Next Article in Journal
Study of Injection Method for Maximizing Oil-Cooling Performance of Electric Vehicle Motor with Hairpin Winding
Previous Article in Journal
Regional Diversification of Potential, Production and Efficiency of Use of Biogas and Biomass in Poland
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermal and Transport Properties of Molten Chloride Salts with Polarization Effect on Microstructure

1
School of Materials Science and Engineering, Sun Yat-Sen University, Guangzhou 510006, China
2
School of Intelligent Systems Engineering, Sun Yat-Sen University, Guangzhou 510006, China
3
School of Data and Computer Science, Sun Yat-Sen University, Guangzhou 510006, China
*
Author to whom correspondence should be addressed.
Energies 2021, 14(3), 746; https://doi.org/10.3390/en14030746
Submission received: 30 December 2020 / Revised: 24 January 2021 / Accepted: 27 January 2021 / Published: 31 January 2021
(This article belongs to the Section D: Energy Storage and Application)

Abstract

:
Molten chloride salt is recognized as a promising heat transfer and storage medium in concentrating solar power in recent years, but there is a serious lack for thermal property data of molten chloride salts. In this work, local structures and thermal properties for molten chloride salt—including NaCl, MgCl2, and ZnCl2—were precisely simulated by Born–Mayer–Huggins (BMH) potential in a rigid ion model (RIM) and a polarizable ion model (PIM). Compared with experimental data, distances between cations, densities, and heat capacities of molten chloride slats calculated from PIM agree remarkably better than those from RIM. The polarization effect brings an extra contribution to screen large repulsive Coulombic interaction of cation–cation, and then it makes shorter distance between cations, larger density and lower heat capacity. For NaCl, MgCl2, and ZnCl2, PIM simulation deviations of distances between cations are respectively 3.8%, 3.7%, and 0.3%. The deviations of density and heat capacity for NaCl between PIM simulation and experiments are only 0.6% and 2.2%, and those for MgCl2 and ZnCl2 are 0.7–10.7%. As the temperature rises, the distance between cations increases and the structure turns into loose state, so the density and thermal conductivity decrease, while the ionic self-diffusion coefficient increases, which also agree well with the experimental results.

1. Introduction

Concentrating solar thermal power (CSP) is a promising technique for high temperature solar energy utilization [1]. Molten salts have the advantages of wide working temperature range, high heat capacity, low viscosity, low vapor pressure, and low material prices [2]; consequently, they are widely used as heat transfer and storage materials in CSP. Molten salts include alkali or alkali earth halides (chloride salts, fluoride salts, and so on), and oxyacid salts (carbonates, nitrates, and so on). In recent years, molten chloride salts including LiCl, NaCl, KCl, MgCl2, and ZnCl2 have been researched as promising heat transfer materials with high operating temperature [3]. Another unique advantage is the tremendous nature resources worldwide, and it is reported that there are 38.81 billion tons magnesium chlorides in Qinghai salt lake of China [4].
Many researchers experimentally reported the microstructure and properties of molten chloride salts. McGreevy [5] investigated the experimental structure of molten alkali halides by the technique of neutron scattering and energy-dispersive X-ray diffraction (EDXRD). Biggin and Enderby [6] researched the structural characteristic of ZnCl2 at melting point by neutron diffraction technique. Besides, high-energy X-ray diffraction method was taken to obtain the structure of ZnCl2 by Zeidler et al. [7], showing the result of corner-sharing and edge-sharing tetrahedral unit. Pavlatou et al. [8] studied the vibrational spectra of network-forming ionic melts (LaCl3 and ZnCl2) by the Raman and infrared spectra. The structure of molten KCl-ZnCl2 system was examined by Takagi and Nakamura [9] with XRD measurement, and the polymer of ZnC12 is dissociated with the increasing mole fraction of KCl. Yoko et al. [10] researched the density of molten ZnCl2-M (M = Li, Na, K) system with the Archimedean method. Differential scanning calorimetry (DSC) tests were conducted by Mohan et al. [11] to measure the heat capacity of ternary NaCl-KCl-MgCl2. Janz et al. [12,13,14] experimentally investigated density, viscosity and other thermal properties of molten chloride salts in a narrow temperature range, and the thermal conductivity of NaCl at 1100 K measured by coaxial cylinder technique is 0.999 W/(m·K). Nagasaka et al. [15] used the forced Rayleigh scattering method to obtain the thermal conductivity of NaCl, which is 0.519 W/(m·K) at 1100 K. According to available experimental results, there is a large discrepancy between the different measurement methods. Moreover, it is difficult to obtain the complete database of experimental microstructures and thermodynamics properties over the whole working temperature due to the high temperature and strong corrosion.
Molecular dynamics simulation (MD) is an effective method to obtain the properties of molten salts. Solar salt has been industrially applied in CSP plants and its thermodynamics properties were calculated by Ding et al. [16] with the MD method. Wang et al. [17] evaluated the local structure and transport properties of LiCl, NaCl, KCl, and other molten chlorides in rigid ion models (RIM) with Born–Mayer–Huggins potential [18]. Pan et al. [19] acquired the thermal and transport properties and the local structure characters of binary molten NaCl-KCl by MD simulations with BMH potential. The first principle molecular dynamics simulation (FPMD) was carried out by Tang et al. [20] to study the properties and phase transition of binary molten NaCl-MgCl2. Liang et al. [21] studied the microstructure evolution of molten KCl-MgCl2 in composition dependence with FPMD method. The reverse non-equilibrium molecular dynamics (RNEMD) method was applied to forecast thermal conductivities of molten salts in recent researches [22,23,24]. Additionally, Huang et al. [25] computed the local structure of MgCl2 at 998 K by MD simulations, which agrees well with the experimental results by neutron diffraction technique except the distance of Mg-Mg. Gardner et al. [26] calculated pair radial distribution functions, coordination numbers and internal pressure of ZnCl2 at 1200 K with different parameters in RIM, but the distances of Zn-Zn in all simulations are in excess of more than 23% of experimental value. Rollet and Salanne [27] studied the polarization effect in the structure of molten metal halides by polarizable ion model (PIM) introduced by Wilson and Madden [28], which can improve the simulated results. The force-matching procedure based on the first-principles density functional theory calculation was performed by Salanne et al. [29] to acquire the potential parameters of molten alkali chlorides and fluorides in PIM. Sharma and Wilson [30] improved the simulation results of structure factors for MgCl2 and ZnCl2 by PIM. However, there is still a lack of precise thermal and transport properties of MgCl2 and ZnCl2 over the whole working temperature.
In order to estimate the properties of molten chloride salts more precisely, MD simulations with PIM are first carried out to obtain thermal and transport properties of molten NaCl, MgCl2, and ZnCl2 with polarizable effect on microstructures. The evolution of microstructure including partial radial distribution function and coordination number is analyzed by considering polarization effect. In addition, the thermal and transport properties of molten chloride salts calculated by PIM and RIM are both compared with experimental data.

2. Numerical Model and Conditions

2.1. Force Field

Among all of the force fields presented in molten salts, it has been recommended that a pair potential can represent the interactions. Tosi and Fumi [18] have come up with Born–Mayer–Huggins potential of rigid ion model (RIM) as follows
U ( r i j ) = z i z j e 2 r i j + A i j e x p [ B i j ( σ i j r i j ) ] C i j r i j 6 D i j r i j 8
where the first term is electrostatic potential, the second term is repulsion potential and the last two terms represent dipole dispersion energies. The potential parameters in RIM of NaCl and ZnCl2 are reported in the research [31,32]. A new series of parameters in RIM of MgCl2 used in this paper are put forward. The interactions between Cl-Cl pairs are adopted from Seo et al. [32] and the rest pair parameters are adopted from Yuen et al. [33], which are listed in Table 1.
In order to improve the force field, the polarization ion model (PIM) is researched by Madden and Salanne [34,35]. The PIM is shown by Wilson and Madden [28] with the shell model. In this model, the polarizability of ion is combined by representing an ion as the two-particle system with a charged core and a charged shell connected by an atomic spring. The induction effect is included in the polarization term. The pair potentials are described as
U ( r i j ) = z i z j e 2 r i j + A i j e x p ( - B i j r i j ) f 6 ( r ) C i j r i j 6 f 8 ( r ) D i j r i j 8 + U p o l
U p o l = i | μ i | 2 α i + i , j [ z i r i j μ j r i j 3 f i j 4 ( r i j ) z j r i j μ i r i j 3 f j i 4 ( r i j ) ] + i , j [ μ i μ j r i j 3 3 ( r i j μ i ) ( r i j μ j ) r i j 5 ]
f i j n ( r i j ) = 1 c i j n e x p ( b i j n r i j ) [ k = 0 n ( b i j n r i j ) k k ! ]
where the αi is the polarizability of the i ion and the fijn(rij) is the Tang–Toennies dispersion damping functions [36]. In Equation (3), the first term is the energy cost for the deformation of electron cloud and the last two terms represent the charge–dipole and dipole–dipole energies.
The polarization effect is the deformation of the outer electron cloud caused by the influence of an external electric field [37]. The deformation of the electron cloud and the induced dipole of anions result in the structural changes of ions. Thus, the potential parameters in PIM of molten chloride slats should be restudy. Salanne et al. [35] used the first-principles calculation method to obtain the PIM parameters of NaCl; also, Sharma and Wilson [30] provided the PIM parameters of MgCl2 and ZnCl2, which are listed in Table 2.

2.2. Simulation Details and Conditions

The RIM simulations were carried out by LAMMPS code [38] and the PIM simulations were performed by CP2K package [39]. All the simulations were carried out under the Periodic Boundary Condition (PBC), removing the side effects. The PPPM solver was used to conduct long-range corrections and the precise was set equal to 1.0 × 10−6. The initial velocity was set up to obey Gaussian distribution. Newtonian equation of motion was solved by Verlet computing method [40]. The PIM simulations include the iterations for induced dipoles in CP2K, so that the computational complexity of PIM is higher than that of RIM. The number of ions, space group and the details of unit cell are listed in Table 3. The sizes of initial simulation boxes in RIM are 45.12 × 45.12 × 90.23 Å, 40.60 × 40.60 × 83.76 Å and 43.19 × 43.19 × 82.65 Å for NaCl, MgCl2, and ZnCl2 respectively, while those in PIM are 22.56 × 22.56 × 45.12 Å, 24.36 × 24.36 × 41.88 Å, and 21.59 × 21.59 × 41.32 Å. The cutoff distance was set to 20 Å and 10 Å for RIM and PIM, which is slightly shorter than the half of edge length. Although the ions in the PIM simulations are less than the RIM simulations, the PIM simulated results have higher accuracy, which will be illustrated in Section 3.
The experimental melting points of NaCl, MgCl2, and ZnCl2 are 1073 K, 987 K, and 591 K [14]. For all the MD simulations, the system was firstly heated to 1700 K and then the system was cooled down to different temperature for simulations in NPT ensemble with the Nosé–Hoover thermostat and barostat at a constant pressure of 0.1 MP. The damping parameters were 100 fs and 500 fs for thermostat and barostat, respectively. After that, the system was simulated on the equilibrated cell volume in NVT ensemble. The simulated time step was set to 1 fs. In order to get precise calculation results, the equilibrium time lasted 1 ns for each simulation.

2.3. Properties Evaluation

2.3.1. Partial Radial Distribution Functions

Partial radial distribution function (RDF) is the typical local structure characteristic of liquid molten salts, which defines the distribution probability of j-type ions centered on i-type ion within a radius r
g i j = 1 4 π ρ j r 2 [ d n i j ( r ) d r ]
where ρj denotes the number density of species j and nij denotes the number of j-type ions locating in a globular of radius r around the center of an i-type ion.

2.3.2. Coordination Number

The quantity of j-type ions locating in a globular of radius rmin around the center of an i-type ion is the definition of coordination number (CN), and rmin is the first peak valley position of RDF. Nij can be calculated as
N i j ( r ) = 4 π ρ j 0 r m i n g i j ( r ) r 2 d r

2.3.3. Density

The densities of molten salts could be computed from the formula
ρ = N M N a V
where M is the molar mass, N is the number of particles, Na is the Avogadro’s constant, and V denotes the equilibrated volume of the simulation system at the given temperature in NPT ensemble.

2.3.4. Specific Heat Capacity

Specific heat capacities were calculated from enthalpies with the formula
C p = ( H T ) P Δ H Δ T = Δ ( U + P V ) Δ T
where U is the total internal energy of the system consisting of kinetic energy and potential energy. Thus, the value of heat capacity can be calculated from the slope of enthalpy variation with temperature.

2.3.5. Thermal Conductivity

The reverse non-equilibrium molecular dynamics (RNEMD) method [22,23,24] can be applied to forecast thermal conductivities of molten salts. RNEMD method executes kinetic energy changes between the center and the bottom slabs and measures the inducing temperature gradient with linear response theory. At first, the system is equilibrated at object temperature for 1 ns. After that, the system is simulated in NVE ensemble to swap kinetic energy between different slabs for 5 ns. Thermal conductivity is calculated from the total transformation of kinetic energy and the ultimate temperature gradient [22] as
j z = λ T z
λ = transfer m 2 ( v hot 2 v cold 2 ) 2 t L x L y ( T / z )
where t is the swap time of kinetic energy, Lx and Ly denote the lengths of simulation box in x and y directions, respectively.

2.3.6. Self-Diffusion Coefficient

Ion self-diffusion coefficient could reveal the ion transport characters of molten salts. In the cubic cell, the time-dependent mean square displacement (MSD) method [41] was carried out to compute the ion self-diffusion coefficient. In the recent researches [42,43], this method has been carried out to obtain the self-diffusion coefficients of molten salts, which agree well with the experiment results. MSD and ion self-diffusion coefficient are calculated as
M ( t ) = 1 N i = 1 N [ r i ( t ) r i ( 0 ) ] 2
D = 1 6 d [ M ( t ) ] d t
where the brackets <···> denote the time averages and ri is the position of ion i.

3. Results and Analyses

3.1. Local Structures in PIM

To study the relation between microstructure and thermodynamics, the simulated partial radial distribution function (RDF) and coordination number (CN) are illustrated as Figure 1 and Figure 2 and Table 4. It can be seen that all the RDFs have comparatively high first peaks and the height of other peaks decrease gradually, which demonstrates the disorder microstructures of molten chloride salts in long distance. For the ion pair of cation–anion, the first peak is sharper and the first valley is lower than that of cation–cation, indicating that cation–anion pair has a stronger interaction.
Figure 1 and Table 4 show the curves and characteristic values of RDFs by PIM simulation. Compared with the experiments by neutron diffraction technique [5], the first peak position of Na+-Na+ from PIM is slightly overestimated by 3.8% at 1100 K. Near the melting point, RDFs of MgCl2 in PIM also have a good agreement with the experimental values [25]. For ZnCl2, the first peak position of Zn2+-Zn2+ at 600 K is 3.80 Å [6], and the PIM simulation result is 3.79 Å. Thus, the PIM can provide the reasonable force field between particles and obtain the accurate structures of molten chloride salts. When the temperature increases, the first peak positions of RDFs for cation–anion decrease while those for cation–cation increase, showing that the distances between cations and anions become closer while the distances of cations become larger.
The coordination number curves of NaCl, MgCl2 and ZnCl2 in diverse working temperature condition are displayed in Figure 2. It can be observed that the coordination number diminishes from low temperature to high temperature in all simulations. For NaCl, the coordination number of Na+-Cl from PIM is 5.06 at 1100 K, which agrees well with the experimental value of 4.83 [5]. As the temperature rises, the coordination number of Mg2+-Cl changes from 5.21 to 4.72, while that of Zn2+-Cl changes from 5.36 to 5.02. According to the analysis of RDF and CN, the structure of molten chloride salts becomes looser under high temperature condition.

3.2. Polarization Effect on Structures

To research polarization effect on structures of molten chloride salts, RIM and PIM simulations are performed with the same conditions in Section 3.2. The simulation results of RDF and CN in RIM and PIM are displayed in Figure 3 and Table 5. The first peak position of Na+-Na+ is 4.05 Å from PIM and 4.10 Å from RIM, which are slightly different from experiment value of 3.90 Å. However, there is a large difference between PIM simulated distances of MgCl2/ZnCl2 and those in RIM. When polarization effect is included, the first peak position of Mg2+-Mg2+ changed from 4.38 Å in RIM to 3.95 Å in PIM, comparing with the experiment value of 3.81 Å. In Figure 3b, the first peak position of Zn2+-Zn2+ in PIM is 3.79 Å and fits very well with the experiment value of 3.80 Å, while that in RIM is 4.30 Å.
In addition, the coordination number of Zn2+-Zn2+ in PIM is larger than that in RIM, indicating that the cation clusters in PIM are more compact than the cation clusters in RIM. That is to say, the large repulsive coulombic interactions between divalent cations lead to the larger distance between ions in RIM. As soon as the induced dipoles of anions are considered, the large Coulombic force of cation–cation can be screened by the polarization effect, which brings an extra contribution to take the shorter distance of cation–cation. The higher polarizability of Zn2+ leads to the stronger polarizable force between ions, so that the PIM performs better to derive the first peak position of Zn2+-Zn2+. In Figure 3, the second peaks of Mg2+-Mg2+ and Zn2+-Zn2+ in PIM show characteristic of the intermediate-ranged order, which reveals the network structure of MgCl2 and ZnCl2. Besides, the snapshots of network structures for molten salts are plotted in Figure 4. Thus, the PIM simulations are more reasonable and correspond better to the physical laws.

3.3. Density

Density is the critical thermodynamics property of molten chloride salts, which reflects the change of the system volume. The experimental melting points of NaCl, MgCl2, and ZnCl2 are 1073 K, 987 K, and 591 K [14]. The systems were heated to molten state at 1700 K in NPT ensemble. After that, the systems in both models were cooled down to the target temperature and equilibrated in the NPT ensemble for 1 ns. Specific heat capacity was also computed with the similar method.
In Figure 5, the simulated densities of molten chloride salts are compared with the experimental results by Yaffe et al. [44], Janz et al. [14] and Angell et al. [45]. It is obvious that the density is in linear negative correlation with temperature. For molten ZnCl2, the PIM simulated densities decrease from 2.50 g/cm3 at 600 K to 2.28 g/cm3 at 1000 K, which is slightly smaller than the experiment values by 1.1%. When the temperature is over 773 K, the RIM simulated densities of ZnCl2 become into a gross difference from experiments. Besides, the slope of density in RIM is different from that in PIM for MgCl2 and ZnCl2. Combined with the analysis of polarization effect on local structure in Section 3.2, the large Coulombic interactions of ions induce the longer distance between cations in RIM for lack of the polarization effect. In addition, the looser coordination structures and longer distance between cations in RIM cause the greatly larger volume of system at high temperature, so the densities from RIM are much smaller than real values. As listed in Table 6, the fit formulas of PIM simulation results agree well from the experimental data, with the mean percentage errors of 0.3%, 9.5%, and 1.1% for NaCl, MgCl2, and ZnCl2 respectively. The error of MgCl2 is larger than others, which may be caused by the error of the fitting potential parameters searched for MgCl2. The polarization effect included in the potential of MgCl2 can bring extra contribution to screen repulsive Coulombic force, but the repulsive force is still slightly larger and induced larger volume for MgCl2.

3.4. Specific Heat Capacity

As shown in Figure 6, the enthalpy variations of molten chloride salts are in linear relation with temperature, which indicates the specific heat capacity is constant from Equation (8). Listed in Table 7, the specific heat capacities of NaCl, MgCl2 and ZnCl2 in PIM are respectively 65.5 J/(mol·K), 99.7 J/(mol·K) and 111.7 J/(mol·K) and their errors are 2.2%, 7.8% and 10.7%. Simulated heat capacities of NaCl in RIM and PIM are both in good agreement with experiment [24], while the results of MgCl2 and ZnCl2 in RIM are 38.8% and 32.9% larger than the experimental data [14,46]. In conclusion, the polarization effect can reflect the accurate ability of thermal storage for molten divalent chloride (MgCl2 and ZnCl2).

3.5. Thermal Conductivity

The RNEMD method was used to compute the thermal conductivities. To input a heat flux and obtain a temperature distribution, the simulation box was evenly split into 20 bins along the ‘z’ axis. The 1st bin was defined as the ‘cool’ bin, and the 11th bin was defined as the ‘hot’ bin. The heat flux was generated by swapping the kinetic energy of an atom in the cool bin and one in the hot bin, which results in the increasing temperature in the hot bin and decreasing temperature in the cool bin [22]. After equilibrating for 1 ns, the simulations transferred the kinetic energy over 5 ns in NVE ensemble. Different swap rates were tried for the simulations, and it was found that kinetic energy swap rates between 100 and 400 time steps brought about the linear temperature gradients. The total average kinetic energy exchange as well as the slopes of linear temperature gradient in Figure 7 were substituted into Equation (10) to calculate the thermal conductivities. The swapping rates are same for all molten chloride salts at 400 fs.
In order to approximate the finite-size effect on thermal conductivity of molten chloride salt by RNEMD method, a series of different size simulations are carried out, and the lengths in x and y dimension are the same while the length of z dimension is changed. The size-dependent thermal conductivities of molten chloride salt are displayed in Figure 8. It is obvious that the simulated thermal conductivity increases non-linearly with increasing length of z dimension and approaches a stable value when the length of z dimension is over 100 Å.
As shown in Figure 9, the thermal conductivities of molten chloride salts are in negative linear correlation with temperature. For molten NaCl, the simulated results of RNEMD decrease from 0.571 W/(m·K) to 0.424 W/(m·K). There is a mean percentage error of 8.1% between simulated results and experiments from Nagasaka et al. [15], so it proves that RNEMD method is suitable for calculating the thermal conductivities. Compared with the experimental values from Singh et al. [47] and Cornwell [48], the RNEMD simulated results of MgCl2 and ZnCl2 agree well and are completely predicted over the whole operating temperature. In general, the simulated thermal conductivities of molten chloride salts in different temperature are in the range of 0.213~0.571 W/(m·K).

3.6. Ionic Self-Diffusion Coefficient

In order to research the transport property of molten chloride salt ions, the ionic self-diffusion coefficients were calculated for 1 ns in NVT ensemble with the mean square displacement (MSD) method. It is evident that the MSD presents positive linear temperature dependence in Figure 10. Compared with experiments from Janz et al. [14], there are mean percentage errors of 5.9% and 5.7% between simulated results and experiments of Na+ and Cl from 1100 K to 1500 K in Figure 11a, respectively. In Figure 11b, both Zn2+ and Cl are difficult to move near melting point with quite small self-diffusion coefficients, because of the glassy state with the network structure [49], while Na+ and Cl can move more easily at melting point without the network structure. When the temperature increases, the ions become easier to move for the break of network structure. It can be seen that self-diffusion coefficients of all cations are larger than Cl in Figure 11, resulting from the limited mobility of Cl by the larger mass and ionic radius.

4. Conclusions

Microstructure and thermopyhsical properties of molten chloride salts (NaCl, MgCl2, and ZnCl2) were researched by molecular dynamics simulations with RIM and PIM. Compared with results from RIM, PIM simulations reveal that the polarization effect brings an extra contribution to screen the large Coulombic interactions, which makes the shorter distance between cations. For divalent chloride, the high charge density of divalent cation leads to more deformation of the electron cloud, so that the polarizable force between anion and cation is stronger. The RIM simulations are unable to precisely predict the densities and heat capacities of divalent chloride (MgCl2 and ZnCl2) for lack of considering polarization effect, while PIM simulations improve the accuracy of simulated densities and heat capacities. Thus, the PIM performs better for molten divalent chloride. There are 0.3%, 1.1%, and 9.5% mean errors between PIM simulated densities and experiments for NaCl, ZnCl2, and MgCl2 respectively, and those of heat capacities are 2.2–10.7%. As the temperature rises, the distance between cations increases and the structure turns into loose state, so the density and thermal conductivity decrease, while the ionic self-diffusion coefficient increases, which agree well with the experimental results. This work provides an alternative way to build the property database of molten chloride salts for the precise design and regulation of high-temperature thermal storage system.

Author Contributions

Conceptualization, J.L.; Methodology, G.P.; Software, S.Y.; Validation, S.Y., G.P., and S.L.; Formal analysis, S.Y.; Investigation, J.L.; Writing—review and editing, J.L.; Project administration, J.D.; Funding acquisition, J.L. and W.W. All authors have read and agreed to the published version of the manuscript.

Funding

This paper is supported by National Natural Science Foundation of China (52036011, U1601215), and Natural Science Foundation of Guangdong Province (2017B030308004).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

ABMH repulsive parameter
BBMH repulsive parameter
CVan der Waals parameter
DVan der Waals parameter
Llength dimension
nnumber of particles
qcharge of ion
rdistance between ions
Ttemperature of the system
tsimulation time
UBorn–Mayer–Huggins–Tosi–Fumi potential
Vvolume of the system
vvelocity
Greek symbols
αpolarizability of the ion
ρnumber density
σBMH repulsive parameters
Subscripts
iion i
jion j
kslab k
xdirection x
ydirection y

Abbreviations

BMHBorn–Mayer–Huggins
CNcoordination number
PIMpolarizable ion model
RDFpartial radial distribution function
RIMrigid ion model

References

  1. Li, C.; Li, P.; Wang, K.; Molina, E.E. Survey of properties of key single and mixture halide salts for potential application as high temperature heat transfer fluids for concentrated solar thermal power systems. AIMS Energy 2014, 2, 133–157. [Google Scholar] [CrossRef]
  2. Cheng, J. Study on Molten Salt Thermophysical Properties for Heat Transfer and Storage. Ph.D. Thesis, Graduate School of Chinese Academy of Sciences, Beijing, China, 2014. [Google Scholar]
  3. Ding, W.; Bonk, A.; Bauer, T. Molten chloride salts for next generation CSP plants: Selection of promising chloride salts & study on corrosion of alloys in molten chloride salts. In Proceedings of the International Conference on Concentrating Solar Power and Chemical Energy Systems, Casablanca, Morocco, 2–5 October 2018. [Google Scholar] [CrossRef]
  4. Long, L. Magnesium resources and utilization of Qinghai salt lake. Guangdong Chem. Ind. 2011, 38, 85–86. [Google Scholar]
  5. McGreevy, R.L. Experimental studies of the structure and dynamics of molten alkali and alkaline earth halides. Solid State Phys. 1987, 40, 247–325. [Google Scholar] [CrossRef]
  6. Biggin, S.; Enderby, J.E. The structure of molten zinc chloride. J. Phys. C Solid State Phys. 2000, 14, 3129. [Google Scholar] [CrossRef]
  7. Zeidler, A.; Salmon, P.S.; Martin, R.A.; Usuki, T.; Mason, P.E.; Cuello, G.J.; Kohara, S.; Fischer, H.E. Structure of liquid and glassy ZnCl2. Phys. Rev. B 2010, 82. [Google Scholar] [CrossRef]
  8. Pavlatou, E.A.; Madden, P.A.; Wilson, M. The interpretation of vibrational spectra of ionic melts. J. Chem. Phys. 1997, 107, 10446–10457. [Google Scholar] [CrossRef]
  9. Takagi, Y.; Nakamura, T. X-Ray-diffraction Analysis of the Molten ZnCl2-KCl system. J. Chem. Soc. 1985, 81, 1901–1912. [Google Scholar] [CrossRef]
  10. Yoko, T.; Crescent, R.; Tsukagoshi, Y.; Ejima, T. Density Measurements in the molten ZnCl2-MCl (M=Li, Na, K, Cs) binary systems. J. Jpn. Inst. Met. 1978, 42, 1179–1186. [Google Scholar] [CrossRef]
  11. Mohan, G.; Venkataraman, M.; Gomez-Vidal, J.; Coventry, J. Assessment of a novel ternary eutectic chloride salt for next generation high-temperature sensible heat storage. Energy Conv. Manag. 2018, 167, 156–164. [Google Scholar] [CrossRef]
  12. Janz, G.J.; Allen, C.B.; Bansal, N.P. Physical properties data compilations relevant to energy storage. II. Molton salts: Data on single and multi-component salt systems. In STIN; National Standard Reference Data System: Gaithersburg, MD, USA, 1979. [Google Scholar]
  13. Janz, G.J.; Tomkins, R. Physical properties data compilations relevant to energy storage. IV. Molton salts: Data on additional single and multi-component salt systems. Natl. Bureau Stand. 1981, 861, 635–647. [Google Scholar]
  14. Janz, J.G. Molten salts handbook. J. Chem. Edu. 1969, 46, A550. [Google Scholar] [CrossRef]
  15. Nagasaka, Y.; Nakazawa, N.; Nagashima, A. Experimental determination of the thermal diffusivity of molten alkali halides by the forced Rayleigh scattering method. I. Molten LiCl, NaCl, KCl, RbCl, and CsCl. Int. J. Thermophys. 1992, 13, 555–574. [Google Scholar] [CrossRef]
  16. Anagnostopoulos, A.; Alexiadis, A.; Ding, Y. Molecular dynamics simulation of solar salt (NaNO3-KNO3) mixtures. Solar Energy Mater. Solar Cells 2019, 200. [Google Scholar] [CrossRef]
  17. Wang, J.; Yu, J.; Sun, Z.; Lu, G. Molecular Dynamics Simulations of the Local Structures and Transport Coefficients of Molten Alkali Chlorides. J. Phys. Chem. B 2014, 118, 10196–10206. [Google Scholar] [CrossRef] [PubMed]
  18. Tosi, M.P.; Fumi, F.G. Ionic sizes and born repulsive parameters in the NaCl-type alkali halides—II The generalized Huggins-Mayer form. Phys. Gem. Solids 1964, 25. [Google Scholar] [CrossRef]
  19. Ding, J.; Pan, G.; Du, L.; Lu, J.; Wei, X.; Li, J.; Wang, W.; Yan, J. Theoretical prediction of the local structures and transport properties of binary alkali chloride salts for concentrating solar power. Nano Energy 2017, 39, 380–389. [Google Scholar] [CrossRef]
  20. Xu, T.; Li, X.; Guo, L.; Wang, F.; Tang, Z. Powerful predictability of FPMD simulations for the phase transition behavior of NaCl-MgCl2 eutectic salt. Solar Energy 2020, 209, 568–575. [Google Scholar] [CrossRef]
  21. Liang, W.; Lu, G.; Yu, J. Composition-dependent microstructure evolution in liquid MgCl2-KCl: A first-principles molecular dynamics study. J. Mol. Liq. 2020, 309. [Google Scholar] [CrossRef]
  22. Müller-Plathe, F. A simple nonequilibrium molecular dynamics method for calculating the thermal conductivity. J. Chem. Phys. 1997, 106, 6082–6085. [Google Scholar] [CrossRef]
  23. Jayaraman, S.; Thompson, A.P.; Lilienfeld, O.A.V.; Maginn, E.J. Molecular simulation of the thermal and transport properties of three alkali nitrate salts. Ind. Eng. Chem. Res. 2010, 49, 559–571. [Google Scholar] [CrossRef]
  24. Pan, G.-C.; Ding, J.; Wang, W.; Lu, J.; Li, J.; Wei, X. Molecular simulations of the thermal and transport properties of alkali chloride salts for high-temperature thermal energy storage. Int. J. Heat Mass Transf. 2016, 103, 417–427. [Google Scholar] [CrossRef]
  25. Huang, S.; Tang, B.; Ma, Y.; Xu, C. Computerized simulation of molten MgCl2 by molecular dynamics method. Acta Metall. Sin. 1994, 8, 156–158. [Google Scholar]
  26. Gardner, P.J.; Heyes, D.M. Molecular dynamics computer simulations of molten zinc chloride. Physical B 1985, 131, 227–233. [Google Scholar] [CrossRef]
  27. Rollet, A.-L.; Salanne, M. Studies of the local structures of molten metal halides. Annu. Rep. Prog. Chem. 2011, Sect. C, 88–123. [Google Scholar] [CrossRef]
  28. Wilson, M.; Madden, P.A. Polarization effects in ionic systems from first principles. J. Phys. Condens. Matter. 1993, 5, 2687–2706. [Google Scholar] [CrossRef]
  29. Salanne, M.; Rotenberg, B.; Jahn, S.; Vuilleumier, R.; Simon, C.; Madden, P.A. Including many-body effects in models for ionic liquids. Theoret. Chem. Accoun. 2012, 131, 1–16. [Google Scholar] [CrossRef] [Green Version]
  30. Sharma, B.K.; Wilson, M. Intermediate-range order in molten network-forming systems. Univ. Coll. London 2009, 73, 060201. [Google Scholar] [CrossRef] [Green Version]
  31. Adams, D.J.; McDonald, I.R. Rigid-ion models of the interionic potential in the alkali halides. J. Phys. C Solid State Phys. 1974, 7, 2761–2775. [Google Scholar] [CrossRef]
  32. Seo, W.G.; Matsuura, H.; Tsukihashi, F. Calculation of phase diagrams for the FeCl2, PbCl2,and ZnCl2 binary system by using molecular dynamics simulation. Metall. Mater. Trans. B 2006, 37, 239–251. [Google Scholar] [CrossRef]
  33. Yuen, P.S.; Murfitt, R.M.; Collin, R.L. Interionic forces and ionic polarization in alkaline earth halide crystals. J. Chem. Phys. 1974, 61, 2383–2393. [Google Scholar] [CrossRef]
  34. Takagi, R.; Hutchinson, F.; Madden, P.A.; Adya, A.K.; Gaune-Escard, M. The structure of molten DyCl3 and DyNa3Cl6 simulated with polarizable- and rigid-ion models. J. Phys. Condens. Matter. 1999, 11, 645. [Google Scholar] [CrossRef]
  35. Ishii, Y.; Kasai, S.; Salanne, M. Transport coefficients and the Stokes–Einstein relation in molten alkali halides with polarisable ion model. Mol. Phys. 2015, 113, 2442–2450. [Google Scholar] [CrossRef]
  36. Tang, K.T.; Toennies, J.P. An improved simple model for the van der Waals potential based on universal damping functions for the dispersion coefficients. J. Chem. Phys. 1984, 132, 154104. [Google Scholar] [CrossRef]
  37. Zhang, W. Extensive influcence of ionic polarization on matter properties. J. Shangrao Teach. Coll. 2000, 20, 55–59. [Google Scholar] [CrossRef]
  38. Plimpton, S. Fast parallel algorithms for short-range molecular dynamics. J. Comput. Phys. 1995, 117, 1–19. [Google Scholar] [CrossRef] [Green Version]
  39. Hutter, J.; Iannuzzi, M.; Schiffmann, F.; Vandevondele, J. CP2K: Atomistic simulations of condensed matter systems. Computat. Mol. Sci. 2014, 4. [Google Scholar] [CrossRef] [Green Version]
  40. Allen, M.P.; Jildesley, T.D. Computer Simulation of Liquids; Oxford University Press: Oxford, UK, 1987. [Google Scholar] [CrossRef] [Green Version]
  41. Lantelme, F.; Turq, P.; Quentrec, B.; Lewis, J.W.E. Application of the molecular dynamics method to a liquid system with long range forces (Molten NaCl). Mol. Phys. 1974, 28, 1537–1549. [Google Scholar] [CrossRef]
  42. Rong, Z.; Pan, G.; Lu, J.; Liu, S.; Ding, J.; Wang, W.; Lee, D.-J. Ab-initio molecular dynamics study on thermal property of NaCl–CaCl2 molten salt for high-temperature heat transfer and storage. Renew. Energy 2020, 163, 579–588. [Google Scholar] [CrossRef]
  43. Yu, S.; Chu, R.; Li, X.; Wu, G.; Meng, X. Nonequilibrium Molecular Dynamics Simulations of Coal Ash. Energies 2020, 14, 11. [Google Scholar] [CrossRef]
  44. Artsdalen, E.R.V.; Yaffe, I.S. Electrical conductance and density of molten salt systems: KCl-LiCl, KCl-NaCl and KCl-KI. Phys. Chem. 1995, 59. [Google Scholar] [CrossRef]
  45. Angell, C.A.; WONG, J. Structure and glass transition thermodynamics of liquid zinc chloride from far-infrared, raman, and probe ion electronic and vibrational spectra. J. Chem. Phys. 1970, 53, 2053. [Google Scholar] [CrossRef]
  46. Cubicciotti, D.; Eding, H. Heat contents of molten zinc chloride and bromide and the molecular constants of the gases. J. Chem. Phys. 1964, 40, 978. [Google Scholar] [CrossRef]
  47. Singh, D.; Kim, T.; Zhao, W.; Yu, W.; France, D.M. Development of graphite foam infiltrated with MgCl2 for a latent heat based thermal energy storage (LHTES) system. Renew. Energy 2016, 94, 660–667. [Google Scholar] [CrossRef] [Green Version]
  48. Cornwell, K. The thermal conductivity of molten salts. J. Phys. D Appl. Phys. 1971, 4, 441–445. [Google Scholar] [CrossRef]
  49. Mackenzie, J.D.; Murphy, W.K. Structure of glass-forming halides. II. Liquid zinc chloride. J. Chem. Phys. 1960, 33, 366–369. [Google Scholar] [CrossRef]
Figure 1. PIM simulated results of RDF: (i) The RDF of cation–anion in PIM (a) NaCl; (b) MgCl2; (c) ZnCl2. (ii) The RDF of cation–cation in PIM (d) NaCl; (e) MgCl2; (f) ZnCl2.
Figure 1. PIM simulated results of RDF: (i) The RDF of cation–anion in PIM (a) NaCl; (b) MgCl2; (c) ZnCl2. (ii) The RDF of cation–cation in PIM (d) NaCl; (e) MgCl2; (f) ZnCl2.
Energies 14 00746 g001
Figure 2. Coordination number in different temperature of molten chloride salts: (a) NaCl; (b) MgCl2; (c) ZnCl2.
Figure 2. Coordination number in different temperature of molten chloride salts: (a) NaCl; (b) MgCl2; (c) ZnCl2.
Energies 14 00746 g002
Figure 3. Simulated RDF and CN of molten chloride salts in PIM and RIM: (a) MgCl2 at 1000 K; (b) ZnCl2 at 600 K.
Figure 3. Simulated RDF and CN of molten chloride salts in PIM and RIM: (a) MgCl2 at 1000 K; (b) ZnCl2 at 600 K.
Energies 14 00746 g003
Figure 4. Snapshots of ion local structures in molten chloride salts: (a) MgCl2 at 1000 K, Mg2+ in orange, Cl in green; (b) ZnCl2 at 600 K, Zn2+ in blue, Cl in green.
Figure 4. Snapshots of ion local structures in molten chloride salts: (a) MgCl2 at 1000 K, Mg2+ in orange, Cl in green; (b) ZnCl2 at 600 K, Zn2+ in blue, Cl in green.
Energies 14 00746 g004
Figure 5. Simulated and experimental density results of molten chloride salts. (a) NaCl; (b) MgCl2; (c) ZnCl2.
Figure 5. Simulated and experimental density results of molten chloride salts. (a) NaCl; (b) MgCl2; (c) ZnCl2.
Energies 14 00746 g005
Figure 6. Enthalpies of MgCl2 and ZnCl2 with temperature in PIM.
Figure 6. Enthalpies of MgCl2 and ZnCl2 with temperature in PIM.
Energies 14 00746 g006
Figure 7. Mean temperature gradient in different temperature with kinetic energy swap rate: (a) Different swap rates; (b) MgCl2 at 400 time steps; (c) ZnCl2 at 400 time steps.
Figure 7. Mean temperature gradient in different temperature with kinetic energy swap rate: (a) Different swap rates; (b) MgCl2 at 400 time steps; (c) ZnCl2 at 400 time steps.
Energies 14 00746 g007
Figure 8. Length dependence of thermal conductivities on Lz.
Figure 8. Length dependence of thermal conductivities on Lz.
Energies 14 00746 g008
Figure 9. Thermal conductivities of molten chloride salts in different temperatures.
Figure 9. Thermal conductivities of molten chloride salts in different temperatures.
Energies 14 00746 g009
Figure 10. PIM simulated results of MSDs in different temperature: (a) Na ion and Cl ion; (b) Mg ion and Cl ion; (c) Zn ion and Cl ion.
Figure 10. PIM simulated results of MSDs in different temperature: (a) Na ion and Cl ion; (b) Mg ion and Cl ion; (c) Zn ion and Cl ion.
Energies 14 00746 g010
Figure 11. PIM simulated results of self-diffusion coefficients in different temperature: (a) Na ion and Cl; (b) MgCl2 and ZnCl2.
Figure 11. PIM simulated results of self-diffusion coefficients in different temperature: (a) Na ion and Cl; (b) MgCl2 and ZnCl2.
Energies 14 00746 g011
Table 1. Potential parameters of RIM for molten chloride salts.
Table 1. Potential parameters of RIM for molten chloride salts.
ModelPairAij (kJ·mol−1)Bij−1)σij(Å)Cij6·kJ·mol−1)Dij8·kJ·mol−1)
RIMNa+-Na+25.44353.15462.340101.17248.1771
Na+-Cl20.35483.15462.755674.479837.088
Cl-Cl15.26613.15463.1706985.6814,031.6
Mg2+-Mg2+0.543927.69201.3202.22300
Mg2+-Cl0.606686.89602.620200.4030
Cl-Cl0.669446.25003.92018,066.10
Zn2+-Zn2+0.502088.33331.20000
Zn2+-Cl0.585767.14292.56000
Cl-Cl0.669446.25003.92018,066.10
Table 2. Potential parameters of PIM for molten chloride salts.
Table 2. Potential parameters of PIM for molten chloride salts.
ModelPairAij
(kJ·mol−1)
Bij
−1)
b i j 6 & b i j 8
−1)
Cij
6·kJ·mol−1)
Dij
8·kJ·mol−1)
b i j 4 & b j i 4
−1)
c j 4 & c j i 4
PIMNa+-Na+2625.59.44863.2125674.536836.27900
Cl-Na+177,2213.26163.21252732.733023.843.3263.0, 0.7
Cl-Cl722,2753.39583.21258071.374520.4300
Mg2+-Mg2+70,888.52.94122.9412126.836000
Mg2+-Cl135,7912.94122.94121091.9403.1181.4, 1.4
Cl-Cl165,4072.94122.941210,550.4000
Zn2+-Zn2+51,485.82.94122.94122306.11000
Cl-Zn2+132,6692.94122.94125154.1402.9291.0, 1.0
Cl-Cl192,9142.94122.941211,530.5000
α N a + = 0.90 α . u . , α M g 2 + = 0.49 α . u . , α Z n 2 + = 2.85 α . u . , α C l = 20.0 α . u .
Table 3. Details of simulations for chloride salts at solid state.
Table 3. Details of simulations for chloride salts at solid state.
SaltsUnit Cell ParametersSpace
Group
Number of IonsCode and Model
a (Å)b (Å)c (Å)α (°)β (°)γ (°)
NaCl5.6425.6425.642909090Fm-3m8192LAMMPS (RIM)
1024CP2K (PIM)
MgCl24.0604.0604.188909090P-4m28640LAMMPS (RIM)
1080CP2K (PIM)
ZnCl25.4005.40010.350909090I-42d6144LAMMPS (RIM)
768CP2K (PIM)
Table 4. First peak positions of RDF for molten chloride salts.
Table 4. First peak positions of RDF for molten chloride salts.
SaltPairT/KPIM/ÅEXP/ÅCalculation Error
NaClNa+-Na+11004.053.90 [5]3.8%
15004.14
Na+-Cl11002.672.76 [5]3.3%
15002.64
MgCl2Mg2+-Mg2+10003.953.81 [25]3.7%
15004.03
Mg2+-Cl10002.522.42 [25]4.1%
15002.50
ZnCl2Zn2+-Zn2+6003.793.80 [6]0.3%
10003.86
Zn2+-Cl6002.492.29 [6]8.8%
10002.45
Table 5. First peak positions of cation–cation for molten chloride salts.
Table 5. First peak positions of cation–cation for molten chloride salts.
PairT/KEXP/ÅRIM/ÅRIM ErrorPIM/ÅPIM Error
Na+-Na+11003.90 [5]4.105.10%4.053.80%
Mg2+-Mg2+10003.81 [25]4.3814.9%3.953.70%
Zn2+-Zn2+6003.80 [6]4.3013.2%3.790.30%
Table 6. Densities of molten chloride salts in experiment and simulation.
Table 6. Densities of molten chloride salts in experiment and simulation.
SaltsRange (K)EXP RIM ErrorPIM Density Function (g/cm3)PIM Error
NaCl1080–1500Ref [44]6.1~6.3% ρ = 2.145 5.439 × 10 - 4 T 0.1~0.6%
MgCl21000–1500Ref [14]12.5~19.4% ρ = 1.976 4.614 × 10 - 4 T 9.2~9.8%
ZnCl2600–1000Ref [45]0.5~8.8% ρ = 2.836 5.538 × 10 - 4 T 0.7~1.6%
Table 7. Heat capacities of molten chloride salts in experiment and simulation.
Table 7. Heat capacities of molten chloride salts in experiment and simulation.
SaltsEXP
(J·mol−1·K−1)
RIM
(J·mol−1·K−1)
Error of
RIM
PIM
(J·mol−1·K−1)
Error of
PIM
NaCl67.0 [24]68.21.9%65.52.2%
MgCl292.5 [14]128.438.8%99.77.8%
ZnCl2100.9 [46]134.132.9%111.710.7%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lu, J.; Yang, S.; Pan, G.; Ding, J.; Liu, S.; Wang, W. Thermal and Transport Properties of Molten Chloride Salts with Polarization Effect on Microstructure. Energies 2021, 14, 746. https://doi.org/10.3390/en14030746

AMA Style

Lu J, Yang S, Pan G, Ding J, Liu S, Wang W. Thermal and Transport Properties of Molten Chloride Salts with Polarization Effect on Microstructure. Energies. 2021; 14(3):746. https://doi.org/10.3390/en14030746

Chicago/Turabian Style

Lu, Jianfeng, Senfeng Yang, Gechuanqi Pan, Jing Ding, Shule Liu, and Weilong Wang. 2021. "Thermal and Transport Properties of Molten Chloride Salts with Polarization Effect on Microstructure" Energies 14, no. 3: 746. https://doi.org/10.3390/en14030746

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop