Processing math: 26%
Xin Song, Shaolong Li, Shanshan Liu, Yong Fan, Jilin He, and Jianxun Song, Coordination states of metal ions in molten salts and their characterization methods, Int. J. Miner. Metall. Mater., 30(2023), No. 7, pp.1261-1277. https://dx.doi.org/10.1007/s12613-023-2608-7
Cite this article as: Xin Song, Shaolong Li, Shanshan Liu, Yong Fan, Jilin He, and Jianxun Song, Coordination states of metal ions in molten salts and their characterization methods, Int. J. Miner. Metall. Mater., 30(2023), No. 7, pp.1261-1277. https://dx.doi.org/10.1007/s12613-023-2608-7
Invited Review

Coordination states of metal ions in molten salts and their characterization methods

Author Affilications
  • Corresponding author:

    Jianxun Song      E-mail: jianxun.song@zzu.edu.cn

  • The macroscopic characteristics of molten salts are governed by their microstructures. Research on the structures of molten salts provides the foundation for a full understanding of the physicochemical properties of molten salts as well as a deeper analysis of the microscopic electrolysis process in molten salts. Information about the microstructure of matter can be obtained with the help of several speculative and experimental procedures. In this review, the advantages and disadvantages of the various test procedures used to determine the microstructures of molten salts are compared. The typical coordination configurations of metal ions in molten salt systems are also summarized. Furthermore, the impact of temperature, anions, cations, and metal oxides (O2−) on the structures of molten salts is discussed in detail. The accuracy and completeness of the information on molten salt structures need to be investigated by the integration of multiple methods and interdisciplinary fields. Information on the microstructure and coordination of molten salts deepens the understanding of the elementary elements of the microstructure of matter. This paper, which is based on the review of the coordination states of metal ions in molten salts, is hoped to inspire researchers to explore the inter-relationship between the microstructure and macroscopic properties of materials.
  • The comprehension of matter and selected materials by researchers is based on the macroscopic physical characteristics of matter. The types of atoms and molecules present in a material and how these structures interact to form the microscopic structure, which in turn determines the macroscopic properties of the material. High-temperature molten salts are utilized frequently because of their high conductivity, wide electrochemical window, quick mass transfer, and other benefits [1]. They are used in a variety of fields, including electrochemical metallurgy, the recovery and utilization of secondary resources, functional materials, and the preparation of rechargeable batteries, wherein they can serve as an excellent medium for chemical reactions, energy transfer, and storage [2]. Numerous techniques for measuring the microstructures of molten salt, which can reflect the microstructures of substances from different angles, exist. The earliest investigation of the architecture of molten salts involved thermodynamic calculations [3]. The thermodynamic calculation method mainly uses calorimetry to determine the thermal properties of a molten salt system to obtain the structural information of the system indirectly. The theory of molten salt structure has been rapidly developed with the application of X-ray diffractometry [4], Raman spectroscopy [5], and nuclear magnetic resonance (NMR) [6]. Further research has been done on the microstructures and macroscopic thermodynamic characteristics of molten salts.

    The study of the structures and characteristics of molten salts has been encouraged by the progressive advancement of spectroscopic technology and computer simulation technology and has led to a progression from a shallow to a profound understanding of molten salts. However, a comprehensive knowledge of the microstructures of molten salts has not yet been attained due to the intricacy of molten salts and the variable ion interactions in each molten salt. The findings of current studies remain limited, and considerable effort needs to be made to understand the structure of molten salts better. The test methods for obtaining the microstructures of molten salts and the comparison of their advantages and disadvantages are summarized in this paper. The typical coordination structures of metal ions in molten salt systems are also listed. In addition, research on the influencing factors of the structures of molten salts, such as temperature, anions, cations, and metal oxides (O2−), will provide a reference for studies on the coordination structures of molten salts.

    Given the particularity of the experimental conditions of high-temperature molten salts, the analytical methods used to study their structures are different from those used to analyze the structures of other substances at room temperature. The main methods are described below.

    The structure of molten salts can be analyzed and studied by using the excellent method of X-ray diffraction (XRD). When salts are melted, volume expansion exerts a diffusion effect on the XRD peaks of the original crystals. Consequently, the structural characteristics of the molten salt cannot be directly determined by XRD. The distribution function of the molten salt can be deduced by measuring the intensity of diffracted light.

    Structural information, such as coordination numbers (CNs), average interatomic distances, average atomic displacements, and ordered domain sizes, can be derived from distribution functions. Thus, the microstructure of molten salts, the formation mechanism of ionic clusters in molten salts, and their evolution can be analyzed by means of high-temperature XRD.

    Iwadate et al. [7] found that the CN of La–Cl was close to six when calculated by XRD and approximately seven when determined by X-ray absorption fine structure spectroscopy. These findings suggested that La3+ is in an octahedral-like structure surrounded by 6 or 7 Cl.

    High-temperature XRD can perform in situ measurement, and its temperature can also reach 1500°C or even 1600°C. However, the amount of information it obtains is limited. For molten disordered states, high-temperature XRD can yield the bond length, chemical bond, CN, and other information of the first nearest neighbor but cannot easily and effectively obtain the information of neighbors further away.

    Raman spectroscopy can acquire the structural information of a substance by detecting the vibration and rotational energy levels of a molecule [5]. It is based on higher-order photon–molecule interactions and is considerably weaker than infrared absorption spectroscopy. It is suitable for the detection of symmetrical molecules without polarity.

    Wang et al. [8] investigated the specifications of FLiNaK–ScF3 and FLiNaK–ScF3–Li2O melts at temperatures above 600°C by using Raman spectroscopy and also conducted density flooding theory (DFT) calculations. Dracopoulos et al. [9] studied a series of KF–LnF3 (Ln = La, Ce, Nd, Sm, Dy, Yb) molten salt mixtures by utilizing Raman spectroscopy. Ma et al. [10] applied a combination of Raman spectroscopy and theoretical calculations to quantify the fluoride and fluorine oxide structures present in a K3AlF6–Al2O3 system and established a method for the quantitative analysis of molten salt structures by Raman spectroscopy.

    In the analysis process, Raman spectroscopy has the advantages of simple operation, short determination time, and high sensitivity. However, it can only study the structural information of the corresponding group qualitatively, and the structural information of the corresponding group also needs further study on the basis of specific signals.

    The neutron diffraction technique (NDT) is a crystallographic method that is used to determine the atomic or magnetic structure of a material. This technique is similar to XRD, with the main difference being that the source of the radiation is different. The two techniques can complement each other. Given that neutrons are uncharged and have a strong penetration ability, samples can be determined in special containers. This approach is suitable for studying unusual samples, such as liquid, at high temperatures. Considering that different isotopes have different scattering lengths, an individual partial structure factor can be determined directly by isotope substitution [11].

    Mitchell et al. [12] analyzed the ion pair distribution functions of RbCl molten salts by employing the NDT isotope substitution technique and found that Rb–Cl, Rb–Rb, and Cl–Cl had CNs of (6.9 ± 0.3), (13.0 ± 2.0), and (14.0 ± 2.0), respectively, and coordination distances of 3.18, 4.86, and 4.80 Å, respectively. Locke et al. [13] measured and reanalyzed the biased radial distribution functions of NaCl, RbCl, and CsCl molten salts by applying the isotope substitution method of neutron diffraction. They found that Na–Cl, Rb–Cl, and Cs–Cl had CNs of 4.83, 7.4, and 5.8, respectively, and ion spacings of 2.76, 3.2, and 3.4 Å, respectively.

    NDT has numerous advantages, including sensitivity to light atoms, the ability to distinguish isotopes, the lack of radiation damage, and a penetration depth of several centimeters. Its main disadvantage is that its crystal cell parameters are inaccurate, and it is inconvenient to use because it requires a nuclear reactor.

    NMR techniques have been proven to be a powerful tool for studying local structures around selected nuclei, cations, or anions. Its recent development for application at high temperatures has now enabled the study of mass melting systems and a more precise description of microstructures based on the properties of different species, average coordination, or first neighbors.

    NMR measures [6] the frequency shift of the nucleus of an isotope due to different chemical environments, namely, the chemical shift δ, which reflects the chemical information of the nucleus and its microscopic environment. It can be used to distinguish the microscopic structure of a substance.

    Liu et al. [14] used 29Si and 27Al magnetic angle spinning NMR spectra to systematically study the thermal activation mechanism of silica alumina materials. Rollet et al. [15] investigated the local structures of molten YF3–LiF and LaF3–LiF binary systems by utilizing NMR technology. They also performed high-temperature NMR (HT NMR) experiments on a LaF3–AF (A = Li, Na, K, Rb, Cs)–CaO mixture [16] to track the modification of the first rare earth coordination shell after oxide addition.

    NMR technology has obvious advantages in resolving the microstructures of molten salts due to its high signal resolution and selective observation of characteristic elements. However, its weakness lies in its very poor sensitivity for structure capture when a sample is melted into the liquid state. Moreover, the rapid dynamic structure information of the liquid state is averaged such that the measured information is actually distorted and the information is inaccurate.

    Extended X-ray absorption fine structure spectra (EXAFS) can reflect structural information, including the near-neighboring atom centered on the absorption of the atom, which can be applied to the determination of amorphous structures [17]. The application of this technique to molten salt systems has been well studied by Okamoto et al. [18]. In addition to using the EXAFS technique directly to determine the microstructure of molten salts, Okamoto [19] also proposed calculating EXAFS spectra by extracting atomic snapshots obtained from molecular dynamics (MD) calculations, pointing out that this method is more rigorous and provides reliable structural information.

    Through EXAFS, Okamoto et al. [20] calculated the CN of La–Cl to be (7.4 ± 0.5), which suggested the presence of LaCl3+6, LaCl47, or LaCl58 groups in the molten salt, with each unit group linked by a bridging Cl angle. Zissi et al. [21] deduced from Raman spectroscopy that La–Cl is a six-coordinated octahedral structure.

    EXAFS has a high sensitivity to local atomic displacement, element characteristics, and vibration mechanics (accuracy above 0.01 Å). This technique is more suitable for in situ studies on structural transformations because it can be performed under a wide range of conditions. However, it is insensitive to stereoscopic structures.

    The main electrochemical methods (EC) used in experimental procedures are cyclic voltammetry (CV), square wave voltammetry (SWV), chronopotentiometry (CP), and chronoamperometry (CA).

    (1) CV is one of the most widely used electrochemical measurement methods [22]. The basic principle is to apply the triangular waveform pulse voltage to the closed loop formed by the working electrode and the opposite electrode, and change the potential on the working electrode/electrolyte interface at a certain rate. By forcing the active substance on the working electrode to undergo oxidation/reduction reaction, the response current of electrochemistry on the electrode can be obtained.

    (2) SWV is a large-scale differential technique that superposes the symmetric square wave and the step voltage on the excitation signal of the working electrode [23]. It is a voltammetry method in which a rapid scanning step voltage is applied to the working electrode, and a small amplitude square wave is superimposed on each step to represent the potential–current relationship.

    (3) CP is a transient measurement method employed to control the potential step size [24]. It controls the current flowing through the electrolytic cell and records the change in the potential with time to enable the study of the reaction mechanism.

    (4) CA records the change in the current flowing through the system with time by controlling the potential of the electrolytic cell and obtains the current–time curve.

    Zhu et al. [25] investigated the morphology and behavior of chromium ions in LiCl–KCl–CrF3 and LiCl–KCl–LiF–CrF3 melts by utilizing CV, SWV, chronopotentiometry, and Raman spectroscopy. Their results showed that the electrochemical reduction of Cr(III) into Cr(0) in LiCl–KCl–CrF3 was a two-step process mediated by Cr(II) generation. In the molten salt medium, Cr(III) ions were present in the form of CrCl36, and CrCl6xF3x (x ≤ 3) ions were less abundant. According to the CV results obtained at different potential scan rates, the diffusion coefficient of chromium ions in LiCl–KCl–CrF3 was higher than that in LiCl–KCl–CrCl3. Wu et al. [26] investigated the electrochemical reduction mechanism and electrocrystallization process of vanadium ions in the NaKClF system by applying CV, chronopotentiometry, and chronocurrent methods. Their results showed that the reduction reaction of VO3 ions in the NaCl–KCl–NaF–V2O5 (abbreviated as NaKClF–V2O5) system was a reversible reaction involving five electrons in one step.

    The advantages of the electrochemical method are its simplicity, high sensitivity, and good real-time performance. However, this method cannot directly detect the microstructure of molten salts and thus needs to be combined with other methods.

    The development of computer technology has enabled the simulation of microstructures on computers. The main computer simulation methods (Sim) are as follows [27].

    (1) Quantum chemical (QC) essentially solves Schrödinger’s equation and thus obtains the particle motion laws and parameters necessary to derive the other physical and chemical properties of molten salts [28]. QC makes direct use of fundamental principles and does not rely on empirical values, and it acquires very accurate results. Li et al. [29] studied the dynamic fluctuations of the U3+ coordination structure in molten LiCl–KCl mixtures by using first-principles MD simulations. They calculated the radial distribution function and probability distribution of the CN. Their results revealed that the coordination structure of U3+ in the molten mixture of LiCl–KCl–UCl3 was dominated by a six-fold coordination structure.

    (2) Monte Carlo (MC) is a statistical method that uses discrete random sampling [30]. It allows atoms to jump randomly in search of the lowest energy state to identify the properties of the material and to deduce the structure of the molten salt. The MC method is a random method, and its statistical results are ensemble-averaged. Baranyai et al. [31] calculated the structural parameters of a series of alkali metal chloride molten salts with MC simulation. They then analyzed the structural characteristics of alkali metal chlorides by combining spectral diffraction techniques and the inverse MC method and found that the CNs of the molten salts LiCl, NaCl, and CsCl were 5.2, 4.85, and 6.1, respectively, which were then used as criteria.

    (3) MD was developed on the basis of MC [32]. After the equation of motion is simulated, the time-dependent properties of the molten salt system can be obtained by solving this equation. However, no way to obtain the nuclear magnetic properties of the molten salt system exists. The MD method is a deterministic method whose statistical results are based on time averages. Bessada et al. [3334] conducted a series of studies on the structures of metal fluorides, such as ZrF4, ThF4, and LaF3, in metal fluorides and analyzed the experimental data with MD simulation calculations. He et al. [35] investigated the microstructure and dispersion behavior of LiF–BeF2 molten salt by using Car–Parrinello MD simulations.

    High-temperature molten salts have numerous microstructures. The microstructures of the material can be viewed from different angles and provide different information on the material. Fig. 1 shows the principles and available information on the characterization methods. All of these techniques enable the characterization of molten salt structures, though at various characteristic times and through differing processes. For example, XRD can only provide the total radial distribution function of a structure; Raman spectroscopy entails the vibration of a unit of known symmetry; NDT yields only the partial radial distribution function of a structure; NMR involves the modification of the magnetic field experienced by a specific nucleus as a result of its neighbors; EXAFS entails the diffraction of a photoelectron by nearby nuclei. These methods are thus complementary, and the local structure can be better described by comparing their results. In practical applications, multiple test sections can be used to verify and fill each other, and the microformation information of molten salt can be perfected on the whole surface. A comparison of the structural characterization methods is shown in Table 1.

    Fig. 1.  Principle and obtained information of characterization methods. In (a), d is the crystal plane spacing, θ is the included angle between the incoming ray, the reflected ray and the reflected crystal plane, λ is the wavelength, n is the reflection order; in (b), λ is the laser wavelength; in (c), α is the scattering angle, and p is the scattering length; in (d), B0 is the main magnetic field; in (e), the red circle is the coordination layer; in (g), Ψ is the wave function of a physical system, i is the imaginary unit, ћ is reduced Planck’s constant, ∂/∂t is partial differential for time, m is the mass, 2 is the Laplace operator, and V is the potential energy distribution in the system.
    Table  1.  Comparison of structural characterization methods
    MethodAdvantagesDisadvantagesRefs.
    XRDIn situ measurement.
    Reveals the formation mechanism and evolution of ion clusters.
    The amount of information, such as valence and structure, obtained is relatively limited.[4]
    RamanDoes not need sample pretreatment.
    Simple operation, short determination time, and high sensitivity.
    Only the structural information of the corresponding group qualitatively.[5]
    NMRHigh signal resolution.
    Selective observation of characteristic elements.
    The sensitivity to capture structure is very poorly measured, and information is actually distorted.[6]
    NDTSensitive to light atoms.
    No radiation damage.
    Inconvenient to use and inaccurate cell parameters.
    Nuclear reactors are needed.
    [11]
    EXAFSHighly sensitive to local atomic displacement, element characteristics, and vibration mechanics.Insensitive to stereoscopic structures.[1719]
    ECSimple, sensitive, and real-time.Cannot direct microstructures directly.[2224]
    QCIndependent of empirical values and yields accurate results.Limited to the simulation of smaller systems (less than 100 atoms) and has high computer requirements.[28]
    MCProblems with statistical properties can be solved directly.
    Does not need to discretize the problem of continuity.
    Deterministic problems need to be transformed into stochastic problems.
    More steps are usually required to calculate N.
    [30]
    MDPhysical quantities related to time can be used to calculate transport properties.The nuclear magnetic properties of molten salt systems cannot be obtained.[32]
    Note: Refs. represents references.
     | Show Table
    DownLoad: CSV

    Molten salts are systems composed of ions, whether they are monomeric or pluralistic. The anions and cations in these systems are wrapped around each other and are randomly distributed. Common molten salts can be broadly classified into four categories: chloride salts, fluoride salts, nitrates, carbonates, and other molten salt systems. The typical coordination structures of metal ions in each molten salt system are discussed in the following sections.

    The microstructures of molten chloride salts have become widely studied with the rapid development of high-temperature detection techniques, such as XRD, Raman spectroscopy, and computer simulation. This section briefly introduces the results of existing research on chloride molten salt systems.

    Levy et al. [36] studied the structures of LiCl, KCl, and other chloride molten salts by using high-temperature liquid neutron diffraction and X-ray techniques and discovered that the ionic spacing of the high-temperature melt was smaller than that of the corresponding crystal at room temperature. They also found that the average CNs of ions were between 4 and 5 and that the ions were anisotropic within the first coordination layer. Førland et al. [37] studied the structural parameters of LiCl molten salts by applying MC computational simulation software and obtained a Li–Cl ionic spacing of 2.40 Å and CN of 4.3. Howe and McGreevy [38] investigated the microstructures of LiCl and NaCl molten salts by utilizing the isotope method and showed that Li–Cl in LiCl molten salts had an approximate tetrahedral structure. In addition, they calculated the CNs of Na–Na, Na–Cl, and Cl–Cl ion pairs to be (13.0 ± 0.5), 5.8, and (13.0 ± 0.5), respectively. Takag et al. [39] measured the radial distribution function of KCl molten salts via XRD and calculated the coordination distance and CN of K–Cl to be 3.05 Å and 4.1, respectively. Fig. 2 shows a snapshot of the chloride molten salt structure.

    Fig. 2.  Snapshot of the chloride molten salt structure: (a) model system for Eu ions solvated in a LiCl–KCl eutectic molten salt [40]; (b) ion distribution in the LiCl–KCl–SmCl3 system [41]; (c) ion distribution in the LiCl–KCl eutectic [42]; (d) ion distribution in molten LiCl–KCl–UCl3 [42]. (a) C. Kwon, S.H. Noh, H. Chun, I.S. Hwang, and B. Ha, Int. J. Energy Res., vol. 42, 2757-2765 (2018) [40]. Copyright John Wiley & Sons, Ltd. Reproduced with permission. (b) Reprinted from J. Mol. Liq., 363, J. Zhao, Z.T. Liu, W.S. Liang, and G.M. Lu, Evaluation of the local structure and electrochemical behavior in the LiCl–KCl–SmCl3 melt, art. No. 119818, Copyright 2022, with permission from Elsevier. (c, d) Reprinted from J. Mol. Liq., 234, J. Song, S.P. Shi, X.J. Li, and L.M. Yan, First-principles molecular dynamics modeling of UCl3 in LiCl–KCl eutectic, 279-286, Copyright 2017, with permission from Elsevier.

    Jiang et al. [43] studied the structural properties of CeCl3 in the LiCl–KCl–CeCl3 system of chloride molten salts by applying MD computational simulation software and obtained the relationship among density, composition, and temperature. Kwon et al. [40] analyzed the basic mechanism of the spontaneous reduction of Eu3+ into Eu2+ in eutectic LiCl–KCl molten salt by first-principles calculation and characterized the main structural features of the solvated shell on the basis of the radial distribution function and CNs. Zhao et al. [41] investigated the local structure of LiCl–KCl and LiCl–KCl–SmCl3 by using the radial distribution function, CN, and structure factor by performing QC simulation. They discovered that the average CNs of Li–Cl, K–Cl, and Cl–Cl pairs at 450°C were 4.33, 7.54, and 10.91, respectively. The first shell CN of Sm coordinated by Cl in LiCl–KCl–SmCl3 melt was 6.56. In addition, SmCl3 had little effect on the short-range order of the LiCl–KCl melt but affected the medium-range order distribution of the Li–Li melt. Akdeniz et al. [4445] studied the ionic structures of ThCl4, ZrCl4, and ThCl4–ZrCl4 molten salts by using Raman scattering.

    Okamoto et al. [4647] studied the microstructure of LaCl3 mixed with LaCl3–ACl (A for alkali metal) molten salts in the molten state by using X-ray absorption fine structure ( XAFS) and MD and found that four ionic structures of LaCl36, LaCl47, LaCl58, and LaCl69 existed simultaneously in the LaCl3 molten salt. They also mentioned that while the CN of the anion cluster decreased as the concentration of La3+ in the molten salt decreased, the ionic clusters formed were all octahedral in structure. Song et al. [42] used QC simulations to study the structures and UCl3 in the LiCl–KCl eutectic and calculated the corresponding CNs of Li–Cl, K–Cl, and Cl–Cl as 3.97–4.29, 6.69–7.32, and 9.99–11.39, respectively. In addition, in the molten mixture of LiCl–KCl–UCl3, the positions of the first peaks of Li–Cl, K–Cl, and U–Cl were 2.230–2.326, 3.024–3.092, and 2.202–2.726 Å, respectively, and the corresponding first CNs were 3.94–4.35, 6.66–7.73, and 6.13–6.35, respectively. A summary of the coordination structures of metal ions in molten chloride salts is shown in Table 2.

    Table  2.  Summary of the coordination structures of metal ions in molten chloride salts
    Molten saltsMethodsResultsRefs.
    LiCl, KCl, and othersND and XRDThe average CN of ions was between 4 and 5, and the ions were anisotropic within the first coordination layer.[36]
    LiClMCLi–Cl ionic spacing of 2.40 Å and CN of 4.3.[37]
    LiCl and NaClMMRLi–Cl in LiCl molten salts had an approximate tetrahedral structure. The CNs of Na–Na, Na–Cl, and Cl–Cl ion pairs were calculated to be (13.0 ± 0.5), 5.8, and (13.0 ± 0.5), respectively.[38]
    KClXRDK–Cl had a coordination distance of 3.05 Å and CN of 4.1.[39]
    LiCl–KCl and LiCl–KCl–SmCl3QCThe CN values of Li–Cl, K–Cl, and Cl–Cl pairs were 4.33, 7.54, and 10.91, and the first shell CN of Sm coordinated by Cl was 6.56.[41]
    ThCl4, ZrCl4, and ThCl4–ZrCl4Raman[4445]
    LaCl3–ACl (A for alkali metal)XAFS and MDLaCl36, LaCl47, LaCl58, and LaCl69 were present
    in the LaCl3 molten salt.
    [4647]
    LiCl–KCl, LiCl–KCl–UCl3QCThe CNs of Li–Cl, K–Cl, and Cl–Cl were calculated as 3.97–4.29, 6.69–7.32, and 9.99–11.39, respectively; the CNs of Li–Cl, K–Cl, and U–Cl were calculated to be 3.94–4.35, 6.66–7.73, and 6.13–6.35, respectively.[42]
     | Show Table
    DownLoad: CSV

    With the application of spectral technology and computer simulation technology to characterize the structures of high-temperature molten salts, the research on the microstructures of molten fluoride salt has made some progress, and the research of scholars on the ionic structure of fluoride molten salt systems in recent years is reviewed in this paper.

    Bulavin et al. [48] investigated the ionic structure of NaF–LiF–LnF3 (Ln = La, Nd) by using a high-temperature XRD technique and concluded that Ln was present in the form of LnF4 complexes. Cui et al. [49] studied the ionic structure present in the LaF3–LiF molten salt system by applying MD and concluded that the LaF36 anion cluster was present within this molten salt system. The structure of the LaF36 ion cluster underwent some deformation with the changes in the melt temperature or the concentration of LaF3 in the molten salt fraction. However, the CN of La–F was very close and was always between 7.33 and 7.75. Bessada et al. [50] investigated LiF–ThF4–UF4 by EXAFS experiments and MD simulations. The structure of the LiF–ThF4 eutectic with 4mol% UF4 is shown in Fig. 3(a). The results revealed a slight increment in the number of free fluoride ions due to the breakdown of the Th–F chain and the increase in the content of the complex (ThXUYFZ)4X+4Y−Z from 0 to 8mol%. In addition, the average CN of Th–F was very close to eight. Rollet et al. [51] investigated the ionic structure of lanthanide fluoride molten salts (LiF–LuF3/LaF3) by using HT NMR and EXAFS. They inferred that a strong LnF3xx ionophore was present in the LiF–LnF3 molten salt. Stefanidaki et al. [52] investigated the ionic structures of LiF–NdF3 by high-temperature Raman technology and concluded that NdF36 ionophores existed in LiF–NdF3. Hatem and Gaune-Escard [53] analyzed the ionic structure of the KF–NdF3 molten salt and discovered that NdF36 complex ionic groups were present in this molten salt. Hu et al. [54] hypothesized that NdF36 and NdF4 complexed anion clusters were present in LiF–NdF3. Dracopoulos et al. [9] utilized high-temperature Raman techniques to study the microstructure of the NdF3–LiF binary melt salt and found that the melted salt also contained ortho-octahedral NdF36 complexed anions.

    Fig. 3.  Snapshot of the fluoride molten salt structure: (a) atomic configuration of the LiF–ThF4 eutectic with 4mol% UF4 (Th4+ in green, U4+ in blue, and F in red) [50]; (b) the local ionic structure in the simulation box of the LiF–NaF–AlF3 molten salt with 9wt% LiF (Li+ in green, Na+ in blue, Al3+ in pink, and F in red) [56]; (c) the stable configuration of the KF–NaF–AlF3 system (18wt% KF; Na+ in blue, K+ in yellow, Al3+ in green and F in red) [57]; (d) the network structure of molten LiF–ZrF4 salts [60]. (a) Reprinted from J. Mol. Liq., 307, C. Bessada, D. Zanghi, M. Salanne, et al, Investigation of ionic local structure in molten salt fast reactor LiF-ThF4–UF4 fuel by EXAFS experiments and molecular dynamics simulations, art. No. 112927, Copyright 2020, with permission from Elsevier. (b) Reprinted from Chem. Phys. Lett., 706, C. Bessada, X.J. Lv, Z.X. Han, J.G. Chen, L.X. Jiang, Z.M. Xu, and Q.S. Liu, First-principles molecular dynamics study of ionic structure and transport properties of LiF–NaF–AlF3 molten salt, 237-242, Copyright 2018, with permission from Elsevier. (c) Reprinted from Chem. Phys. Lett., 730, H. Guo, J. Li, H.L. Zhang, et al., First-principles molecular dynamics investigation on KF–NaF–AlF3 molten salt system, 587-593, Copyright 2019, with permission from Elsevier. (d) Reprinted with permission from [O. Pauvert, D. Zanghi, M. Salanne, et al., J. Phys. Chem. B, vol. 114, 6472-6479 (2010) [60]]. Copyright 2010 American Chemical Society

    Dai et al. [55] investigated the ionic structure and vibrational spectra of LiF–BeF2 molten salts by the DFT method and discovered that the most likely ionic groups present in the molten salt were BeF24, Be2F37, and Be3F410. Moreover, the average F–Be and Be–Be distances were within the ranges of (1.564 ± 0.100) and (3.025 ± 0.200) Å, respectively, and the CN of F–Be was maintained at 4. He et al. [35] investigated the microstructure of LiF–BeF2 molten salts by MD. The LiF–BeF2 molten salt was observed to mainly consist of a network structure formed by the tetrahedral structure BeF24 and the aggregation of Li+. Multiple BeF24 formed Be–F–Be bonds through point sharing, yielding structures, such as Be2F37 and Be3F410. In addition, the Be–F ion pair had an average first peak radius of 1.58 Å and a CN of 4; the results obtained were consistent with the findings of Dai et al. [55].

    Lv et al. [56] studied the microstructure of LiF–NaF–AlF3 molten salt by using first-principles calculation and discovered that the complex anion groups present in the melt were AlF4, AlF25, and AlF36. Fig. 3(b) shows the local ionic structure in the simulation box for LiF–NaF–AlF3 molten salt with 9wt% LiF. Guo et al. [57] applied First principles molecular dynamics (FPMD) to investigate the ionic structure and electronic properties of the KF–NaF–AlF3 molten salt. The configuration is depicted in Fig. 3(c). The results illustrated that the interaction between Al–F was strong and mainly occurred through covalent bonding and formed a large number of AlF4, AlF25, and AlF36 complex ion groups.

    Researchers have also conducted some studies on the structure of zirconium in fluoride molten salts, and the structure of ZrF4 in alkali metal fluoride molten salts has been experimentally studied via Raman spectroscopy, NMR, and EXAFS; ZrF26, ZrF37, and ZrF48 were the three main anions present [3334,5862]. The network structure of the LiF–ZrF4 molten salt is shown in Fig. 3(d).

    In addition, the crystal structure of K2TaF7, the structure of tantalum fluoride in FLiNaK, belonged to the triangular system with the P21/C space group [6365]. The central Ta atom was surrounded by seven fluoride atoms in the form of the TaF27 anion. Wang and Duan [66] identified the species of Nb(V) in the FLiNaK melt, and their results indicated that the NbF27 complex ion was the dominant ion in the Nb(V)–FLiNaK solution. In the perfluorinated melt, Ti3+ or Ti4+ was stable in the forms of TiF36 and TiF26, respectively [6768]. A summary of coordination structures in molten fluoride salts is shown in Table 3.

    Table  3.  Summary of coordination structures in fluoride molten salt
    Molten saltMethodResultsRefs.
    NaF–LiF–LnF3 (Ln = La, Nd)High-temperature XRDLnF4 existed in the molten salt.[48]
    LaF3–LiFMDLaF36; the CN of La–F was always between 7.33 and 7.75.[49]
    LiF–ThF4–UF4EXAFS and MDThe content of the complex (ThXUYFZ)4X+4Y−Z increased from 0 to 8mol%, and the average CN of Th−F was very close to 8.[50]
    LiF–LuF3/LaF3NMR and EXAFSLnF3xx[51]
    LiF–NdF3High-temperature RamanNdF36[52]
    KF–NdF3NdF36[53]
    LiF–NdF3NdF36 and NdF4[54]
    LiF–NdF3High-temperature Raman techniquesNdF36[8]
    LiF–BeF2DFTBeF24, Be2F37, and Be3F410 existed in the molten salt; the average F–Be and Be–Be distances were (1.564 ± 0.100) and (3.025 ± 0.200) Å, respectively; and the CN of F−Be was maintained at 4.[55]
    LiF–BeF2MD { {\mathrm{B}\mathrm{e} }_{2}\mathrm{F} }_{7}^{3-}{\boldsymbol{}} , { {\mathrm{B}\mathrm{e} }_{3}\mathrm{F} }_{10}^{4-} ; the Be−F ion pair had the average first peak radius of 1.58 Å and the CN of 4.[35]
    LiF–NaF–AlF3MD{\mathrm{A}\mathrm{l}\mathrm{F} }_{4}^{-}, {\mathrm{A}\mathrm{l}\mathrm{F} }_{5}^{2-} , and {\mathrm{A}\mathrm{l}\mathrm{F} }_{6}^{3-} [56]
    KF–NaF–AlF3QC{\mathrm{A}\mathrm{l}\mathrm{F} }_{4}^{-} , {\mathrm{A}\mathrm{l}\mathrm{F} }_{5}^{2-} , and {\mathrm{A}\mathrm{l}\mathrm{F} }_{6}^{3-} [57]
    ZrF4 in alkali metal fluoride meltRaman, NMR, and EXAFS {\mathrm{Z}\mathrm{r}\mathrm{F}}_{6}^{2-} , {\mathrm{Z}\mathrm{r}\mathrm{F}}_{7}^{3-} , and {\mathrm{Z}\mathrm{r}\mathrm{F}}_{8}^{4-} [3334,
    5862]
    K2TaF7–FLiNaKTriangular system[6365]
    Nb(V)–FLiNaK {\mathrm{N}\mathrm{b}\mathrm{F}}_{7}^{2-} [66]
    Perfluorinated melt {\mathrm{T}\mathrm{i}\mathrm{F}}_{6}^{3-} and {\mathrm{T}\mathrm{i}\mathrm{F}}_{6}^{2-} [6768]
     | Show Table
    DownLoad: CSV

    At present, researchers at home and abroad have conducted considerable research on the structure of molten nitrate by using various experimental and computer simulation methods and obtained a series of important conclusions.

    Iwadate et al. [69] investigated the structures of NaNO2 and KNO2 monosalts by applying the neutron scattering method. The precise analysis of structural parameters, such as CNs, interatomic distances, and temperature factors, yielded the following conclusions. In the presence of stable {\mathrm{N}\mathrm{O}}_{2}^{-} anions in the melt, the behavior of {\mathrm{N}\mathrm{O}}_{2}^{-} anions was similar to that of {\mathrm{N}\mathrm{O}}_{3}^{-} anions due to the presence of lone pairs of electrons. The short-range structure of the KNO2 monosalt was similar to that of NaNO2. However, for the cation (Na+/K+), its position relative to the anion was slightly different. Fang et al. [7071] studied the microstructure of molten salts, such as Ca(NO3)2. The radial distribution function of the molten salts was obtained through data processing, and a local structure model of molten salts was constructed to investigate the connection between the melt structure and the crystal growth process. Perelygin and Mikhailov [72] investigated the interaction between ions in nitrate ions by applying infrared and Raman spectroscopy, focusing on ionic association forms.

    Berg et al. [73] performed a phase diagram study on the NaNO2–NaNO3 binary system by combining differential scanning calorimetry and Raman spectroscopy. Xu and Chen [74] studied the Raman spectra of the NaNO3–KNO3 system at different temperatures. Jayaraman et al. [75] calculated the thermal properties of Li, Na, and K alkali metal nitrates and the lattice vibrations of NaNO2 by exploiting MD methods.

    Ponyatenko and Radchenko [7677] used Raman spectroscopy to study the interaction and rotational motion of {\mathrm{N}\mathrm{O}}_{3}^{-} ions in LiNO3, NaNO3, KNO3, RbNO3, CsNO3, AgNO3, and TlNO3 monovalent nitrates and monovalent Ca(NO3)2, Ba(NO3)2, and Sr(NO3)2 over a relatively wide temperature range. Gao et al. [78] investigated the structure of the binary mixed molten salts of nitrates. The ionic structure of NaNO3–KNO3–NaNO2 molten salts was investigated by Ultra-Violet laser Raman spectroscopy. {\mathrm{N}\mathrm{O}}_{3}^{-} and {\mathrm{N}\mathrm{O}}_{2}^{-} were found to be present in the NaNO3–KNO3–NaNO2 molten salt without complex anions. The strength of the N–O bond increased when NaNO3 was added to KNO2–NaNO2.

    Zhao et al. [79] acquired the structural configurations of the four clusters of NaNO3, KNO3, NaNO2, and KNO2 and Hitec ternary molten salts by applying DFT calculations. Their results revealed that in the microstructures of the four monosalts NaNO3, KNO3, NaNO2, and KNO2, Na+/K+ and {\mathrm{N}\mathrm{O}}_{3}^{-} / {\mathrm{N}\mathrm{O}}_{2}^{-} formed a conformation dominated by monodentate ligands and supplemented by dentate ligands. The higher the content of monodentate ligands, the more stable the molecules. In Hitec mixed molten salts (53wt%KNO3–7wt%NaNO3–40wt%NaNO2), Na+ and K+ were bonded in a single–double dentate mixture, and the structure was more stable when the monodentate content was relatively high and unstable when the double dentate content was relatively high.

    As of now, only a few studies on the ionic structure of molten carbonate salts exist. In this paper, the relevant research work of scholars in recent years is discussed.

    Hou et al. [80] studied the Raman spectra of Li2CO3, Na2CO3, and K2CO3 in solid and molten states at different temperatures (up to 1000°C). The wave number shifts and half-height width changes of the {\mathrm{C}\mathrm{O}}_{3}^{2-} symmetric stretching vibrational modes with the increase in temperature were analyzed.

    Chen et al. [81] used in situ Raman spectroscopy to detect the peroxycarbonate species present in molten carbonates. Their experimental results confirmed that the main oxygen species existing in the Li/K2CO3 (62:38, mass ratio) molten salt electrolyte under acidic conditions (1 atm O2 + CO2) were {\mathrm{C}\mathrm{O}}_{4}^{2-} or {{\mathrm{C}}_{2}\mathrm{O}}_{6}^{2-} .

    Ohata et al. [82] used DFT to derive the structures with the minimum interaction energy of lithium carbonate monomers, dimers, trimers, and tetramers and optimized these structures. Koura et al. [83] studied the equilibrium structures of lithium carbonate and potassium carbonate with ab initio molecular orbital calculations. Among the four structures of lithium and potassium carbonate, the most stable structure is the one in which all five atoms are in the same plane.

    Numerous factors, including temperature, affect the structure of molten salts. This section reviews the effect of current temperature on the molten salt structure.

    Wang et al. [84] used the MD method to study the effect of temperature on the structure and properties of cryolite and cryolite–alumina molten salt systems. Their results revealed that the CN of F–Al and F–F decreased with the increase in temperature and that the CN of Al–Al first increased and then decreased. Therefore, with the increase in the temperature, the energy of ions in the system increased, the interaction between ions weakened, and mutual attraction between ions was gradually eliminated.

    Zhao et al. [79] used high-temperature Raman experiments to investigate the effect of temperature on the structure of a molten salt system of nitric acid. Their results revealed that the half-peak widths of the Raman peaks of the mono, di-, and ternary molten salts of NaNO3, KNO3, and NaNO2 increased with the increase in temperature. The effect of temperature on the microstructure of Hitec ternary molten salts was obtained via X-ray scattering experiments. The microstructure of the Hitec molten salt became increasingly unstable as the temperature increased.

    Similarly, Hu et al. [85] used Raman spectroscopy to study the ionic structures of molten salts at different temperatures. Three Raman characteristic peaks of KF–KBF4 at room temperature were found near 360, 533, and 775 cm−1, which were the characteristic Raman peaks of KBF4 corresponding to the υ2, υ4, and υ1 characteristic peaks of {\mathrm{B}\mathrm{F}}_{4}^{-} , respectively. When the mixture melted, the Raman shifts of the three characteristic peaks decreased significantly and further redshifted with the increase in temperature due to the strengthening of the thermal motion of the atoms and the weakening of interatomic forces. In addition, the half-height widths of the above three characteristic peaks increased when the mixture melted due to the increase in the vibrational disorder of the structure.

    As can be concluded from the above results, temperature affects the structure of molten salts. As the temperature increases, the thermal motion of atoms intensifies, the structure of each vibrational group loosens; and interatomic forces weaken, resulting in an increase in atomic spacing as well as a broadening of bond angle distribution, an increase in vibrational disorder, and the destabilization of the microstructure of molten salts.

    The change in composition affects the structures of molten salts to some extent. Therefore, this section first discusses the influence of cations.

    Jiang et al. [85] studied the form of Sm(III) in molten LiCl–KCl–SmCl3 by Raman spectroscopy. Fig. 4 shows that the Raman spectra of quenched LiCl–KCl salts dissolved diverse amounts of SmCl3 and XRD pattern of quenched salt. In the quenched salt with SmCl3 content less than 5.17mol%, Sm(III) species are mainly {\mathrm{S}\mathrm{m}\mathrm{C}\mathrm{l}}_{6}^{3-} and a small amount are {\mathrm{S}\mathrm{m}\mathrm{C}\mathrm{l}}_{7}^{4-} . As the concentration of SmCl3 increases, {\mathrm{S}\mathrm{m}\mathrm{C}\mathrm{l}}_{7}^{4-} gradually exceeds {\mathrm{S}\mathrm{m}\mathrm{C}\mathrm{l}}_{6}^{3-} in the melt.

    Fig. 4.  (a) Raman spectra of quenched LiCl–KCl salts containing various amounts of SmCl3 at ambient temperature; (b) XRD patterns of quenched LiCl–KCl–SmCl3 salts with SmCl3 molar fractions of 12.6 and 25.2 [85]. Reprinted from Electrochim. Acta, 439, S.L. Jiang, C.M. Ye, Y.L. Liu, et al., nsights into the effects of fluoride anions on the electrochemical behavior and solution structure of trivalent samarium in LiCl–KCl molten salt, art. No. 141733, Copyright 2023, with permission from Elsevier.

    You et al. [86] studied the in situ Raman spectra of NaF–AlF3 molten salt with different molar ratios (CR = NaF : AlF3) and obtained the types of cluster structures. The law of their variation with composition in high-temperature molten salts was acquired through comparison with theoretical calculations. The molten salt of cryolite (CR = 3.0) contained aluminous fluorine tetrahedra (Q2) with two bridge fluorines in addition to its main structure of isolated aluminous fluorine octahedra (H0). In the molten salt of subcryolite (CR = 1.5), a five-coordinated aluminous fluorine structure type was absent, and although (H0) and (Q2) continued to coexist, their relative content changed. The molten salt of monocryolite (CR = 1.0) showed a significantly changed structure, and isolated aluminous fluorine tetrahedra (Q0) were present with a small amount of aluminous fluorine tetrahedra dimers (Q1). When CR = 0.5, the main structure was dominated by aluminous fluorine tetrahedra dimers (Q1) with a small amount of isolated aluminous fluorine tetrahedra (Q0). These findings indicated that the type of aluminofluorine tetrahedra and the number of bridge fluorines changed with the increase in AlF3 content.

    The microstructures of NaCl–KCl and NaCl–RbCl binary mixed molten salts at 827°C were calculated by Wang and Liu [87] by using MD to investigate the effect of NaCl content on the microstructure of molten salts.

    In the NaCl–KCl mixed molten salts, the first peak heights of the radial distribution functions between the Na–Cl and K–Cl ions gradually decreased with the gradual increase in NaCl content, and the CNs of Na–Cl and K–Cl gradually increased. The first peak height gradually increased with the increase in NaCl content, and the Cl–Cl ion spacing and CN gradually decreased. The first peak heights of the Na–Na, Na–K, and K–K radial distribution functions decreased slightly with increasing NaCl content, the ion spacing did not show any obvious trend, and the corresponding equivalent CNs gradually increased. In the NaCl–RbCl mixed molten salts, the first peak heights of the Na–Cl and Rb–Cl radial distribution functions decreased gradually with increasing NaCl content, and the ion spacings of Na–Cl and Rb–Cl and their CNs increased gradually with increasing NaCl content.

    With the increase in cation content, the ionic spacing and ionic CN between dissimilar ions and cations gradually increased. At the same time, the microscopic arrangement between ions was affected, promoting the close arrangement between dissimilar ions and between cations and inhibiting the close arrangement between anions.

    In titanium molten salts, anions also have a great influence in addition to the cations in molten salt electrolytes that can affect ion morphology and microstructure.

    Guo et al. [57] used FPMD to study the Al–F bond group and CN in the KF–NaF–AlF3 system at different KF concentrations. Fig. 5 illustrates the Al–F bond population and CN of the system at different KF concentrations. The results indicated that with the increase in KF concentration, the bond population of Al–F first increased and then decreased, which resulted in the same change in the CN of the Al–F complex ion groups.

    Fig. 5.  Bond population and CN of Al–F in KF–NaF–AlF3 at different KF concentrations [57]. Reprinted from Chem. Phys. Lett., 730, H. Guo, J. Li, H.L. Zhang, et al., First-principles molecular dynamics investigation on KF–NaF–AlF3 molten salt system, 587-593, Copyright 2019, with permission from Elsevier.

    Hu et al. [85] used UV laser Raman spectroscopy to investigate the effect of the different molar fractions of NaF on the ionic structure of acidic NaF–AlF3 molten salt. Their results demonstrated that the F content was low at the measured temperature and intervals of molten salt molar fractions. Furthermore, the molar fraction of {\mathrm{A}\mathrm{l}\mathrm{F}}_{4}^{-} decreased with the increase in the molar fraction of the electrolyte NaF, whereas that of {\mathrm{A}\mathrm{l}\mathrm{F}}_{6}^{3-} increased. When the molar fraction of NaF was 0.6, the molar fraction of {\mathrm{A}\mathrm{l}\mathrm{F}}_{4}^{-} was approximately 0.75, whereas that of {\mathrm{A}\mathrm{l}\mathrm{F}}_{6}^{3-} was only approximately 0.25. When the molar fraction of NaF increased to 0.71, that of {\mathrm{A}\mathrm{l}\mathrm{F}}_{4}^{-} decreased to approximately 0.25, whereas that of {\mathrm{A}\mathrm{l}\mathrm{F}}_{6}^{3-} increased to approximately 0.75. Liu et al. [88] investigated the electrochemical behavior and coordination properties of uranium by combining electrochemical and spectral techniques and ab initio MD simulations. Their results revealed that F ions can more easily coordinate with U(IV) than with U(III) and had a smaller ionic radius and higher charge density. The involvement of F ions made the U(IV) complex more stable, and the equilibrium potential of U(IV)/U(III) moved in a negative direction, approaching the equilibrium potential of U(III)/U(0). When Cl ions were abundant, F ions cannot completely replace the role of Cl ions such that U(IV) and U(III) were more inclined to form {\mathrm{U}\mathrm{C}\mathrm{l}}_{3}{\mathrm{F}}_{3}^{2-} and {\mathrm{U}\mathrm{C}\mathrm{l}}_{5}{\mathrm{F}}^{3-} complexes, respectively. Wu et al. [89] studied the cathodic reduction mechanism of Hf(IV) ions in a molten NaCl–KCl–NaF–K2HfF6 salt system at different NaF concentrations. F gradually replaced Cl with the increase in NaF concentration. As a result of the high thermodynamic stability of {\mathrm{H}\mathrm{f}\mathrm{F}}_{6}^{2-} , {\mathrm{H}\mathrm{f}\mathrm{C}\mathrm{l}}_{m}{\mathrm{F}}_{n}^{2-} (m + n = 6) was transformed into {\mathrm{H}\mathrm{f}\mathrm{F}}_{6}^{2-} .

    Our team has also done considerable research on the influence of F on molten salt structures. Song et al. [90] added KF as a fluoride ion source to NaCl–KCl–TiCl4 molten salt. They found that almost all Ti3+ and Ti4+ formed coordination complexes and the CN equaled 6, {\mathrm{T}\mathrm{i}\mathrm{F}}_{6}^{3-} and {\mathrm{T}\mathrm{i}\mathrm{F}}_{6}^{2-} in high-concentration fluoride melt. By applying a combination of spectral, electrochemical, and mathematical analyses, Liu et al. [91] investigated the effect of F on the electrochemical behavior and coordination characteristics of titanium ions. They also performed Raman spectral analysis on samples with different proportions. The results are provided in Fig. 6, which illustrates that F can reduce the reduction steps of titanium ions and affect the proportion of the valence states of titanium ions. Raman analysis and X-ray photoelectron spectroscopy (XPS) results revealed that fluoride ions formed {\mathrm{T}\mathrm{i}\mathrm{C}\mathrm{l}}_{j}{\mathrm{F}}_{i}^{m-} with titanium ions in molten salts containing titanium ions.

    Fig. 6.  Raman spectra of samples under different α (molar ratios of [F]/[Tin+]) conditions [91]. Reprinted by permission from Springer Nature: Int. J. Miner. Metall. Mater., Effect of fluoride ions on coordination structure of titanium in molten NaCl–KCl, S.S. Liu, S.L. Li, C.H. Liu, J.L. He, and J.X. Song, Copyright 2023.

    Yuan et al. [92] used XPS and Raman to analyze the complexes in LiCl–KCl eutectic salts containing VCl3 and KF. They demonstrated that when fluoride was added to the molten salt, the V–Cl bond was replaced by the V–F bond and combined with V(III) to form {\mathrm{V}\mathrm{F}}_{6}^{3-} , as shown in Fig. 7. Moreover, with the addition of fluoride ions, the particle size of the product decreased. Bai et al. [93] revealed the chemical coordination mechanism of tantalum ions in the above molten salts. When fluoride ions were added to the NaCl–KCl–TaCl5 molten salt, the Ta–Cl bond in the system was gradually replaced by the Ta–F bond such that the tantalum and fluoride ions in the molten salt formed a new complex, namely, \mathrm{T}\mathrm{a}{\mathrm{C}\mathrm{l}}_{x}{\mathrm{F}}_{y}^{n-} , with stronger stability. Finally, \mathrm{T}\mathrm{a}{\mathrm{F}}_{y}^{n-} was formed.

    Fig. 7.  (a) Raman spectra of VCl3 in LiCl–KCl–KF molten salt with various molar ratios of [F]/[Vn+]; (b) the geometric structure of {\mathbf{V}\mathbf{F}}_{6}^{3-} [92]. Reprinted from Trans. Monferrous Met. Soc. China, 32, R. Yuan, C. Lü, H.L. Wan, et al., Effect of fluoride addition on electrochemical behaviors of V(III) in molten LiCl−KCl, 2736-2745, Copyright 2022, with permission from Elsevier.

    F is a single base ligand with a stronger coordination ability than Cl. The addition of F to the chloride molten salt will result in coordination substitution and change the coordination structure of ions in the molten salt to form fluoro–chloride. The fluorine-containing ligand is stronger than the chlorine-containing ligand and has a more stable structure.

    The influence of oxygen ions on the structure and properties of molten salt has rarely been reported. Therefore, this section summarizes the influence of the addition of oxygen ions on the structure of molten salts.

    Guo et al. [94] directly studied the ion microstructure of the KF–NaF–AlF3–Al2O3 system by using FPMD. Their results indicated that with the increase in Al2O3 concentration, Al–F–Al, Al–O–Al, and Al–O–F formed complex ion groups, and the ion microstructure of the system became more complex.

    Wang and Duan [66] investigated the effect of O2− on the structure of molten salts by adding Na2O to K2NbF7–FLiNaK and showed that {\mathrm{N}\mathrm{b}\mathrm{O}\mathrm{F}}_{6}^{3-} is a stable monooxyfluoro complex anion present in FLiNaK. The number of {\mathrm{N}\mathrm{b}\mathrm{O}\mathrm{F}}_{6}^{3-} species increased with the increase in O2− when the molar ratio of O2−/Nb(V) was less than 1.0. Wang et al. [95] investigated the effect of O2− on the structure of metal fluorides with the addition of Li2O by applying a combination of Raman spectroscopy and QC theoretical calculations. Their results showed that for vanadium group elements, the addition of Li2O resulted in the formation of {\mathrm{T}\mathrm{a}\mathrm{O}\mathrm{F}}_{5}^{2-} / {\mathrm{N}\mathrm{b}\mathrm{O}\mathrm{F}}_{5}^{2-} and {\mathrm{T}\mathrm{a}\mathrm{O}\mathrm{F}}_{6}^{3-} / {\mathrm{N}\mathrm{b}\mathrm{O}\mathrm{F}}_{6}^{3-} structures and the depletion of {\mathrm{T}\mathrm{a}\mathrm{F}}_{7}^{2-} / {\mathrm{N}\mathrm{b}\mathrm{F}}_{7}^{2-} and {\mathrm{T}\mathrm{a}\mathrm{F}}_{8}^{3-} / {\mathrm{N}\mathrm{b}\mathrm{F}}_{8}^{3-} fluoride structures.

    Chen et al. [96] used Raman spectroscopy to study the effect of oxide ions on the morphology of KF molten liquid containing ZrF4 and HfF4. Zirconium and hafnium fluoride existed in the forms of {\mathrm{M}\mathrm{F}}_{6}^{2-} and {\mathrm{M}\mathrm{F}}_{7}^{3-} (M = Zr, Hf) in KF melts with lower MF4 concentrations, whereas {\mathrm{M}\mathrm{F}}_{6}^{2-} was dominant at higher MF4 concentrations. When 3mol%–5mol% Li2O was added to KF–MF4 (10mol%) melt, {\mathrm{M}\mathrm{O}\mathrm{F}}_{5}^{3-} formed oxyfluorides with characteristic terminal M–O stretching bands at 775 (Zr) and 763 (Hf) cm−1. However, when 5mol% Li2O was added to the KF melt containing 30mol%–40mol% ZrF4 and HfF4, these fluoroxides were not observed. Consistent with the formation of {\mathrm{M}\mathrm{O}\mathrm{F}}_{5}^{3-} , the mean binding energy of {\mathrm{M}\mathrm{F}}_{7}^{3-} was considerably lower than that of {\mathrm{M}\mathrm{F}}_{6}^{2-} .

    Wang et al. [97] investigated the effect of Nd2O3 on the properties and structure of an AlF3–(Na/Li)F–Al2O3 melt by measuring the density, viscosity, and conductivity of the molten salt system by applying the Archimedes, rotational, and continuously varying conductivity cell constants methods. Their results showed that the addition of Nd2O3 increased the ionic gap of the system and produced ionic groups with low stacking density. In addition, the introduction of Nd3+ may lead to the formation of the Nd–O–F complex ion groups {\mathrm{N}\mathrm{d}\mathrm{O}\mathrm{F}}_{3}^{2-} , {\mathrm{N}\mathrm{d}\mathrm{O}\mathrm{F}}_{5}^{4-} , {\mathrm{N}\mathrm{d}}_{2}{\mathrm{O}\mathrm{F}}_{6}^{2-} , {\mathrm{N}\mathrm{d}}_{2}{\mathrm{O}\mathrm{F}}_{8}^{4-} , and {\mathrm{N}\mathrm{d}}_{2}{\mathrm{O}\mathrm{F}}_{4}^{2-} as a result of the substitution reaction with the existing Al–O–F in the molten salt.

    Liu et al. [98] studied the microstructure of fused FLiNaK–LuF3 and FLiNaK–LuF3–Li2O systems by combining Raman spectroscopy and density functional theory calculations. The anion {\mathrm{L}\mathrm{u}}_{2}{\mathrm{O}\mathrm{F}}_{8}^{4-} was formed by the addition of Li2O to molten FLiNaK–LuF3 (20mol%), which has a linear Lu–O–Lu geometry with one oxygen atom bridging two LuF4 groups. When the content of Li2O was 10mol%, {\mathrm{L}\mathrm{u}}_{2}{{\mathrm{O}}_{2}\mathrm{F}}_{4}^{2-} and {\mathrm{L}\mathrm{u}}_{2}{{\mathrm{O}}_{2}\mathrm{F}}_{6}^{4-} were formed.

    In summary, the above researchers studied the effects of metal oxides (O2−) on molten salt systems by employing spectroscopic techniques and computer simulations. Their results demonstrated that in molten salts, complex oxygen ions changed the structure of the molten salts with the addition of metal oxides (O2−). This finding provides a theoretical basis for clarifying the electrolytic mechanism. The influence of cations and anions on the structural aspects of molten salts is summarized in Table 4, and the influence of O2− on the structural aspects of molten salts is shown in Table 5.

    Table  4.  Influence of cations and anions on the structural aspects of molten salts
    IonMolten saltMethodMain conclusionRef.
    CationSm3+LiCl–KCl– SmCl3Raman, XRDAs the concentration of SmCl3 increases, {\mathrm{S}\mathrm{m}\mathrm{C}\mathrm{l}}_{7}^{4-} gradually exceeds {\mathrm{S}\mathrm{m}\mathrm{C}\mathrm{l}}_{6}^{3-} in the melt. [85]
    Al3+NaF–AlF3In situ RamanCR = 3.0: H0, Q2; CR = 1.5: coexistence of H0 and Q2; CR = 1.0: Q0, mainly; a small amount of Q1;
    CR = 0.5: Q1, mainly; a small amount of Q0.
    [86]
    Na+NaCl –KClMDThe CN of Na–Cl and K–Cl gradually increased; the Cl–Cl ion spacing and CN gradually decreased; the corresponding equivalence CN gradually increased.[87]
    Na+NaCl –RbClMDNa–Cl and Rb–Cl ion spacing and CN gradually increased; Cl–Cl ion spacing and CN gradually decreased; Na–Na, Na–Rb, and Rb–Rb ion spacing had no obvious trend.[87]
    AnionFKF–NaF–AlF3MDWith the increase in KF concentration; the bond population of Al–F first increased and then decreased; the same change of CN of Al–F complex ion groups.[57]
    FNaF–AlF3UV, RamanThe molar fraction of {\mathrm{A}\mathrm{l}\mathrm{F} }_{4}^{-} decreased with the increase in the molar weight fraction of the electrolyte NaF, whereas the molar fraction of {\mathrm{A}\mathrm{l}\mathrm{F}}_{6}^{3-} increased.[85]
    FLiCl–KCl–UCl4–LiFMDF ion easily coordinated with U(IV), its ionic radius was smaller, and its charge density was higher. When Cl ions were abundant, U(IV) and U(III) were more inclined to form {\mathrm{U}\mathrm{C}\mathrm{l}}_{3}{\mathrm{F}}_{3}^{2-} and UCl5F3− complexes, respectively.[88]
    FNaCl–KCl–NaF–K2HfF6CV, SWV, and XRDWith the increase in NaF concentration, [F] gradually replaced [Cl], and [ {\mathrm{H}\mathrm{f}\mathrm{C}\mathrm{l}}_{m}{\mathrm{F}}_{n}^{2-} ] (m + n = 6)] transformed into [ {\mathrm{H}\mathrm{f}\mathrm{F}}_{6}^{2-} ].[89]
    FNaCl–KCl–TiCl4–KFIn high-concentration fluoride melts, almost all Ti3+ and Ti4+ formed coordination complexes, and CN = 6.[90]
    FTi2+/Ti3+/Ti4+–NaCl–KCl–KFEC, mathematical analysis, and spectral techniquesF shortened the reduction step of titanium ions and affected the proportion of the valence states of titanium ions. Fluoride ions and titanium ions formed {\mathrm{T}\mathrm{i}\mathrm{C}\mathrm{l}}_{j}{\mathrm{F}}_{i}^{m-} in molten salt.[91]
    FLiCl–KCl–VCl3–KFXPS and RamanWith the addition of fluoride ions, the V–Cl bond was replaced by the V–F bond and combined with V(III) to form {\mathrm{V}\mathrm{F}}_{6}^{3-} , and the particle size of the product decreased.[92]
    FNaCl–KCl–TaCl5–NaFXPS and RamanThe Ta–Cl bond in the system was replaced by the Ta–F bond, and tantalum and fluoride ions formed the new complex {\mathrm{T}\mathrm{a}\mathrm{C}\mathrm{l}}_{x}{\mathrm{F}}_{y}^{n-} with stronger stability, then \mathrm{T}\mathrm{a}{\mathrm{F}}_{y}^{n-} formed.[93]
     | Show Table
    DownLoad: CSV
    Table  5.  Influence of O2− on the structural aspects of molten salts
    Molten saltOxides formMethodMain conclusionRef.
    KF–NaF–AlF3–Al2O3Al2O3FPMDWith the increase in Al2O3 concentration, Al–F–Al, Al–O–Al, and Al–O–F formed complex ion groups.[94]
    K2NbF7–FLiNaKNa2O {\mathrm{N}\mathrm{b}\mathrm{O}\mathrm{F}}_{6}^{3-} was a stable mono-oxyfluoro complex anion present in FLNAK, and the number of {\mathrm{N}\mathrm{b}\mathrm{O}\mathrm{F}}_{6}^{3-} species increased with the increase in O2−.[66]
    FLiNaK–K2TaF7/ K2NbF7Li2ORaman and QC
    theoretical calculations
    {\mathrm{T}\mathrm{a}\mathrm{O}\mathrm{F}}_{5}^{2-} / {\mathrm{N}\mathrm{b}\mathrm{O}\mathrm{F}}_{5}^{2-} and {\mathrm{T}\mathrm{a}\mathrm{O}\mathrm{F}}_{6}^{3-} / {\mathrm{N}\mathrm{b}\mathrm{O}\mathrm{F}}_{6}^{3-}
    formed, and O2− caused the polymerization of fluorine oxides in the molten salt and the appearance of solid immiscible material, forming an oxide-like structure.
    [95]
    KF–ZrF4/HfF4Li2ORaman {\mathrm{M}\mathrm{O}\mathrm{F}}_{5}^{3-} was observed. Raman bands formed by zirconium and hafnium oxides were formed.[96]
    AlF3–(Na/Li)F–Al2O3Nd2O3Spectroscopy and MDNd2O3 led to the increase in the ionic gap of the system and the generation of ionic groups with small stacking density. {\mathrm{N}\mathrm{d}\mathrm{O}\mathrm{F}}_{3}^{2-} , {\mathrm{N}\mathrm{d}\mathrm{O}\mathrm{F}}_{5}^{4-} , {\mathrm{N}\mathrm{d}}_{2}{\mathrm{O}\mathrm{F}}_{6}^{2-} , {\mathrm{N}\mathrm{d}}_{2}{\mathrm{O}\mathrm{F}}_{8}^{4-} ,
    and {\mathrm{N}\mathrm{d}}_{2}{\mathrm{O}\mathrm{F}}_{4}^{2-} formed.
    [97]
    FLiNaK–LuF3–Li2OLi2ORaman and DFT {\mathrm{L}\mathrm{u}}_{2}{\mathrm{O}\mathrm{F}}_{8}^{4-} formed, and {\mathrm{L}\mathrm{u}}_{2}{{\mathrm{O}}_{2}\mathrm{F}}_{4}^{2-} and {\mathrm{L}\mathrm{u}}_{2}{{\mathrm{O}}_{2}\mathrm{F}}_{6}^{4-}
    formed when more Li2O was added.
    [98]
     | Show Table
    DownLoad: CSV

    The molten salt structure is of crucial importance as the foundation for the study of molten salt theory. Nevertheless, it is difficult to obtain because microstructures cannot be investigated directly by experimental means and can only be described qualitatively on the basis of experimental data. However, with the continuous improvement in simulation algorithms and the increasing development of computer technology, computer simulation techniques, which are represented by MD simulations, have emerged and are widely used. Studying the interaction between system particles and the microstructures of molten salts by using MD simulation technologies, which can simulate the behavior of real fluids, is possible. In this paper, a variety of characterization methods for studying the structure of molten salts are listed. These methods mainly included XRD, Raman, NDT, NMR, EXAFS, electrochemical analysis, and computer simulation. At the same time, the results of existing research on the system structures of chloride salt, fluoride salt, nitrate, and molten carbonate salt were introduced.

    The effects of temperature, anions and cations, and metal oxides (O2−) on the structure of molten salts are summarized below.

    (1) The increment in temperature loosens the structure of molten salts. With the increase in temperature, the thermal motion of atoms strengthens, interatomic forces weaken, vibrational disorder increases, and the microstructures of molten salts destabilize.

    (2) The increase of cation content in the composition of molten salt affects the microscopic arrangement between ions, which makes the arrangement between different ions and cations closer but also inhibits the tight arrangement between anions. It also changes the structural types of clusters in molten salts. Anions also affect the structure of molten salts. Given the strong coordination ability of F, F addition leads to coordination substitution in the molten chlorine salt and changes the coordination structure of the clusters in the molten salt. Considering that the coordination bond strength of fluorine is greater than that of chlorine, the structure of fluorine is also more stable.

    (3) With the increase in metal oxide (O2−) concentration, more complex oxygen-containing ions form in molten salt. This effect will change the composition and properties of the molten salt.

    Studies on the microstructure of molten salts have not yet reached a consensus, particularly regarding complex molten salt systems, such as complex mixture systems that involve multibody interactions. This situation requires more abundant research tools and more research work. Any kind of structural research method only reflects the side of the overall structural information of matter and gradually improves with its development. Thus, combining theoretical and experimental methods as well as different methods in theory and experiments is necessary to obtain relatively complete microstructure information. The multiscale analysis and characterization of molten salt structures can be carried out by combining various in situ and non-in situ techniques to explore the microstructure and coordination of molten salts deeply. In the future, more accurate and intuitive molten salt structures can be revealed with the aid of evolving scientific and technological methods.

    This work was financially supported by the National Key Research and Development Program of China (Nos. 2021YFC2901600 and 2021YFC2902305), the National Natural Science Foundation of China (No. 52274356), the Natural Science Foundation of Henan Province, China (No. 222300420545), the State Key Laboratory of Special Rare Metal Materials, China (No. SKL2020K004), the Northwest Rare Metal Materials Research Institute, China, and the State Key Laboratory of Complex Nonferrous Metal Resources Clean Utilization, China (No. CNMRCUKF2008).

    The authors declare no conflict of interest.

  • [1]
    S.Q. Jiao, M.Y. Wang, and W.L. Song, Editorial for special issue on high-temperature molten salt chemistry and technology, Int. J. Miner. Metall. Mater., 27(2020), No. 12, p. 1569. DOI: 10.1007/s12613-020-2225-7
    [2]
    X.L. Xi, M. Feng, L.W. Zhang, and Z.R. Nie, Applications of molten salt and progress of molten salt electrolysis in secondary metal resource recovery, Int. J. Miner. Metall. Mater., 27(2020), No. 12, p. 1599. DOI: 10.1007/s12613-020-2175-0
    [3]
    V. van Speybroeck, R. Gani, and R.J. Meier, The calculation of thermodynamic properties of molecules, Chem. Soc. Rev., 39(2010), No. 5, p. 1764. DOI: 10.1039/b809850f
    [4]
    W. Jia, Molecular Dynamics Study on Structure and Properties of Molten Salt in Alkali Metal Chloride System [Dissertation], East China University of Science and Technology, Shanghai, 2016, p. 145.
    [5]
    M.H. Brooker, R.W. Berg, J.H. von Barner, and N.J. Bjerrum, Matrix-isolated {\mathrm{A}\mathrm{l}}_{2}{\mathrm{O}\mathrm{F}}_{6}^{2-} ion in molten and solid LiF/NaF/KF, Inorg. Chem., 39(2000), No. 21, p. 4725. DOI: 10.1021/ic000489x
    [6]
    Y.J. Jin and Y.B. Ai, The technology of nuclear magnetic resonance and its applications, Phy. Eng., 12(2002), No. 1, p. 47.
    [7]
    Y. Iwadate, K. Suzuki, N. Onda, et al., Local structure of molten LaCl3 analyzed by X-ray diffraction and La–LIII absorption-edge XAFS technique, J. Alloys Compd., 408-412(2006), p. 248. DOI: 10.1016/j.jallcom.2005.04.037
    [8]
    C.Y. Wang, X.T. Chen, R. Wei, and Y. Gong, Raman spectroscopic and theoretical study of scandium fluoride and oxyfluoride anions in molten FLiNaK, J. Phys. Chem. B, 124(2020), No. 30, p. 6671. DOI: 10.1021/acs.jpcb.0c06021
    [9]
    V. Dracopoulos, B. Gilbert, and G.N Papatheodorou, Vibrational modes and structure of lanthanide fluoride-potassium fluoride binary melts LnF3–KF (Ln = La, Ce, Nd, Sm, Dy, Yb), J. Chem. Soc., Faraday Trans., 94(1998), No. 17, p. 2601. DOI: 10.1039/a802812e
    [10]
    N. Ma, J.L. You, L.M. Lu, J. Wang, M. Wang, and S.M. Wan, Micro-structure studies of the molten binary K3AlF6–Al2O3 system by in situ high temperature Raman spectroscopy and theoretical simulation, Inorg. Chem. Front., 5(2018), No. 8, p. 1861. DOI: 10.1039/C8QI00226F
    [11]
    F.G. Edwards, J.E. Enderby, R.A. Howe, and D.I. Page, The structure of molten sodium chloride, J. Phys. C Solid State Phys., 8(1975), No. 21, p. 3483. DOI: 10.1088/0022-3719/8/21/018
    [12]
    E.W.J. Mitchell, P.F.J. Poncet, and R.J. Stewart, The ion pair distribution functions in molten rubidium chloride, Philos. Mag., 34(1976), No. 5, p. 721. DOI: 10.1080/14786437608222045
    [13]
    J. Locke, S. Messoloras, R.J. Stewart, R.L. McGreevy, and E.W.J. Mitchell, The structure of molten CsCl, Philos. Mag. B, 51(1985), No. 3, p. 301. DOI: 10.1080/13642818508240576
    [14]
    X.M. Liu, H.H. Sun, X.P. Feng, and N. Zhang, Relationship between the microstructure and reaction performance of aluminosilicate, Int. J. Miner. Metall. Mater., 17(2010), No. 1, p. 108. DOI: 10.1007/s12613-010-0119-9
    [15]
    A.L. Rollet, C. Bessada, A. Rakhmatoulline, et al., In situ high temperature NMR and EXAFS experiments in rare-earth fluoride molten salts, C. R. Chim., 7(2004), No. 12, p. 1135. DOI: 10.1016/j.crci.2004.02.021
    [16]
    A.L. Rollet, E. Veron, and C. Bessada, Fission products behavior in molten fluoride salts: Speciation of La3+ and Cs+ in melts containing oxide ions, J. Nucl. Mater., 429(2012), No. 1-3, p. 40. DOI: 10.1016/j.jnucmat.2012.05.010
    [17]
    A.L. Rollet and M. Salanne, Studies of the local structures of molten metal halides, Annu. Rep. Prog. Chem., Sect. C: Phys. Chem., 107(2011), p. 88. DOI: 10.1039/C1PC90003J
    [18]
    Y. Okamoto, H. Shiwaku, T. Yaita, S. Suzuki, and M. Gaune-Escard, High-energy EXAFS study of molten GdCl3 systems, J. Mol. Liq., 187(2013), p. 94. DOI: 10.1016/j.molliq.2013.05.018
    [19]
    Y. Okamoto, XAFS simulation of highly disordered materials, Nucl. Instrum. Methods Phys. Res. Sect. A, 526(2004), No. 3, p. 572. DOI: 10.1016/j.nima.2004.02.025
    [20]
    Y. Okamoto, H. Shiwaku, T. Yaita, H. Narita, and H. Tanida, Local structure of molten LaCl3 by K-absorption edge XAFS, J. Mol. Struct., 641(2002), No. 1, p. 71. DOI: 10.1016/S0022-2860(02)00329-0
    [21]
    G.D. Zissi, A. Chrissanthopoulos, and G.N. Papatheodorou, Vibrational modes and structure of the LaCl3–CsCl melts, Vib. Spectrosc., 40(2006), No. 1, p. 110. DOI: 10.1016/j.vibspec.2005.07.005
    [22]
    G. Bontempelli, F. Magno, and S. Daniele, Simple relationship for calculating backward to forward peak-current ratios in cyclic voltammetry, Anal. Chem., 57(1985), No. 7, p. 1503. DOI: 10.1021/ac00284a081
    [23]
    J.J. O'Dea, J. Osteryoung, and R.A. Osteryoung, Theory of square wave voltammetry for kinetic systems, Anal. Chem., 53(1981), No. 4, p. 695. DOI: 10.1021/ac00227a028
    [24]
    F. Lantelme and M. Chemla, Chronoamperometry for the determination of metallic interdiffusion coefficients. Rapid transport processes in the first atomic layers, J. Electroanal. Chem., 396(1995), No. 1-2, p. 203. DOI: 10.1016/0022-0728(95)03933-8
    [25]
    T.J. Zhu, C.Y. Wang, H.Y. Fu, W. Huang, and Y. Gong, Electrochemical and Raman spectroscopic investigations on the speciation and behavior of chromium ions in fluoride doped molten LiCl–KCl, J. Electrochem. Soc., 166(2019), No. 10, p. H463. DOI: 10.1149/2.1221910jes
    [26]
    M.Y. Wu, Dissolution and Electrochemical Reduction of Multivalent Refractory Metal Oxides in NaKClF Molten Salt [Dissertation], North China University of Science and Technology, Tangshan, 2016, p. 85.
    [27]
    J. Hetmańczyk, Ł. Hetmańczyk, A. Migdał-Mikuli, and E. Mikuli, Vibrational and reorientational dynamics, crystal structure and solid–solid phase transition studies in [Ca(H2O)6]Cl2 supported by theoretical (DFT) calculations, J. Raman Spectrosc., 47(2016), No. 5, p. 591. DOI: 10.1002/jrs.4863
    [28]
    Y.C. Chen, Principle and application of quantum chemical computation, Technol. Manage. Sci., (2020), p. 110.
    [29]
    X.J. Li, J. Song, S.P. Shi, et al., Dynamic fluctuation of U3+ coordination structure in the molten LiCl–KCl eutectic via first principles molecular dynamics simulations, J. Phys. Chem. A, 121(2017), No. 3, p. 571. DOI: 10.1021/acs.jpca.6b10193
    [30]
    N. Metropolis and S. Ulam, The Monte Carlo method, J. Am. Stat. Assoc., 44(1949), No. 247, p. 335. DOI: 10.1080/01621459.1949.10483310
    [31]
    A. Baranyai, I. Ruff, and R.L. McGreevy, Monte Carlo simulation of the complete set of molten alkali halides, J. Phys. C: Solid State Phys., 19(1986), No. 4, p. 453. DOI: 10.1088/0022-3719/19/4/008
    [32]
    B.J. Alder and T.E. Wainwright, Studies in molecular dynamics. I. General method, J. Chem. Phys., 31(1959), No. 2, p. 459. DOI: 10.1063/1.1730376
    [33]
    C. Bessada, O. Pauvert, D. Zanghi, et al., In situ experimental approach of the speciation in molten lanthanide and actinide fluorides combining NMR, EXAFS and molecular dynamics, ECS Trans., 33(2010), No. 7, p. 361. DOI: 10.1149/1.3484794
    [34]
    C. Bessada, O. Pauvert, L. Maksoud, et al., In situ experimental approach of speciation in molten fluorides: A combination of NMR, EXAFS, and molecular dynamics, [in] M. Gaune-Escard and G.M. Haarberg eds., Molten Salts Chemistry Technololgy, Weley, 2014, p. 219.
    [35]
    G.D. He, R. Tang, X.Z. Duan, et al., Molecular dynamics study on microstructure and diffusion characteristics of LiF–BeF2 molten salt, Chem. Eng. J., 71(2020), No. 8, p. 3565.
    [36]
    H.A. Levy, P.A. Agron, M.A. Bredig, and M.D. Danford, X-ray and neutron diffraction studies of molten alkali halides, Ann. N. Y. Acad. Sci., 79(1960), No. 11, p. 762.
    [37]
    T. Førland, T. Østvold, J. Krogh-Moe, Monte Carlo studies on fused salts. I. Calculations for a two-dimensional ionic model liquid, Acta Chem. Scand., 22(1968), No. 8, p. 2415.
    [38]
    M.A. Howe and R.L. McGreevy, A neutron-scattering study of the structure of molten lithium chloride, Philos. Mag. B, 58(1988), No. 5, p. 485. DOI: 10.1080/13642818808208460
    [39]
    R. Takagi, H. Ohno, and K. Furukawa, Structure of molten KCl, J. Chem. Soc., Faraday Trans. 1, 75(1979), art. No. 1477. DOI: 10.1039/f19797501477
    [40]
    C. Kwon, S.H. Noh, H. Chun, I.S. Hwang, and B. Han, First principles computational studies of spontaneous reduction reaction of Eu(III) in eutectic LiCl–KCl molten salt, Int. J. Energy Res., 42(2018), No. 8, p. 2757. DOI: 10.1002/er.4064
    [41]
    J. Zhao, Z.T. Liu, W.S. Liang, and G.M. Lu, Evaluation of the local structure and electrochemical behavior in the LiCl–KCl–SmCl3 melt, J. Mol. Liq., 363(2022), art. No. 119818. DOI: 10.1016/j.molliq.2022.119818
    [42]
    J. Song, S.P. Shi, X.J. Li, and L.M. Yan, First-principles molecular dynamics modeling of UCl3 in LiCl–KCl eutectic, J. Mol. Liq., 234(2017), p. 279. DOI: 10.1016/j.molliq.2017.03.099
    [43]
    T. Jiang, N. Wang, C.M. Cheng, S.M. Peng, and L.M. Yan, Molecular dynamics simulation on the structure and thermodynamics of molten LiCl–KCl–CeCl3, Acta Phys. Chim. Sin., 32(2016), No. 3, p. 647. DOI: 10.3866/PKU.WHXB201601042
    [44]
    G. Pastore, Z. Akdeniz, and M.P. Tosi, Structure of molten yttrium chloride in an ionic model, J. Phys. Condens. Matter, 3(1991), No. 42, p. 8297. DOI: 10.1088/0953-8984/3/42/024
    [45]
    Z. Akdeniz and M.P. Tosi, Structure and binding of ionic clusters in Th and Zr chloride melts, Z. Naturforsch. A: Phys. Sci., 56(2001), No. 11, p. 717. DOI: 10.1515/zna-2001-1102
    [46]
    Y. Okamoto and P.A. Madden, Structural study of molten lanthanum halides by X-ray diffraction and computer simulation techniques, J. Phys. Chem. Solids, 66(2005), No. 2-4, p. 448. DOI: 10.1016/j.jpcs.2004.06.038
    [47]
    Y. Okamoto, S. Suzuki, H. Shiwaku, A. Ikeda-Ohno, T. Yaita, and P.A. Madden, Local coordination about La3+ in molten LaCl3 and its mixtures with alkali chlorides, J. Phys. Chem. A, 114(2010), No. 13, p. 4664. DOI: 10.1021/jp910637p
    [48]
    L. Bulavin, V. Sokol’skii, O. Roik, et al., Structure and physical properties of ternary NaF–LiF–LnF3 (Ln = La, Nd) systems of eutectic compositions, Phys. Chem. Liq., 54(2016), No. 6, p. 717. DOI: 10.1080/00319104.2016.1149176
    [49]
    H. Cui, Molecular Dynamics Simulation of Melt Structure of Rare Earth Metal Halide Molten Salt System [Dissertation], Kunming University of Science and Technology, 2002, Kumming, p. 136.
    [50]
    C. Bessada, D. Zanghi, M. Salanne, et al., Investigation of ionic local structure in molten salt fast reactor LiF–ThF4–UF4 fuel by EXAFS experiments and molecular dynamics simulations, J. Mol. Liq., 307(2020), art. No. 112927. DOI: 10.1016/j.molliq.2020.112927
    [51]
    A.L Rollet, A. Rakhmatullin, and C. Bessada, Local structure analogy of lanthanide fluoride molten salts, Int. J. Thermophys., 26(2005), No. 4, p. 1115. DOI: 10.1007/s10765-005-6688-6
    [52]
    E. Stefanidaki, G.M. Photiadis, C.G. Kontoyannis, A.F. Vik, and T. Østvold, Oxide solubility and Raman spectra of NdF3–LiF–KF–MgF2–Nd2O3 melts, J. Chem. Soc., Dalton Trans., 2002, No. 11, p. 2302. DOI: 10.1039/b111563b
    [53]
    G. Hatem and M. Gaune-Escard, Calorimetric investigation of {xKF+(1−x)NdF3}(l), J. Chem. Thermodyn., 25(1993), No. 2, p. 219. DOI: 10.1006/jcht.1993.1021
    [54]
    X.W. Hu, Z.W. Wang, B.L. Gao, Z.N. Shi, F.G. Liu, and X.Z. Cao, Density and ionic structure of NdF3–LiF melts, J. Rare Earths, 28(2010), No. 4, p. 587. DOI: 10.1016/S1002-0721(09)60159-9
    [55]
    J.X. Dai, H. Han, Q.N. Li, and P. Huai, First-principle investigation of the structure and vibrational spectra of the local structures in LiF–BeF2 molten salts, J. Mol. Liq., 213(2016), p. 17. DOI: 10.1016/j.molliq.2015.10.053
    [56]
    X.J. Lv, Z.X. Han, J.G. Chen, L.X. Jiang, Z.M. Xu, and Q.S. Liu, First-principles molecular dynamics study of ionic structure and transport properties of LiF–NaF–AlF3 molten salt, Chem. Phys. Lett., 706(2018), p. 237. DOI: 10.1016/j.cplett.2018.06.005
    [57]
    H. Guo, J. Li, H.L. Zhang, et al., First-principles molecular dynamics investigation on KFNaF–AlF3 molten salt system, Chem. Phys. Lett., 730(2019), p. 587. DOI: 10.1016/j.cplett.2019.06.060
    [58]
    C. Bessada, D. Zanghi, O. Pauvert, et al., High temperature EXAFS experiments in molten actinide fluorides: The challenge of a triple containment cell for radioactive and aggressive liquids, J. Nucl. Mater., 494(2017), p. 192. DOI: 10.1016/j.jnucmat.2017.07.023
    [59]
    Y. Qiao, C.M. Pedersen, Y.X. Wang, and X.L. Hou, NMR insights on the properties of ZnCl2 molten salt hydrate medium through its interaction with SnCl4 and fructose, ACS Sustainable Chem. Eng., 2(2014), No. 11, p. 2576. DOI: 10.1021/sc5004693
    [60]
    O. Pauvert, D. Zanghi, M. Salanne, et al., In situ experimental evidence for a nonmonotonous structural evolution with composition in the molten LiF–ZrF4 system, J. Phys. Chem. B, 114(2010), No. 19, p. 6472. DOI: 10.1021/jp912195j
    [61]
    C. Bessada and A.L. Rollet, In situ spectroscopy in molten fluoride salts, [in] F. Lantelme and H. Grou eds., Molten Salts Chemistry, Elsevier, 2013, p. 33.
    [62]
    O. Pauvert, M. Salanne, D. Zanghi, et al., Ion specific effects on the structure of molten AF–ZrF4 systems (A+ = Li+, Na+, and K+), J. Phys. Chem. B, 115(2011), No. 29, p. 9160. DOI: 10.1021/jp203137h
    [63]
    J.L. Hoard, Structures of complex Fluorides.1 Potassium heptafluocolumbate and potassium heptafluotantalate. The configuration of the heptafluocolumbate and heptafluotantalate ions, J. Am. Chem. Soc., 61(1939), No. 5, p. 1252. DOI: 10.1021/ja01874a073
    [64]
    R.B. English, A.M. Heyns, and E.C. Reynhardt, An X-ray, NMR, infrared and Raman study of K2TaF7, J. Phys. C Solid State Phys., 16(1983), No. 5, p. 829. DOI: 10.1088/0022-3719/16/5/010
    [65]
    C.C. Torardi, L.H. Brixner, and G. Blasse, Structure and luminescence of K2TaF7 and K2NbF7, J. Solid State Chem., 67(1987), No. 1, p. 21. DOI: 10.1016/0022-4596(87)90333-1
    [66]
    X.D. Wang and S.Z. Duan, Identification of complex ions of Nb(V) in FLINAK–O2− System by infrared spectra, Rare Met., 12(1993), No. 3, p. 209.
    [67]
    G.S. Chen, M. Okido, and T. Oki, Electrochemical studies of titanium ions (Ti4+) in equimolar KCl–NaCl molten salts with 1 wt% K2TiF6, Electrochim. Acta, 32(1987), No. 11, p. 1637. DOI: 10.1016/0013-4686(87)90017-X
    [68]
    G.S. Chen, M. Okido, and T. Oki, Electrochemical studies of titanium in fluoride–chloride molten salts, J. Appl. Electrochem., 18(1988), No. 1, p. 80. DOI: 10.1007/BF01016208
    [69]
    Y. Iwadate, T. Harada, T. Ohkubo, et al., Pulsed neutron diffraction study of NaNO2 and KNO2 pure melts, Electrochemistry, 77(2009), No. 8, p. 741. DOI: 10.5796/electrochemistry.77.741
    [70]
    Y. Fang, C.H. Fang, L.J. Lin, and X.F. Qin, A study on the structure of calcium nitrate tetrahydrate molten salt, Salt Lake Res., 2007, No. 1, p. 39.
    [71]
    H.X. Zhou, F.Y. Zhu, W.C. Fang, Y.G. Zhou, H.Y. Liu, C.H. Fang, and Y. Fang, Effect of additives on phase transition temperature and undercooling temperature of CaCl2·6H2O–Ca(NO3)2·4H2O and their molten salt structure, J. Mater. Sci. Eng., 37(2019), No. 1, p. 160.
    [72]
    I.S. Perelygin and G.P. Mikhailov, Manifestations of ion-ion interaction in the Raman spectra of the nitrate ion, J. Appl. Spectrosc., 48(1988), No. 5, p. 516. DOI: 10.1007/BF00663465
    [73]
    R.W. Berg, D.H. Kerridge, and P.H. Larsen, NaNO2 + NaNO3 phase diagram: New data from DSC and Raman spectroscopy, J. Chem. Eng. Data, 51(2006), No. 1, p. 34. DOI: 10.1021/je050105n
    [74]
    K.C. Xu and Y. Chen, Temperature-dependent Raman spectra of mixed crystals of NaNO3–KNO3: Evidence for limited solid solutions, J. Raman Spectrosc., 30(1999), No. 3, p. 173. DOI: 10.1002/(SICI)1097-4555(199903)30:3<173::AID-JRS366>3.0.CO;2-2
    [75]
    S. Jayaraman, A.P. Thompson, O.A. von Lilienfeld, and E.J. Maginn, Molecular simulation of the thermal and transport properties of three alkali nitrate salts, Ind. Eng. Chem. Res., 49(2010), No. 2, p. 559. DOI: 10.1021/ie9007216
    [76]
    N.A. Ponyatenko and I.V. Radchenko, Study of structure of binary fused monovalent nitratesd with Ca(NO3)2, Ba(NO3)2 and Sr(NO3)2 by the method of Raman scattering, Ukr. J. Phys., 14(1969), No. 1, p. 20.
    [77]
    N.A. Ponyatenko and I.V. Radchenko, Orientational interaction and rotational motion of the {\mathrm{N}\mathrm{O}}_{3}^{-} ion in monovalent nitrate melts, Opt. Spectrosc., 26(1969), p. 353.
    [78]
    B.L. Gao, F.G. Liu, and Z.W. Wang, Raman spectra of NaNO2–KNO3–NaNO3 ternary molten salts, Rare Met., 28(2009), No. S1, p. 581.
    [79]
    J.Y. Zhao, Structure Study on the Ternary NaNO3KNO3NaNO2 Molten Sallts [Dissertation], Qinghai Normal University, Xining, 2020, p. 77.
    [80]
    H.Y. Hou, J.L. You, Y.Q. Wu, H. Chen, and G.C. Jiang, Raman spectroscopic study of alkali carbonates, Chin. J. Light. Scatt., 13(2001), No. 3, p. 162.
    [81]
    L.J. Chen, J. Zuo, and C.J. Lin, In-situ Raman spectroscopy studies on the electrode process of cathode in MCFC, [in] Proceedings of the 13th Chinese Symposium on Molecular Spectroscopy, Xiamen, 2004, p. 199.
    [82]
    H. Ohata, K. Takeuchi, K. Ui, and N. Koura, The structure of molten lithium carbonate calculated by DFT and MD simulations, ECS Trans., 6(2007), No. 14, p. 57. DOI: 10.1149/1.2811943
    [83]
    N. Koura, S. Kohara, K. Takeuchi, et al., Alkali carbonates: Raman spectroscopy, ab initio calculations, and structure, J. Mol. Struct., 382(1996), No. 3, p. 163. DOI: 10.1016/0022-2860(96)09314-3
    [84]
    Y.Z. Wang, Influence of Additives on the Structure and Properties of Aluminum Electrolysis Molten Salt system [Dissertation], Kunming University of Science and Technology, Kunming, 2015, p. 93.
    [85]
    S.L. Jiang, C.M. Ye, Y.L. Liu, et al., Insights into the effects of fluoride anions on the electrochemical behavior and solution structure of trivalent samarium in LiCl–KCl molten salt, Electrochim. Acta, 439(2023), art. No. 141733. DOI: 10.1016/j.electacta.2022.141733
    [86]
    J.L. You, T. Zhao, S. Petrik, Y.Y. Wang, and X.W. Liu, In-situ high temperature Raman spectroscopy and quantum chemical Ab initio simulation on species in molten NaF–AlF3 fluorides, [in] Proceedings of the 17th National Symposium on Molecular Spectroscopy, Shaoguan, 2012, p. 223.
    [87]
    J. Wang and C.L. Liu, Temperature and composition dependences of shear viscosities for molten alkali metal chloride binary systems by molecular dynamics simulation, J. Mol. Liq., 273(2019), p. 447. DOI: 10.1016/j.molliq.2018.10.062
    [88]
    Y.L. Liu, J.H. Lan, L. Wang, et al., The influence of F ion on the electrochemical behavior and coordination properties of uranium in LiCl–KCl molten salt, Electrochimica Acta, 404(2022), art. No. 139573. DOI: 10.1016/j.electacta.2021.139573
    [89]
    Y.K. Wu, G.Q. Yan, S. Chen, and L.J. Wang, Electrochemistry of Hf(IV) in NaCl–KCl–NaF–K2HfF6 molten salts, Int. J. Miner. Metall. Mater., 27(2020), No. 12, p. 1644. DOI: 10.1007/s12613-020-2083-3
    [90]
    J.X. Song, Q.Y. Wang, J.Y. Wu, S.Q. Jiao, and H.M. Zhu, The influence of fluoride ions on the equilibrium between titanium ions and titanium metal in fused alkali chloride melts, Faraday Discuss., 190(2016), p. 421. DOI: 10.1039/C6FD00007J
    [91]
    S.S. Liu, S.L. Li, C.H. Liu, J.L. He, and J.X. Song, Effect of fluoride ions on coordination structure of titanium in molten NaCl–KCl, Int. J. Miner. Metall. Mater., 30(2023), No. 5, p. 868. DOI: 10.1007/s12613-022-2527-z
    [92]
    R. Yuan, C. Lü, H.L. Wan, et al., Effect of fluoride addition on electrochemical behaviors of V(III) in molten LiCl−KCl, Trans. Monferrous Met. Soc. China, 32(2022), No. 8, p. 2736. DOI: 10.1016/S1003-6326(22)65980-6
    [93]
    X. Bai, S. Li, J. He, and J. Song, Effect of fluoride-ion on the electrochemical behavior of tantalum ion in NaCl–KCl molten salt, J. Electrochem. Soc., 169(2022), No. 8, art. No. 082504. DOI: 10.1149/1945-7111/ac8379
    [94]
    H. Guo, J. Li, H.L. Zhang, et al., Study on micro-structure and transport properties of KF–NaF–AlF3–Al2O3 system by first-principles molecular dynamics simulation, J. Fluorine Chem., 235(2020), art. No. 109546. DOI: 10.1016/j.jfluchem.2020.109546
    [95]
    C.Y. Wang, Raman Spectroscopy and Theoretical Calculation of Metal Fluoride and Fluoride Oxides in Fluoride Molten Salt [Dissertation], University of Chinese Academy of Sciences, Beijing, 2021, p. 135.
    [96]
    X. Chen, H. Fu, and C. Wang, Influence of oxide ions on the speciation in molten KF–ZrF4 and KF–HfF4: A Raman spectroscopic and theoretical investigation, J. Mol. Liq., 342(2021), art. No. 117476. DOI: 10.1016/j.molliq.2021.117476
    [97]
    X. Wang, C.F. Liao, and L.S. Luo, Effect of Nd2O3 on the properties and structure of AlF3–(Na/Li)F–Al2O3 melt, Chin. Rare Earth, 38(2017), No. 5, p. 1.
    [98]
    X.Y. Liu, Y.J. Li, B.Z. Wang, and C.Y. Wang, Raman and density functional theory studies of lutecium fluoride and oxyfluoride structures in molten FLiNaK, Spectrochim. Acta A: Mol. Biomol. Spectrosc., 251(2021), art. No. 119435. DOI: 10.1016/j.saa.2021.119435
  • Related Articles

    [1]Yuhui Liu, Meng Tang, Shuang Zhang, Yuling Lin, Yingcai Wang, Youqun Wang, Ying Dai, Xiaohong Cao, Zhibin Zhang, Yunhai Liu. U(VI) adsorption behavior onto polypyrrole coated 3R-MoS2 nanosheets prepared with the molten salt electrolysis method [J]. International Journal of Minerals, Metallurgy and Materials, 2022, 29(3): 479-489. DOI: 10.1007/s12613-020-2154-5
    [2]Tai-qi Yin, Yun Xue, Yong-de Yan, Zhen-chao Ma, Fu-qiu Ma, Mi-lin Zhang, Gui-ling Wang, Min Qiu. Recovery and separation of rare earth elements by molten salt electrolysis [J]. International Journal of Minerals, Metallurgy and Materials, 2021, 28(6): 899-914. DOI: 10.1007/s12613-020-2228-4
    [3]Xiao-li Xi, Ming Feng, Li-wen Zhang, Zuo-ren Nie. Applications of molten salt and progress of molten salt electrolysis in secondary metal resource recovery [J]. International Journal of Minerals, Metallurgy and Materials, 2020, 27(12): 1599-1617. DOI: 10.1007/s12613-020-2175-0
    [4]Shu-qiang Jiao, Han-dong Jiao, Wei-li Song, Ming-yong Wang, Ji-guo Tu. A review on liquid metals as cathodes for molten salt/oxide electrolysis [J]. International Journal of Minerals, Metallurgy and Materials, 2020, 27(12): 1588-1598. DOI: 10.1007/s12613-020-1971-x
    [5]George Z. Chen. Interactions of molten salts with cathode products in the FFC Cambridge Process [J]. International Journal of Minerals, Metallurgy and Materials, 2020, 27(12): 1572-1587. DOI: 10.1007/s12613-020-2202-1
    [6]Ying Liu, Yong-an Zhang, Wei Wang, Dong-sheng Li, Jun-yi Ma. Microstructure and electrolysis behavior of self-healing Cu-Ni-Fe composite inert anodes for aluminum electrowinning [J]. International Journal of Minerals, Metallurgy and Materials, 2018, 25(10): 1208-1216. DOI: 10.1007/s12613-018-1673-9
    [7]Le Wang, Bing Li, Miao Shen, Shi-yan Li, Jian-guo Yu. Corrosion resistance of steel materials in LiCl-KCl melts [J]. International Journal of Minerals, Metallurgy and Materials, 2012, 19(10): 930-933. DOI: 10.1007/s12613-012-0649-4
    [8]Ze-quan Li, Li-yue Ru, Cheng-guang Bai, Na Zhang, Hai-hua Wang. Effect of sintering temperature on the electrolysis of TiO2 [J]. International Journal of Minerals, Metallurgy and Materials, 2012, 19(7): 636-641. DOI: 10.1007/s12613-012-0606-2
    [9]Zhuo-fei Cai, Zhi-mei Zhang, Zhan-cheng Guo, Hui-qing Tang. Direct electrochemical reduction of solid vanadium oxide to metal vanadium at low temperature in molten CaCl2-NaCl [J]. International Journal of Minerals, Metallurgy and Materials, 2012, 19(6): 499-505. DOI: 10.1007/s12613-012-0586-2
    [10]Huimin Lu, Zuxian Qiu, Keming Fang, Fuming Wang, Yanruo Hong. Current Efficiency of Low Temperature Aluminum Electrolysis Studied by Neural Network [J]. International Journal of Minerals, Metallurgy and Materials, 1999, 6(2): 107-110.
  • Cited by

    Periodical cited type(8)

    1. Huakui Zhang, Zepeng Lv, Shaolong Li, et al. Electrochemical behavior of molybdenum ions and their coordination structures in NaCl-KCl melt containing fluoride ions. Journal of Molecular Liquids, 2025, 426: 127375. DOI:10.1016/j.molliq.2025.127375
    2. Hang Liu, Chao-yun Yang, Xing Li, et al. Preparation of low-oxygen La/Ce mischmetal by molten salt electrolysis. Journal of Iron and Steel Research International, 2025. DOI:10.1007/s42243-024-01422-z
    3. Shaolong Li, Zepeng Lv, Jilin He, et al. Anodic dissolution of consumable ZrC O and cathodic deposition of zirconium in NaCl-KCl based melts. Journal of Molecular Liquids, 2025, 420: 126831. DOI:10.1016/j.molliq.2024.126831
    4. Huakui Zhang, Zepeng Lv, Shaolong Li, et al. Electrochemical behavior and cathodic nucleation mechanism of molybdenum ions in NaCl-KCl. Separation and Purification Technology, 2024, 329: 125121. DOI:10.1016/j.seppur.2023.125121
    5. Xinda Wang, Liwen Zhang, Liwen Ma, et al. Multiscale modeling of Na2WO4–WO3–CoO melt diffusion layer for molecular dynamics combined with finite element analysis. Materials Chemistry and Physics, 2024, 311: 128589. DOI:10.1016/j.matchemphys.2023.128589
    6. Yifan Zhao, Zhiyuan Li, Shijie Li, et al. A review of in-situ high-temperature characterizations for understanding the processes in metallurgical engineering. International Journal of Minerals, Metallurgy and Materials, 2024, 31(11): 2327. DOI:10.1007/s12613-024-2891-y
    7. Junchao Wu, Zhaoyang Yin, Qichi Le, et al. Molecular dynamics study on molten salt structure and density for magnesium alloy flux system. Materials Today Communications, 2024, 40: 109821. DOI:10.1016/j.mtcomm.2024.109821
    8. Junchao Wu, Zhaoyang Yin, Qichi Le, et al. Molecular dynamics study on the thermophysical properties of KCl-CaCl2-NaCl ternary salt for magnesium alloy smelting. Materials Today Sustainability, 2024, 28: 100980. DOI:10.1016/j.mtsust.2024.100980

    Other cited types(0)

Catalog

    Figures(7)  /  Tables(5)

    Share Article

    Article Metrics

    Article views (1279) PDF downloads (159) Cited by(8)

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return