You are on page 1of 8

Journal of Molecular Liquids 323 (2021) 114610

Contents lists available at ScienceDirect

Journal of Molecular Liquids

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

Atomic interactions between rock substrate and water-sand mixture


with and without graphene nanosheets via molecular
dynamics simulation
Amirhosein Mosavi a,b, Maboud Hekmatifar c, Davood Toghraie c, Roozbeh Sabetvand d, As’ad Alizadeh e,f,
Zahra Sadeghi g, Aliakbar Karimipour h,⁎
a
Environmental Quality, Atmospheric Science and Climate Change Research Group, Ton Duc Thang University, Ho Chi Minh City, Vietnam
b
Faculty of Environment and Labour Safety, Ton Duc Thang University, Ho Chi Minh City, Vietnam
c
Department of Mechanical Engineering, Khomeinishahr Branch, Islamic Azad University, Khomeinishahr, Iran
d
Department of Energy Engineering and Physics, Department of Energy Engineering and Physics, Amirkabir University of Technology, Tehran, Iran
e
Department of Mechanical Engineering, Urmia University, Urmia, Iran
f
Department of Mechanical Engineering, College of Engineering, University of Zakho, Zakho, Iraq
g
Department of Chemical Engineering, Isfahan University of Technology, Isfahan, Iran
h
Institute of Research and Development, Duy Tan University, Da Nang 550000, Vietnam.

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

Article history: In this work, we describe the atomic effects of graphene nanosheets adding to the hydraulic fracturing process by
Received 8 September 2020 molecular dynamics method. In our simulations, we study the nanosheets type effect on the fracturing process.
Received in revised form 6 October 2020 For this purpose, we reported physical parameters such as temperature, potential energy, joint force, and the
Accepted 16 October 2020
number of lost atoms from atomic substrates. In our approach, nanofluid, which used in hydraulic fracturing, is
Available online 20 October 2020
exactly simulated by various interatomic force fields. Our simulations show that adding graphene nanoparticles
Keywords:
to the first fluid causes maximum atomic removal from simulated rock substrate. Numerically, after 2.5 ns, the
Hydraulic fracturing process departed atoms from rock substrate reach to 74 atoms.
Molecular dynamic simulation Furthermore, the atomic rate of graphene nanosheets is another important parameter in hydraulic fracturing. Our
Nanosheet molecular dynamics results show that, by 5% atomic rate of graphene nanosheets to initial fluid, the departed
Graphene atoms reach to the maximum rate (101 atoms). Finally, by adding graphene nanosheets with 4 nm length, the
Armchair edge top quality of particles departed from the rock substrate.
Zig-zag edge © 2020 Elsevier B.V. All rights reserved.

1. Introduction tight gas and shale gas [4]. In certain veins or dikes, naturally, some
hydraulic fractures can form [5]. The United States changed to a most
Hydraulic fracturing is a standard and appropriate method contain- important exporter of crude oil in 2019 by Drilling and hydraulic frac-
ing the fracturing of bedrock formations. This procedure consists of sand turing [6], but the leakage of methane, has dramatically enhanced [7].
or other proppants suspended with the aid of thickening agents, includ- Enriched gas and oil production from the decade long fracking boom
ing the high-pressure injection of fracking fluid, which conclude of has flowed to lower costs for consumers [8,9].
water, brine and petroleum will flow more freely as depicted in Fig. 1. Elements of fracking discuss that these are outweighed by the envi-
Small grains of hydraulic fracturing fractions hold the fractures created ronmental influence, which concludes noise and air pollution, and
when the hydraulic pressure was removed [1]. Historically, the hydrau- groundwater and surface water contamination, along with the resulting
lic fracturing process began in 1947. About 2.5 million performances hazards to the environment and public health [10–13]. But these harm-
had been performed worldwide on gas and oil wells since 2012. Over ful effects can vanish with optimized of the hydraulic fracturing process.
1.000.000 of those within the united states of America [2,3]. Such treat- Wigwe et al. [14] research about the influences of fluid viscosities, injec-
ment is essential to obtain a flow rate in coal seam gas wells, tight oil, tion rates on hydraulic fracturing parameters and proppant densities.
These researchers express that the injection rate of 13 bpm yielded
⁎ Corresponding author.
the propped, useful prop and longest fracture length for both fractions.
E-mail addresses: amirhosein.mosavi@tdtu.edu.vn (A. Mosavi), The low-density proppant produced the longest effective fracture
aliakbarkarimipour@duytan.edu.vn (A. Karimipour). length while the high-density proppant gave the longer fracture length.

https://doi.org/10.1016/j.molliq.2020.114610
0167-7322/© 2020 Elsevier B.V. All rights reserved.
A. Mosavi, M. Hekmatifar, D. Toghraie et al. Journal of Molecular Liquids 323 (2021) 114610

work, we study the graphene nanosheets with armchair and zig-zag


Nomenclature
edges effects on the hydraulic fracturing process with molecular dy-
namics (MD) simulations. MD approach is one of the significant sorts
Fij interparticle force on particle i by particle j (eV/Å)
of computer simulations that can be predicting the atomic manner of
Vij interparticle potential (eV)
various structures [27–32].
Fext external applied force (eV/Å)
m particle mass (u)
2. Computational method
rc cut-off distance (Å)
rij distance between particle i and j (Å)
In MD simulations, atoms are permitted to interact together and give
t time step (ps)
a view of the dynamic changes of the overall atomic system. In this
T temperature (K)
paper, all molecular dynamics simulations are done by using LAMMPS
v velocity of particle (Å/ps)
software presented by Sandia National Laboratories [33–36]. To using
Natom number of particles
LAMMPS software to measure the atomic behaviour of water/sand mix-
ture in the vicinity of rock substrate, rock substrate with Silicon, Iron,
Greek symbols
Aluminum, Oxygen, Sodium, Calcium, Potassium, and Magnesium
ε energy parameter in Lennard-Jones (LJ) potential
atoms modelled by coarse-graining method and fixed in the bottom of
σ length parameter of LJ potential
the MD box, while water and atomic sand structures filled the sur-
ϕ interaction potential
rounding position of volume. Computationally, sand particles with the
θ0 Equilibrium angle
coarse-graining method described as one particle. This sand model con-
sists of atomic properties of each constituent atoms such as atomic
mass, atomic interaction, etc. Fig. 2 displays the simulation package at
Cao et al. [15] used implementing a different of the simultaneous per- the perspective faces, top and front, which were visualized by Open Vi-
turbation stochastic approximation algorithm to optimizing the hy- sualization Tool [37]. The first atomic position prepared by Packmol
draulic fracturing design. atomic modeling software [38].
Furthermore, Jahandideh et al. [16] optimized the hydraulic fractur- Technically, in this molecular dynamics simulations, periodic
ing design to maximize the net present value of the shale asset. Using of boundary conditions were used in x and y directions and fix one im-
nanotechnology in various processes leads to their improvement and plemented in the z-direction. For creating the initial temperature of
development [17]. Nanotechnology is another effective way to optimize simulated structures, the NVT ensemble used in equilibrates the ini-
the hydraulic fracturing process. Nanotechnology is the manipulation of tial temperature of atomic structures at 300 K with dt = 1 fs time
materials on a nano-metric (atomic) scale. The earliest, common report step and 0.01 temperature damping rate. This rate of temperature
of nanotechnology referred to the specific technological aim of precisely damping appropriate with the previous report [39]. After 2.5 ns, the
manipulating atoms for the fabrication of macroscale structures, also simulated structure was equilibrated at the beginning temperature,
now referred to as molecular nanotechnology [18,19]. Carbon Nano- then the NVT effect cancelled, and NVE one was used to the time evo-
tubes, Graphene, Graphene quantum dot, and Graphene nanosheet are lution of simulated structures. To simulate the rock substrate, we use
some of the important atomic structure in nanotechnology [20,21]. Universal Force Field (UFF) [40]. In this force field were shown as a
These compounds have many applications in electrochemical sensors, superposition of non-bonded and bonded forces-non-bond interac-
electromagnetic shields, energy storage, improvement in electrical con- tion between atoms described by the Lennard-Jones (LJ) potential.
ductivity, boiling, hydraulic fracturing process, etc. [22–26]. In this The interaction between a pair of atoms approximates by the LJ po-
tential formula. This interatomic potential was proposed by John
Lennard-Jones in 1924 [41]. This simple atomic interaction stated
as below:
"   6 #
σ 12 σ
U ðrÞ ¼ 4ε − r ≪ rc ð1Þ
rij r ij

Where ε is the depth of the potential well, σ is the distance at which


the potential is zero; rij is the distance between the 2 atoms. In MD sim-
ulations, both σ and ε parameters depend upon the sort of atoms in the
simulation box. The length scale parameter and energy for different
atoms in atomic substrate simulation are listed in Table 1 [40].
The bonded forces consist of bond angle bend and bond strength
terms. Harmonic oscillator equations calculate the angle and bond
strength stretch in UFF force-field as below:

1
Er ¼ kr ðr−r 0 Þ ð2Þ
2

1
Eθ ¼ kθ ðθ−θ0 Þ ð3Þ
2

In these formulas, Kθ and Kr are harmonic oscillator constants. θ0 and


r0 are the atomic bond length and equilibrium value of angles. For H2O
molecules, we used the SPC model in our MD simulations. The interac-
tion between water molecules is simulated using an LJ potential.
These interaction parameters for atomic water structure reported in
Fig. 1. The schematic of the common hydraulic fracturing process. below [42,43]:

2
A. Mosavi, M. Hekmatifar, D. Toghraie et al. Journal of Molecular Liquids 323 (2021) 114610

Oxygen mass = 15.9994 u.


Hydrogen mass = 1.008 u.
Oxygen charge = −0.820.
Hydrogen charge = 0.410.
LJ interaction epsilon for Oxygen-Oxygen = 0.1553 kcal/mol.
LJ interaction sigma for Oxygen-Oxygen = 3.166 Å.
LJ interaction epsilon and sigma of Oxygen- Hydrogen and Hydrogen
- Hydrogen = 0.0.
Equilibrium bond length of Oxygen- Hydrogen = 1.0 Å.
Equilibrium angle of Hydrogen -Oxygen- Hydrogen = 109.47°.
Furthermore, graphene nanosheets in our MD simulations described
by Tersoff potential [44,45]. The most common expression of the Tersoff
potential represented as below,

1
E¼ ∑∑V ð4Þ
2 i i≠j ij

     
V ij ¼ f C r ij f R r ij þ bij f A r ij ð5Þ

In Eq. (5), fR is a two-body term, and fA includes three-body interac-


tions. The summations in the formula are overall neighbours j and k of
atom i within a cut-off distance Rc = R + D. Furthermore, the
potential between the particles V(rN) is assumed for each pair of parti-
cles. Therefore, for N particles,
   
V r N ¼ ∑∑V r ij ð6Þ
i<j

Previous reports show that the UFF force field can be used for various
structures and elements of the periodic table [30]. Further, Tersoff po-
tential is an appropriate choice for carbon nanostructures such as
graphene nanosheets, carbon nanotube, etc. [44,45]. To calculate the
particle movements by simulation time, Newton's second law's equa-
tion in atomic level is applied as the gradient of interatomic potential
in below equations,

2
d ri dvi
F i ¼ ∑F ij ¼ mi 2
¼ mi ð7Þ
i≠j dt dt

F ij ¼ −grad V ij ð8Þ

From the above equations, the momentum Pi can be defined as fol-


low,

P i ¼ mi vi ð9Þ

So, Total energy (E) of the atomic system can be displayed in the for-
mation of Hamilton as below equation,

1
Hðr, P Þ ¼ ∑P 2 þ V ðr 1 þ r 2 þ . . . þ r n Þ ¼ E ð10Þ
2m i i
Fig. 2. Schematic of rock substrate and water /sand mixture, which provided with Packmol
package at a) Front, b) Top, and c) Perspective views.
Furthermore, in MD simulations, Gaussian distribution is used for
calculating the temperature of particles that are shown in the following
Table 1 formula:
The length scale and energy parameters for LJ interaction in simulated structures [40].
3 1 N 1
Atom ε (kcal/mol) σ (Å) kB T ¼ ∑ mv2i ð11Þ
2 Natom i¼1 2
Oxygen 0.060 3.500
Silicon 0.402 4.295
Aluminum 0.505 4.499 The instantaneous temperature fluctuates obtained from the below
Iron 0.013 2.912 equation,
Calcium 0.238 3.399
Sodium 0.030 2.983 N mi v2i ðt Þ
Potassium 0.035 3.812 T ðt Þ ¼ ∑ ð12Þ
i kB Nsf
Magnesium 0.111 3.021

3
A. Mosavi, M. Hekmatifar, D. Toghraie et al. Journal of Molecular Liquids 323 (2021) 114610

-3 0 0
Where Nsf is the degree of freedom of the atomic structures. Finally,
we can say that MD simulations of this work carried out in these two
steps:
Step A: rock substrate with water/sand mixture and graphene nano- -4 0 0
particles were simulated at T = 300 K with 1 fs time step. The simula-

Potential Energy (eV)


tion package in our molecular dynamics calculations has 100*50*75 Å3
length in x, y, and z directions, respectively. In the temperature equili-
-5 0 0
bration process, we use the NVT ensemble for 2.5 ns [46,47].
Step B: In the second step of our MD simulation, the atomic interac-
tion between substrate and defined atomic mixtures was carried out for
2.5 ns later. For analyzing the interaction between these structures, -6 0 0
physical parameters such as mutual atomic force and the number of de-
parted atoms from rock substrate were reported.

-7 0 0
0 1 2 3
3. Discussion and results T im e (n s )

3.1. Atomic equilibration of simulated structures Fig. 4. Potential energy variation of simulated structures with/without graphene
nanosheets as a function of molecular dynamics simulation time.
In the first step of this molecular dynamics simulations, the atomic
structure of base fluid (consist of water molecules and sand particles)
was studied at T = 300 K. Fig. 2 shows the atomic arrangement of
atomic structures in the simulation box. Our MD results in the equilib-
rium process showed that the initial position of atoms in the hydraulic improve the water /sand mixture mechanical properties. By this im-
fracturing process is adopted with UFF and Tersoff force-fields. Numer- provement, we predict that the hydraulic fracturing process done effec-
ically, structures atomic stability described by reporting temperature tively by adding graphene nanoparticles to water/sand mixture.
and potential energy of them. Total temperature variation of the simu-
lated structure depicted in Fig. 3. From this Fig, we can say that all struc- 3.2. Dynamical evolution of simulated structures
tures equilibrated after 2.5 ns. Fig. 4 displays the potential energy of
atomic structures as a function of MD simulation time. According to After initial molecular dynamics equilibration and simulations pro-
this Fig, we can say that the potential energy of structures converged cess of simulated structure, the NVE ensemble implemented for 2.5 ns
after 2.5 ns, too. Numerically, the potential energy of simulated struc- later (see Fig. 5). Then, the joint force of two groups of atomic structures
ture reaches to −416 eV after 2.5 ns. From our MD result, we can say (group A: water molecules and sand particles - group B: rock substrate)
that simulated structures mean distance increases after the equilibra- calculated. Fig. 6 shows that the mutual force between these structures
tion process and reach to atomic stability after 2.5 ns. By adding varies from 10.11 eV/ Å to 13.57 eV/ Å by MD simulation time passing.
graphene nanosheets to base structures by a 1% atomic rate, the simu- Physically, the non-zero rate of the joint force rate of these structures
lated structures don't disrupt, and so, the equilibration process can be shows that the water/sand mixture and atomic substrate distance is
detected after 2.5 ns. From Fig. 3, we can express that the temperature lesser than the cut off the radius of them at 300 K, and these structures
of atomic structures in the presence of armchair and zig-zag nanosheets interacted with each other. By adding graphene nanosheets by a 1%
converged to 300 K. furthermore, atomic stability of simulated struc- atomic rate, the mutual force of these structures rises. Numerically, by
tures after adding graphene nanostructures rises. Numerically, by adding armchair and zig-zag graphene nanosheets, the mutual force of
adding armchair and zig-zag graphene nanosheets to water molecules initial structures increases from 13.57 eV/ Å to 14.41 eV/ Å and
and sand particles, the potential energy of these structures increases 16.45 eV/ Å, respectively (Table 2). This atomic behaviour displays
to −599 eV and − 559 eV from −416 eV. By potential energy increases, that, by adding graphene nanoparticles to initial structure, the hydraulic
the atomic stability of structures rises, and so these nanostructures fracturing process can occur effectively, and so nanoparticles are the
positive parameter in this mechanical procedure. Furthermore, we can
say that using armchair graphene nanosheets has more atomic effects
350
on the substrate, and by using this nanostructure, the hydraulic fractur-
ing process done more effectively.
The number of separated atoms from a simulated rock substrate is
proportional to its stability. Our MD simulations show that by MD sim-
300 ulation time passing, the number of these atoms increases (see Fig. 7).
Temperature (K)

Numerically, after 2.5 ns, the number of departed atoms increases to


57 rates. By adding 1% graphene nanosheets to water/sand mixture,
the departing atoms from substrate increases. By adding armchair and
zig-zag graphene nanosheet to the original mix, the departed atoms
250
W ith o u t G ra p he ne
rate after 2.5 ns increases to 63 and 74 rates, respectively (Table 3).
W ith G ra p h e n e -A rm c h a ir From these calculations, we can say that adding graphene nanosheets
W ith G ra p h e n e -Z ig z a g
to initial water/sand mixture cause improve this atomic structure me-
chanical strength. Furthermore, nanoparticles adding to atomic struc-
tures cause an increase in the atomic interaction between two groups
200 of atoms. So, the more number of atomic substrate interacts with
0 1 2 3
T im e(ns ) water/sand mixture, and more number of rock substrate atoms de-
parted of them.
Fig. 3. Temperature variation of simulated structures with/without graphene nanosheets The atomic rate of graphene nanosheets, which added to our simu-
as a function of molecular dynamics simulation time. lated structures, has an essential effect on the atomic manner of

4
A. Mosavi, M. Hekmatifar, D. Toghraie et al. Journal of Molecular Liquids 323 (2021) 114610

15
water/sand and substrate. In this section of our computational study, we
describe this subject by reporting departed atoms rate from the sub-
strate as a function of the graphene nanosheet atomic rate. In this sec-
tion, we add armchair graphene nanosheets by 1%, 2.5%, 5%, and 10%

Mutual Force (eV/Angstrom)


atomic rates to water/sand mixture and calculate the departed atoms 12
from rock substrate (Fig. 8). From Table 3 and Fig. 9, we can conclude

9 W ithout G raphene
W ith G raphene-A rm chair
W ith G raphene-Z igzag

0 1 2 3
Tim e (ns)

Fig. 6. Mutual force variation of water/sand mixture and rock substrate with/without
graphene nanosheets as a function of MD simulation time.

Table 2
Mutual force variation of water /sand mixture and rock substrate with/without
graphene nanosheets after 2.5 ns.

Atomic sample Mutual force (eV/ Å)

Without Graphene 13.57


With Armchair Graphene 14.41
With Zigzag Graphene 16.45

that, by 5% graphene nanosheet adding to water/sand mixture, the max-


imum number of departed rock atoms detectable. Numerically, by
adding graphene nanostructures to water /sand mixture by 1%, 2.5%,
5%, and 10%, the number of departed rock atoms from atomic substrate

80

60
Number of Atoms

40

20
W ith o u t G r a p h e n e
W ith G r a p h e n e - A r m c h a ir
W ith G r a p h e n e - Z ig z a g

0
0 1 2 3
T im e ( n s )

Fig. 7. The number of departed atoms from rock substrate with/without graphene
nanosheets as a function of MD simulation time.

Table 3
The number of departed atoms from rock substrate with/without graphene nano-
sheets after 2.5 ns.

Atomic sample Number of lost atoms

Without Graphene 57
With Armchair Graphene 63
With Zigzag Graphene 74
Fig. 5. Time evolution of rock substrate after a) 0, b) 1 ns and c) 2.5 ns.

5
A. Mosavi, M. Hekmatifar, D. Toghraie et al. Journal of Molecular Liquids 323 (2021) 114610

Fig. 8. Water /sand mixture with a) 1%, b) 2.5%, c) 5%, and d) 10% graphene nanosheets.

reaches 74, 89, 101, and 91 rates, respectively. So, we can say that the Table 4
The number of departed atoms from rock substrate as a function of armchair
hydraulic fracturing process reaches to maximum efficiency by 5%
graphene nanosheet atomic rate after 2.5 ns.
nanostructure adding to the initial atomic arrangement. Practically, by
using the reported results from this section of our MD calculations, Nanoparticle rate (%) Number of lost atoms
one can optimize the hydraulic fracturing process in actual cases 1 74
(Table 4). 2.5 89
The size of graphene nanosheets is another important parameter on 5 101
10 91
the dynamical manner of water /sand mixture and atomic substrate. In
the last section of this molecular dynamics simulations, we estimate this
atomic parameter effect by calculating the number of departed particles Table 5
from the substrate as a function of the graphene nanosheet size. For The number of departed atoms from rock substrate as a function of graphene
doing this procedure, we add armchair graphene nanosheets by 20 Å, nanosheet size.
30 Å, 40 Å, and 50 Å size in x and y directions to water /sand mixture. Nanoparticle size (Å) Number of lost atoms
From Table 5 and Fig. 10, we can say by graphene nanosheet with
20 74
40 Å length adding to simulated structure, the maximum number 30 82
of departed particles detectable. Numerically, by adding graphene 40 93
50 86

120
100

90 90
Number of Atoms

Number of Atoms

80
60

70

30 A r m c h a ir - 1 %
A r m c h a ir - 2 .5 %
60
A r m c h a ir - 5 %
A r m c h a ir - 1 0 %

0 50
0 1 2 3 20 30 40 50 60
T im e s ( n s ) G r a p h e n e L e n g th ( Å )

Fig. 9. The number of departed atoms from rock substrate as a function of armchair Fig. 10. The number of departed atoms from rock substrate as a function of graphene
graphene nanosheet atomic rate. nanosheet size.

6
A. Mosavi, M. Hekmatifar, D. Toghraie et al. Journal of Molecular Liquids 323 (2021) 114610

nanostructures with various sizes (20 Å, 30 Å, 40 Å, and 50 Å) to simu- [4] D. Blundell, N. Arndt, P.R. Cobbold, C. Heinrich, 9: processes of tectonism,
magmatism and mineralization: lessons from Europe, Ore Geol. Rev. 27 (1–4)
lated sand /water mixture by 1%, atomic rate, the number of departed (2005) 333–349.
particles from rock substance reaches to 74, 82, 93, and 86 particles, [5] D.T. Allen, et al., Measurements of methane emissions at natural gas production sites
respectively. in the United States, Proc. Natl. Acad. Sci. 110 (44) (2013) 17768–17773.
[6] C.D. Kassotis, D.E. Tillitt, J.W. Davis, A.M. Hormann, S.C. Nagel, Estrogen and andro-
gen receptor activities of hydraulic fracturing chemicals and surface and ground
4. Conclusion water in a drilling-dense region, Endocrinology 155 (3) (2014) 897–907.
[7] S. Chalupka, Occupational silica exposure in hydraulic fracturing, Workplace Health
Safe 60 (2012) 460.
The present study investigates the atomic interaction between rock
[8] V.J. Brown, Industry Issues: Putting the Heat on Gas, National Institute of Environ-
substrate and water /sand mixture with/without graphene nanostruc- mental Health Sciences, 2007.
tures via molecular dynamics simulations. In our molecular dynamics [9] V.J. Brown, Radionuclides in Fracking Wastewater: Managing a Toxic Blend, National
simulations, rock substrate is modelled by a coarse-graining method. Institute of Environmental Health Sciences, 2014.
[10] A.M. Bamber, et al., A systematic review of the epidemiologic literature assessing
Also, to simulate of interatomic force, Universal Force Field and Tersoff health outcomes in populations living near oil and natural gas operations: study
potential have been implemented to various atomic structures. The re- quality and future recommendations, Int. J. Environ. Res. Public Health 16 (12)
sult of our simulations shows that graphene nanostructures have a mea- (2019) 2123.
[11] R. Wright, R.D. Muma, High-volume hydraulic fracturing and human health out-
surable effect on optimizing the hydraulic fracturing process, which can comes: a scoping review, J. Occup. Environ. Med. 60 (5) (2018) 424–429.
be used for petroleum applications. Generally, our computational out- [12] H. Karimi-Maleh, M. Shafieizadeh, M.A. Taher, F. Opoku, E. Muriithi Kiarii, P. Penny
comes from molecular dynamics simulations are as following: Govender, S. Ranjbari, M. Rezapour, Y. Orooji, The role of magnetite/graphene
oxide nano-composite as a high-efficiency adsorbent for removal of
phenazopyridine residues from water samples, an experimental/theoretical investi-
• Universal Force Field and Tersoff potentials are the appropriate gation, J. Mol. Liq. 298 (2020), 112040, .
interatomic potential for molecular dynamics simulation of defined [13] H. Karimi-Maleh, B.G. Kumar, S. Rajendran, J. Qin, S. Vadivel, D. Durgalakshmi, M.
structures. Numerically, the potential energy of our simulated struc- Soto-Moscoso, Y. Orooji, F. Karimi, Tuning of metal oxides photocatalytic perfor-
mance using Ag nanoparticles integration, J. Mol. Liq. 314 (2020), 113588, .
tures converged to −416 eV and − 599 eV after 2.5 ns at 300 K [14] M. Wigwe, O. Kolawole, M. Watson, I. Ispas, W. Li, Influence of fracture treatment
with and without graphene nanosheets, respectively. parameters on hydraulic fracturing optimization in unconventional formations,
• The mutual force between rock substrate and H2O/sand mixture in- ARMA-CUPB Geothermal International Conference, American Rock Mechanics Asso-
ciation, 2019.
creases by graphene nanosheets adding to the MD simulation box. [15] K. Cao, P. Siddhamshetty, Y. Ahn, R. Mukherjee, J.S.-I. Kwon, Economic model-based
Numerically, the defined joint force in our simulated structures controller design framework for hydraulic fracturing to optimize shale gas produc-
reached to 13.57 eV/ Å and 16.45 eV/ Å after 2.5 ns with and without tion and water usage, Ind. Eng. Chem. Res. 58 (27) (2019) 12097–12115.
[16] A. Jahandideh, B. Jafarpour, Optimization of hydraulic fracturing design under spa-
graphene nanosheets, respectively.
tially variable shale fracability, J. Pet. Sci. Eng. 138 (2016) 174–188.
• The departed atoms from rock substrate increase by graphene nano- [17] Y.-l. Huang, S.-h. Shi, Phylogenetics of Lythraceae sensu lato: a preliminary analysis
sheet, adding to H2O/sand mixture. Numerically, the number of de- based on chloroplast rbc L gene, psa A-ycf 3 spacer, and nuclear rDNA internal tran-
parted atoms in the hydraulic fracturing process increased to 74 scribed spacer (ITS) sequences, Int. J. Plant Sci. 163 (2) (2002) 215–225.
[18] A.W. Hübler, O. Osuagwu, Digital quantum batteries: energy and information stor-
atoms from 57 atoms with a 1% graphene nanosheet presence. age in nanovacuum tube arrays, Complexity 15 (5) (2010) 48–55.
• Graphene nanosheets with an armchair edge have more impact on [19] E. Shinn, A. Hübler, D. Lyon, M.G. Perdekamp, A. Bezryadin, A. Belkin, Nuclear energy
the hydraulic fracturing process rather than zig-zag edge one. conversion with stacks of graphene nanocapacitors, Complexity 18 (3) (2013)
24–27.
• By adding graphene nanosheet with a 5% atomic ratio to water /sand [20] H. Karimi-Maleh, F. Karimi, M. Alizadeh, A.L. Sanati, Electrochemical sensors, a
mixture, the hydraulic fracturing process optimized and the number bright future in the fabrication of portable kits in analytical systems, Chem. Rec.
of departed atoms from rock substrate reach to the maximum rate. 20 (2019) 682–692.
[21] H. Karimi-Maleh, K. Cellat, K. Arıkan, A. Savk, F. Karimi, and F. Şen, "Palladium–nickel
• The maximum rate of departed atoms from rock substrate (93 parti-
nanoparticles decorated on functionalized-MWCNT for high precision non-
cles) occurs by adding graphene nanoparticles with 40 Å length. enzymatic glucose sensing," Mater. Chem. Phys. 123042, 2020.
[22] Y. Orooji, et al., Preparation of mullite-TiB2-CNTs hybrid composite through spark
plasma sintering, Ceram. Int. 45 (13) (2019) 16288–16296.
CRediT authorship contribution statement [23] Y. Orooji, M.R. Derakhshandeh, E. Ghasali, M. Alizadeh, M.S. Asl, T. Ebadzadeh, Ef-
fects of ZrB2 reinforcement on microstructure and mechanical properties of a
spark plasma sintered mullite-CNT composite, Ceram. Int. 45 (13) (2019)
Amirhosein Mosavi: Writing - review & editing, Writing - original 16015–16021.
draft, Investigation. Maboud Hekmatifar: Methodology, Software, [24] H. Karimi-Maleh, F. Karimi, Y. Orooji, G. Mansouri, A. Razmjou, A. Aygun, F. Sen, A
Validation, Writing - original draft. Davood Toghraie: Methodology, new nickel-based co-crystal complex electrocatalyst amplified by NiO dope Pt nano-
structure hybrid; a highly sensitive approach for determination of cysteamine in the
Software, Validation, Writing - original draft, Investigation. Roozbeh presence of serotonin, Sci. Rep. 10 (1) (2020) 1–13.
Sabetvand: Methodology, Software, Validation, Writing - original [25] Y. Peng, M. Zarringhalam, M. Hajian, D. Toghraie, S.J. Tadi, M. Afrand, Empowering
draft. As’ad Alizadeh: Writing - review & editing. Zahra Sadeghi: Writ- the boiling condition of Argon flow inside a rectangular microchannel with
suspending Silver nanoparticles by using of molecular dynamics simulation,
ing - review & editing, Writing - original draft. Aliakbar Karimipour:
CJournal Mol. Liquids 295 (2019) 111721.
Writing - review & editing, Writing - original draft, Investigation. [26] H. Karimi-Maleh, O.A. Arotiba, Simultaneous determination of cholesterol, ascorbic
acid and uric acid as three essential biological compounds at a carbon paste elec-
Declaration of Competing Interest trode modified with copper oxide decorated reduced graphene oxide nanocompos-
ite and ionic liquid, J. Colloid Interface Sci. 560 (2020) 208–212.
[27] N.A. Jolfaei, et al., Investigation of thermal properties of DNA structure with precise
The authors declare that they have no known competing financial atomic arrangement via equilibrium and non-equilibrium molecular dynamics ap-
interests or personal relationships that could have appeared to influ- proaches, Comput. Methods Prog. Biomed. 185 (2020), 105169, .
[28] D. Toghraie, M. Mokhtari, M. Afrand, Molecular dynamic simulation of copper and
ence the work reported in this paper. platinum nanoparticles Poiseuille flow in a nanochannels, Physica E 84 (2016)
152–161.
[29] A. Razmjou, G. Eshaghi, Y. Orooji, E. Hosseini, A. Habibnejad Korayem, F.
Mohagheghian, Y. Boroumand, A. Noorbakhsh, M. Asadnia, V. Chen, Lithium ion-
References selective membrane with 2D subnanometer channels, Water Res. 159 (2019)
313–323.
[1] L. Gandossi, U. Von Estorff, An overview of hydraulic fracturing and other formation [30] A. Mosavi, M. Hekmatifar, A. Alizadeh, D. Toghraie, R. Sabetvand, A. Karimipour, The
stimulation technologies for shale gas production, Eur. Commisison Jt. Res. Cent. molecular dynamics simulation of thermal manner of Ar/Cu nanofluid flow: the ef-
Tech. Reports 26347 (2013). fects of spherical barriers size, J. Mol. Liquids. 114183 (2020).
[2] T. Colborn, C. Kwiatkowski, K. Schultz, M. Bachran, Natural gas operations from a [31] D. Toghraie, M. Hekmatifar, Y. Salehipour, M. Afrand, Molecular dynamics simula-
public health perspective, Hum. Ecol. Risk Assess. 17 (5) (2011) 1039–1056. tion of Couette and Poiseuille Water-Copper nanofluid flows in rough and smooth
[3] L. B.-E. E. MAP, Methane in Pennsylvania Water Wells Unrelated to Marcellus Shale nanochannels with different roughness configurations, Chem. Phys. 527 (2019),
Fracturing, 2011. 110505, .

7
A. Mosavi, M. Hekmatifar, D. Toghraie et al. Journal of Molecular Liquids 323 (2021) 114610

[32] M. Hekmatifar, et al., The study of asphaltene desorption from the iron surface with [40] A.K. Rappé, C.J. Casewit, K. Colwell, W.A. Goddard III, W.M. Skiff, UFF, a full periodic
molecular dynamics method, J. Mol. Liq. 318 (2020), 114325, . table force field for molecular mechanics and molecular dynamics simulations, J.
[33] W.M. Brown, A. Kohlmeyer, S.J. Plimpton, A.N. Tharrington, Implementing molecular Am. Chem. Soc. 114 (25) (1992) 10024–10035.
dynamics on hybrid high performance computers–Particle–particle particle-mesh, [41] J.E. Jones, On the determination of molecular fields.—II. From the equation of state of
Comput. Phys. Commun. 183 (3) (2012) 449–459. a gas, Proc. Royal Soc. Lond. 106 (738) (1924) 463–477.
[34] W.M. Brown, P. Wang, S.J. Plimpton, A.N. Tharrington, Implementing molecular dy- [42] K. Toukan, A. Rahman, Molecular-dynamics study of atomic motions in water, Phys.
namics on hybrid high performance computers–short range forces, Comput. Phys. Rev. B 31 (5) (1985) 2643.
Commun. 182 (4) (2011) 898–911. [43] H. Berendsen, J. Grigera, T. Straatsma, The missing term in effective pair potentials, J.
[35] S. Plimpton, Fast parallel algorithms for short-range molecular dynamics, J. Comput. Phys. Chem. 91 (24) (1987) 6269–6271.
Phys. 117 (1) (1995) 1–19.
[44] J. Tersoff, Modeling solid-state chemistry: interatomic potentials for multicompo-
[36] R.M. Mukherjee, P.S. Crozier, S.J. Plimpton, K.S. Anderson, Substructured molecular
nent systems, Phys. Rev. B 39 (8) (1989) 5566.
dynamics using multibody dynamics algorithms, Int. J. Non-Linear Mech. 43 (10)
(2008) 1040–1055. [45] J. Tersoff, New empirical approach for the structure and energy of covalent systems,
[37] A. Stukowski, Visualization and analysis of atomistic simulation data with OVITO– Phys. Rev. B 37 (12) (1988) 6991.
the Open Visualization Tool, Model. Simul. Mater. Sci. Eng. 18 (1) (2009) 015012. [46] S. Nosé, A unified formulation of the constant temperature molecular dynamics
[38] L. Martínez, R. Andrade, E.G. Birgin, J.M. Martínez, PACKMOL: a package for building methods, J. Chem. Phys. 81 (1) (1984) 511–519.
initial configurations for molecular dynamics simulations, J. Comput. Chem. 30 (13) [47] W.G. Hoover, Canonical dynamics: equilibrium phase-space distributions, Phys. Rev.
(2009) 2157–2164. A 31 (3) (1985) 1695.
[39] A.Z. Ashkezari, et al., Calculation of the thermal conductivity of human serum albu-
min (HSA) with equilibrium/non-equilibrium molecular dynamics approaches,
Comput. Methods Prog. Biomed. 188 (2020), 105256, .

You might also like