Ampdf

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Journal of The Electrochemical

Society

Solution Synthesis of Sb2S3 and Na3SbS4 Solid-State Electrolyte


To cite this article: Saeed Ahmadi Vaselabadi et al 2021 J. Electrochem. Soc. 168 110533

Manuscript version: Accepted Manuscript


Accepted Manuscript is “the version of the article accepted for publication including all changes made as a result of the peer review process,
and which may also include the addition to the article by IOP Publishing of a header, an article ID, a cover sheet and/or an ‘Accepted
Manuscript’ watermark, but excluding any other editing, typesetting or other changes made by IOP Publishing and/or its licensors”

This Accepted Manuscript is© .

During the embargo period (the 12 month period from the publication of the Version of Record of this article), the Accepted Manuscript is fully
protected by copyright and cannot be reused or reposted elsewhere.

As the Version of Record of this article is going to be / has been published on a subscription basis, this Accepted Manuscript will be available for
reuse under a CC BY-NC-ND 3.0 licence after the 12 month embargo period.

After the embargo period, everyone is permitted to use copy and redistribute this article for non-commercial purposes only, provided that they
adhere to all the terms of the licence https://creativecommons.org/licences/by-nc-nd/3.0

Although reasonable endeavours have been taken to obtain all necessary permissions from third parties to include their copyrighted content
within this article, their full citation and copyright line may not be present in this Accepted Manuscript version. Before using any content from this
article, please refer to the Version of Record on IOPscience once published for full citation and copyright details, as permissions may be required.
All third party content is fully copyright protected, unless specifically stated otherwise in the figure caption in the Version of Record.

View the article online for updates and enhancements.

This content was downloaded from IP address 152.59.167.245 on 04/03/2024 at 07:37


Journal of The Electrochemical Society

Solution Synthesis of Sb2S3 and Na3SbS4 Solid-State


Electrolyte

Journal: Journal of The Electrochemical Society

Manuscript ID JES-105370.R1

Manuscript Type: Research Paper

Date Submitted by the


20-Oct-2021
Author:

Complete List of Authors: Vaselabadi, Saeed; Colorado School of Mines, Chemical & Biological
Engineering
Fo
Smith, William; Colorado School of Mines, Chemical & Biological
Engineering
Wolden, Colin; Colorado School of Mines, Chemical & Biological
Engineering
rR

Keywords: Batteries, Energy Storage, Solid-State Ionics, Solution synthesis


ev
iew
On
ly

https://mc04.manuscriptcentral.com/jes-ecs
Page 1 of 15 Journal of The Electrochemical Society

1
2
3
Solution Synthesis of Sb2S3 and Na3SbS4 Solid-State Electrolyte
4
5 Saeed Ahmadi Vaselabadi, William H. Smith,* and Colin A. Wolden**,z
6 Department of Chemical and Biological Engineering, Colorado School of Mines, Golden, Colorado,
7
8 USA
9
10 *Electrochemical Society Student Member.
11
**Electrochemical Society Member.
12
13
14
15 z
E-mail: cwolden@mines.edu
16
17
18 Abstract
19 Sodium thioantimonate (Na3SbS4) is an attractive solid-state electrolyte for sodium-ion
20
21 batteries due to its high ionic conductivity and stability in protic solvents. Herein, we describe
Fo

22 solution-based routes for its synthesis. First, we demonstrate the synthesis of the Sb2S3
23 precursor via thermodynamically favorable metathesis between Na2S and SbCl3. This solution-
24
based approach is further extended to couple the resulting Sb2S3 with Na2S for the synthesis of
rR

25
26 Na3SbS4. It is shown that ethanol is a superior solvent to water for solution-based synthesis of
27 Na3SbS4 with respect to yield, morphology, and performance. Amorphous Sb2S3 synthesized
28
ev

from low-temperature metathesis produced highly crystalline Na3SbS4 with a room


29
30 temperature Na+ conductivity of 0.52 mS/cm and low activation energy, comparable to leading
31 values reported in the literature.
iew

32
33
34 Introduction
35 Sodium-ion batteries are attractive for low-cost energy storage necessary to manage
36 fluctuations in renewable electricity due to sodium’s abundance, low cost, and reduced
On

37
Environ. impact relative to their high-performance lithium counterparts.1–4 Replacement of the
38
39 organic liquid electrolyte with a thin solid-state electrolyte (SSE) can further enhance
40 performance and specific capacity. Moreover, the use of SSEs improves reliability by
ly

41 preventing safety concerns such as flammability and leakage. Among various classes of SSEs,
42
43 sulfides have received significant attention in the research community due to their high ionic
44 conductivity (> 0.1 mS cm-1) as well as their ability to efficiently contact electrodes based on
45 their deformability.5–7 Synthesis of cubic Na3PS4 via ball-milling with a conductivity of 0.2 mS
46
47 cm-1 was the starting point in the development of highly conductive chalcogenide-based
48 sodium conductors.8 Since then, extensive research has led to development of new SSEs in
49 NPS class such as Na3PSe4 (1.16 mS cm-1)9, Na2.9375PS3.9375Cl0.0625 (1 mS cm-1)10,
50
Na3P0.62As0.38S4 (1.46 mS cm-1)11, and Na2.730Ca0.135PS4 (~1 mS cm-1)12.
51
52 A drawback of thiophosphates’ is their reactivity that requires extra handling
53
54 considerations such as the need of air-free environment that manifest by increasing the
55 production costs. The phosphorus cation typically reacts with oxygen and water vapor in
56 ambient air and release toxic H2S gas. Hard and soft acid and base (HSAB) theory suggests
57
that substitution of P5+ with a softer acid such as Sb5+ would minimize this undesired
58
59 reaction.13,14 As a result, antimony-based electrolytes such as sodium thioantimonate (Na3SbS4)
60 have been proposed as an air-stable solid electrolyte with achievable conductivity of more than
1

https://mc04.manuscriptcentral.com/jes-ecs
Journal of The Electrochemical Society Page 2 of 15

1
2
3
1 mS cm-1. The tetrathioantimonate (SbS44-) group is stable in presence of moisture because of
4
5 the strong bonding between Sb5+ and S2-. As such, Na3SbS4 can withstand water and other
6 protic solvents while maintaining high ionic conductivity.13–15
7
8 Sulfide-based SSEs are often synthesized through high-temperature reactions between
9 solid-state precursors conducted in quartz ampoules followed by ball-milling, which is energy
10
and time intensive. This has prompted the development of solution-based synthesis approaches
11
12 that offer low thermal budgets and are more amenable for scaleup for the synthesis of both
13 binary precursors and compound SSEs.16–18 Yoon et al.13 reported the dissolution-precipitation
14 synthesis of Na3SbS4 through producing the SSE via solid-state reaction at high temperatures
15
16 as the first step. Next, Na3SbS4 was dissolved in solvents such as water and methanol and
17 recrystallized while retaining a high ionic conductivity (0.1-0.3 mS cm-1) proving its scalable
18 solution processability. Furthermore, the same group demonstrated the aqueous synthesis of
19
20
Na3SbS4 from binary metal sulfide precursors. SSE recovery was performed through an extra
21 step of acetone addition as an anti-solvent (0.21 mS cm-1) or ethanol wash of the vacuum-dried
Fo

22 solution (0.15 mS cm-1) followed by mild thermal annealing. The formation of toxic H2S during
23
the recovery process was examined as a drawback of this approach.15 Tatsumisago et al.19 also
24
rR

25 followed a similar aqueous route while 2-propanol was used as the anti-solvent. High
26 conductivity of 1.2 mS cm-1 was achieved through a series of drying, pelletizing, and further
27 annealing processes at a relatively high fabrication pressure (720 MPa).
28
ev

29 In the case of Na3SbS4, the key binary precursors are sodium sulfide (Na2S) and antimony
30
trisulfide (Sb2S3). Our group recently described protocols for upgrading low-cost, technical
31
iew

32 grade sodium hydrate flakes into anhydrous Na2S powder of exceptional purity.17 Sb2S3 is a
33 precursor to many thioantimonates SSEs based on sodium and lithium cations.20,21 It has also
34 been employed directly as an anode due to its high theoretical capacity (946 mAhg-1).22–24
35
36 Moreover, antimony trisulfide is a material of interest for solar energy conversion including
On

37 thin film solar cells and photoelectrochemical water splitting.25–27


38
39 Solution chemistry approaches reported to date for Sb2S3 synthesis include chemical bath
40 deposition (CBD)28,29 and solvothermal30–34 methods. CBD is typically employed to form thin
ly

41 films of metal sulfides used in photovoltaic applications. In a typical CBD procedure antimony
42
43 trichloride (SbCl3) and sodium thiosulfate (Na2S2O3) precursors are combined in a solution of
44 acetone and water, and CBD needs to be performed at low temperatures (<10 °C) to avoid the
45 immediate precipitation of products.29 Solvothermal methods also combine antimony salt (i.e.,
46
47
SbCl3) with a sulfur source such as Na2S.9H2O30, Na2S34, Na2S2O3·5H2O31, and L-cysteine33,
48 elemental sulfur32 reacted in different solvents. However, this method requires the use of an
49 autoclave at high pressure and temperature for a prolonged time to complete the formation of
50
crystalline Sb2S3. In addition, a co-solvent like acetone28 and ethylene glycol34 is normally
51
52 needed to buffer the undesired hydrolyses.
53
54
In this work, we demonstrate the formation of phase pure Sb2S3 via a simple metathesis
55 reaction between antimony trichloride and sodium sulfide. Subsequently, the synthesized Sb2S3
56 is converted to Na3SbS4 via liquid-state reaction to validate the applicability that produces ionic
57
conductivities (σ = 0.52 mS cm-1) comparable with reported literature (Table 1).15,19 Previous
58
59 solution-based synthesis of Na3SbS4 employed aqueous solution followed by an ethanol
60 wash15. It is shown that use of ethanol simplifies the process, increases the yield, improves
2

https://mc04.manuscriptcentral.com/jes-ecs
Page 3 of 15 Journal of The Electrochemical Society

1
2
3
morphology and electrolyte performance. The composition and morphology of Sb2S3 and
4
5 Na3SbS4 produced in this work are extensively characterized using a variety of techniques
6 (XRD, TGA/DSC, Raman spectroscopy, TOF-SIMS, SEM-EDX) and correlated with ion
7 conductivity measurements using impedance spectroscopy.
8
9 Experimental
10
Materials Sodium sulfide hydrate flakes (60%, Sigma Aldrich) were converted to anhydrous
11
12 Na2S and purified using the dehydration/H2 reduction procedure described in the literature.17
13 SbCl3 (99+%, Alfa Aesar) and elemental sulfur (S, Sigma Aldrich, 99.998% trace metal basis)
14 were used as received. Solvents employed included DI water, methanol (MeOH, Sigma-
15
16 Aldrich, anhydrous, 99.8%), and ethanol (Ethanol, Sigma Aldrich, anhydrous, ≥99.5%).
17 Ambients employed for various evaporation and annealing processes included vacuum
18 evacuated tube furnace, UHP grade argon, or a 10% H2S/Ar specialty gas mixture (Matheson).
19
20
Characterization. Thermogravimetric analysis and differential scanning calorimetry
21 (TGA/DSC) was performed on a TA Instruments SDT-Q600 model. For a typical run, 10 mg
Fo

22 of sample was loaded into a pre-cleaned alumina pan and heated under flowing Ar from RT to
23
600 °C at 10 °C min-1 rate and then cooled down naturally. X-ray diffraction (XRD) was
24
performed with a Philips X’Pert X-ray diffractometer with Cu Kα radiation (λ = 0.15405 nm)
rR

25
26 between 10 and 60° at a scan rate of 2-5° min−1. Samples were prepared on a glass slide with a
27 protective tape covering the material to prevent undesired reactions with ambient air. XRD
28
ev

29 background subtraction was performed using HighScore software. Confocal Raman


30 spectroscopy was performed with a WiTec Alpha300 laser confocal microscope equipped with
31 a Raman spectrometer. The samples were prepared by placing 0.1 mm glass covers on powder
iew

32
33
samples secured on glass slides. Spectragryph software was used to subtract the Raman
34 background.
35 Fourier-transform infrared spectroscopy (FTIR) was performed with a Nicolet Summit
36
FTIR spectrometer using an attenuated total reflection (ATR) accessory equipped with a
On

37
38 diamond crystal. Field emission scanning electron microscopy (FESEM) images were
39 collected on a JEOL JSM-7000F FESEM instrument equipped with energy-dispersive X-ray
40 spectroscopy (EDX) for compositional analysis. To prepare the samples for SEM and EDX
ly

41
42 measurements, powder samples were immobilized onto an aluminum stub by using a double-
43 sided carbon tape. An accelerating voltage of 5 kV was used for taking the SEM image while
44 a higher voltage of 12-25 V was employed for EDX spectra collection. Compositional analysis
45
46
was performed on Na3SbS4 pellets using time of flight secondary ion mass spectrometry (TOF-
47 SIMS, IONTOF.SIMS 5). The TOF-SIMS sputter parameters were 0.5 keV and 1.17 × 10 Cs+
48 cm-2 on a 400 * 400 μm2 scanning area for depth profiling. All sample preparation was done in
49
an Ar glovebox, though samples were briefly exposed to ambient upon transfer to various
50
51 characterization instrumentation.
52 Pellets of Na3SbS4 SSE were prepared for conductivity measurement via conventional
53 uniaxial pressing. For this purpose, 150-250 mg of the electrolyte was loaded into a PEEK split
54
55 cell with 12 mm diameter, and pellets 0.9-1.1 mm in thickness were formed by cold pressing
56 using a hydraulic press (YLJ-15L, MTI Corp.). The electrolyte powders were pressed gradually
57 by increasing pressure in 60 MPa increments (30-second stabilization interval) and finally were
58
59
held for 5 min at the target pressure of 240 MPa (Pfab). After reducing the pressure to zero,
60 different stack pressure (Pst) was applied to the pellet for 2 min. The SSE pellets were contacted
3

https://mc04.manuscriptcentral.com/jes-ecs
Journal of The Electrochemical Society Page 4 of 15

1
2
3
using stainless steel plungers as the ion-blocking electrodes for electrochemical
4
5 characterization. A Gamry Interface 1000E potentiostat was used to perform electrochemical
6 impedance spectroscopy (EIS) measurements across a frequency range of 1 Hz to 1 MHz.
7 Temperature-dependent measurements were performed by contacting the cell assembly with a
8
9 temperature-controlled heating coil connected to a PID controller. DC polarization
10 measurements were performed by applying three step potentials (0.5, 0.75, and 1 V) to the
11 sample and recording the transient current. The steady-state current was recorded after 2 hrs at
12
13
each step potential, and the electrical conductivity was calculated using Ohm’s law.
14
15 Results and Discussion
16
Metathesis of Sb2S3
17
18 Metathesis is a simple chemical process that involves counterion exchange between two
19 chemical species where one of the products often precipitates, providing favorable energetics.
20 The salts employed are typically low cost and widely available. The first reaction of interest is
21
Fo

22 the thermodynamically favorable metathesis reaction between Na2S and SbCl335,36:


23
3𝑁𝑎2 𝑆(𝑠𝑜𝑙) + 2𝑆𝑏𝐶𝑙3 (𝑠𝑜𝑙) → 𝑆𝑏2 𝑆3 (𝑠) + 6𝑁𝑎𝐶𝑙(𝑠𝑜𝑙) , 𝛥𝐺° = −1160.2 𝑘𝐽/𝑚𝑜𝑙 (1)
24
rR

25
26
One of the important aspects of metathesis is appropriately choosing a solvent that both
27 facilitates the reaction kinetics and acts as a separation medium for the resulting products. Also,
28
ev

the potential of side reactions between the solvent and precursors can play an important role in
29
the formation of impurities. For instance, some metal sulfides may get hydrolyzed with protic
30
31 solvents. While SbCl3 is moisture sensitive and needs to be handled in an air-free environment,
iew

32 its full dissolution in the solvent is the driving force for the metathesis reaction. As highly polar
33 H2O causes hydrolysis of the starting precursor, we chose methanol as the reaction solvent that
34
35 provides both high precursor solubility as well as acceptable separation access for both Sb 2S3
36 and NaCl products. NaCl is expected to remain fully dissolved in methanol while sparingly
On

37 insoluble Sb2S3 will precipitate.


38
39
Figure 1a demonstrates the procedure used for metathesis reaction of Sb2S3: First, SbCl3
40 was dissolved in MeOH forming a 0.4 M solution that was allowed to equilibrate at room
ly

41 temperature. A stoichiometric amount of a 0.6 M Na2S solution was added gradually. The
42
solution color turned brown instantaneously upon Na2S addition indicating fast reaction
43
44 kinetics, though the mixture was stirred for one day to ensure the reaction went to completion.
45 The dark brown solution was centrifuged and decanted. The precipitate was washed initially
46 with MeOH. The precipitate and supernatant were dried at room temperature under vacuum.
47
48 An extra step of grinding for 5 min in a pestle and mortar and H2O washing was added to ensure
49 the full separation of metathesis products. Figure 1b displays the XRD patterns of the Sb2S3
50 metathesis precipitate and solids recovered from the supernatant. The precipitate is amorphous
51
52
with no discernable crystalline features while the solids recovered from the supernatant match
53 the cubic phase of NaCl. The extra grinding and H2O washing were required as XRD patterns
54 of the precipitate after just the MeOH wash contained residual NaCl (Fig. S1a). The precipitate
55
composition was analyzed using EDX, which supported the formation of Sb2S3. EDX analysis
56
57 found a S/Sb ratio of ~1.3-1.4 that is very close to the ideal value of 1.5 considering the
58 uncertainties in this technique (Fig. S1b).
59
60 Figure 1c displays TGA/DSC scans of amorphous Sb2S3 to identify possible crystallization
4

https://mc04.manuscriptcentral.com/jes-ecs
Page 5 of 15 Journal of The Electrochemical Society

1
2
3
points. The heat trace shows a weak exothermic peak at ~233 °C coincided with ~1.5% weight
4
5 loss. The slight weight reduction might correspond to the volatilization of compounds with
6 excess sulfur such as polysulfides. The identified crystallization temperature is in good
7 agreement with the reported temperature for the thermal activation of Sb2S3 thin films.28 The
8
9 TGA trace indicates that Sb2S3 is thermally stable before reaching its melting point at ~550 °C.
37–39
10
11
12 Annealing in the presence of elemental sulfur40 and H2S41 have been shown to improve the
13 crystallinity of metal sulfides, and the use of H2S especially provides a well-controlled and
14 reproducible annealing process by avoiding the difficulties of handling sulfur. The Sb2S3
15
16 powder was annealed at 200, 250, and 300 °C using a packed bed setup under flowing of 10%
17 H2S/Ar gas mixture (40 sccm) in a tube furnace at reduced pressure (~70 torr). The XRD
18 patterns of the thermally annealed Sb2S3 (Fig. 2a) samples demonstrate significant crystalline
19
20
phase with diffraction peaks matching those for an orthorhombic structure with Pnma (62)
21 space group (PDF 01-070-9254).29 As manifested by more intense Bragg peaks, increasing the
Fo

22 temperature improves the degree of crystallinity. Annealing under Ar flow at 250 °C also
23
showed a similar degree of crystallinity in comparison to H2S-annealed Sb2S3 (Fig. S2). Figure
24
rR

25 2b shows the Raman spectra of the annealed Sb2S3 that further confirms the presence of Sb2S3.
26 As described in detail by Parize and coworkers.28, the onset of crystallization at 200 °C has
27 coincided with the formation of weak broad peaks centered at 280 and 310 cm-1 while further
28
ev

29 increasing the temperature solidifies the crystallization process and gives rise to more
30 characteristic peaks corresponding to Sb2S3. A minor formation of residual senarmontite Sb2O3
31 was observed in Raman spectra as the annealing temperature increased. The characteristic
iew

32
33
peaks corresponding to Sb-S bonds are observed at ~190, ~240, ~285, and ~310 cm-1.27,28
34 To investigate the impact of Na2S reagent purity on Sb2S3 products, Sb2S3 was synthesized
35 following the same procedure using technical grade Na2S (Na2S.xH2O, x~3) that contains 40%
36
water. Figure S3 displays the FTIR spectra of resulted Sb2S3 from purified and hydrated Na2S
On

37
38 reagent dried at room temperature without further processing. Sb2S3 derived from purified Na2S
39 exhibits a lower level of impurity. The main peaks include symmetric stretching band42 of SO32-
40 and residual water remaining (O-H bending and stretching modes at 1570 and 3200 cm-1,
ly

41
42 respectively).43 Further drying can desolvate the residual water to improve the Sb2S3 impurity
43 profile. On the other hand, Sb2S3 derived from hydrated precursors mainly carries on the
44 insoluble impurities already present in the sodium sulfide such as oxysulfurs (Na2SOx) and
45
46
residual carbon bindings.17 Detailed descriptions of the observed impurity peaks are
47 summarized in Table S1. These results confirm that the presence of impurities in the Na2S
48 reagent can negatively affect the functionality of the sulfide product.
49
50
51 Solution synthesis of Na3SbS4
52 Synthesized Sb2S3 was then used as a reagent with Na2S to form Na3SbS4 SSEs using both
53 water and ethanol solvents. First, Na3SbS4 was synthesized by mixing Sb2S3, Na2S, and
54
55 elemental sulfur in an aqueous solution following the procedure described by Yoon and
56 coworkers.15 After mixing water was evaporated at 140 °C and the recovered powder was
57 washed with ethanol multiple times to remove any unreacted precursors and further vacuum
58
59
dried at room temperature to produce a light green powder (Fig. S4). The following mechanism
60 was suggested for the formation of Na3SbS4 in aqueous solutions:
5

https://mc04.manuscriptcentral.com/jes-ecs
Journal of The Electrochemical Society Page 6 of 15

1
2
3
+ − −
4 3𝑁𝑎2 𝑆(𝑠) + 3𝐻2 𝑂(𝑙) → 6𝑁𝑎(𝑎𝑞) + 3𝐻𝑆(𝑎𝑞) + 3𝑂𝐻(𝑎𝑞) (2)
5
6 − −
7
𝑆𝑏2 𝑆3 (𝑠) + 𝐻𝑆(𝑎𝑞) + 𝑂𝐻(𝑎𝑞) → 𝑆𝑏2 𝑆42−(𝑎𝑞) + 𝐻2 𝑂(𝑙) (3)
8
9 𝑆𝑏2 𝑆42−(𝑎𝑞) + 2𝐻𝑆(𝑎𝑞)
− −
+ 2𝑂𝐻(𝑎𝑞) → 2𝑆𝑏𝑆33−(𝑎𝑞) + 2𝐻2 𝑂(𝑙) (4)
10
11
12 2𝑆𝑏𝑆33−(𝑎𝑞) + 2𝑆(𝑠) → 2𝑆𝑏𝑆43−(𝑎𝑞) (5)
13
14
3𝑁𝑎2 𝑆(𝑠) + 𝑆𝑏2 𝑆3 (𝑠) + 2𝑆(𝑠) → 2𝑁𝑎3 𝑆𝑏𝑆4 (𝑠) (6)
15
16
The formation of the bisulfide anion (HS-) is the key to driving this reaction; however,
17
18 competition with the Na2S hydrolysis pathway below compromises yield and releases H2S44:
19
20
𝑁𝑎2 𝑆 (𝑠) + 𝐻2 𝑂(𝑙) → 𝑁𝑎𝐻𝑆(𝑎𝑞) + 𝑁𝑎𝑂𝐻(𝑎𝑞) (7)
21
Fo

22 𝑁𝑎𝐻𝑆(𝑎𝑞) + 𝐻2 𝑂(𝑙) → 𝐻2 𝑆 (𝑔) + 𝑁𝑎𝑂𝐻(𝑎𝑞) (8)


23
24 Therefore, H2S formation, as well as handling material losses, contribute to the low yield
rR

25
26
of Na3SbS4 synthesis in water. We hypothesized that ethanol would be a viable option for the
27 single-step synthesis of Na3SbS4 that could minimize the H2S formation due to its lower
28 polarity and the insolubility of Na3SbS4 in ethanol.15 Figure 3a displays the schematic of
ev

29
Na3SbS4 synthesis in ethanol. In this process, 0.075 M of Sb2S3, 0.15 M of elemental sulfur,
30
31 and 0.450 M (twice the stoichiometric concentration) of Na2S were dissolved in ethanol and
iew

32 stirred overnight to fully react. Extra Na2S was used to ensure the Sb2S3 consumption in the
33 reaction.15 To recover Na3SbS4, the solution was centrifuged and washed with ethanol several
34
35 times. A light brown powder was obtained after vacuum drying the precipitate. The powder
36 was further dried at 140 °C to remove any solvated complexes. The Na3SbS4 recovered from
On

37 ethanol showed a near stoichiometric yield of 91% resulting in a significant improvement in


38
39
recovery compared to the aqueous solution (~72%).
40 Figure 3b demonstrates the XRD patterns of the Na3SbS4 recovered from ethanol and water.
ly

41 Despite their color difference (Fig. 1a and S4), both ethanol and water products are highly
42
crystalline with a tetragonal structure. Typically, it has been observed that Na3SbS4 follows the
43
44 powder pattern in which the (110) peak is most prominent. In this work for both ethanol and
45 water, there is a degree of preferential orientation in the (211) and (220) directions. The
46 ethanol-recovered Na3SbS4 exhibits broadened peaks with higher full-width half-maximum
47
48 (FWHM) compared to sharp split peaks of the tetragonal phase from aqueous synthesis. The
49 differences in FWHM are consistent with the morphology revealed by SEM. The SEM images
50 of Na3SbS4 synthesized in the aqueous solution (Fig. 3c) features fairly large individual crystals
51
52
(~1 micron) that also assemble into large agglomerates. In contrast, ethanol-derived SSE (Fig.
53 3d) displays uniform nanocrystals on the order of ~100 nm. The difference in grain size is
54 consistent with the observation of Scherrer broadening in the XRD patterns of the ethanol-
55
derived Na3SbS4. Zhang et al.45 reported production of tetragonal Na3SbS4 with less than 30
56
57 nm grain size and higher conductivity achieved through a ball-milling step followed by solid-
58 state synthesis. On the other hand, Tatsumisago et al19 also showed that the solution approach
59 creates a smaller grain size of Na3SbS4 (< 1 μm) compared to the mechanochemical method;
60
6

https://mc04.manuscriptcentral.com/jes-ecs
Page 7 of 15 Journal of The Electrochemical Society

1
2
3
however, that could explain the lower ionic conductivity of the solution method due to the
4
5 presence of a higher number of grain boundaries.
6 The composition of the two materials is compared in Figure 4. Figure 4a shows the Raman
7 spectra of Na3SbS4 samples recovered from ethanol and water. The observation of symmetric
8
9 (νs=368 cm-1), and asymmetric (νa=389, 409 cm-1) stretching vibration peaks of SbS44- confirms
10 the formation of tetrathioantimonate anions in both samples. The attenuated peaks at 168 and
11 201 cm-1 are attributed to the deformational vibrations of SbS44-, δF and δE, respectively.46,47
12
13
Figure 4b displays the negative ion intensity profiles obtained from TOF-SIMS profiling of
14 Na3SbS4 derived from both solvents as a function of sputter time. The signals from ethanol-
15 and water-derived electrolytes are shown as blue and red, respectively. For the principal
16
elements (Na, Sb, S) there is no discernible difference based on solvent. Although present in a
17
18 3:1:4 molar ratio the dramatic difference in intensity reflects each element's ability to form
19 negative ions, which is why the intensity of electronegative S is more than 3 orders of
20 magnitude greater than electropositive Na. The only discernible difference between the two
21
Fo

22 materials is the level of the oxygen impurities, with ethanol-derived material containing about
23 twice as much oxygen as water-derived material. Without standards, absolute values cannot be
24 assessed but we note that oxygen is significantly more electronegative than sulfur, which is
rR

25
26
why its intensity is relatively high despite likely being present at levels less than 0.1%. The
27 only other element detected was trace amounts of Cl that were present at identical levels for
28
ev

both solvents (not shown), presumably introduced during Sb2S3 metathesis (Figure S1b). The
29
profiles suggest that the surface may be slightly enriched with O, but this may be an artifact of
30
31 sputter initiation or reflect contamination from sample transfer in air. The bulk composition
iew

32 appears homogeneous in both samples. Figure S5 shows the homogenous distribution of


33 constituent atoms from the 3D TOF-SIMS profiles recovered from water and ethanol solutions.
34
35 To understand the difference in oxygen content TGA scans were performed of Na3SbS4
36 recovered from ethanol and water (Fig. S6). The H2O-derived sample shows a very stable
On

37 weight trace with no significant weight loss; however, the ethanol-recovered SSE displays a
38
39
small weight loss at 62.5 °C that is attributed to the presence of ethanol solvated moieties. This
40 is consistent with the differences in residual oxygen observed in TOF-SIMS. Annealing at
ly

41 higher temperatures with better control over the mass transfer of the solvent evaporation, in a
42
fluidized regime, for instance, needs to be explored to fully eradicate the solvent impurities.
43
44 EIS measurements were performed to measure the ionic conductivity of Na3SbS4. Nyquist
45 plots of various stacking pressure and their fitted curves are shown in Figure 5a, b. The spectra
46 were fitted with an equivalent circuit consisting of a parallel resistor (R1) and constant phase
47
48 element (CPE1) arrangement that is in series with an additional CPE (CPE2).48,49 The use of
49 stainless steel blocking electrodes contributed significant contact resistance that was overcome
50 through the application of stacking pressure. Stacking pressure helps with obtaining nominal
51
52
conductivity values close to innate bulk values and ultimately decreases the chances of
53 delamination during battery cycling operation.50–52All samples have been constructed by
54 constant fabricated pressure of 240 MPa while stacking pressure is variant. Figure5c shows the
55
ionic conductivity of Na3SbS4 recovered from ethanol and water as a function of stacking
56
57 pressure. For both solvents, pressure-induced contact during EIS measurement improves the
58 conductivity before reaching a constant plateau at pressure more than 100 MPa. The
59 conductivity of ethanol-derived Na3SbS4 was consistently greater than its aqueous counterparts
60
7

https://mc04.manuscriptcentral.com/jes-ecs
Journal of The Electrochemical Society Page 8 of 15

1
2
3
by nearly a factor of 2. We speculate that the small improvement in conductivity of ethanol-
4
5 derived Na3SbS4 is related to the homogenous, nanocrystalline morphology revealed in Figure
6 3. Under compression we speculate that this morphology leads to lower porosity in the final
7 pellet than is achieved with the larger, aggregated particles formed from water. Such a change
8
9 would be expected to enhance deformability upon cold pressing and reduce grain boundary
10 resistance and improve conductivity.53–55
11 To investigate the impact of Sb2S3 crystallinity on subsequent Na3SbS4 structure and
12
13
functionality, Na3SbS4 was synthesized from both amorphous and crystalline Sb2S3 reagents in
14 ethanol. The differences in crystallinity and conductivity were very minor (Fig. S7). This
15 suggests that annealing of the Sb2S3 precursor is not required, and in fact amorphous Sb2S3 may
16
be preferred for Na3SbS4 synthesis.
17
18 The Na3SbS4 synthesized from H2O solvent achieved a room temperature ionic conductivity
19 of σw= 0.33 mS cm-1 while the sample recovered from ethanol demonstrated a higher
20 conductivity σe = 0.52 mS cm-1 (Pst = 120 MPa). The kinetics of Na+ transport was investigated
21
Fo

22 by performing temperature-dependent EIS. Figure 5a shows the Arrhenius plots of Na3SbS4


23 samples recovered from ethanol and water that fitted to the following equation56,57:
24 𝐸𝑎
rR

25 𝜎𝑇 = 𝜎0 e(−𝑅𝑇) (11)
26
27 The activation energy of Na+ conduction calculated accordingly is ~0.14 eV for both
28 synthetic routes, which is lower than previously reported values in the literature (Table 1).15,56
ev

29
The reasons for the lower activation energy are unclear. It may reflect the preferential
30
31 orientation of the crystals (Fig. 3b) or the trace amounts of Cl, which has been correlated with
iew

32 lower activation energy in similar systems.10 Figure S8 displays the Nyquist plots of Na3SbS4
33 recovered from ethanol and water solution at different temperatures. DC polarization
34
35 measurements (Fig. S9) showed an insignificant electronic conductivity of 1.57*10-7 mS cm-1
36 which is 6 orders of magnitude lower than the recorded ionic conductivity.
On

37 Table 1 summarizes the findings of this work compared with the various reports in the
38
39
literature. The conductivity achieved in this work using water is comparable or slightly higher
40 than most solution-based reports, and the material derived from ethanol has the highest
ly

41 conductivity reported to date with the exception of the work of Tatsumisago et al.19, who
42
employed extremely high pressure for pellet formation (720 MPa) as well as additional
43
44 annealing treatments. The high conductivity achieved in this work may reflect the use of
45 purified Na2S, whose purity has been shown to be superior to commercial Na2S17 and this was
46 reflected in the purity of the Sb2S3. The simplification of using a single solvent for synthesis
47
48 and washing, the higher yield, mitigation of H2S formation, improved morphology and higher
49 conductivity clearly demonstrate the benefits of ethanol over water, which has been used
50 exclusively for solution-based synthesis of Na3SbS4 to date.
51
52
53 Conclusions
54 This work reports a simple and scalable synthesis route to produce antimony trisulfide
55
applicable for energy storage applications such as sodium-ion battery anodes and as a precursor
56
57 to solid-state electrolytes. XRD, EDX, and Raman corroborated the formation of amorphous
58 Sb2S3 at room temperature, while H2S annealing at a mild temperature of 250 °C resulted in a
59 highly crystalline material. The utility of Sb2S3 was validated as a precursor reagent via solution
60
8

https://mc04.manuscriptcentral.com/jes-ecs
Page 9 of 15 Journal of The Electrochemical Society

1
2
3
synthesis of air-stable tetragonal Na3SbS4 with room temperature ionic conductivity of 0.52
4
5 mS cm-1. It was shown that ethanol was the preferred solvent over water for this process, as it
6 improves yield, mitigates hazardous byproducts, and delivers higher ionic conductivity. The
7 findings of this work further prove the advantages of the reagent chemistry tuning in liquid-
8
9 phase approaches that can be utilized for large scale fabrication of all-solid-state batteries.
10
11
12 Acknowledgments
13 This work was supported by the US National Sci. Foundation through Award 1825470 and an
14 Advanced Industries grant from the Colorado Office for Economic Development and
15
16 International Trade. We thank Dr. Michael Walker for conducting TOF-SIMS measurements
17 employing instrumentation supported by the NSF MRI award 1726898.
18
19
20
References
21 1. H. Pan, Y. S. Hu, and L. Chen, Energy Environ. Sci., 6, 2338–2360 (2013).
Fo

22 2. J. Y. Hwang, S. T. Myung, and Y. K. Sun, Chem. Soc. Rev., 46, 3529–3614 (2017).
23 3. K. Kubota and S. Komaba, J. Electrochem. Soc., 162, A2538–A2550 (2015).
24
rR

25 4. F. Wei, Q. Zhang, P. Zhang, W. Tian, K. Dai, L. Zhang, J. Mao, and G. Shao, J. Electrochem. Soc., 168,
26 050524 (2021).
27 5. Z. Zhang, Y. Shao, B. Lotsch, Y. S. Hu, H. Li, J. Janek, L. F. Nazar, C. W. Nan, J. Maier, M. Armand, and
28
ev

29 L. Chen, Energy Environ. Sci., 11, 1945–1976 (2018).


30 6. H. Jia, L. Peng, C. Yu, L. Dong, S. Cheng, and J. Xie, J. of Mater. Chem. A, 5134–5148 (2021).
31 7. M. Li, C. Wang, Z. Chen, K. Xu, and J. Lu, Chem. Rev. (2020).
iew

32
8. S. Yubuchi, A. Hayashi, and M. Tatsumisago, Chem. Lett., 44, 884–886 (2015).
33
34 9. L. Zhang, K. Yang, J. Mi, L. Lu, L. Zhao, L. Wang, Y. Li, and H. Zeng, Adv. Energy Mater., 5 (2015).
35 10. I. H. Chu, C. S. Kompella, H. Nguyen, Z. Zhu, S. Hy, Z. Deng, Y. S. Meng, and S. P. Ong, Sci. Rep., 6,
36 1–10 (2016).
On

37
38 11. Z. Yu, S. L. Shang, J. H. Seo, D. Wang, X. Luo, Q. Huang, S. Chen, J. Lu, X. Li, Z. K. Liu, and D. Wang,
39 Adv. Mater., 29, 0–6 (2017).
40 12. C. K. Moon, H. J. Lee, K. H. Park, H. Kwak, J. W. Heo, K. Choi, H. Yang, M. S. Kim, S. T. Hong, J. H.
ly

41
42 Lee, and Y. S. Jung, ACS Energy Lett., 3, 2504–2512 (2018).
43 13. A. Banerjee, K. H. Park, J. W. Heo, Y. J. Nam, C. K. Moon, S. M. Oh, S.-T. T. Hong, and Y. S. Jung,
44 Angewandte Chem. - Int. Ed., 55, 9634–9638 (2016).
45
14. H. Wang, Y. Chen, Z. D. Hood, G. Sahu, A. S. Pandian, J. K. Keum, K. An, and C. Liang, Angewandte
46
47 Chemie - Int. Ed., 55, 8551–8555 (2016).
48 15. T. W. Kim, K. H. Park, Y. E. Choi, J. Y. Lee, and Y. S. Jung, J. Mater. Chem. A, 6, 840–844 (2018).
49 16. Y. Zhao, Y. Yang, and C. A. Wolden, ACS Appl. Energy Mater., 2, 2246–2254 (2019).
50
51 17. W. H. Smith, J. Birnbaum, and C. A. Wolden, J. Sulfur Chem. (2021).
52 18. Y. Zhao, W. Smith, and C. A. Wolden, J. Electrochem. Soc., 167, 070520 (2020).
53 19. S. Yubuchi, A. Ito, N. Masuzawa, A. Sakuda, A. Hayashi, and M. Tatsumisago, J. Mater. Chem. A, 8,
54
55 1947–1954 (2020).
56 20. S. Yubuchi, A. Ito, N. Masuzawa, A. Sakuda, A. Hayashi, and M. Tatsumisago, J. Mater. Chem. A, 8,
57 1947–1954 (2020).
58
21. L. Zhang, D. Zhang, K. Yang, X. Yan, L. Wang, J. Mi, B. Xu, and Y. Li, Adv. Sci., 3, 1–6 (2016).
59
60 22. W. Zhao, M. Li, Y. Qi, Y. Tao, Z. Shi, Y. Liu, and J. Cheng, J. Colloid Interface Sci., 586, 404–411 (2021).
9

https://mc04.manuscriptcentral.com/jes-ecs
Journal of The Electrochemical Society Page 10 of 15

1
2
3 23. X. Zhou, Z. Zhang, P. Yan, Y. Jiang, H. Wang, and Y. Tang, Mater. Chem. Phys., 244, 122661 (2020).
4
5 24. C. Li, H. Song, C. Mao, H. Peng, and G. Li, J. of Alloys and Compounds, 786, 169–176 (2019).
6 25. H. S. Lin and L. Y. Lin, Electrochim. Acta, 252, 235–244 (2017).
7 26. H. Lei, G. Yang, Y. Guo, L. Xiong, P. Qin, X. Dai, X. Zheng, W. Ke, H. Tao, Z. Chen, B. Li, and G. Fang,
8
9 Phys. Chem. Chem. Phys., 18, 16436–16443 (2016).
10 27. A. D. DeAngelis, K. C. Kemp, N. Gaillard, and K. S. Kim, ACS Appl. Mater. Interfaces, 8, 8445–8451
11 (2016).
12
28. R. Parize, T. Cossuet, O. Chaix-Pluchery, H. Roussel, E. Appert, and V. Consonni, Mater. Des., 121, 1–
13
14 10 (2017).
15 29. Y. C. Choi, B. Roose, A. Sadhanala, and H. J. Snaith, ChemComm, 51, 8640–8643 (2015).
16 30. Y. Cheng, Z. Yao, Q. Zhang, J. Chen, W. Ye, S. Zhou, H. Liu, and M. S. Wang, Adv. Funct. Mater.,
17
18 2005417, 1–11 (2020).
19 31. J. Zhang, Z. Liu, and Z. Liu, ACS Appl. Mater. Interfaces, 8, 9684–9691 (2016).
20 32. H. Lei, T. Lin, X. Wang, S. Zhang, Q. Cheng, X. Chen, Z. Tan, and J. Chen, Mater. Lett., 233, 90–93
21
Fo

22 (2018).
23 33. P. Ge, H. Hou, X. Ji, Z. Huang, S. Li, and L. Huang, Mater. Chem. Phys., 203, 185–192 (2018).
24 34. G. Q. Zhu, P. Liu, H. Y. Miao, J. P. Zhu, X. B. Bian, Y. Liu, B. Chen, and X. B. Wang, Mater. Res. Bull.,
rR

25
43, 2636–2642 (2008).
26
27 35. G. Lindberg, A. Larsson, M. Råberg, D. Boström, R. Backman, and A. Nordin, J. Chem. Thermodyn.,
28
ev

39, 44–48 (2007).


29 36. G. A. Sehmel, 224 (1989).
30
31 37. Y. Li, L. Wei, R. Zhang, Y. Chen, L. Mei, and J. Jiao, Nanoscale Res. Lett., 8, 1–7 (2013).
iew

32 38. H. Liu, W. Dong, H. Wang, L. Lu, Q. Ruan, Y. S. Tan, R. E. Simpson, and J. K. W. Yang, Sci. Adv., 6
33 (2020).
34
35 39. N. Tigau, Cryst. Res.Technol., 42, 281–285 (2007).
36 40. G. Ham, S. Shin, J. Park, J. Lee, H. Choi, S. Lee, and H. Jeon, RSC Adv., 6, 54069–54075 (2016).
On

37 41. S. Lee, S. Shin, G. Ham, J. Lee, H. Choi, H. Park, and H. Jeon, AIP Adv., 7 (2017).
38
42. K. Nakamoto, Infrared and Raman Spectra of Inorganic and Coordination Compounds: Part A: Theory
39
40 and Applications in Inorganic Chemistry: Sixth Edition, 1–419 (2008).
ly

41 43. M. Kizilyalli, M. Bilgin, and H. M. Kizilyalli, J. Solid State Chem., 85, 283–292 (1990).
42 44. R. H. Arntson, F. W. Dickson, and G. Tunell, Sci. (New York, N.Y.), 153, 1673–1674 (1966).
43
44 45. D. Zhang, X. Cao, D. Xu, N. Wang, C. Yu, W. Hu, X. Yan, J. Mi, B. Wen, L. Wang, and L. Zhang,
45 Electrochim. Acta, 259, 100–109 (2018).
46 46. W. Mikenda and A. Prennger, Spectmchim. Acta, 36, 0–5 (1980).
47
48 47. S. A. Wood, Geochim. et Cosmochim. Acta, 53, 237–244 (1989).
49 48. L. Zhou, A. Assoud, Q. Zhang, X. Wu, and L. F. Nazar, J. American Chem. Soc., 141, 19002–19013
50 (2019).
51
49. T. Fuchs, S. P. Culver, P. Till, and W. G. Zeier, ACS Energy Lett., 5, 146–151 (2020).
52
53 50. J. M. Doux, Y. Yang, D. H. S. Tan, H. Nguyen, E. A. Wu, X. Wang, A. Banerjee, and Y. S. Meng, J.
54 Mater. Chem. A, 8, 5049–5055 (2020).
55 51. J. Cannarella and C. B. Arnold, J. Power Sources, 245, 745–751 (2014).
56
57 52. T. B. W. B. R. Marvin Cronau, Marvin Szabo, Christopher Konig, ACS Energy Lett., 6, 3072–3077 (2021).
58 53. D. Pérez-Coll, E. Sánchez-López, and G. C. Mather, Solid State Ionics, 181, 1033–1042 (2010).
59 54. K. H. Park, D. Y. Oh, Y. E. Choi, Y. J. Nam, L. Han, J. Y. Kim, H. Xin, F. Lin, S. M. Oh, and Y. S. Jung,
60
10

https://mc04.manuscriptcentral.com/jes-ecs
Page 11 of 15 Journal of The Electrochemical Society

1
2
3 Adv. Mater., 28, 1874–1883 (2016).
4
5 55. C. R. Mariappan, C. Yada, F. Rosciano, and B. Roling, J. Power Sources, 196, 6456–6464 (2011).
6 56. T. Fuchs, S. P. Culver, P. Till, and W. G. Zeier, ACS Energy Lett., 5, 146–151 (2020).
7 57. Y. Gao, N. Li, Y. Wu, W. Yang, and S. H. Bo, Adv. Energy Mater., 11 (2021).
8
9 58. H. Cao, M. Yu, L. Zhang, Z. Zhang, X. Yan, P. Li, and C. Yu, J. Mater. Sci. Technol., 70, 168–175 (2021).
10 59. H. Wan, J. P. Mwizerwa, F. Han, W. Weng, J. Yang, C. Wang, and X. Yao, Nano Energy, 66, 104109
11 (2019).
12
13
14
15
16
17
18
19
20
21
Fo

22
23
24
rR

25
26
27
28
ev

29
30
31
iew

32
33
34
35
36
On

37
38
39
40
ly

41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
11

https://mc04.manuscriptcentral.com/jes-ecs
Journal of The Electrochemical Society Page 12 of 15

1
2
3
Figures
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
Fo

22
23
24
rR

25
26
27
28
ev

Figure 1 a) Schematic illustrating the Sb2S3 synthesis procedure in methanol, b) XRD patterns of the
29
metathesis precipitate and supernatant, c) TGA/DSC traces of Sb2S3 recovered from methanol.
30
31
iew

32
33
34
35
36
On

37
38
39
40
ly

41
42
43
44
45
46
47 Figure 2 a) XRD, and b) Raman spectra of H2S annealed Sb2S3.
48
49
50
51
52
53
54
55
56
57
58
59
60
12

https://mc04.manuscriptcentral.com/jes-ecs
Page 13 of 15 Journal of The Electrochemical Society

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
Fo

22
23
24
rR

25
26
27
28
ev

29
30
31 Figure 3 a) Schematic illustrating the Na3SbS4 synthesis procedure in ethanol, b) XRD patterns of
iew

32 Na3SbS4 recovered from methanol and water. SEM micrographs of Na3SbS4 recovered from c) H2O
33 (2,000X), d) ethanol (10,000X).
34
35
36
On

37
38
39
40
ly

41
42
43
44
45
46
47
48
49
50
51
52
53
54
55 Figure 4 a) Raman patterns, and b) plot of negative ion intensities for selected elements from TOF-
56 SIMS profiling of Na3SbS4 derived from ethanol (blue curves) and water (red curves).
57
58
59
60
13

https://mc04.manuscriptcentral.com/jes-ecs
Journal of The Electrochemical Society Page 14 of 15

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
Fo

22
23
24
rR

25
26
27
28
ev

29
30
31
iew

32
33
34
35
36 Figure 5 Nyquist plots and equivalent circuit used to fit data (inset) of Na3SbS4 recovered from a)
On

37 ethanol and b) water solution at different stacking pressure; c) the room temperature ionic conductivity
38
of Na3SbS4 as a function of solvent and stacking pressure; d) Arrhenius plots for the ionic conductivity
39
40 of Na3SbS4 recovered from ethanol and water (Pst = 120MPa).
ly

41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
14

https://mc04.manuscriptcentral.com/jes-ecs
Page 15 of 15 Journal of The Electrochemical Society

1
2
3
4
5 Table 1 Summary of Na3SbS4 synthesis parameters and performance reports in the literature
6
7 Synthesis Trxn Tdry Tan Pfab Blocking Pstack σNa+ Ea
8 Solvent Co-solvent Precursors Ref.
method (°C) (°C) (°C) (MPa) ion contacts (MPa) (mS/cm) (eV)
9
10 Hydrate C-coated Al
11

Fo
14
- - Na3SbS4.9H2O RT 150 - 320 - 1.05 0.22
12 Purification foils
13

rR
14 Na2S,
Solid-state - - RT - 550 370 Ti - 1.10 0.20
15 Sb2S3, S

ev
13
16

17 Dissolution- MeOH - Na2S , 0.23 0.37
RT 100 - 370 Ti -

iew
18 precipitation H2O - Sb2S3, S 0.26 0.32
19
20 Ethanol Na2S, 0.15 0.35
15
21 Solution H2O RT 200 - 370 Ti 75
Sb2S3, S

On
Acetone 0.21 0.27
22
23 Solution Na2S, 80 150 - 0.95 0.24 19
24 H2O 2-propanol 720 Au -

ly
Solution Sb2S3, S 80 150 100 1.20 0.24
25
26 Na2S.9H2O, 58
27 Solution H2O Ethanol RT 140 - 350 In foils - 0.12 -
Sb2S3, S
28
29 Na2S, Carbon 59
30 Solution ACN - 50 80 450 240 - 0.25 0.30
Sb2S3, S paste
31
32 H2O Ethanol Na2S, Sb2S3, Stainless 0.33 0.14 This
Solution RT 140 - 240 120
33 Ethanol - S Steel 0.52 0.13 work
34
35

36 Sigma Aldrich, > 99.5%,  Alfa Aesar, > 99.5%, Nagao Co,, 99.1%, Aladdin, 99.99%, In-house synthesis
37
38
39
40
41 15
42
43 https://mc04.manuscriptcentral.com/jes-ecs
44
45
46

You might also like