Jurnal Kimia

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Subscriber access provided by Argonne Research Library

B: Liquids, Chemical and Dynamical Processes in Solution, Spectroscopy in Solution


Temperature Dependence of Short and Intermediate
Range Order in Molten MgCl2 and its Mixture with KCl
Fei Wu, Shobha Sharma, Santanu Roy, Phillip Halstenberg, Leighanne C. Gallington, Shannon M.
Mahurin, Sheng Dai, Vyacheslav S. Bryantsev, Alexander S. Ivanov, and Claudio Javier Margulis
J. Phys. Chem. B, Just Accepted Manuscript • DOI: 10.1021/acs.jpcb.0c00745 • Publication Date (Web): 09 Mar 2020
Downloaded from pubs.acs.org on March 23, 2020

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a service to the research community to expedite the dissemination
of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in
full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully
peer reviewed, but should not be considered the official version of record. They are citable by the
Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore,
the “Just Accepted” Web site may not include all articles that will be published in the journal. After
a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web
site and published as an ASAP article. Note that technical editing may introduce minor changes
to the manuscript text and/or graphics which could affect content, and all legal disclaimers and
ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or
consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W.,


Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 28 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8 Temperature Dependence of Short and
9
10
11
12
Intermediate Range Order in Molten MgCl2
13
14
15 and its Mixture with KCl
16
17
18
19
Fei Wu,†,§ Shobha Sharma,†,§ Santanu Roy,‡ Phillip Halstenberg,‡ Leighanne C.
20
21 Gallington,¶ Shannon M. Mahurin,‡ Sheng Dai,‡ Vyacheslav S. Bryantsev,∗,‡
22
23
24
Alexander S. Ivanov,∗,‡ and Claudio J. Margulis∗,†
25
26
†Department of Chemistry, University of Iowa, Iowa City, IA 52242
27
28
‡Chemical Sciences Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831
29
30
¶X-ray Science Division, Advanced Photon Source, Argonne National Laboratory, Argonne, IL
31
32
60439
33
34
§These authors contributed equally
35
36
37
E-mail: bryantsevv@ornl.gov; ivanova@ornl.gov; claudio-margulis@uiowa.edu
38
39
Phone: +1 (319) 335-0615
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 1
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 2 of 28

1
2
3
Abstract
4
5
6 This article addresses the non-Debye-Waller temperature behavior in the intermediate range
7
8 order for molten MgCl2 and its mixtures with KCl from a theory, Molecular Dynamics and ex-
9
10 perimental X-ray scattering perspective and puts these findings in the context of discussions
11
and controversies extending at least four decades. We find that these liquids are defined by two
12
13 structural motifs. The first motif is associated with chains of positive-negative charge alterna-
14
15 tion; the second motif, which results in a prepeak in the structure function S(q), is associated
16
17 with the interaction of Mg2+ and Cl− ions that do not belong to the same charge alternation
18
19 chain or aggregate. Our complementary X-ray scattering and computational results provide
20
21 a quantitative explanation for the increase in intensity of the prepeak with temperature as op-
22
23 posed to the behavior of other peaks following normal Debye-Waller behavior. Temperature
24
25 has opposite effects on the prevalence of each of the two structural motifs, and the enhance-
26
ment of one pattern appears to be at the detriment of the other. Whereas the intensity in S(q)
27
28 associated with the charge alternation motif is diminished at higher temperature the opposite
29
30 is true for the prepeak associated with intermediate range order due to the second structural
31
32 motif.
33
34
35
36
37 Introduction
38
39
40 The quest for cleaner, safer and less costly sources of energy has sparked intense renewed interest in
41
42 inorganic molten salts for use as coolants in new molten salt reactors. In such complex and extreme
43
44 environments a thorough understanding of structural, dynamical and thermodynamic properties of
45
46 the melt is fundamentally important. Significant recent work has been done on fluoride-based
47
48 molten salt mixtures, 1–9 but it is the chlorides –including those subject of this article–, which are
49
50 seen as particularly relevant to the next-generation nuclear reactors 10–12 and concentrated solar
51
52 power. 13 If we consider the early work of Roshenhein 14 and Zachariasen 15 there is almost a cen-
53
54 tury of work attempting to address the structure and dynamics of networked liquids. MgCl2 has for
55
56 decades been linked to ZnCl2 and more generally to AX2 -type systems such as SiO2 or GeSe2 that
57
58
59 2
60 ACS Paragon Plus Environment
Page 3 of 28 The Journal of Physical Chemistry

1
2
3
are known or expected to form tetrahedral or in cases square planar units 16–32 with the possibility
4
5
of sharing counterions to form chains, clusters or networks. However, not all prior work agrees
6
7
on the fact that MgCl2 is four-coordinated in the liquid phase, 22,33,34 and our own recent work
8
9 combining ab-initio molecular dynamics simulations, recently fit classical polarizable force fields,
10
11 and X-ray scattering experiments paint a more nuanced picture of the MgCl2 -KCl mixtures in the
12
13 liquid phase. 35 In this recent work, we found that the coordination number of Mg2+ can be either
14
15 four, five or six with five (square pyramidal or a octahedron with a missing Cl− ) being the more
16
17 dominant unit when the KCl fraction is low; four-coordination only becomes more dominant as
18
19 MgCl2 is significantly diluted with KCl and the networks templated by Mg2+ become significantly
20
21 disrupted. 35 Caution is therefore required when making general comparisons with other AX2 -type
22
23 systems as some of the structural features are likely to be the same but others may not, in particular
24
25 when invoking tetrahedrality.
26
27 Then, there is the issue of the first sharp diffraction peak (FSDP) or prepeak in X-ray and
28
29 neutron scattering experiments in these systems that is associated with intermediate range order
30
31 and which is often but not always seen to have non-Debye-Waller behavior (its intensity increases
32
33 with increasing temperature). 22–27,32–34,36–56 Here also, MgCl2 is commonly taken as yet another
34
35 case in the family of AX2 systems where tetrahedral structures dominate the liquid and glassy
36
37 states. For these tetrahedral systems, multiple explanations have been put forth to account for
38
39 the FSDP, including voids caused by the specific geometry of tetrahedral systems, 46,49,53,55–61
40
41 layered structures 25,26,36,39,41 that swell at higher temperature, and low density regions between
42
43 clusters. 45,50,52 Our simulations, which are consistent with our experimental results, imply that it
44
45 is not necessary to invoke the properties of tetrahedral units to arrive at the MgCl2 and MgCl2 -KCl
46
47 prepeak with non Debye-Waller behavior. Yet, we find that many of the concepts put forth in the
48
49 past for AX2 -type systems as well as other networked systems in the liquid and glassy states still
50
51 properly apply to MgCl2 and its mixtures with KCl and, as we describe our findings, we attempt
52
53 to make connections with these earlier interpretations.
54
55 As cleanly explained in a recent review article by Wilson, 32 prepeaks can form for multiple
56
57
58
59 3
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 4 of 28

1
2
3
reasons; for example, ionic size mismatch or charge mismatch can cause the main peak to split
4
5
into two (the main peak is what in our prior articles we have commonly referred to as the charge
6
7
alternation peak 35 ), another possibility is the effect of anionic polarizability resulting in divalent
8
9 cations being at distances closer than they normally would be and forming networks, clusters,
10
11 rings or other possible edge-, corner- or in general anion-shared structures. In the case of room-
12
13 temperature molten salts –the ionic liquids–, there is also a prepeak that appears because of charge
14
15 networks that are alternated by apolar domains. For MgCl2 , there is consensus that the origin of
16
17 the prepeak stems from the effects of polarizability. 29,30,35,62 Without polarizability, it appears to
18
19 be impossible to capture the correct Mg2+ -Mg2+ distance in the condensed phase 29,35 as well as
20
21 the overall structure of the liquid. 35,62 It is the networked structure of the liquid (or more precisely
22
23 the spacing between chains) that we have recently argued is related to the prepeak in MgCl2 and
24
25 its mixtures with KCl. In this article, we attempt to describe in a more quantitative way, what
26
27 happens to these networks as temperature is increased (and the density concomitantly decreases)
28
29 and how this results in the non Debye-Waller behavior observed for MgCl2 and its mixture with
30
31 KCl without the need of arguments associated with tetrahedral structures.
32
33
34
35
36 Methods
37
38
39 In a recent article, 35 we presented in the context of prior work 16–19,21,23,31,38,54,63 a thorough anal-
40
41 ysis of high-energy X-ray scattering data using first-principles and classical Molecular Dynamics
42
43 (MD) simulations that employ a novel polarizable ion model. 64,65 Although simulations includ-
44
45 ing polarizable ions are computationally expensive (simulations in the current study took several
46
47 months to run), the inclusion of polarizability is fundamentally important among other things to
48
49 capture the correct inter-cationic distance for divalent ions. 29,30,35,62 For this study, X-ray diffrac-
50
51 tion measurements were performed at the Advanced Photon Source (beamline 11-ID-B) using a
52
53 versatile sample holder composed of a quartz tube located between two resistive heaters 66 (see
54
55 Figure 1); data collected for a 50% mole fraction mixture of MgCl2 and KCl at 873 K, 973 K,
56
57
58
59 4
60 ACS Paragon Plus Environment
Page 5 of 28 The Journal of Physical Chemistry

1
2
3
and 1073 K allowed for the observation of structural progressions in S(q). A full description of
4
5
protocols was already given in reference 35 but we include an abbreviated description here.
6
7
8
9 Preparation of the MgCl2-KCl Salt Mixture
10
11
12 Commercial MgCl2 (Sigma Aldrich, anhydrous, 99.9%) was distilled using a quartz distillation
13
14 chamber, to achieve ultrahigh purity. 67 The detailed distillation procedures are described in our
15
16 prior work. 35 The obtained anhydrous MgCl2 salt was mixed with ultradry, high purity KCl beads
17
18 (Sigma Aldrich) to a 50/50 mole ratio under inert atmosphere conditions. The mixture was fused
19
20 at 1073 K resulting in a transparent homogeneous liquid (Figure 1(a)), cooled down, crushed to a
21
22 fine powder and loaded into a thin-walled quartz capillary (Charles Supper, 1.0 mm O.D., 0.010
23
24 mm wall thickness). Prior to loading, a 6 mm quartz tube was fused to the capillary enabling
25
26 us to connect a vacuum fitting to the capillary. A valve on the vacuum fitting was closed before
27
28 removing from the glovebox. The capillary was slowly evacuated and then flame sealed near the
29
30 top for the synchrotron X-ray scattering experiments.
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47 Figure 1: (a) 50% MgCl2 /KCl molten salt mixture at ∼1073 K. (b) Experimental setup for X-ray
48 scattering experiments: 1- quartz NMR tube, 2 - thermocouple, 3 - heater, 4 - salt.
49
50
51
52
53
54
55
56
57
58
59 5
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 6 of 28

1
2
3
4
X-ray Scattering Experiments
5
6 High-energy X-ray diffraction data for the 50% MgCl2 - KCl mixture were collected at the Ad-
7
8 vanced Photon Source (APS), beamline 11-ID-B (0.2113 Å (58.7 keV) wavelength, 0.5×0.5 mm2
9
10 beam size). An amorphous silicon area detector (PerkinElmer XRD 1621) was placed 170 mm
11
12 behind the sample, giving an accessible q range of 0.5-24 Å−1 . A versatile sample holder (Fig-
13
14 ure 1(b)) originally designed by Chupas et al. 66 at the 11-ID-B beamline was used to collect the
15
16 data at 873 K, 973 K, and 1073 K. The empty quartz capillary scattering was also measured at 873
17
18 K, 973 K, and 1073 K to account for its possible thermal expansion in the background subtraction
19
20 procedure.
21
22 Calibration of the precise sample to detector distance, detector tilt and rotation, and beam
23
24 center was performed in GSAS II 68 using crystalline CeO2 powder as the standard; reduction of
25
26 2D images to 1D diffraction patterns was also performed in GSAS II. The background subtracted
27
28 scattering intensity was corrected in the PDFgetX2 69 software for inelastic Compton scattering,
29
30 sample self-attenuation, and detector effects and further converted to the structure function, S(q),
31
32 using the q-dependent atomic form factors corresponding to the elemental composition of the salt.
33
34
35 Icoh (q) − ∑i xi fi2 (q)
36 S(q) = (1)
[∑i xi fi (q)]2
37
38
39 Here Icoh (q) is the coherent scattering intensity, xi and fi (q) are the molar fraction and X-ray atomic
40
41 form factor of species i, and q denotes the magnitude of the scattering vector (q = 4π sin θ
42 λ , 2θ is
43 the scattering angle, and λ is the incident X-ray wavelength). Our experimentally derived S(q)
44
45 functions are shown in Figure S3 as a function of temperature up to 20 Å−1 .
46
47
48
49 Simulation protocol
50
51
52 To study the effect of temperature on the structure of MgCl2 and its mixture with KCl, two simu-
53
54 lation boxes were generated. The first box contained 1000 Mg2+ ions and 2000 Cl− ions, whereas
55
56 the second contained 500 Mg2+ ions, 500 K+ ions and 1500 Cl− ions. At all target temperatures
57
58
59 6
60 ACS Paragon Plus Environment
Page 7 of 28 The Journal of Physical Chemistry

1
2
3
(1073 K, 1173 K, 1273 K for MgCl2 and 873 K, 973 K, 1073 K, 1173 K for the 50% MgCl2 KCl
4
5
mixture) the systems were equilibrated using the following protocol: 50 ps initial equilibration at
6
7
target temperature and ambient pressure in the NPT ensemble (constant pressure and constant tem-
8
9 perature) followed by 200 ps simulated annealing where the temperature was increased to 1640 K
10
11 and then decreased to target values. Following this, the system was further equilibrated in the NPT
12
13 ensemble for 2 ns at target temperature. The last 1 ns of this run was used to compute S(q) and its
14
15 subcomponents as well as for the analysis of aggregate sizes using the code by Bernardes and col-
16
17 laborators. 70–72 As in our prior study 35 cations and anions were represented by a polarizable ion
18
19 model (PIM). 64,65 Simulations were carried out using the FIST module of the CP2K software. 73
20
21 In particular the parameters for MgCl2 in this model are those introduced in our prior study. 35 A
22
23 non-bonded cutoff of 15 Å was used for all simulation. Electrostatic interactions were handled us-
24
25 ing the Ewald summation as coded in CP2K. The Nose Hoover thermostat and barostat 74,75 with a
26
27 time constant of 1 ps were used to control the temperature and pressure (1 bar). A timestep of 1 fs
28
29 was used to integrate the equations of motion. Using methodology we have thoroughly described
30
31 in prior work, 76–81 S(q) can be computed and decomposed as described in Eqns. S1.
32
33
34
35
36 Results and Discussion
37
38
39 For pure MgCl2 and the 50% mole fraction mixture of this salt with KCl, Fig. 2 shows S(q) as
40
41 a function of temperature. The range in which we choose to plot these functions highlights the
42
43 temperature-dependent phenomena we want to discuss (see also Figs. S1 through S3 for extended
44
45 range computational and experimental results). Whereas it is easy to determine the origin of the
46
47 peak at about 2Å−1 which we associate with positive-negative charge alternation (vide infra), the
48
49 origin of the peak below 1Å−1 associated with intermediate range order is more obscure. Notice
50
51 that the behavior in Fig. 2 is the same as that observed for many AX2 systems 26,43–45,50,56 in that
52
53 concomitant with an increase in temperature (and molar volume) there is a decrease in intensity
54
55 for the charge alternation peak but an increase of intensity for the prepeak. This anomalous or
56
57
58
59 7
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 8 of 28

1
2
3 Simulated Simulated Experimental
4 100% MgCl2 50% MgCl2/KCl 50% MgCl2/KCl
5 0.4
6
7 0.2
8
9 0
10
S(q)

11
-0.2
12 873 K
-0.4 1073 K 973 K 873 K
13
14 -0.6 1173 K 1073 K 973 K
15 1273 K 1173 K 1073 K
16 -0.8
17 1 2 3 1 2 3 1 2 3
18 -1
19
q (Å )
20 (a) (b) (c)
21
22 Figure 2: S(q) at different temperatures in the molten state for (a) pure MgCl2 based on MD
23
24 simulations, (b) mixture of 50% MgCl2 and KCl based on MD simulations, and (c) mixture of
25 50% MgCl2 and KCl based on experimental X-ray scattering data. Arrows highlight the decrease
26 in intensity at q-values associated with charge alternation and the anomalous increase in intensity
27 associated with the first sharp diffraction peak. Figures S1 – S3 show the temperature-dependent
28 S(q) in an extended q-range.
29
30
31
non Debye-Waller behavior of the prepeak is in fact common, and for AX2 -type systems in which
32
33
this is not observed such as vitreous SiO2 , there is zero or little density decrease with increase
34
35
of temperature. 82 For example, in the case of GeSe2 , Vashishta showed that if in simulation one
36
37 forces the volume to remain constant but increases the temperature, there is no anomalous change
38
39 in the prepreak. 43,44 This implies that an increase in molar volume with temperature appears to be
40
41 important to the structural rearragements required for non Debye-Waller behavior of the prepeak.
42
43 Computationally, there are two routes to computing S(q); the first one is via Fourier trans-
44
45 formation of pair distribution functions 76–81 and the second one is via the direct summation in
46
47 reciprocal space over atomic pairs. 35,83,84 Results from molecular dynamics simulations shown in
48
49 Figure 2 use the first method. The sum in reciprocal space over atom pairs can instead be very ad-
50
51 vantageous when attempting to gauge from simulation the pattern of ionic interactions giving rise
52
53 to features at specific q values in the overall S(q) or its subcomponents; the reader is encouraged to
54
55 revisit section 2.3 of reference 83 for an in-detail example of this (see also references 35 and 84).
56
57
58
59 8
60 ACS Paragon Plus Environment
Page 9 of 28 The Journal of Physical Chemistry

1
2
3
In the current study, we use this approach to find the type of ion-ion correlations that for MgCl2
4
5
contribute significantly to each of the two highlighted peaks in Figure 2 while at the same time
6
7
having inter-ionic distances consistent with the Bragg condition d ≈ 2π/q peak . For MgCl2 , sorting
8
9 through a simulation frame, one finds the typical patterns shown in Figure 3 which can be linked
10
11 to the two peaks in Figure 2.
12
13 On the left in Figure 3 we find two Mg2+ ions decorated by Cl− counterions. At the waist of
14
15 the two cations there are two "shared" anions (this is what is commonly referred to as edge-sharing
16
17 units). Such pattern of cations alternated by counterions is the hallmark of all ionic liquids and
18
19 molten salts and gives rise to the charge alternation peak (or main peak) at around 2Å−1 . This
20
21 assignment is clearly confirmed by Figure 4 where cationic and anionic partial subcomponents of
22
23 S(q) have peaks whereas cationic-anionic subcomponents have antipeaks. 76–81,85–89 Notice that
24
25 the charge alternation regime (q ≈ 2 Å−1 ) follows "normal" Debye-Waller behavior in that cor-
26
27 relations and anticorrelations diminish in magnitude with higher temperature. Instead, patterns
28
29 on the right in Figure 3 importantly contribute to the peak at q ≈ 0.85 Å−1 while satisfying the
30
31 Bragg condition d ≈ 7 or 8 Å. The highlighted interactions (red dashes) are between cations and
32
33
34
35
36
37
38
39
40
41
42 Figure 3: (Left) As blue surface and not to scale, two Mg2+ ions decorated by Cl− counterions;
43
44 at their waist, two of the Cl− are "shared". For MgCl2 the charge alternation peak at around
45 q ≈ 2 Å−1 is associated with the typical distance between two same-charge ions alternated by
46 "shared" opposite-charge ions. This typical separation is highlighted with a solid-red bracket.
47 (Right) As blue surfaces and not to scale, Cl− -decorated Mg2+ ions separated from other Cl− -
48 decorated Mg2+ ions that do not belong to the same charge network or aggregate. Cl− belonging
49
50 to different networks or aggregates are highlighted in green and teal. Interactions across networks
51 or aggregates give rise to the first sharp diffraction peak. At the typical distance 2π/q prepeak , Mg2+
52 ions in different aggregates do not share Cl− counterions. Example ion distances important to the
53 prepeak are highlighted with red dashes. In all cases, structures are from simulations at 1273K.
54
55
56
57
58
59 9
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 10 of 28

1
2
3
4 Mg-Mg -0.4 Mg-Cl 0.9 Cl-Cl
5 0.2
6 -0.5 0.8
7 0.1
8 -0.6 0.7
S(q)

9 0
10 -0.7 0.6
11 1073 K
12
-0.1
1173 K -0.8 0.5
13
-0.2 1273 K
14
15 1.5 1.75 2 2.25 2.5 1.5 1.75 2 2.25 2.5 1.5 1.75 2 2.25 2.5
16 -1
17 q (Å )
18
19 (a) (b) (c)
20
21 Figure 4: For neat MgCl2 as a function of temperature, computationally derived (a) cation-cation,
22 (b) cation-anion and (c) anion-anion partial subcomponents of S(q) in the charge alternation region.
23 Same-charge species show peaks and opposite-charge species show antipeaks in this q-range; this
24
is the signature of liquid-phase charge alternation. These peaks and antipeaks show normal tem-
25
26 perature behavior where at higher temperature structural correlations diminish. Figure S1 shows
27 the temperature-dependent S(q) in an extended q-range.
28
29
30 cations, cations and anions as well as anions and anions "across" aggregates or networks as op-
31
32 posed to "along" them. Specifically, charge alternation "along" networks giving rise to the peak at
33
34 q ≈ 2Å−1 requires anionic "sharing" by cations. Instead, the prepeak –associated with intermedi-
35
36 ate range order– is based on the repetitive pattern of interactions of ions in different but adjacent
37
38 aggregates/networks that are randomly distributed (see Figs. 2bc in reference 35). The key distinc-
39
40 tion between the prepeak and the charge alternation peak is that interactions associated with the
41
42 former have Mg2+ ions in different but adjacent aggregates that do not share Cl− (the unshared
43
44 Cl− ions for the two adjacent aggregates are highlighted as green and teal respectively for clarity
45
46 in Figure 3). This distinction between cations that edge-share anions with those that do not edge-
47
48 share anions is already clear in the crystal structure of MgCl2 . 90 In our recent work, 35 we show
49
50 how Miller planes associated with the two types of interactions in the crystal 90 nicely coincide
51
52 with the prepeak and the charge alternation peak in the molten phase. Similar types of arguments
53
54 on the origin of the prepeak are ubiquitous in the literature, particularly while describing AX2 sys-
55
56 tems. 26,32,33,36,39,43–45,50 The reader should understand that in the molten state not all Mg2+ ions
57
58
59 10
60 ACS Paragon Plus Environment
Page 11 of 28 The Journal of Physical Chemistry

1
2
3
form chains and some of the interactions are with anion-decorated single cations forming what
4
5
looks like a "complex"; this is seen in two of the examples shown in Figure 3.
6
7
Similar to the case of other AX2 type systems, 26,32,33,36,39,43–45,50 both MgCl2 and its mixture
8
9 with KCl display intermediate range order (manifested as a prepeak) that follows non Debye-
10
11 Waller behavior; this is in contrast to what is observed for the charge alternation peak at around
12
13 2Å−1 . The prepeak in Figure 2 and corresponding subcomponents in Figure 5 become more in-
14
15 tense at higher temperature; this implies that the type of intermediate-range order associated with
16
17 it is more pronounced at higher temperature. But, why is this so and what structural rearrage-
18
19 ments make this possible? Our hypothesis is that temperature and concomitant density changes
20
21
22
23 Mg-Mg Mg-Cl -0.64 Cl-Cl
24
25
-0.06 0.24
26 -0.66
27 -0.08 0.22
28 -0.68
S(q)

29 -0.1 0.2
30 -0.7
31 -0.12 1073 K 0.18
32 1173 K -0.72
33 -0.14 1273 K 0.16
34
35 0.5 0.75 1 1.25 0.5 0.75 1 1.25 0.5 0.75 1 1.25
36 -1
q (Å )
37
38 (a) (b) (c)
39
40 Figure 5: For neat MgCl2 as a function of temperature, computationally derived (a) cation-cation,
41
42
(b) cation-anion and (c) anion-anion partial subcomponents of S(q) in the first sharp diffraction
43 peak region. The change in intensity as a function of temperature for cationic correlations is small,
44 but cationic-anionic as well as anionic-anionic intensities change significantly. The change in this
45 region is "anomalous" in the sense that correlations appear more significant (larger peaks) at higher
46 temperature. This is opposite to the behavior for charge alternation observed in Figure 4. Figure S1
47
48
shows the temperature-dependent S(q) in an extended q-range.
49
50 enhance one of two the liquid patterns at the detriment of the other. Specifically, we propose that
51
52 longer, better structured edge-sharing Mg2+ charge networks should be favored at lower temper-
53
54 ature (higher density) resulting in prominent charge alternation features in S(q) but less intense
55
56 prepeaks. Instead, at higher temperature the aggregates should be smaller emphasizing the order
57
58
59 11
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 12 of 28

1
2
3
due to non-anion sharing aggregates that results in the prepeak. This argument is akin to what
4
5
Busse 36,39 proposed for As2 Se3 and Asx S1−x type glasses and Ellis for mixtures of ZnCl2 with
6
7
alkali metal halides. 57
8
9 To quantitatively test if high temperature (and concomitant lower densities) results in shorter
10
11 charge networks, we used the "Aggregates" software by Bernardes, 70 which can establish the prob-
12
13 ability distribution of aggregate sizes as defined by the user in its input. 71,72,91 To choose a rea-
14
15 sonable maximum cutoff defining the typical in-network Mg2+ -Mg2+ distance, we simply notice
16
17 from Figure 4(a) that the lowest q value consistent with the charge alternation peak is on the order
18
19 of 1.5 Å−1 . The reader is reminded that this peak is associated with the distance in real space
20
21 between two Mg2+ ions that share counterions as in Figure 3 (left) and 1.5 Å−1 corresponds to a
22
23 real space periodicity distance of 2π/q ≈ 4.2Å. We find that the quantitative aggregate size dis-
24
25 tribution depends on the cutoff choice but any reasonable distance between 3.8 and 4.2 Å leads
26
27 to the same conclusion, –higher temperature results in shorter networks–; this can be clearly seen
28
29 from Figure 6(a) and (b). In other words, consistent with the diminished peaks and antipeaks in
30
31 Figure 4 we find from the aggregates analysis that charge alternation networks are shorter at higher
32
33 temperature. Another way to see that high temperature disrupts the Cl− -decorated Mg2+ network
34
35 is by analyzing the 2-D free energy as a function of Mg2+ -Mg2+ distance and number of shared
36
37 Cl− (W (r, nshared )) in Figures 6 (c) and (d). 35 From these two figures it is clear that what at 1073
38
39 K was a prominent minimum at r ≈ 3.8 Å has become a much shallower minimum at 1273 K.
40
41 W (r = 3.8 Å, nshared = 2) corresponds to the case of two Mg2+ ions that edge-share two Cl− ions
42
43 as in Figure 3 (left). In other words, higher temperature (and lower density) disrupts the structural
44
45 motif that causes the charge alternation peak and the formation of Cl− -decorated Mg2+ networks.
46
47 If Mg2+ ions are less committed to Cl− edge-sharing networks, they must instead form shorter
48
49 aggregates –including the most common aggregate of one or "complex"–, that are necessarily sep-
50
51 arated by Cl− ions that are not shared like the ones depicted in green and teal in Figure 3 (right);
52
53 these are the structures giving rise to the prepeak. Hence, it is not surprising that a diminished set
54
55 of charge alternation peaks and antipeaks (normal behavior) at q ≈ 2 Å−1 go hand-in-hand with a
56
57
58
59 12
60 ACS Paragon Plus Environment
Page 13 of 28 The Journal of Physical Chemistry

1
2
3
higher intensity first sharp diffraction peak (anomalous behavior).
4
5
We now analyze the temperature dependence of the prepeak for the much more complex case
6
7
of the 50% mixture of MgCl2 and KCl with total S(q) from experiments and MD simulations
8
9 shown in Figure 2. The reader is reminded that the first sharp diffraction peak is solely due to
10
11 the MgCl2 component; molten KCl does not have a prepeak. 46–48,92 In fact, mixtures at differ-
12
13 ent ratios of MgCl2 and KCl have prepeaks of diminishing intensity when the proportion of KCl
14
15 is large and the fraction of Mg2+ ions being able to form adjacent aggregates that do not share
16
17 Cl− ions becomes small due to entropy or other considerations (see Figs. 2.abc in reference 35).
18
19 The effect of alkali halides on network-forming liquids such as ZnCl2 has been thoroughly dis-
20
21 cussed before; 34,46,48,49,57 for example Wilson and Madden 49 point to the fact that KCl acts as a
22
23 network-breaking component that can change strong liquids into fragile ones, 46 and in the case
24
25 of ZnCl2 the viscosity decreases due to network disruption. 46 Yet, the temperature effect on the
26
27 structural properties of the different individual mixture subcomponents as described by their partial
28
29 subcomponents in S(q) is perhaps less well understood and deserves scrutiny.
30
31 For our 50% mixture where MgCl2 can accommodate the same type of networks as in its neat
32
33 form, 35 Figure 7 shows that the behavior of the Mg2+ -Mg2+ , Mg2+ -Cl− , and Cl− -Cl− subcompo-
34
35 nents of S(q) is similar to that in neat MgCl2 . Both the Mg2+ -Cl− , and Cl− -Cl− subcomponents
36
37 are enhanced resulting in a larger prepeak at higher temperature. The effect of KCl on the interme-
38
39 diate range structure for the mixture is quite interesting and can be derived from careful analysis
40
41 of Figure 7. Normally, at the distance one expects to find two Mg2+ ions that are non edge-sharing
42
43 (in different but adjacent networks) the likelihood of one of these being replaced by K+ is low.
44
45 This –independent of temperature– is the reason for the antipeak in the Mg2+ -K+ subcomponent
46
47 of S(q) in Figure 7. The same can be said about the K+ -Cl− subcomponent; where one finds
48
49 the "type" of Cl− associated with two adjacent Mg2+ aggregates or networks (see the right-most
50
51 example in Figure 3(right)) it is highly unlikely that one of those may be replaced by K+ . This
52
53 is the reason for the antipeak in the K+ -Cl− subcomponent of S(q) in the prepeak region. Notice
54
55 that both the Mg2+ -K+ and K+ -Cl− antipeaks in Figure 7 become less intense as temperature in-
56
57
58
59 13
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 14 of 28

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
Figure 6: For neat MgCl2 as a function of temperature the distribution of in-network Mg2+ ag-
48 gregate sizes (in units of number of Mg2+ ions) from our molecular dynamics trajectories. Higher
49 temperatures favor shorter networks including the most common "network of one" Mg2+ ion. Sub-
50 figure (a) shows the shorter network side of the distribution whereas (b) the longer network side.(c)
51 and (d) show the free energy W (r, nshared ) at 1073 K and 1273 K respectively.
52
53
54
55
56
57
58
59 14
60 ACS Paragon Plus Environment
Page 15 of 28 The Journal of Physical Chemistry

1
2
3
creases. In other words, due to increased temperature (and lower density), the existence of K+ in
4
5
these "structurally-forbidden" regions is less unlikely. Perhaps counter-intuitively, these intensity-
6
7
diminished K+ -related antipeaks associated with a more disorganized (higher-temperature) molten
8
9 salt structure enhance the overall prepeak, as one is adding larger partial subcomponents to S(q).
10
11 In Figure 8 we show all subcomponents of S(q) for our mixture in the charge alternation (qcharge )
12
13 region. Just as in the case for pure MgCl2 in Figure 4, the behavior for the 50% mixture in the
14
15
16
17 -0.08 -0.04
18 0.04 Mg-Mg Mg-K K-K
19 -0.1 -0.06
20 0.02
21 -0.08
-0.12
S(q)

22 0
23
-0.14 -0.1 873 K
24 -0.02 973 K
25
-0.16 -0.12 1073 K
26 1173 K
27
-0.04
28 (a) (b) (c)
29
30
31
Mg-Cl K-Cl Cl-Cl
32
0.24 -0.12 -0.46
33
34 0.22 -0.14 -0.48
35
S(q)

36 0.2 -0.16 -0.5


37
38 0.18 -0.18 -0.52
39
40 0.16 -0.2 -0.54
41 0.5 0.75 1 1.25 0.5 0.75 1 1.25 0.5 0.75 1 1.25
42 -1 -1 -1
43 q (Å ) q (Å ) q (Å )
44 (d) (e) (f)
45
46 Figure 7: For our 50% mixture of MgCl2 and KCl as a function of temperature, computationally
47
48
derived partial subcomponents of S(q) in the first sharp diffraction peak region. In each case an
49 arrow highlights the direction of change with increasing temperature (see Fig. S2 for extended
50 q-range).
51
52
53 qcharge region is "normal" with peaks and antipeaks that are less pronounced at higher temperature.
54
55 This indicates that in the charge alternation region, both pure MgCl2 and its mixture with KCl are
56
57
58
59 15
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 16 of 28

1
2
3
less structured.
4
5
6 0.15
7 0.15 Mg-Mg Mg-K 0.15 K-K
8 0.1
9 0.1 0.1
10 873 K
S(q)

11 0.05 973 K
12 0.05 0.05
1073 K
13
0 1173 K
14 0 0
15
16 -0.05
17 (a) (b) (c)
18
19 -0.2 Mg-Cl K-Cl 0.45 Cl-Cl
20
21 -0.25
22 -0.25 0.4
23 -0.3
-0.3 0.35
S(q)

24
25 -0.35
26 -0.35 0.3
27 -0.4
28 -0.4 0.25
29
30 1.5 1.75 2 2.25 2.5 1.5 1.75 2 2.25 2.5 1.5 1.75 2 2.25 2.5
31 -1 -1 -1
32 q (Å ) q (Å ) q (Å )
33
(d) (e) (f)
34
35 Figure 8: Decomposition of computationally derived S(q) into its subcomponents in the charge
36
alternation region. In all cases the behavior is "normal" in that for all ion pairs, peaks and antipeaks
37
38 become less pronounced as temperature is increased. Notice that the charge alternation region
39 occurs at different q values for MgCl2 and KCl. This is to be expected due to the different nature
40 of the cation-anion interactions (see Fig. S2 for extended q-range).
41
42
43
44
45
46 Conclusions
47
48
49 Molten MgCl2 and its 50% mixture with KCl show structural motifs associated with (1) charge
50
51 alternation networks or aggregates where Mg2+ ions edge-share Cl− counterions and (2) interme-
52
53 diate range order associated with interactions of adjacent aggregates where Mg2+ ions belonging
54
55 to the different aggregates do not share Cl− counterions. We find that with increasing temperature
56
57
58
59 16
60 ACS Paragon Plus Environment
Page 17 of 28 The Journal of Physical Chemistry

1
2
3
one necessarily disrupts the MgCl2 charge network and both peaks and antipeaks in the charge
4
5
alternation region of S(q) are less pronounced (normal behavior). At higher temperature and lower
6
7
density, where shorter Mg2+ -templated charge networks or aggregates are favored, one necessarily
8
9 gets more structural correlations across aggregates that do not share Cl− ions (as opposed to along
10
11 them); this is the pattern associated with intermediate range order and an increase in the number of
12
13 such interactions results in an enhanced prepeak (i.e. anomalous behavior). Even though MgCl2
14
15 is not dominated by tetrahedral units, the system is clearly networked and the early concepts as-
16
17 sociated with across-layer spacing and interactions are not inconsistent with what we observe.
18
19 Interestingly, and counterintuitively, more disorder in the case when K+ is present also favors a
20
21 larger prepeak. How these structural changes may affect the solvation structure and solubility of
22
23 solutes such as fuel (U and Pu), fission products and corrosion products, or reactivity at interfaces,
24
25 is not well understood and should be the subject of future studies.
26
27
28
29
30 Acknowledgement
31
32
33 This work was supported as part of the Molten Salts in Extreme Environments (MSEE) Energy
34
35 Frontier Research Center, funded by the U.S. Department of Energy Office of Science. MSEE
36
37 work at Iowa was supported via subcontract from Brookhaven National Laboratory (BNL). BNL
38
39 and ORNL are operated under DOE contracts DE-SC0012704 and DE-AC05-00OR22725, respec-
40
41 tively. This research used resources of the Advanced Photon Source operated by Argonne National
42
43 Laboratory under Contract No. DE-AC02-06CH11357 and of the Oak Ridge Leadership Com-
44
45 puting Facility at the Oak Ridge National Laboratory supported by the Office of Science of the
46
47 U.S. Department of Energy under contract No. DE-AC05-00OR22725. FW, SS and CJM also
48
49 acknowledge the University of Iowa High Performance Computing Facility.
50
51
52
53
54
55
56
57
58
59 17
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 18 of 28

1
2
3
4
Supporting Information Available
5
6
The Supporting Information is available free of charge at XXX.
7
8
Decompositon of S(q) into ionic subcomponents, simulated S(q) in extended q-range as a function
9
10
of temperature for pure MgCl2 , simulated S(q) in extended q-range as a function of temperature
11
12
for the 50% mixture of MgCl2 and KCl, experimental S(q) in extended q-range as a function of
13
14
temperature for the 50% mixture of MgCl2 and KCl.
15
16
17
18
19 References
20
21
22 (1) Olson, L. C.; Ambrosek, J. W.; Sridharan, K.; Anderson, M. H.; Allen, T. R. Materials corro-
23
24 sion in molten LiF-NaF-KF salt. J. Fluor. Chem. 2009, 130, 67–73.
25
26
27 (2) Delpech, S.; Cabet, C.; Slim, C.; Picard, G. S. Molten fluorides for nuclear applications.
28
29 Mater. Today 2010, 13, 34–41.
30
31
(3) An, X.-H.; Cheng, J.-H.; Yin, H.-Q.; Xie, L.-D.; Zhang, P. Thermal conductivity of high
32
33
temperature fluoride molten salt determined by laser flash technique. Int. J. Heat Mass Transf.
34
35
2015, 90, 872–877.
36
37
38 (4) Nishiumi, R.; Fukada, S.; Yamashita, J.; Katayama, K.; Sagara, A.; Yagi, J. Hydrogen per-
39
40 meation through fluoride molten salt mixed with Ti powder. Fusion Sci. Technol. 2017, 72,
41
42 747–752.
43
44
45 (5) Song, J.; Huang, X.; Wu, J.; Zhang, X. Electrochemical behaviors of Ti(III) in molten NaCl-
46
47 KCl under various contents of fluoride. Electrochim. Acta 2017, 256, 252–258.
48
49
50 (6) Zong, G.; Zhang, Z.-B.; Sun, J.-H.; Xiao, J.-C. Preparation of high-purity molten FLiNaK
51
52 salt by the hydrofluorination process. J. Fluor. Chem. 2017, 197, 134–141.
53
54
(7) Raiman, S. S.; Lee, S. Aggregation and data analysis of corrosion studies in molten chloride
55
56
and fluoride salts. J. Nucl. Mater. 2018, 511, 523–535.
57
58
59 18
60 ACS Paragon Plus Environment
Page 19 of 28 The Journal of Physical Chemistry

1
2
3
(8) Goh, B.; Carotti, F.; Scarlat, R. O. A review of electrochemical and non-electrochemical
4
5
approaches to determining oxide concentration in molten fluoride salts. ECS Trans. 2018, 85,
6
7
1459–1471.
8
9
10 (9) Xue, W.; Yang, X.; Ye, X.-X.; Han, L.; Wang, J.; Ignatiev, V.; Zhou, X. Effects of silicon
11
12 carbide on the corrosion of metallic materials in molten LiF-NaF-KF salt. Corros. Sci. 2018,
13
14 143, 157–165.
15
16
17 (10) Forsberg, C. W. Thermal- and fast-spectrum molten salt reactors for actinide burning and
18
19 fuel production; 2007.
20
21
22 (11) Gehin, J. C.; Holcomb, D. E.; Flanagan, G. F.; Patton, B. W.; Howard, R. L.; Harrison, T. J.
23
24 Fast spectrum molten salt reactor options; 2011.
25
26
27 (12) Diamond, D. J.; Brown, N. R.; Denning, R.; Bajorek, S. Phenomena important in molten salt
28
29 reactor simulations; 2018.
30
31
(13) Ding, W.; Bonk, A.; Gussone, J.; Bauer, T. Electrochemical measurement of corrosive impu-
32
33
rities in molten chlorides for thermal energy storage. J. Energy Storage 2018, 15, 408–414.
34
35
36 (14) Rosenhain, W. The structure and constitution of glass. J. Soc. Glass Technol. Trans. 1927, 11,
37
38 77.
39
40
41 (15) Zachariasen, W. H. The atomic arrangement in glass. J. Am. Chem. Soc. 1932, 54, 3841–3851.
42
43
44 (16) Maroni, V. A.; Hathaway, E. J.; Cairns, E. J. Structural studies of magnesium halide-
45
46 potassium halide melts by Raman spectroscopy. J. Phys. Chem. 1971, 75, 155–159.
47
48
49 (17) Maroni, V. A. Vibrational frequencies and force constants for tetrahedral MgX−2
4 (X=Cl, Br,
50
51 and I) in MgX2 –KX melts. J. Chem. Phys. 1971, 55, 4789–4792.
52
53
54
(18) Capwell, R. Raman spectra of crystalline and molten MgCl2 . Chem. Phys. Lett. 1972, 12,
55
56
443–446.
57
58
59 19
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 20 of 28

1
2
3
(19) Brooker, M. H. A Raman spectroscopic study of the structural aspects of K2 MgCl4 and
4
5
Cs2 MgCl4 as solid single crystals and molten salts. J. Chem. Phys. 1975, 63, 3054.
6
7
8 (20) Huang, C.-H.; Brooker, M. Raman spectrum of molten MgCl2 . Chem. Phys. Lett. 1976, 43,
9
10 180 – 182.
11
12
13 (21) Brooker, M. H.; Huang, C.-H. Raman spectroscopic studies of structural properties of solid
14
15 and molten states of the magnesium chloride – alkali metal chloride system. Can. J. Chem.
16
17 1980, 58, 168–179.
18
19
20 (22) Bunten, R. A. J.; Mcgreevy, R. L.; Mitchell, E. W. J.; Raptist, C.; Walker, P. J. Collective
21
22 modes in molten alkaline-earth chlorides: I. Light scattering. J. Phys. C: Solid State Phys.
23
24 1984, 17, 4705–4724.
25
26
27 (23) Biggin, S.; Gay, M.; Enderby, J. E. The structures of molten magnesium and manganese (II)
28
29 chlorides. J. Phys. C: Solid State Phys. 1984, 17, 977–985.
30
31
(24) Newport, R. J.; Howe, R. A.; Wood, N. D. The structure of molten nickel chloride. J. Phys.
32
33
C: Solid State Phys. 1985, 18, 5249–5257.
34
35
36 (25) McGreevy, R. L.; Mitchell, E. W. J. Collective modes in molten alkaline earth chlorides:
37
38 III. Inelastic neutron scattering from molten MgCl2 and CaCl2 . J. Phys. C: Solid State Phys.
39
40 1985, 18, 1163–1178.
41
42
43 (26) Day, S. E.; McGreevy, R. L. Structure factors of molten CaCl2 and MgCl2 at low Q. Phys.
44
45 Chem. Liq. 1985, 15, 129–136.
46
47
48 (27) Gardner, P. J.; Heyes, D. M. Molecular dynamics computer simulations of molten zinc chlo-
49
50 ride. Physica B+C 1985, 131, 227–233.
51
52
53 (28) Ferrante, A.; Li, W.; Tosi, M. P. Interatomic force modelling of local coordination and ex-
54
55 tended structure in ionic-covalent liquids. Philos. Mag. A 1988, 58, 13–25.
56
57
58
59 20
60 ACS Paragon Plus Environment
Page 21 of 28 The Journal of Physical Chemistry

1
2
3
(29) Wilson, M.; Madden, P. A. Short- and intermediate-range order in MCl2 melts: the impor-
4
5
tance of anionic polarization. J. Phys.: Condens. Matter 1993, 5, 6833–6844.
6
7
8 (30) Wilson, M.; Madden, P. A. Anion polarization and the stability of layered structures in MX2
9
10 systems. J. Phys.: Condens. Matter 1994, 6, 159–170.
11
12
13 (31) Dai, S.; Begun, G. M.; Young, J. P.; Mamantov, G. Application of chemometric methods
14
15 in raman spectroscopic studies of molten salt system containing MgCl2 -KCl: Experimental
16
17 evidence for existence of Mg2 Cl3−
7 dimer and its Raman spectrum. J. Raman Spectrosc. 1995,
18
19 26, 929–932.
20
21
22 (32) Wilson, M. Structure and dynamics in network-forming materials. J. Phys.: Condens. Matter
23
24 2016, 28, 503001.
25
26
27 (33) McGreevy, R. L.; Pusztai, L. The structure of molten salts. Proc. R. Soc. Lond. A 1990, 430,
28
29 241–261.
30
31
(34) Tosi, M. Ordering in metal halide melts. Annu. Rev. Phys. Chem. 1993, 44, 173–211.
32
33
34 (35) Wu, F.; Roy, S.; Ivanov, A. S.; Gill, S. K.; Topsakal, M.; Dooryhee, E.; Abeykoon, M.;
35
36 Kwon, G.; Gallington, L. C.; Halstenberg, P. et al. Elucidating ionic correlations beyond
37
38 simple charge alternation in molten MgCl2 –KCl mixtures. J. Phys. Chem. Lett. 2019, 10,
39
40 7603–7610.
41
42
43 (36) Busse, L. E.; Nagel, S. R. Temperature dependence of the structure factor of As2 Se3 glass up
44
45 to the glass transition. Phys. Rev. Lett. 1981, 47, 1848–1851.
46
47
48 (37) Biggin, S.; Enderby, J. E. The structure of molten zinc chloride. J. Phys. C: Solid State Phys.
49
50 1981, 14, 3129–3136.
51
52
53 (38) Biggin, S.; Enderby, J. E. Comments on the structure of molten salts. J. Phys. C: Solid State
54
55 Phys. 1982, 15, L305–L309.
56
57
58
59 21
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 22 of 28

1
2
3
(39) Busse, L. E. Temperature dependence of the structures of As2 Se3 and Asx S1−x glasses near
4
5
the glass transition. Phys. Rev. B 1984, 29, 3639–3651.
6
7
8 (40) Elliott, S. R. Medium-range order in amorphous materials: documented cases. J. Non-Cryst.
9
10 Solids 1987, 97-98, 159–162.
11
12
13 (41) Wood, N. D.; Howe, R. A. The structure of the molten nickel halides. J. Phys. C: Solid State
14
15 Phys. 1988, 21, 3177–3190.
16
17
18 (42) Price, D. L.; Moss, S. C.; Reijers, R.; Saboungi, M.-L.; Susman, S. Intermediate-range order
19
20 in glasses and liquids. J. Phys. C: Solid State Phys. 1988, 21, L1069–L1072.
21
22
23 (43) Vashishta, P.; Kalia, R. K.; Antonio, G. A.; Ebbsjö, I. Atomic correlations and intermediate-
24
25 range order in molten and amorphous GeSe2 . Phys. Rev. Lett. 1989, 62, 1651–1654.
26
27
28
(44) Vashishta, P.; Kalia, R. K.; Ebbsjö, I. Structural correlations and phonon density of states in
29
GeSe2 : A molecular-dynamics study of molten and amorphous states. Phys. Rev. B 1989, 39,
30
31
6034–6047.
32
33
34 (45) Elliott, S. R. Origin of the first sharp diffraction peak in the structure factor of covalent
35
36 glasses. Phys. Rev. Lett. 1991, 67, 711–714.
37
38
39 (46) Allen, D. A.; Howe, R. A.; Wood, N. D.; Howells, W. S. The structure of molten zinc chloride
40
41 and potassium chloride mixtures. J. Phys.: Condens. Matter 1992, 4, 1407–1418.
42
43
44 (47) Salmon, P. S. The structure of molten and glassy 2:1 binary systems: An approach using the
45
46 Bhatia-Thornton formalism. Proc. R. Soc. Lond. A 1992, 437, 591–606.
47
48
49 (48) Badyal, Y. S.; Allen, D. A.; Howe, R. A. Structural modification in molten metal chloride and
50
51 alkali chloride mixtures. J. Phys.: Condens. Matter 1993, 5, 7189.
52
53
54
(49) Wilson, M.; Madden, P. A. Prepeaks and first sharp diffraction peaks in computer simulations
55
56
of strong and fragile ionic liquids. Phys. Rev. Lett. 1994, 72, 3033–3036.
57
58
59 22
60 ACS Paragon Plus Environment
Page 23 of 28 The Journal of Physical Chemistry

1
2
3
(50) Elliott, S. R. Extended-range order, interstitial voids and the first sharp diffraction peak of
4
5
network glasses. J. Non-Cryst. Solids 1995, 182, 40–48.
6
7
8 (51) Swenson, J.; Börjesson, L.; Howells, W. S. Structure of borate glasses from neutron-
9
10 diffraction experiments. Phys. Rev. B 1995, 52, 9310–9319.
11
12
13 (52) Taraskin, S. N.; Elliott, S. R.; Klinger, M. I. Void structure in models of vitreous silica. J.
14
15 Non-Cryst. Solids 1995, 192-193, 263–266.
16
17
18 (53) Wilson, M.; Madden, P. A. Voids, layers, and the first sharp diffraction peak in ZnCl2 . Phys.
19
20 Rev. Lett. 1998, 80, 532–535.
21
22
23 (54) Pusztai, L.; McGreevy, R. L. The structure of molten ZnCl2 and MgCl2 . J. Phys.: Condens.
24
25 Matter 2001, 13, 7213–7222.
26
27
28
(55) Salmon, P. S.; Martin, R. A.; Mason, P. E.; Cuello, G. J. Topological versus chemical ordering
29
in network glasses at intermediate and extended length scales. Nature 2005, 435, 75–78.
30
31
32 (56) Crupi, C.; Carini, G.; González, M.; D’angelo, G. Origin of the first sharp diffraction peak in
33
34 glasses. Phys. Rev. B 2015, 92, 134206.
35
36
37 (57) Ellis, R. B. Raman spectra of molten mixtures of zinc halides with alkali metal halides. J.
38
39 Electrochem. Soc. 1966, 113, 485–490.
40
41
42 (58) Clare, A. G.; Wright, A. C.; Sinclair, R. N.; Galeener, F. L.; Geissberger, A. E. A neutron
43
44 diffraction investigation of the structure of vitreous As2 O3 . J. Non-Cryst. Solids 1989, 111,
45
46 123 – 138.
47
48
49 (59) Marians, C. S.; Hobbs, L. W. Network properties of crystalline polymorphs of silica. J. Non-
50
51 Cryst. Solids 1990, 124, 242–253.
52
53
54
(60) Fischer-Colbrie, A.; Fuoss, P. H. X-ray scattering studies of intermediate-range order in amor-
55
56
phous GeSe2 . J. Non-Cryst. Solids 1990, 126, 1–34.
57
58
59 23
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 24 of 28

1
2
3
(61) Mei, Q.; Benmore, C. J.; Sen, S.; Sharma, R.; Yarger, J. L. Intermediate range order in
4
5
vitreous silica from a partial structure factor analysis. Phys. Rev. B 2008, 78, 144204.
6
7
8 (62) Salanne, M.; Madden, P. A. Polarization effects in ionic solids and melts. Mol. Phys. 2011,
9
10 109, 2299–2315.
11
12
13 (63) Balasubrahmanyam, K. Raman spectra of liquid MgCl2 and liquid MgCl2 –KCl system. J.
14
15 Chem. Phys. 1966, 44, 3270–3273.
16
17
18 (64) Ohtori, N.; Salanne, M.; Madden, P. A. Calculations of the thermal conductivities of ionic
19
20 materials by simulation with polarizable interaction potentials. J. Chem. Phys. 2009, 130,
21
22 104507.
23
24
25 (65) Ishii, Y.; Kasai, S.; Salanne, M.; Ohtori, N. Transport coefficients and the Stokes–Einstein
26
27 relation in molten alkali halides with polarisable ion model. Mol. Phys. 2015, 113, 2442–
28
29 2450.
30
31
(66) Chupas, P. J.; Chapman, K. W.; Kurtz, C.; Hanson, J. C.; Lee, P. L.; Grey, C. P. A versatile
32
33
sample-environment cell for non-ambient x-ray scattering experiments. J. Appl. Crystallogr.
34
35
2008, 41, 822–824.
36
37
38 (67) Kurley, J. M.; Halstenberg, P. W.; McAlister, A.; Raiman, S.; Dai, S.; Mayes, R. T. En-
39
40 abling chloride salts for thermal energy storage: implications of salt purity. RSC Adv. 2019,
41
42 9, 25602–25608.
43
44
45 (68) Toby, B. H.; Dreele, R. B. V. GSAS-II: the genesis of a modern open-source all purpose
46
47 crystallography software package. J. Appl. Crystallogr. 2013, 46, 544–549.
48
49
50 (69) Qiu, X.; Thompson, J. W.; Billinge, S. J. L. PDFgetX2: a GUI-driven program to obtain the
51
52 pair distribution function from x-ray powder diffraction data. J. Appl. Crystallogr. 2004, 37,
53
54 678.
55
56
57
58
59 24
60 ACS Paragon Plus Environment
Page 25 of 28 The Journal of Physical Chemistry

1
2
3
(70) Bernardes, C. E. S. AGGREGATES: Finding structures in simulation results of solutions. J.
4
5
Comput. Chem. 2017, 38, 753–765.
6
7
8 (71) Shimizu, K.; Bernardes, C. E. S.; Canongia Lopes, J. N. Structure and aggregation in the
9
10 1-alkyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ionic liquid homologous se-
11
12 ries. J. Phys. Chem. B 2014, 118, 567–576.
13
14
15 (72) Bernardes, C. E. S.; Shimizu, K.; Lobo Ferreira, A. I. M. C.; Santos, L. M. N.
16
17 B. F.; Canongia Lopes, J. N. Structure and aggregation in the 1,3-dialkyl-imidazolium
18
19 bis(trifluoromethylsulfonyl)imide ionic liquid family: 2. From single to double long alkyl
20
21 side chains. J. Phys. Chem. B 2014, 118, 6885–6895.
22
23
24 (73) Hutter, J.; Iannuzzi, M.; Schiffmann, F.; VandeVondele, J. CP2K: atomistic simulations of
25
26 condensed matter systems. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2013, 4, 15–25.
27
28
29 (74) Nosé, S. A molecular dynamics method for simulations in the canonical ensemble. Mol. Phys.
30
31 1984, 52, 255–268.
32
33
(75) Nosé, S. A unified formulation of the constant temperature molecular dynamics methods. J.
34
35
Chem. Phys. 1984, 81, 511–519.
36
37
38 (76) Annapureddy, H. V. R.; Kashyap, H. K.; Biase, P. M. D.; Margulis, C. J. What is the origin
39
40 of the prepeak in the x-ray scattering of imidazolium-based room-temperature ionic liquids?
41
42 J. Phys. Chem. B 2010, 114, 16838–16846.
43
44
45 (77) Kashyap, H. K.; Hettige, J. J.; Annapureddy, H. V. R.; Margulis, C. J. SAXS anti-peaks reveal
46
47 the length-scales of dual positive–negative and polar–apolar ordering in room-temperature
48
49 ionic liquids. Chem. Comm. 2012, 48, 5103–5105.
50
51
52 (78) Kashyap, H. K.; Margulis, C. J. (Keynote) Theoretical deconstruction of the x-ray structure
53
54 function exposes polarity alternations in room temperature ionic liquids. ECS Trans. 2013,
55
56 50, 301–307.
57
58
59 25
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 26 of 28

1
2
3
(79) Kashyap, H. K.; Santos, C. S.; Daly, R. P.; Hettige, J. J.; Murthy, N. S.; Shirota, H.; Castner
4
5
Jr., E. W.; Margulis, C. J. How does the ionic liquid organizational landscape change when
6
7
nonpolar cationic alkyl groups are replaced by polar isoelectronic diethers? J. Phys. Chem. B
8
9 2013, 117, 1130–1135.
10
11
12 (80) Kashyap, H. K.; Santos, C. S.; Murthy, N. S.; Hettige, J. J.; Kerr, K.; Ramati, S.;
13
14 Gwon, J.; Gohdo, M.; Lall-Ramnarine, S. I.; Wishart, J. F. et al. Structure of 1-alkyl-1-
15
16 methylpyrrolidinium bis(trifluoromethylsulfonyl)amide ionic liquids with linear, branched,
17
18 and cyclic alkyl groups. J. Phys. Chem. B 2013, 117, 15328–15337.
19
20
21 (81) Santos, C. S.; Annapureddy, H. V. R.; Murthy, N. S.; Kashyap, H. K.; Castner
22
23 Jr., E. W.; Margulis, C. J. Temperature-dependent structure of methyltributylammonium
24
25 bis(trifluoromethylsulfonyl)amide: X ray scattering and simulations. J. Chem. Phys. 2011,
26
27 134, 064501.
28
29
30 (82) Susman, S.; Volin, K. J.; Montague, D. G.; Price, D. L. Temperature dependence of the first
31
32 sharp diffraction peak in vitreous silica. Phys. Rev. B 1991, 43, 11076–11081.
33
34
35 (83) Hettige, J. J.; Araque, J. C.; Margulis, C. J. Bicontinuity and multiple length scale ordering
36
37 in triphilic hydrogen-bonding ionic liquids. J. Phys. Chem. B 2014, 118, 12706–12716.
38
39
(84) Araque, J. C.; Hettige, J. J.; Margulis, C. J. Modern room temperature ionic liquids, a simple
40
41
guide to understanding their structure and how it may relate to dynamics. J. Phys. Chem. B
42
43
2015, 119, 12727–12740.
44
45
46 (85) Kashyap, H. K.; Santos, C. S.; Annapureddy, H. V. R.; Murthy, N. S.; Margulis, C. J.; Castner
47
48 Jr., E. W. Temperature-dependent structure of ionic liquids: X-ray scattering and simulations.
49
50 Faraday Discuss. 2012, 154, 133–143.
51
52
53 (86) Hettige, J. J.; Kashyap, H. K.; Annapureddy, H. V. R.; Margulis, C. J. Anions, the reporters
54
55 of structure in ionic liquids. J. Phys. Chem. Lett. 2013, 4, 105–110.
56
57
58
59 26
60 ACS Paragon Plus Environment
Page 27 of 28 The Journal of Physical Chemistry

1
2
3
(87) Dhungana, K. B.; Faria, L. F. O.; Wu, B.; Liang, M.; Ribeiro, M. C. C.; Margulis, C. J.;
4
5
Castner Jr., E. W. Structure of cyano-anion ionic liquids: X-ray scattering and simulations. J.
6
7
Chem. Phys. 2016, 145, 024503.
8
9
10 (88) Amith, W. D.; Hettige, J. J.; Castner Jr., E. W.; Margulis, C. J. Structures of ionic liquids
11
12 having both anionic and cationic octyl tails: Lamellar vacuum interface vs sponge-like bulk
13
14 order. J. Phys. Chem. Lett. 2016, 7, 3785–3790.
15
16
17 (89) Wu, F.; Karunaratne, W. V.; Margulis, C. J. Ionic liquid mixture at the vacuum interface and
18
19 the peaks and antipeaks analysis of X-ray reflectivity. J. Phys. Chem. C 2019, 123, 4914–
20
21 4925.
22
23
24 (90) Partin, D.; OKeeffe, M. The structures and crystal chemistry of magnesium chloride and
25
26 cadmium chloride. J. Solid State Chem. 1991, 95, 176–183.
27
28
29 (91) Bernardes, C. E. S.; Minas da Piedade, M. E.; Canongia Lopes, J. N. The structure of
30
31 aqueous solutions of a hydrophilic ionic liquid: The full concentration range of 1-ethyl-3-
32
33 methylimidazolium ethylsulfate and water. J. Phys. Chem. B 2011, 115, 2067–2074.
34
35
(92) Revere, M.; Tosi, M. P. P. Structure and dynamics of molten salts. Rep. Prog. Phys 1986, 49,
36
37
1001–1081.
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 27
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 28 of 28

1
2
3
4
TOC Graphic
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 28
60 ACS Paragon Plus Environment

You might also like