophilic states. The results for pressure drop, axial density and axial velocity of water flow in CNT in presence of two reservoirs were presented. Based on the MD numerical simulation, plug flows were observed in CNTs compared to the parabolic profile for a water flow under continuum flow condition. More importantly, the results showed that with the increasing of IEC and stronger interactions between water molecules and wall carbon atoms, pressure drops quickly increased due to the increased friction forces. However, when the IEC reached 1.192 kJ/mol, the CNT surface became hydrophilic and the pressure drop reached to a maximum and stable value. The results were consistent with the minimum contact angle obtained by the MD simulation for a water drop on the graphene surface.

References [1] M. Sweetman, S. May, N. Mebberson, et al., Activated carbon,

[2]

[3]

[4] [5] [6]

[7]

[8]

carbon nanotubes and graphene: materials and composites for advanced water purification, C. 3 (2017) 18. B.C. Satishkumar, L.O. Brown, Y. Gao, et al., Reversible fluorescence quenching in carbon nanotubes for biomolecular sensing, Nature Nanotechnology 2 (2007) 560–564. A. Bianco, K. Kostarelos, M. Prato, Applications of carbon nanotubes in drug delivery, Current Opinion in Chemical Biology 9 (2005) 674–679. H.M. Cheng, Q.H. Yang, C. Liu, Hydrogen storage in carbon nanotubes, Carbon 39 (2001) 1447–1454. R.C. Mani, X. Li, M.K. Sunkara, et al., Carbon nanopipettes, Nano Letters 3 (2003) 671–673. H. Liu, G. Cao, Reusable energy absorption performance based on nanofluidic systems, The Journal of Physical Chemistry C 120 (2016) 5213–5220. R. Das, M.E. Ali, S.B.A. Hamid, et al., Carbon nanotube membranes for water purification: A bright future in water desalination, Desalination 336 (2014) 97–109. V. Rastogi, P. Yadav, S.S. Bhattacharya, et al., Carbon nanotubes: An emerging drug carrier for targeting cancer cells, Journal of Drug Delivery 2014 (2014) 670815.

H. Esmaeilzadeh et al. / Theoretical & Applied Mechanics Letters 8 (2018) 284-290

289

[9] A. Barati Farimani, N.R. Aluru, Existence of multiple phases of

[27] J.A. Thomas, A.J.H. McGaughey, Reassessing fast water trans-

water at nanotube interfaces, The Journal of Physical Chemistry C 120 (2016) 23763–23771. H. Esmailzadeh, M. Passandideh-Fard, Numerical and experimental analysis of the fluid-structure interaction in presence of a hyperelastic body, Journal of Fluids Engineering 136 (2014) 111107–111112. H. Esmailzadeh, M. Passandideh-Fard, Numerical and experimental analysis of the motion of a hyperelastic body in a fluid, in: 21st Annual International Conference on Mechanical Engineering-ISME2013, 2013 A.I. Podolska, N.I. Zhmakin, Water flow in micro- and nanochannels, Molecular dynamics simulations, Journal of Physics: Conference Series 461 (2013) 012034. T.G. Myers, Why are slip lengths so large in carbon nanotubes?, Microfluidics and Nanofluidics 10 (2011) 1141–1145. Y. Li, J. Xu, D. Li, Molecular dynamics simulation of nanoscale liquid flows, Microfluidics and Nanofluidics 9 (2010) 1011–1031. S. Santhosh, Fluid flow through carbon nanotubes and nanopipes, SSRG International Journal of Mechanical Engineering (SSRG-IJME) 4 (2017) 51–57. H. Aminfar, N. Razmara, M. Mohammadpourfard, On flow characteristics of liquid-solid mixed-phase nanofluid inside nanochannels, Applied Mathematics and Mechanics 35 (2014) 1541–1554. J.C.T. Eijkel, A.V.D. Berg, Nanofluidics: what is it and what can we expect from it?, Microfluidics and Nanofluidics 1 (2005) 249–267. B.Y. Cao, J. Sun, M. Chen, et al., Molecular momentum transport at fluid-solid interfaces in MEMS/NEMS: a review, International Journal of Molecular Sciences 10 (2009) 4638–4706. F. Calabrò, K.P. Lee, D. Mattia, Modelling flow enhancement in nanochannels: Viscosity and slippage, Applied Mathematics Letters 26 (2013) 991–994. H. Esmaeilzadeh, J. Su, C. Su, et al., Molecular dynamic simulation of Couette flow of liquid argon in nanochannel, in: ASME 2016 Fluids Engineering Division Summer Meeting collocated with the ASME 2016 Heat Transfer Summer Conference and the ASME 2016 14th International Conference on Nanochannels, Microchannels, and Minichannels, 2016, pp. V01BT19A001V01BT19A001 M. Barisik, A. Beskok, Surface–gas interaction effects on nanoscale gas flows, Microfluidics and Nanofluidics 13 (2012) 789–798. T.H. Yen, Molecular dynamics simulation of fluid containing gas in hydrophilic rough wall nanochannels, Microfluidics and Nanofluidics 17 (2014) 325–339. K.G.S. Dilrukshi, M.A.N. Dewapriya, U.G.A. Puswewala, Size dependency and potential field influence on deriving mechanical properties of carbon nanotubes using molecular dynamics, Theor. Appl. Mech. Lett. 5 (2015) 167–172. Z. Song, Y. Wang, Z. Xu, Mechanical responses of the bio-nano interface: A molecular dynamics study of graphene-coated lipid membrane, Theor. Appl. Mech. Lett. 5 (2015) 231–235. M.P. Allen, D.J. Tildesley, Computer Simulation of Liquids, Clarendon Press, 1989, p. 385 G. Hummer, J.C. Rasaiah, J.P. Noworyta, Water conduction through the hydrophobic channel of a carbon nanotube, Nature 414 (2001) 188–190.

port through carbon nanotubes, Nano Letters 8 (2008) 2788–2793. L. Joly, Capillary filling with giant liquid/solid slip: Dynamics of water uptake by carbon nanotubes, The Journal of Chemical Physics 135 (2011) 214705. W.D. Nicholls, M.K. Borg, D.A. Lockerby, et al., Water transport through (7.7) carbon nanotubes of different lengths using molecular dynamics, Microfluidics and Nanofluidics 12 (2012) 257–264. B. Liu, R. Wu, J.A. Baimova, et al., Molecular dynamics study of pressure-driven water transport through graphene bilayers, Physical Chemistry Chemical Physics 18 (2016) 1886–1896. S.H. Ganjiani, A. Hossein Nezhad, Molecular dynamics simulation of the effects of the carbon–water interaction parameters on the nanofluidic energy absorption system, The Journal of Physical Chemistry C 120 (2016) 11864–11870. G. Cao, Working mechanism of nanoporous energy absorption system under high speed loading, The Journal of Physical Chemistry C 116 (2012) 8278–8286. H. Liu, G. Cao, Interaction between mechanical wave and nanoporous energy absorption system, The Journal of Physical Chemistry C 117 (2013) 4245–4252. Y. Wu, N.R. Aluru, Graphitic carbon–water nonbonded interaction parameters, The Journal of Physical Chemistry B 117 (2013) 8802–8813. I.W. Sudiarta, D.J.W. Geldart, Interaction energy of a water molecule with a single-layer graphitic surface modeled by hydrogen- and fluorine-terminated clusters, The Journal of Physical Chemistry A 110 (2006) 10501–10506. J. Ma, A. Michaelides, D. Alfè, et al., Adsorption and diffusion of water on graphene from first principles, Physical Review B 84 (2011) 033402. G.R. Jenness, O. Karalti, K.D. Jordan, Benchmark calculations of water-acene interaction energies: Extrapolation to the watergraphene limit and assessment of dispersion-corrected DFT methods, Physical Chemistry Chemical Physics 12 (2010) 6375–6381. M. Rubeš, P. Nachtigall, J. Vondrášek, et al., Structure and stability of the water-graphite complexes, The Journal of Physical Chemistry C 113 (2009) 8412–8419. H. Li, X.C. Zeng, Wetting and interfacial properties of water nanodroplets in contact with graphene and monolayer boron–nitride sheets, ACS Nano 6 (2012) 2401–2409. E. Voloshina, D. Usvyat, M. Schutz, et al., On the physisorption of water on graphene: a CCSD(T) study, Physical Chemistry Chemical Physics 13 (2011) 12041–12047. L. Liu, J. Zhao, C.Y. Yin, et al., Mechanisms of water infiltration into conical hydrophobic nanopores, Physical Chemistry Chemical Physics 11 (2009) 6520–6524. L. Liu, J. Zhao, C. Yin, et al., Mechanisms of water infiltration into conical hydrophobic nanopores, Physical Chemistry Chemical Physics 11 (2009) 6520. F. Li, B. Tang, J. Xiu, et al., Hydrophilic modification of multiwalled carbon nanotube for building photonic crystals with enhanced color visibility and mechanical strength, Molecules 21 (2016) 547. X. Dai, F. Yang, R. Yang, et al., Biphilic nanoporous surfaces enabled exceptional drag reduction and capillary evaporation enhancement, Applied Physics Letters 105 (2014) 191611.

[10]

[11]

[12]

[13] [14]

[15]

[16]

[17]

[18]

[19]

[20]

[21]

[22]

[23]

[24]

[25] [26]

[28]

[29]

[30]

[31]

[32]

[33]

[34]

[35]

[36]

[37]

[38]

[39]

[40]

[41]

[42]

[43]

[44]

290

H. Esmaeilzadeh et al. / Theoretical & Applied Mechanics Letters 8 (2018) 284-290

[45] I. Moskowitz, M.A. Snyder, J. Mittal, Water transport through

[46]

[47]

[48]

[49]

[50]

[51]

[52]

[53]

[54] [55] [56] [57]

[58]

[59]

[60]

functionalized nanotubes with tunable hydrophobicity, The Journal of Chemical Physics 141 (2014) 18C532. M. Toupin, D. Bélanger, Spontaneous functionalization of carbon black by reaction with 4-nitrophenyldiazonium cations, Langmuir 24 (2008) 1910–1917. J. Liu, L. Cheng, B. Liu, et al., Covalent modification of a glassy carbon surface by 4-aminobenzoic acid and its application in fabrication of a polyoxometalates-consisting monolayer and multilayer films, Langmuir 16 (2000) 7471–7476. M.T. Alam, J.J. Gooding, Modification of Carbon Electrode Surfaces, Electrochemistry of Carbon Electrodes, Wiley: Hoboken, NJ, USA, 2016 G. Raabe, R.J. Sadus, Molecular dynamics simulation of the effect of bond flexibility on the transport properties of water, The Journal of Chemical Physics 137 (2012) 104512. H. Esmaeilzadeh, J. Su, M. Charmchi, et al., Molecular dynamic simulation of pressure drop of water flow in carbon nanotube in presence of two reservoirs and movable pistons, in: the ASME 2017 15th International Conference on Nanochannels, Microchannels, and Minichannels, 2017, V001T15A001 Y. Liu, Q. Wang, L. Zhang, et al., Dynamics and density profile of water in nanotubes as one-dimensional fluid, Langmuir 21 (2005) 12025–12030. S. Kuwajima, A. Warshel, The extended Ewald method: A general treatment of long-range electrostatic interactions in microscopic simulations, The Journal of Chemical Physics 89 (1988) 3751–3759. A. Aqel, K.M.M.A. El-Nour, R.A.A. Ammar, et al., Carbon nanotubes, science and technology, part I: structure, synthesis and characterisation, Arabian Journal of Chemistry 5 (2012) 1–23. J. Andrew, Available from: http://www.moltemplate.org/ S. Plimpton, Fast parallel algorithms for short-range molecular dynamics, Journal of Computational Physics 117 (1995) 1–19. A. Alexiadis, S. Kassinos, The density of water in carbon nanotubes, Chemical Engineering Science 63 (2008) 2047–2056. J. Wang, Y. Zhu, J. Zhou, et al., Diameter and helicity effects on static properties of water molecules confined in carbon nanotubes, Physical Chemistry Chemical Physics 6 (2004) 829–835. H. Esmaeilzadeh, J. Su, C. Su, et al., Molecular dynamic simulation of water flow in carbon nanotube, in: the ASME 2017 15th International Conference on Nanochannels, Microchannels, and Minichannels, 2017, V001T11A001 Y.S. Xu, X.L. Tian, M. Lv, et al., Effects of water-channel attractions on single-file water permeation through nanochannels, Journal of Physics D: Applied Physics 49 (2016) 285302. S.D. Hong, M.Y. Ha, S. Balachandar, Static and dynamic con-

[61]

[62] [63]

[64]

[65]

[66]

[67]

[68]

[69]

[70]

[71]

[72]

[73] [74]

tact angles of water droplet on a solid surface using molecular dynamics simulation, Journal of Colloid and Interface Science 339 (2009) 187–195. T. Werder, J.H. Walther, R.L. Jaffe, et al., On the water-carbon interaction for use in molecular dynamics simulations of graphite and carbon nanotubes, The Journal of Physical Chemistry B. 107 (2003) 1345–1352. J. Ritchie, Contact angle of a nano-drop on a heterogeneous surface, [PhD Thesis], Virginia Commonwealth University, 2010 E. Santiso, C. Herdes, E. Müller, On the calculation of solid-fluid contact angles from molecular dynamics, Entropy 15 (2013) 3734–3745. J.W. Choi, M. Alexandrova, H. Park, Carbon nanotube nanofluidics, in: Carbon Nanotubes Applications on Electron Devices, InTech, 2011 J. Zielkiewicz, Structural properties of water: Comparison of the SPC, SPCE, TIP4P, and TIP5P models of water, The Journal of Chemical Physics 123 (2005) 104501. W. Jorgensen, C. Jenson, Temperature dependence of TIP3P, SPC, and TIP4P water from NPT Monte Carlo simulations: Seeking temperatures of maximum density, Journal of Computational Chemistry 19 (1998) 1179–1186. N. Zhang, C. Chen, Y. Feng, et al., Molecular dynamics simulation of the hydrogen bonding structure of water molecules inside carbon nanotube, in: the ASME 2013 11th International Conference on Nanochannels, Microchannels, and Minichannels, 2013, V001T12A002 M.C. Gordillo, J. Martí, Hydrogen bond structure of liquid water confined in nanotubes, Chemical Physics Letters 329 (2000) 341–345. J.Y. Li, Z.Q. Wu, J.J. Xu, et al., Water transport within carbon nanotubes on a wave, Physical Chemistry Chemical Physics 18 (2016) 33204–33210. M. Ma, F. Grey, L. Shen, et al., Water transport inside carbon nanotubes mediated by phonon-induced oscillating friction, Nature Nanotechnology 10 (2015) 692–695. M.H. Köhler, J.R. Bordin, L.B. da Silva, et al., Structure and dynamics of water inside hydrophobic and hydrophilic nanotubes, Physica A: Statistical Mechanics and its Applications 490 (2018) 331–337. I. Hanasaki, A. Nakatani, Flow structure of water in carbon nanotubes: Poiseuille type or plug-like?, The Journal of Chemical Physics 124 (2006) 144708. S. Joseph, N.R. Aluru, Why are carbon nanotubes fast transporters of water?, Nano Letters 8 (2008) 452–458. X. Chen, G. Cao, A. Han, et al., Nanoscale fluid transport: size and rate effects, Nano Letters 8 (2008) 2988–2992.