Solvation structure of supercritical CO2 and brine mixture for CO2 plume geothermal applications: A molecular dynamics study

Solvation structure of supercritical CO2 and brine mixture for CO2 plume geothermal applications: A molecular dynamics study

Journal Pre-proof Solvation structure of supercritical CO2 and brine mixture for CO2 plume geothermal applications: A molecular dynamics study Zonglia...

4MB Sizes 0 Downloads 57 Views

Journal Pre-proof Solvation structure of supercritical CO2 and brine mixture for CO2 plume geothermal applications: A molecular dynamics study Zongliang Qiao, Yue Cao, Yuming Yin, Lingling Zhao, Fengqi Si

PII:

S0896-8446(20)30034-6

DOI:

https://doi.org/10.1016/j.supflu.2020.104783

Reference:

SUPFLU 104783

To appear in:

The Journal of Supercritical Fluids

Received Date:

12 July 2019

Revised Date:

5 February 2020

Accepted Date:

5 February 2020

Please cite this article as: Qiao Z, Cao Y, Yin Y, Zhao L, Si F, Solvation structure of supercritical CO2 and brine mixture for CO2 plume geothermal applications: A molecular dynamics study, The Journal of Supercritical Fluids (2020), doi: https://doi.org/10.1016/j.supflu.2020.104783

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2020 Published by Elsevier.

Solvation structure of supercritical CO2 and brine mixture for CO2 plume geothermal applications: A molecular dynamics study Zongliang Qiao*, Yue Cao, Yuming Yin, Lingling Zhao, Fengqi Si Key Laboratory of Energy Thermal Conversion and Control of Ministry of Education,

ro of

Southeast University, Nanjing, 210096, China

* Corresponding author: Zongliang Qiao*

Key Laboratory of Energy Thermal Conversion and Control of Ministry of Education,

E-mail address: [email protected]

lP

Jo

ur

na

Graphical abstract

re

Tel: +86 13770829247

-p

Southeast University, Nanjing, 210096, China

Highlights 

Solvation structure of scCO2/brine mixture produced by CPG was investigated.



Molecular dynamics simulations were performed for different compositions.



Temperature and ion concentration affected the interaction of H2O and ion.



The scCO2/brine mixture was heterogeneous in the CPG application.

Abstract: This paper reports an investigation for solvation structure of supercritical CO2 (scCO2) and brine mixture for CO2 plume geothermal (CPG) applications. Molecular dynamics simulations are performed to calculate radial distribution function, coordination number and hydrogen-bond number for mixtures of different compositions. This study can help reveal the solvation structures of scCO2/brine

ro of

mixture under different temperatures (373.15 K to 473.15 K) and ion concentrations (1 wt% to 10 wt%). Results show that H2O molecules bind more tightly to Na+ ion

than Cl- ion while scCO2 molecules seem have a uniform distribution in the

-p

scCO2/brine mixture. Both the interaction of Na+-H2O pairs and that of Cl--H2O pairs

re

become stronger under higher temperature conditions. Besides, the increase of ion concentration not only reduces the number of H2O molecules in the solvation shell of

lP

Na+ ion, but also weakens the interaction of hydrogen bonds. Moreover, it seems that

na

the scCO2/brine mixture is heterogeneous in the CPG application.

ur

KEYWORDS: Molecular dynamics simulation; scCO2/brine mixture; Solvation

Jo

structure; Ion pair; CO2 plume geothermal.

Nomenclature

g(r)

radial distribution function

k

fraction of CO2 molecules

N

number of particles

number of hydrogen bonds

n(r)

coordination number

P(r)

probability

q

charge of particle (C)

r

distance (m)

r1

first minimum of RDF (m)

Tc,model

model critical temperature (K)

Tr,sim

reduced simulation temperature

Tsim

simulation temperature (K)

V

volume (m3)

Vc

Coulomb term of non-bonded interaction energy (J)

VLJ

Lennard-Jones term of non-bonded interaction energy (J)

Vij

non-bonded interaction energy (J)

XNaCl

NaCl mass fraction

Jo



-p

vaccum permittivity (C2 N-1 m-2)

ur



lP

re

0

na

Greek letters



parameter of Lennard-Jones potential probability density parameter of Lennard-Jones potential

Subscripts A

solvent atom A

a

anion

ro of

nHB

solute ion B

c

cation

Hw

hydrogen atom of H2O

i

particle i

j

particle j

Oc

oxygen atom of CO2

Ow

oxygen atom of H2O

ro of

B

-p

1. Introduction

Modern coal-fired power plants are increasingly being combined with carbon

re

capture and storage (CCS) systems, typically geological sequestration [1]. As

lP

supercritical carbon dioxide (scCO2) is a better working fluid for geothermal energy utilization than water [2], a novel CO2 plume geothermal (CPG) [3] power generation

na

system is proposed to sequestrate CO2 and utilize geothermal energy simultaneously. The produced scCO2 has a pressure of (8 to 26) MPa and a temperature of (333.15 to

ur

473.15) K. However, this production may contain brine (about 10% mass fraction [4]),

Jo

which is harm to the turbine and heat exchangers of the power generation system [5]. To solve this issue, properties of scCO2/brine mixture should be studied. A wide range of analyses for scCO2/brine mixture have been performed in recent

years, focusing on different aspects of properties. Duan and Zhang [6] studied the PVTx properties of CO2/H2O mixtures and proposed equations of state for H2O, CO2 and CO2/H2O. Yang et al. [7] conducted a molecular dynamics study to show PVT

properties of CO2/H2O mixture, in which the H2O is supercritical or near-critical. Chapoy et al. [8] focused on the thermophysical properties and phase behavior of a CO2-rich system. By their experimentally and theoretically investigation, the effect of water on density and viscosity of CO2/H2O mixture could be obtained. Zhao et al. [9, 10] presented investigations on the interfacial tension between supercritical CO2 and brine. They concluded these investigations would help directly predict interfacial

ro of

tension in scCO2-complex electrolyte solution systems for practical applications.

Besides, other researchers mainly focused on the solubility of scCO2/brine mixture. Caumona et al. [11] developed an experimental platform for solubility measurement

-p

and obtained the solubility of water in CO2 with the pressure from 0.5 MPa to 20 MPa.

re

Wang et al. [12] investigated the solubility of water in CO2 varying from 10 MPa to 50 MPa and 313.15 K to 473.15 K by the experimental method of in-situ quantitative

lP

Raman spectroscopy. This study indicated that the solubility of water in CO2 was

na

greater under high temperature conditions. Aavatsmark and Kaufmann [13] proposed a function for solubility calculation of water in dense-phase CO2 by experimental data.

ur

It showed that the solubility of water first declined and then rose with the increase of pressure when the temperature was between 373.15 K and 473.15 K.

Jo

To understand the microscopic characteristics of the scCO2/brine mixture, the

solvation structure of brine in scCO2 should also be studied. Keshri and Tembe [14] investigated the ion association in water/CO2 mixture by molecular dynamics simulations. Their study focused on the mixture of supercritical conditions of 673 K and 39.94 MPa while the temperature and pressure of CPG production were much

lower. Under the conditions of CPG applications, water is subcritical which means that the solvation structure may be different. Keshri et al. [15] also conducted molecular dynamics simulations for Na+-Cl- ion pair in both supercritical methanol and water-methanol mixtures. Moreover, similar study was carried out by Hidayat et al. [16] to investigate the preferential solvation of K(I) ion in water/ammonia mixture. Callsen et al. [17] studied the solvation structure of lithium ions in ether solution.

ro of

Petrenko et al. [18] investigated solvate structures of salicylic acid and its derivatives

in supercritical carbon dioxide by molecular dynamics simulations. Prasetyo and Hofer [19] studied hydration free energy of carbon dioxide in aqueous solution and

-p

described the first hydration shell of CO2 in aqueous solution. Therefore, it indicates

re

researchers have already focused on study solvation structure of mixtures. Moreover, based on former references, it seems that the molecular dynamics

lP

(MD) simulation is a viable option to investigate the solvation characteristics of

na

scCO2/brine mixture. This is because molecular dynamics simulations consider the interactions between atoms and molecules at microscopic level, which are unavailable

ur

by experimental investigations. Yoon et al. [20] presented a molecular dynamics simulation on the local density distribution and solvation structure of supercritical

Jo

carbon dioxide around naphthalene. Silveira et al. [21] conducted a molecular dynamics study on the solvation of carbon dioxide in the ionic liquids. Xue et al. [22] used the spatial distribution functions to describe the solvation structure of supercritical carbon dioxide around the ether group by molecular dynamics simulations. Qi et al. [23] investigated the solvation free energy of carbon dioxide in

the mixture of brine and CH4 through molecular dynamics simulations. Besides, a molecular dynamics simulation was conducted by Wang et al. [24] to show the difference of solvation structure between benzaldehyde and cinnamaldehyde by using radial distribution functions at different CO2 densities. Vaz et al. [25] presented a molecular dynamics simulation for structural properties of ketones in supercritical carbon dioxide at infinite dilution, which means each ketone was surrounded by

ro of

carbon dioxide molecules only.

Although much literature is available on solvation structure investigation for both carbon dioxide in solutions and solute in dense-phase supercritical carbon

-p

dioxide, limited work has been published on brine solvation structure in supercritical

re

carbon dioxide. Moreover, within the temperature and pressure range of scCO2/brine mixture produced by CPG, carbon dioxide is supercritical while water is subcritical.

lP

This makes the solvation structure of scCO2/brine mixture more complicated and

na

worthy of study. As solubility is a function of temperature and pressure, the mass fraction of solute brine in scCO2 will change with its temperature and pressure, which

ur

may result in different solvation structures. Therefore, to better understand this issue, the effects of thermodynamic parameters should be considered.

Jo

In the present study, molecular dynamics simulations are performed to predict

solvation structure of scCO2/brine mixture produced by CPG. Results from this paper may present why solute brine in scCO2 has such solvation structure and offer what the effects of thermodynamic parameters on the solvation structure are. The investigation in this paper may help solve issues of CPG applications. The organization of this

paper is as follows. In Section 2, the computational methods are described and simulation conditions are presented, respectively. In Section 3, results for molecular dynamics simulations are provided and further discussion is made on the influence of simulation conditions. Then Section 4 presents concise conclusions.

2. Computational Methods

ro of

2.1 Force field parameters

In the present MD simulation, non-bonded interactions of the force field include two terms: Lennard-Jones term (for repulsion and dispersion, i.e. van der Waals

-p

attractions) and Coulomb term (for electrostatic interactions). The equation for

re

non-bonded interaction energy can be expressed as follows

lP

  ij Vij (rij )  VLJ (rij )  Vc (rij )  4 ij    rij 

12 6    ij   1 qi q j         rij   4 0 rij

(1)

na

where rij is the distance between particle i and particle j, q is charge of particle and ε0 is the vacuum permittivity. As the Lorentz-Berthelot rules used, εij and σij can be

Jo

ur

calculated by

 ij  ( ii jj )1/2 1 2

 ij  ( ii   jj )

(2) (3)

The cut-off radius for the Lennard-Jones attractions is set to be 1.2 nm. For Coulomb interactions, a shifted potential is applied using the fourth order Particle Mesh Ewald (PME) with a cut-off radius of 1.2 nm.

Bonded interactions for CO2 and H2O are estimated by EPM2 [26] model and TIP3P [27] model, respectively. To valid these molecular models, the phase boundaries of CO2 and H2O were reproduced by simulation data, experimental data and equation of state (EOS) data, as shown in Table 1. Besides, Na+ and Cl- (main ions in brine) are modeled using the OPLS force field [28].

ro of

2.2 Data analyses The radial distribution function (RDF) is a key evaluation parameter to reveal the solvation structure of scCO2/brine mixture, as it can present how the normal number

-p

density of solvent atoms (B) distribute around the solute ion (A) at different distances. Therefore, RDF gAB(r) is shown as

N A NB 1 V  P(r ) 4 r 2 iA iB

re

g AB (r ) 

(4)

lP

where V is volume, P(r) is the probability of finding a solvent atom at distance r from

na

a solute ion.

Then the coordination number (CN) nAB, which represents the number of solvent

ur

atoms or molecules bonded to a central solute ion, is defined as r1

nAB  4  r 2 g AB (r ) B dr 0

(5)

Jo

where r1 is the first minimum of RDF. Therefore, nAB is the area under the first peak of gAB(r).

The definition of hydrogen bond is based on geometric rules. Two water molecules are chosen as being hydrogen bonded only if their interoxygen distance is less than 3.5 Å, and simultaneously the O-H...O angle is less than 30° [35].

2.3 MD simulation system In this paper, the solvation structure of scCO2/brine mixture with variable temperature (373.15-473.15 K) and fixed pressure (20 MPa) was first studied. Under such simulation condition, CO2 is within supercritical region (critical point of 304.13 K and 7.377 MPa [31]) while H2O is within liquid region (critical point of 647.1 K

ro of

and 22.064 MPa [34]). In CPG applications, brine was saturated in dense-phase

supercritical carbon dioxide. Therefore, compositions of scCO2/brine mixture would change with its temperature because brine solubility in scCO2 varied. Table 2 lists

-p

compositions of present MD simulation systems under different temperatures. For

re

Case 1-3, the numbers of H2O molecules and CO2 molecules were determined by solubility of brine in scCO2 and mass fraction of NaCl in brine [12]. For Case 4-7,

lP

different amounts of ions were added into simulation system to investigate what the

na

effect of ion concentration on the solvation structure of scCO2/brine mixture was.

2.4 Solution procedure

ur

All MD simulations were performed in the platform of GROMACS 5.1.4 [36].

Jo

The simulation box for Case 1-3 was of a geometric parameter of 3.4 nm × 3.4 nm × 3.4 nm, as shown in Fig. S1. Case 4-7 had larger simulation boxes of 8 nm × 8 nm × 8 nm as more molecules were added. Initial positions for ions, CO2 and H2O molecules were random. The Newtonian equations of motion were integrated using leap-frog algorithm with a time step of 1 fs. Before MD simulations conducted, the process of energy minimization (EM) was done by the steepest decent algorithm to ensure the

MD simulation system had a reasonable starting structure. Then MD simulations were carried out under isothermal-isobaric NPT ensemble. The temperature and pressure for each simulation were controlled by the Nosé-Hoover thermostat [37, 38] and Parrinello-Rahman barostat [39], respectively. All MD simulations in this study were

3. Results and discussion 3.1 Radial distribution functions & coordination numbers

ro of

calculated for 15 ns.

In this section, a MD simulation is first performed for Case 1 to show the radial

-p

distribution functions and coordination numbers of Na+ and Cl- ions at different

re

distances. Figure 1 presents that the interactions of ions and CO2 molecules are quite different from those of ions and H2O molecules. The RDFs of Na+-Ow and Cl--Ow are

lP

maximum at 0.238 nm and 0.324 nm, respectively. Then they decline rapidly with the

na

increase of particle distance. Although the RDFs of Na+-Oc and Cl--Oc have maximal values, the RDFs at r=1.6 nm vary less than 20% from their maxima. This is because

ur

polar molecules (H2O) gather around ions while nonpolar molecules (CO2) disperse in the whole space. Moreover, a greater maximum of g(r) and a smaller corresponding

Jo

particle distance of Na+-Ow indicate that H2O molecules seem to bind more tightly to Na+ ion than Cl- ion in the scCO2/brine mixture. Results for coordination numbers of different particles are illustrated in Fig. 2. The variation trend of nNa-Ow(r) implies that there is a solvation shell for Na+ ion. Comparing with the particle distances of Na+-Ow and Cl--Ow at low CNs, it seems that

H2O molecules could gather closer to Na+ than Cl- in the scCO2/brine mixture. The variations of CNs of Na+-Oc and Cl--Oc show cubic polynomial increase trends. It is likely that ions have limited effect on the distribution of supercritical CO2 molecules. Therefore, it is inferred that scCO2 molecules may have a uniform distribution in the simulation space.

ro of

3.2 Effect of thermodynamic parameter

In the application of CPG, the temperature not only influences compositions of

the scCO2/brine mixture but also affects its solvation structure. The compositions

-p

have already been calculated by the solubility of brine in scCO2 under three different

re

temperatures varying from 393.15 K to 473.15 K, which are typical temperatures of CPG productions. Then three MD simulations are conducted to present what the effect

lP

of temperature on the solvation structure is. Since the critical temperature may be

na

different for molecular models, the simulation temperatures are modified by the critical temperature of EPM2 CO2 and TIP3P H2O in this study. Therefore, the

ur

reduced simulation temperature Tr,sim is defined as the ratio of simulation temperature Tsim and model critical temperature Tc,model, as presented in Table S1.

Jo

Figure 3 shows the variations of RDFs of Na+-Ow and Na+-Oc with different

temperatures. The results for Na+-Ow indicate the decline of temperature can reduce the maximum of g(r), which is further presented by the inset of Figure 3(a). As a consequence, the coordination numbers may be changed which will be discussed later. Figure 3(b) illustrates the decrease of temperature is favorable to enlarge the

maximum of RDF while this peak will appear at a larger particle distance. More supercritical CO2 molecules are added to the simulation system when the simulation temperature decreases. This is able to explain why a greater peak appears with the decrease of temperature. Moreover, the results for particle distances of these peaks imply the interactions of Na+ ion and H2O molecules are stronger under higher temperature conditions.

ro of

The variations of radial distribution functions of Cl- ion with the decrease of temperature are presented in Fig. 4. There exists a bias between the blue line and the

other two lines from r=0.4 nm to 1.0 nm in Fig. 4(a). This helps to explain that both

-p

Cl- ion and Na+ ion have stronger ability to bond H2O molecules under higher

re

temperature conditions. In Fig. 4(b), the peaks of gCl-Oc(r) with the reduced simulation temperatures of 1.385 and 1.257 appear at a smaller particle distance comparing with

lP

those of gNa-Oc(r). If the particle distance differences of these peaks in Figure 4(b) are

na

discussed, it can find that the temperature has limited effect on them. Former simulation results have already indicated that only Na+-Ow has an

ur

obvious solvation shell. Therefore, the numbers of H2O molecules in the first solvation shell of Na+ ion with different temperatures are presented in Table 3. The

Jo

increase of temperature results in larger first solvation shell radius. The molecular thermal motion may be the main reason for this variation. It is definitely that molecules move more intensely under higher temperature conditions. Besides, more H2O molecules will stabilize at the first solvation shell of Na+ ion under low temperature conditions. These all imply the first solvation shell of Na+ ion will

become weaker with the increase of temperature.

Coordination numbers of the first solvation shell for the other three particle interactions in the scCO2/brine mixture are not presented because their first solvation shells are not obvious. But we can still discuss how the temperature affects their coordination numbers at specific radius. Table 3 further lists the results for these

ro of

coordination numbers with different temperatures at the radius of 1.6 nm. It is

concluded that the coordination numbers of Na+-Oc and Cl--Oc are almost equal at any temperature. One reason for this phenomenon is that polar ions (Na+ and Cl-) have

-p

limited influence on nonpolar molecules of CO2, therefore CO2 molecules will

re

distribute uniformly in the simulation space. The other reason is that Na+ ion and Clion may associate to an ion pair, which means they have approximate coordination

lP

numbers. Moreover, a fraction of CO2 molecules k is defined as the ratio of

na

coordination number at 1.6 nm and total CO2 molecule number. Although the variations of coordination numbers indicate that CO2 molecules have intense

ur

distribution under low temperature conditions, the results of kNa-Oc and kCl-Oc imply more CO2 molecules distribute out of the sphere with a radius of 1.6 nm. The

Jo

snapshot of scCO2/brine mixture under the condition of 393.15 K is illustrated in Fig. 5. This further validates a Na+-Cl- ion pair does associate under such simulation condition and there exists a stable solvation shell of H2O molecules around this ion pair. Therefore, it is concluded that the increase of temperature will make the first solvation shell of Na+-Cl- ion pair become weaker.

3.3 Effect of ion concentration Brine composition is another key parameter which may affect the solvation structure of scCO2/brine mixture. As the NaCl mass fraction of brine in CPG production may vary, the investigation of ion concentration effects becomes further significant. Firstly, a simulation is performed for Case 4, which is just a mixture of

ro of

supercritical CO2 and water. This is selected as a reference for those cases with different ion concentrations.

Figure 6 shows the interactions of H2O molecules in the scCO2/H2O mixture.

-p

The variation curve of gOw-Ow(r) has an obvious peak and finally drops to zero. This

re

implies H2O molecules may be clustered in the supercritical CO2 environment. The results for nOw-Ow(r) further prove this inference because nOw-Ow(r) is almost achieved

lP

300 at the radius of 3.2 nm. This is an interesting conclusion considering water is able

na

to solvate in supercritical CO2. Therefore, it seems that the scCO2/H2O mixture is heterogeneous under the conditions of CPG application. Moreover, the results for

ur

gOw-Hw(r) in Fig. 6(a) can helps to explain the interaction of hydrogen bond (H-bond) in the scCO2/H2O mixture. The first peak of H-bond appears at the radius of 0.184 nm,

Jo

which is consistent with the result of Ref. [40]. The first minimum of gOw-Hw(r) indicates that the number of hydrogen bonds per oxygen atom is 3.579. Then simulations are carried out for Case 5 and Case 7 to present what the effect of ion concentration on the interactions of Ow-Ow and Ow-Hw is. In Fig. 7(a), the peak of gOw-Ow(r) for Case 7 is much smaller, as well as the area under the variation curve

of gOw-Ow(r). This is because H2O molecules around the Na+ ion and Cl- ion will bind to them. Therefore, the coordination number of Ow-Ow will become smaller at the same time. The results for gOw-Hw(r) imply the interaction of hydrogen bond in the scCO2/H2O mixture will become weaker with the increase of ion concentration. Furthermore, the first minimum of gOw-Hw(r) for Case 7 indicates its number of hydrogen bonds per oxygen atom decreases to 3.194.

ro of

The effects of ion concentration on the RDFs of Na+ ion and Cl- ion are illustrated in Figs. 8 and 9, respectively. A lower peak of gNa-Ow(r) achieves around

0.24 nm when the ion concentration goes up. Besides, the first minimum of gNa-Ow(r)

-p

reaches at larger radius with the increase of ion concentration. These both indicate

re

that less H2O molecules are contained in the first solvation shell of Na+ ion. Table 4 lists the results for coordination numbers of Na+-Ow with different ion concentrations.

lP

This also suggests the number of H2O molecules in the first solvation shell of Na+ ion

na

decreases with the increase of ion concentration. However, it cannot be inferred that H2O molecules bind to ions more tightly under higher ion concentration, because

ur

interactions between ions have not been discussed yet. Figure 8(b) indicates CO2 molecules are not able to form a solvation shell under any simulated ion concentration

Jo

because of their nonpolarity. The RDF of Na+-Oc for Case 7 has a lower peak. As more Na+-Cl- ion pairs are contained in Case 7, it seems that the Na+-Cl- ion pairs are clustered together. Thus, ion pairs in the center may have fewer CO2 molecules around them while other ion pairs may be surrounded by more CO2 molecules. In general, their mean RDF of Na+-Oc shows such variation trend as Figure 8(b).

The results for gCl-Ow(r) present the first solvation shell of H2O molecules around Cl- ion only appears under quite low ion concentration conditions, as shown in Fig. 9(a). Considering the discussion above for H2O molecules clustering, it seems that the amount of H2O molecules is great enough so that the Na+-Cl- ion pair is dissolved into single Na+ ion and Cl- ion. This is the reason why the first solvation shell of H2O

ro of

forms under such conditions. Besides, it also helps to explain why the first maxima of Case 5 in Figs. 8(b) and 9(b) are lower than those of Case 6. Furthermore, the curves

of gNa-Oc(r) and gCl-Oc(r) for Case 7 have similar variation trend, which implies ion

-p

pairs exist under high ion concentration conditions.

re

Table 4 further presents the coordination numbers of cases with different ion concentrations at the radius of 3.2 nm. The variations of these coordination numbers

lP

indicate that ion interactions affect the interactions of Na+-H2O pairs and Cl--H2O

na

pairs, as well as those between H2O molecules, resulting in smaller CNs at high ion concentration condition. Specifically, it seems the decrease of CN number has an

ur

approximate linear relation with the increase of ion concentration. Moreover, the solvation structure of Na+-Cl- ion pair with different ion

Jo

concentration may be different when comparing results for RDF in Figs. 8(a) and 9(a). The peak value of RDF of Na+-Ow is quite different for Case 5 and Case 7. With the decrease of ion concentration, this peak value falls from 426.454 to 205.241. Results in Fig. 9(a) present the peak values of RDF of Cl--Ow for Case 5 and Case 7 are 227.003 and 76.927, respectively. These rapid decreases show more H2O molecules

occupy the space around Na+ ion and Cl- ion under low ion concentration condition. Therefore, it is likely that the distance of Na+-Cl- pair varies with the ion concentration. Fig. 10 illustrates the snapshot of scCO2/brine mixture for Case 7. It further indicates that the scCO2/brine mixture is heterogeneous under the condition of CPG application. Snapshot for other simulated cases are further presented in Fig. S2.

ro of

3.4 Hydrogen bond analysis

Results for gOw-Hw(r) show ion concentration may influence the interactions of

hydrogen bonds. Therefore, further investigation is performed to present the

-p

relationship between ion concentration and number of hydrogen bonds, as listed in

re

Table 5. The NaCl mass fraction XNaCl is calculated by compositions of scCO2/brine mixture. The results indicate the number of total hydrogen bonds nHB in the simulation

lP

system declines monotonically with the increase of XNaCl.

na

By fitting these simulation results, it is likely that the nHB has a quadratic polynomial decrease trend when the XNaCl goes up, which can be expressed as (6)

ur

nHB  1.8491( X NaCl )2  6.3494 X NaCl  368.3 1wt%  X NaCl  10wt%

The rapid decrease of nHB implies the electrostatic interactions of ions and H2O

Jo

molecules are stronger than the hydrogen bonding interactions between H2O molecules. Thus, much fewer hydrogen bonds form between H2O molecules under higher ion concentration condition. Moreover, MD simulation results for Case 4 show its number of hydrogen bonds is 359, which is less than that of Case 5. This maybe because the addition of ion pair makes the cluster of H2O molecules associate closely.

It means more H2O molecules are able to reach within the hydrogen bonding range of their neighbor H2O molecules. Under the combined effect of electrostatic interaction and hydrogen bonding interaction, the number of total hydrogen bonds in the scCO2/brine mixture presents the former variation trend.

4. Conclusions

ro of

Molecular dynamics simulation was performed for the solvation structure of

scCO2/brine mixture produced by the CO2 plume geothermal system. Based on the analyses conducted in this investigation, the following conclusions may be drawn:

-p

(1) H2O molecules bind more tightly to Na+ ion than Cl- ion in the scCO2/brine

re

mixture while supercritical CO2 molecules seem have a uniform distribution in the whole space. Besides, there is a solvation shell of H2O molecules around Na+ ion.

lP

(2) Both Na+ ion and Cl- ion have stronger ability to bind H2O molecules under higher

na

temperature conditions. The increase of temperature will make the first solvation shell of Na+-Cl- ion pair become weaker.

ur

(3) Less H2O molecules are contained in the first solvation shell of Na+ ion under higher ion concentration conditions. Moreover, a first solvation shell of Cl- ion

Jo

will appear under quite low ion concentration condition.

(4) It seems that the scCO2/brine mixture is heterogeneous under the conditions of CPG application. (5) The interaction of hydrogen bond in the scCO2/brine mixture becomes weaker with the increase of ion concentration. It is likely that the number of total

hydrogen bonds has a quadratic polynomial decrease trend when the ion concentration increases.

Declaration of interests The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

ro of

Acknowledgement The authors gratefully acknowledge the financial support by the National Natural

Science Foundation of China (Grant No. 51976031) and Qinglan Project of Jiangsu

Jo

ur

na

lP

re

-p

Province, China.

References [1] R.S. Jayne, H. Wu, R.M. Pollyea, Geologic CO2 sequestration and permeability uncertainty in a highly heterogeneous reservoir, Int. J. Greenhouse Gas Control, 83 (2019) 128-139. https://doi.org/10.1016/j.ijggc.2019.02.001. [2] E.K. Levy, X. Wang, C. Pan, C.E. Romero, C.R. Maya, Use of hot supercritical CO2 produced from a geothermal reservoir to generate electric power in a gas turbine generation

system,

J.

CO2

Util.,

https://doi.org/10.1016/j.jcou.2017.11.001.

23

(2018)

20-28.

ro of

power

[3] J.B. Randolph, M.O. Saar, J. Bielicki, Geothermal energy production at geologic

-p

CO2 sequestration sites: Impact of thermal drawdown on reservoir pressure, Energy

re

Procedia, 37 (2013) 6625-6635. https://doi.org/10.1016/j.egypro.2013.06.595. [4] X. Wang, E.K. Levy, C. Pan, C.E. Romero, A. Banerjee, C. Rubio-Maya, L. Pan,

lP

Working fluid selection for organic Rankine cycle power generation using hot

na

produced supercritical CO2 from a geothermal reservoir, Appl. Therm. Eng., 149 (2019) 1287-1304. https://doi.org/10.1016/j.applthermaleng.2018.12.112.

ur

[5] J. Nogara, S.J. Zarrouk, Corrosion in geothermal environment: Part 1: Fluids and their impact, Renewable Sustainable Energy Rev., 82 (2018) 1333-1346.

Jo

https://doi.org/10.1016/j.rser.2017.06.098. [6] Z. Duan, Z. Zhang, Equation of state of the H2O, CO2, and H2O–CO2 systems up to 10 GPa and 2573.15 K: Molecular dynamics simulations with ab initio potential surface,

Geochim.

Cosmochim.

https://doi.org/10.1016/j.gca.2006.02.009.

Acta,

70

(2006)

2311-2324.

[7] X. Yang, J. Xu, S. Wu, M. Yu, B. Hu, B. Cao, J. Li, A molecular dynamics simulation study of PVT properties for H2O/H2/CO2 mixtures in near-critical and supercritical regions of water, Int. J. Hydrogen Energy, 43 (2018) 10980-10990. https://doi.org/10.1016/j.ijhydene.2018.04.214. [8] A. Chapoy, M. Nazeri, M. Kapateh, R. Burgass, C. Coquelet, B. Tohidi, Effect of impurities on thermophysical properties and phase behaviour of a CO2 -rich system in Int.

J.

Greenhouse

Gas

Control,

https://doi.org/10.1016/j.ijggc.2013.08.019.

19

(2013)

92-100.

ro of

CCS,

[9] L. Zhao, J. Ji, L. Tao, S. Lin, Ionic Effects on Supercritical CO2-Brine Interfacial

Temperature

and

Pressure,

Langmuir,

32

(2016)

9188.

re

Strength,

-p

Tensions: Molecular Dynamics Simulations and a Universal Correlation with Ionic

https://doi.org/10.1021/acs.langmuir.6b02485.

lP

[10] L. Zhao, S. Lin, J.D. Mendenhall, P.K. Yuet, D. Blankschtein, Molecular

na

dynamics investigation of the various atomic force contributions to the interfacial tension at the supercritical CO2-water interface, J. Phys. Chem. B, 115 (2011)

ur

6076-6087. https://doi.org/10.1021/jp201190g. [11] M.C. Caumona, J. Sterpenich, A. Randi, J. Pironon, Experimental mutual

Jo

solubilities of CO2 and H2O in pure water and NaCl solutions, Energy Procedia, 114 (2017) 4851-4856. https://doi.org/10.1016/j.egypro.2017.03.1625. [12] Z. Wang, Q. Zhou, H. Guo, P. Yang, W. Lu, Determination of water solubility in supercritical CO2 from 313.15 to 473.15 K and from 10 to 50 MPa by in-situ quantitative Raman spectroscopy, Fluid Phase Equilib., 476 (2018) 170-178.

https://doi.org/10.1016/j.fluid.2018.08.006. [13] I. Aavatsmark, R. Kaufmann, A simple function for the solubility of water in dense-phase carbon dioxide, Int. J. Greenhouse Gas Control, 32 (2015) 47-55. https://doi.org/10.1016/j.ijggc.2014.11.001. [14] S. Keshri, B.L. Tembe, Ion association in binary mixtures of water-CO2 in supercritical conditions through classical molecular dynamics simulations, J. Mol.

ro of

Liq., 257 (2018) 82-92. https://doi.org/10.1016/j.molliq.2018.02.052.

[15] S. Keshri, A. Sarkar, B.L. Tembe, Molecular dynamics simulations of Na+-Cl−

https://doi.org/10.1016/j.supflu.2015.04.018.

-p

ion-pair in supercritical methanol, J. Supercrit. Fluids, 103 (2015) 61-69.

re

[16] Y. Hidayat, H.D. Pranowo, R. Armunanto, 2018, Revisiting structure and dynamics of preferential solvation of K(I) ion in aqueous ammonia using QMCF-MD Chem.

Phys.

Lett.,

lP

simulation,

699,

S0009261418302628.

na

https://doi.org/10.1016/j.cplett.2018.03.067.

[17] M. Callsen, K. Sodeyama, Z. Futera, Y. Tateyama, I. Hamada, The Solvation

ur

Structure of Lithium Ions in an Ether Based Electrolyte Solution from First-Principles Molecular

Dynamics,

J.

Phys.

Chem.

B,

121

(2016)

180-188.

Jo

https://doi.org/10.1021/acs.jpcb.6b09203. [18] V.E. Petrenko, M.L. Antipova, D.L. Gurina, Salicylic acid, acetylsalicylic acid, methyl salicylate, salicylamide, and sodium salicylate in supercritical carbon dioxide: Solute – cosolvent hydrogen bonds formation, J. Supercrit. Fluids, 116 (2016) 62-69. https://doi.org/10.1016/j.supflu.2016.05.010.

[19] N. Prasetyo, T.S. Hofer, Structure, Dynamics, and Hydration Free Energy of Carbon Dioxide in Aqueous Solution: A Quantum Mechanical/Molecular Mechanics Molecular Dynamics Thermodynamic Integration (QM/MM MD TI) Simulation Study,

J.

Chem.

Theory

Comput.,

14

(2018)

6472-6483.

https://doi.org/10.1021/acs.jctc.8b00557. [20] T.J. Yoon, Y.H. Min, W.B. Lee, Y.W. Lee, Molecular dynamics simulation on the

naphthalene,

J.

Supercrit.

Fluids,

https://doi.org/10.1016/j.supflu.2017.07.012.

ro of

local density distribution and solvation structure of supercritical CO2 around 130

(2017).

-p

[21] A.J. Silveira, S. Pereda, F.W. Tavares, C.R.A. Abreu, A molecular dynamics

re

study of the solvation of carbon dioxide and other compounds in the ionic liquids [emim][B(CN)4] and [emim][NTf2], Fluid Phase Equilib., 491 (2019) 1-11.

lP

https://doi.org/10.1016/j.fluid.2019.03.007.

na

[22] P. Xue, J. Shi, X. Cao, S. Yuan, Molecular Dynamics Simulation of Thickening Mechanism of Supercritical CO2 Thickener, Chem. Phys. Lett., 706 (2018) 658-664.

ur

https://doi.org/10.1016/j.cplett.2018.07.006. [23] C.L. Qi, S.C. Wee, B. Maulianda, R. Barati, A.Z.b. Bahruddin, E. Padmanabhan,

Jo

Determination of solvation free energy of carbon dioxide (CO2) in the mixture of brine, Alfa Olefin Sulfonate (AOS) and CH4 after foam fracturing in the shale reservoirs on enhanced shale gas recovery (ESGR), J. Nat. Gas Sci. and Eng., 54 (2018) 102-109. https://doi.org/10.1016/j.jngse.2018.03.027. [24] J. Wang, F. Zhao, Z. Wu, A study of solvation of benzaldehyde and

cinnamaldehyde in CO2 by molecular dynamics simulation, Chem. Phys. Lett., 492 (2010) 49-54. https://doi.org/10.1016/j.cplett.2010.04.046. [25] R.V. Vaz, J.R.B. Gomes, C.M. Silva, 2015, Molecular dynamics simulation of diffusion coefficients and structural properties of ketones in supercritical CO2 at infinite

dilution,

J.

Supercrit.

Fluids,

107,

S0896844615300784.

https://doi.org/10.1016/j.supflu.2015.07.025.

ro of

[26] J.G. Harris, K.H. Yung, Carbon Dioxide's Liquid-Vapor Coexistence Curve And

Critical Properties as Predicted by a Simple Molecular Model, J. Phys. Chem., 99 (1995) 12021-12024. https://doi.org/10.1021/j100031a034.

-p

[27] W.L. Jorgensen, J. Chandrasekhar, J.D. Madura, R.W. Impey, M.L. Klein,

re

Comparison of simple potential functions for simulating liquid water, J. Phys. Chem., 79 (1983) 926-935. https://doi.org/10.1063/1.445869.

lP

[28] W.L. Jorgensen, D.S.M. And, J. Tiradorives, Development and Testing of the

Liquids,

J.

na

OPLS All-Atom Force Field on Conformational Energetics and Properties of Organic Am.

Chem.

Soc.,

118

(1996)

11225-11236.

ur

https://doi.org/10.1021/ja9621760.

[29] J.G. Harris, K.H. Yung, Carbon dioxide's liquid-vapor coexistence curve and

Jo

critical properties as predicted by a simple molecular model, J. Phys. Chem., 99 (1995) 12021-12024. https://doi.org/10.1021/j100031a034. [30] F. Din, Thermodynamic Functions of Gases: Air, acetylene, ethylene, propane and argon, Butterworths Sci. Publ., 1956. [31] R. Span, W. Wagner, A new equation of state for carbon dioxide covering the

fluid region from the triple-point temperature to 1100 K at pressures up to 800 MPa, J. Phys. Chem. Ref. Data, 25 (1996) 1509-1596. https://doi.org/10.1063/1.555991. [32] C. Vega, J.L. Abascal, Simulating water with rigid non-polarizable models: a general perspective, Phys. Chem. Chem. Phys., 13 (2011) 19663-19688. https://doi.org/10.1039/C1CP22168J. [33] A. Saul, W. Wagner, A fundamental equation for water covering the range from

18 (1989) 1537-1564. https://doi.org/10.1063/1.555836.

ro of

the melting line to 1273 K at pressures up to 25 000 MPa, J. Phys. Chem. Ref. Data,

[34] W. Wagner, A. Pruss, The IAPWS formulation 1995 for the thermodynamic

-p

properties of ordinary water substance for general and scientific use, J. Phys. Chem.

re

Ref. Data, 31 (2002) 387-535. https://doi.org/10.1063/1.1461829.

[35] A.K. Soper, M.G. Phillips, A new determination of the structure of water at 25 ℃,

lP

Chem. Phys., 107 (1986) 47-60. https://doi.org/10.1016/0301-0104(86)85058-3.

na

[36] M.J. Abraham, T. Murtola, R. Schulz, S. Páll, J.C. Smith, B. Hess, E. Lindahl, GROMACS: High performance molecular simulations through multi-level parallelism laptops

to

supercomputers,

Softwarex,

1-2

(2015)

19-25.

ur

from

https://doi.org/10.1016/j.softx.2015.06.001.

Jo

[37] S. Nosé, A unified formulation of the constant temperature molecular dynamics methods, J. Chem. Phys., 81 (1984) 511-519. https://doi.org/10.1063/1.447334. [38] W.G. Hoover, Canonical dynamics: Equilibrium phase-space distributions, Phys. Rev. A, 31 (1985) 1695-1697. https://doi.org/10.1103/physreva.31.1695. [39] M. Parrinello, A. Rahman, Polymorphic transitions in single crystals: A new

molecular

dynamics

method,

J.

Appl.

Phys.,

52

(1981)

7182-7190.

https://doi.org/10.1063/1.328693. [40] A. Botti, F. Bruni, R. Mancinelli, M.A. Ricci, F. Celso, Lo, R. Triolo, F. Ferrante, A.K. Soper, Study of percolation and clustering in supercritical water-CO2 mixtures, J.

Jo

ur

na

lP

re

-p

ro of

Chem. Phys., 128 (2008) 6-48. https://doi.org/10.1063/1.2898538.

-p

ro of

Figure captions

Jo

ur

na

lP

re

(a)

(b) Fig. 1. Results for radial distribution functions of ions and molecules in scCO2/brine mixture of Case 1. (a) Blue line for Na+-Ow, red line for Cl--Ow; (b) blue line for

-p

ro of

Na+-Oc, red line for Cl--Oc.

Jo

ur

na

lP

re

(a)

(b)

Fig. 2. Results for coordination numbers of ions and molecules in scCO2/brine mixture of Case 1. (a) Blue line for Na+-Ow, red line for Cl--Ow; (b) blue line for Na+-Oc, red line for Cl--Oc.

ro of

Jo

(b)

ur

na

lP

re

-p

(a)

Fig. 3. Results for radial distribution functions of Na+ ion in scCO2/brine mixture with different reduced simulation temperatures. (a) Na+-Ow, (b) Na+-Oc. Blue solid line for Case 1; red dash line for Case 2; magenta dot line for Case 3.

ro of

Jo

ur

na

lP

re

-p

(a)

(b)

Fig. 4. Results for radial distribution functions of Cl- ion in scCO2/brine mixture with different reduced simulation temperatures. (a) Cl--Ow, (b) Cl--Oc. Blue solid line for Case 1; red dash line for Case 2; magenta dot line for Case 3.

ro of

Fig. 5. Snapshot of Na+-Cl- ion pair and its first solvation shell in scCO2/brine mixture

ur

Jo

(a)

na

lP

re

-p

for Case 3.

ro of

(b)

-p

Fig. 6. Results for interactions of H2O molecules in scCO2/H2O mixture. (a) Radial

re

distribution functions, (b) coordination numbers. Blue solid line for Ow-Ow; red dash

Jo

ur

na

lP

line for Ow-Hw.

(a)

ro of

(b)

-p

Fig. 7. Effect of ion concentration on radial distribution functions of H2O molecules

re

in scCO2/brine mixture. (a) Ow-Ow, (b) Ow-Hw. Blue solid line for Case 4; red dash

Jo

ur

na

lP

line for Case 5; magenta dot line for Case 7.

(a)

ro of

(b)

Fig. 8. Effect of ion concentration on radial distribution functions of Na+ ion in

-p

scCO2/brine mixture. (a) Na+-Ow, (b) Na+-Oc. Blue solid line for Case 5; red dash line

Jo

ur

na

lP

re

for Case 6; magenta dot line for Case 7.

(a)

ro of -p

(b)

Fig. 9. Effect of ion concentration on radial distribution functions of Cl- ion in

Jo

ur

na

lP

for Case 6; magenta dot line for Case 7.

re

scCO2/brine mixture. (a) Cl--Ow, (b) Cl--Oc. Blue solid line for Case 5; red dash line

Fig. 10. Snapshot of Na+-Cl- ion pair in scCO2/brine mixture for Case 7.

Table 1 Comparison among simulation data, experimental data and EOS data for critical parameters of CO2 and H2O. ρcrit/g·

pcrit/ Item

Tcrit/K cm-3

MPa 7.34 Simulation data [29]

312.8

0.453 0

Experimental

data

7.38

304.21 CO2

[30]

ro of

EPM2

0.467

2

7.37

EOS data [31]

304.13

0.468

re

-p

7

TIP3P

Experimental [33]

na

H2O

lP

Simulation data [32]

Jo

ur

EOS data [34]

12.6

578

0.272 0

data

22.0 647.1

0.322 64 22.0

647.1

0.322 64

Table 2 Compositions of MD simulation systems under different temperatures and ion concentrations. T/

p/

No.

NH2 Nc

K

NC

Ntot

Na

MPa

O

O2

al

473 1

20

1

1

20

1

1

20

1

1

30

116

148

433 2 .15 393 3

20

0

473 5

lP

.15

20

na

.15

re

473 4

30

275

307

30

646

678

116

146

-p

.15

ro of

.15

1

0

300 0 116

1

20

ur

.15

0

Jo

.15

2 116

3

3

146

300 0

473

7

146

300

473

6

0

6 116

20

10

10

148

300 0

0

Table 3 MD simulation results for interactions of ions and molecules at first solvation shell and radius of 1.6 nm in Case 1, Case 2 and Case 3.

First solvation shell

Case 1

Case 2

Case 3

r1/nm

0.338

0.330

0.328

n(r)

3.856

4.075

4.294

NCO2

116

275

646

nNa-Oc(r)

83.791

102.816

121.383

kNa-Oc

0.722

0.374

0.188

nCl-Oc(r)

83.627

102.612

122.343

kCl-Oc

0.721

Radius of

Jo

ur

na

lP

re

-p

1.6 nm

ro of

Item

0.373

0.189

Table 4 Results for coordination numbers of ions and molecules at first solvation shell and radius of 3.2 nm in Cases 4-7. Case

Case

Case

Case

Item 4

solvation shell

-

0.330

0.330

0.344

n(r)

-

5.259

5.028

4.056

nOw-Ow

296.1

296.0

294.8

279.3

84

59

nNa-Ow

81

298.5 -

(r)

32

298.5

-

67

Jo

ur

na

lP

r)

297.9

04

re

nCl-Ow(

85

-p

3.2 nm

7

r1/nm

(r) Radium of

6

ro of

First

5

290.2

22

297.9

49

290.3 11

Table 5 Results for number of hydrogen bonds of different ion concentrations in Cases 5-7. Case 5

Case 6

Case 7

Ion pair number

1

3

10

XNaCl (wt%)

1.071

3.145

9.765

nHB

373

370

254

Jo

ur

na

lP

re

-p

ro of

Item