You are on page 1of 8

Article

Cite This: Chem. Mater. 2018, 30, 990−997 pubs.acs.org/cm

Mechanism of Formation of Li7P3S11 Solid Electrolytes through Liquid


Phase Synthesis
Yuxing Wang,† Dongping Lu,*,† Mark Bowden,‡ Patrick Z. El Khoury,‡ Kee Sung Han,‡
Zhiqun Daniel Deng,† Jie Xiao,† Ji-Guang Zhang,† and Jun Liu*,†

Energy and Environment Directorate, Pacific Northwest National Laboratory, Richland, Washington 99354, United States

Environmental Molecular Sciences Laboratory, Pacific Northwest National Laboratory, Richland, Washington 99354, United States
*
S Supporting Information

ABSTRACT: Crystalline Li7P3S11 is a promising solid electro-


lyte for all solid-state lithium/lithium ion batteries. A
controllable liquid phase synthesis of Li7P3S11 is more desirable
than conventional mechanochemical synthesis, but recent
attempts suffer from reduced ionic conductivities. Here we
elucidate the mechanism of formation of crystalline Li7P3S11
synthesized in the liquid phase [acetonitrile (ACN)]. We
conclude that crystalline Li7P3S11 forms through a two-step
reaction: (1) formation of solid Li3PS4·ACN and amorphous
“Li2S·P2S5” phases in the liquid phase and (2) solid-state
conversion of the two phases. The implication of this two-step
reaction mechanism for morphology control and the transport
properties of liquid phase synthesized Li7P3S11 is identified and
discussed.

■ INTRODUCTION
Energy storage devices are pivotal to vehicle electrification and
Despite many advantages, most solid-state electrolyte
systems suffer from issues such as relatively low ionic
renewable energy storage, which are key steps for meeting conductivity, difficulty in forming and retaining intimate
decarbonization targets around the world. Lithium/lithium ion contact with electrode materials, high processing costs,
batteries (LIBs) are the primary candidates especially in etc.19,20 Compared with oxides and phosphate systems,
applications such as electric vehicles (EVs) and portable sulfide-based solid electrolytes in general have higher ionic
electronic devices where high energy densities are required. conductivities and lower elastic moduli (softer), making them a
However, the high cost and inadequate energy density of state- more practical substitution for liquid electrolytes without much
of-the-art LIBs hinder further market penetration of EVs. It is deviation from the current battery manufacturing process.5,20 It
widely accepted that further enhancement of energy density has been reported that some sulfide systems even have ionic
requires fundamental changes in battery chemistries.1−6 These conductivity exceeding that of liquid electrolytes.17,21−23 Most
changes may arise from electrode materials such as a silicon notable examples are Li10GeP2S12 (12 mS cm−1)21 and Li7P3S11
anode,7 a lithium metal anode,8,9 Li−S,10 Li−air,11 etc., or from (17 mS cm−1)22,23 glass ceramics. Li7P3S11 have been shown to
electrolytes.12 Solid-state electrolytes (SSEs), in particular, form interfaces with electrode materials more kinetically stable
inorganic lithium ion conductors, are generally considered safer than those of Li10GeP2S12.24,25 Conventionally, glass ceramic
and more thermally and chemically/electrochemically stable, Li7P3S11 is synthesized by crystallization of 70Li2S·30P2S5
although this assessment needs to be made on an individual glass,26 which may be prepared by mechanochemical syn-
basis.5,13 Given the well-recognized issues with Li−metal,14 Li− thesis27 or the melt quenching method,28 the former being the
S,15 and Li−air batteries in liquid systems, an all-solid-state preferred method. The mechanochemical synthesis involves
design may be the solution to enable these systems.16 Other
planetary ball-milling of precursor powders under dry or wet
underappreciated advantages of solid-state electrolyte systems
conditions (no chemical interaction between the liquid medium
include the following. (1) Better thermal stability allows battery
operation at elevated temperatures and simplification of and the powder). Although it is effective, there are questions
thermal management, which is translated to enhanced pack- about whether the process is scalable. Also, milled powders
level energy/power densities.17 (2) Dimensional stability allows tend to aggregate into micrometer-sized particles, and an
the use of multiple electrolyte systems with distinct
functionalities.18 (3) The single-ion conduction nature and Received: November 18, 2017
high carrier density improve power performance at low Revised: January 2, 2018
temperatures and ultrahigh current densities.17 Published: January 3, 2018

© 2018 American Chemical Society 990 DOI: 10.1021/acs.chemmater.7b04842


Chem. Mater. 2018, 30, 990−997
Chemistry of Materials Article

additional pulverization treatment may be necessary for use in substance was obtained after the solvent had evaporated from the
composite electrodes.29 supernatant. The dried powder precipitate and the gel from the
Liquid phase synthesis has proven to be an effective method supernatant are denoted as PP and SP, respectively. The DP, PP, and
for synthesizing nanoparticles with controllable size and SP samples were then heat-treated at 200 or 260 °C in a sealed PTFE
container filled with argon gas for 1 h. The sample notation is
morphology.30 The richness of solution chemistry will endow
summarized in Table 1.
great flexibility and tunability to the material preparation
process. For instance, composites of SE and active materials or
SE and conductive carbon can be obtained directly from the Table 1. Samples and Treatment Conditions
liquid phase.18,31 From a manufacturing point of view, the samplea treatment
liquid phase process is readily scalable and more compatible DP deposited precursor from the precipitate and solution mixture
with a conventional electrode preparation process such as after acetonitrile evaporation
composite cathode mixing and slurry coating. The liquid phase PP powder precipitate after centrifugation and decanting
synthesis of sulfide solid electrolytes can be further divided into SP dried solution phase from the supernatant after acetonitrile
two categories. In the first method, precursor powders are evaporation
completely dissolved in organic solvents (methanol, ethanol, N- DP-200 DP annealed at 200 °C for 1 h
methylformamide, hydrazine, etc.)32−35 to form a homoge- DP-260 DP annealed at 260 °C for 1 h
a
neous solution; in the second method, reactions between PP-200, PP-260, SP-200, and SP-260 are defined in the same manner.
precursor powders are mediated by polar, aprotic organic
solvents [tetrahydrofuran (THF), acetonitrile (ACN), 1,2- Thermogravimetric analysis (TGA) and differential thermal analysis
dimethoxyethane (DME), ethyl propionate, etc.] to generate (DTA) were performed on samples DP, PP, and SP with a Netzsch
precipitates.36−38 In both methods, final solid electrolyte STA 449F1 instrument. Sample loading and weighing were performed
products are obtained by drying off the solvent and subsequent inside glovebox. The sample was transferred quickly into the TG
heat treatment for crystallization. While homogeneous instrument in a sealed container, and the chamber was flushed
solutions are suitable for coating electrode particles, the promptly. The samples were heated from 25 to 300 °C under argon.
The morphology of the samples was observed with a dual-focus ion
precipitation method is conducive to producing small, uniform beam (FIB) scanning electron microscope (Environmental, FEI
particles. Unfortunately, reported ionic conductivities of solid Helios) at 5 kV.
electrolytes using either method are lower than those obtained Powder X-ray diffraction (PXRD) was used for phase character-
by mechanochemical or solid-state methods.18,37,39 ization. Samples were sealed in thin-walled glass capillary tubes (500
Recently, various sulfide-based solid electrolytes (β-Li3PS4, μm diameter, 10 μm wall thickness, Charles Supper Co.) under argon.
Li7P3S11, Li7P2S8I, etc.) have been synthesized in the liquid A Rigaku D/Max Rapid II microdiffraction system with a rotating Cr
phase.18,36,37,40 Li7P3S11 is of particular interest because of its target (λ = 2.2910 Å) operated at 35 kV and 25 mA was used to collect
extraordinarily high conductivity in its crystalline form; the diffraction patterns. A parallel X-ray beam collimated to a 300 μm
crystalline Li7P3S11 has been synthesized in THF, acetonitrile, diameter was directed onto the specimen, and the diffracted intensities
were recorded on a large two-dimensional image plate during a 10 min
and DME with a large variation in the reported transport exposure. Alternatively, a desktop diffractometer (Rigaku MiniFlex II)
properties.18,37,39 It is unclear exactly how the Li7P3S11 was employed with a scan speed of 2° min−1 and a step size of 0.05°.
crystalline phase forms from the deposited electrolyte The sample was covered with an 8 μm Kapton film during the
precursor. In this research, we improve our understanding of measurement.
the mechanism of formation of the Li7P3S11 phase in The Raman spectra were collected using a Raman spectrometer
acetonitrile by tracking the phase change of not only the (Horiba LabRAM HR) coupled with an inverted optical microscope
precipitates but also the dissolved phase in the supernatant (Nikon Ti-E). The incident CW laser light source (633 nm) was
liquids. We found that the deposited electrolyte precursor is attenuated using a variable neutral density filter wheel (to ∼5 μW/
actually a mixture of crystalline Li3PS4·ACN and amorphous μm2), reflected off a dichroic beam splitter, and focused onto the
sample using a 10× microscope objective. The back-scattered light was
“Li2S·P2S5·ACN”, which then convert into crystalline Li7P3S11 collected through the same objective, transmitted though the beam
through solid-state reaction. Implications of this formation splitter cube, and dispersed through a 600 g/mm grating onto a CCD
mechanism for the transport property and morphology of detector. Spectra were acquired as time series (10 sequentially
liquid phase synthesized Li7P3S11 solid electrolytes are recorded spectra, each of which was time-integrated for 5 s) to ensure
discussed. the integrity of our sample. As such, the final spectra shown are time-


averaged and otherwise not subjected to further data analysis
procedures.
EXPERIMENTAL SECTION 31
P solid-state magic angle spinning (MAS) nuclear magnetic
Because of the extreme sensitivity of the sulfide compounds to resonance (NMR) spectra were obtained at a spinning speed of 20
moisture, all operations were performed in an Ar-filled glovebox unless kHz and 295 K with a 3.2 mm HXY probe on a 600 MHz NMR
otherwise noted. The Li7P3S11 powder was synthesized from Li2S (Alfa spectrometer (Bruker). The spectra were obtained by the Fourier
Aesar, 99.9%) and P2S5 (Sigma-Aldrich, 99%) in acetonitrile transformation of free induction decay after a single-pulse excitation
(Selectilyte BASF, battery grade). All the raw materials were used with a 90° pulse length of 4 μs and a repetition delay of 100 s. The 31P
without further treatment. Stoichiometric amounts of Li2S and P2S5 chemical shift (δ) was calibrated using 0 ppm of 85% H3PO4 as an
were ground first in an agate mortar and poured into the acetonitrile external reference. Spinning sidebands at multiples of the spinning
solvent. The powder:solvent ratio is 1:20 in grams per milliliter. The speed were determined by comparison between the spectra obtained
mixture was heated at 50 °C on a hot plate and stirred for 3 days. A at various spinning speeds (15, 20, and 23 kHz).
slightly greenish solution containing white precipitate was obtained. Electrochemical impedance spectroscopy (EIS) was employed to
The solvent was then allowed to evaporate on a hot plate at 150 °C. characterize the transport properties of the samples; 100 mg of the
The deposited precursor is denoted as DP. Alternatively, the solution sample powder was pelletized by cold pressing in a 10 mm diameter
mixture was sealed in a tube and centrifuged at 4400 rpm for 6 min. pressing die at 380 MPa. Indium electrodes were formed by pressing
The powder precipitate and the supernatant were separated after In foils onto both sides of the pellet at 130 MPa. The pellet was
decanting. Both were then dried under vacuum. A clear gel-like sandwiched by two stainless steel rods inside a Swagelok setup for

991 DOI: 10.1021/acs.chemmater.7b04842


Chem. Mater. 2018, 30, 990−997
Chemistry of Materials Article

Figure 1. (a) Images of the deposited precursor (DP), the powder precipitate (PP), and the supernatant containing the solution phase. (b) TGA and
DTA data of PP, DP, and SP samples.

impedance measurement. The measured temperature range was −40 appears to be a small exothermic peak in the DTA curve, and a
to 100 °C. Low-temperature testing was performed inside an slight change in slope in the TG curve at 260 °C. Similar
environmental chamber using an electrochemical interface (Solartron features can be seen in the TG and DTA curves of sample DP.
1287, Solartron Analytical) and a frequency response analyzer The origin of these peaks will be discussed below. In addition,
(Solartron 1260, Solartron Analytical); high-temperature testing was
performed inside a heating oven using a Biologic potentiostat sample DP continued to lose weight after the major thermal
(VMP3). The frequency range was 1 MHz to 1 Hz. event, similar to sample SP but to a lesser degree. The TGA


and DTA results suggest sample DP has characteristics of
samples PP and SP.
RESULTS AND DISCUSSION The morphologies of PP-260, DP-260, and SP-260 were
White precipitates appear immediately after the mixture of Li2S characterized by scanning electron microscopy (SEM) (Figure
and P2S5 (originally yellowish) is added to acetonitrile. The 2). PP-260 and DP-260 samples were in the powder form; SP-
colorless solution turns bluish within 1 min. The bluish color is
preserved after the mixture is stirred at 50 °C for 3 days. The
obtained DP sample (without centrifugation) appears slightly
yellowish, whereas the PP sample (with centrifugation) appears
entirely white (Figure 1a). Interestingly, the color of the
supernatant changes from bluish to yellowish slowly after
decanting. It is consistently found that the yields of DP and PP
are ∼0.7 of ∼1.15 g, respectively; a clear gel-like substance
remains after the evaporation of acetonitrile. These pieces of Figure 2. Scanning electron microscopy images of samples (a) PP-260,
evidence clearly suggest that some of the Li2S and P2S5 (b) DP-260, and (c) SP-260.
precursors remain in the supernatant and the DP is actually a
combination of PP and SP. 260 was obtained by casting the supernatant onto an aluminum
According to TGA and DTA (Figure 1b), both samples PP substrate followed by annealing. Sample PP-260 consists of
and DP exhibit one major thermal event around 200 °C. Close submicrometer primary particles (Figure 2a). Determination of
inspection reveals that the onset and peak temperatures of the the exact particle size is complicated by slight agglomeration of
DP sample are 10 °C higher than those of the PP sample (200 primary particles. Nevertheless, these particles are much more
and 220 °C vs 190 and 210 °C, respectively). Both samples uniform and smaller than those synthesized via mechanochem-
experience large weight losses during the event. Rangasamy et ical synthesis.29,41 Sample SP-260 before annealing (Figure
al. reported that Li2S and P2S5 at a 3:1 molar ratio combine S1d) shows a featureless, complete coverage of an amorphous
with acetonitrile to form an unknown crystalline phase, which phase on the substrate; after annealing, the amorphous feature
the authors assigned as Li3PS4·2ACN.40 For sample PP, the remains in most of the regions while cracks and particles appear
weight loss below 100 °C can be attributed to absorbed (Figure 2c), probably due to shrinkage induced by acetonitrile
acetonitrile; minimal weight loss was observed between 100 evaporation. The morphology of DP-260 is similar to that of
and 190 °C and above 230 °C, suggesting that the complex PP-260, but there seems to be more agglomeration or
phase is stable below 190 °C but decomposes and releases all secondary amorphous phase binding of the primary particles
acetonitrile above 190 °C in an endothermic reaction. This is in (Figure 2b). From the low-magnification images (Figure S1), it
stark contrast to Li3PS4·3THF (tetrahydrofuran), which can be seen that some primary particles agglomerate into very
decomposes at ∼100 °C.36 However, on the basis of the large secondary particles in the extreme case. In contrast,
weight loss of the PP sample at 200 °C, we believe the formula particles in sample PP-260 seem to be more dispersible. These
of the acetonitrile complex should be Li3PS4·ACN rather than results indicate the amorphous solution phase has a critical
Li3PS4·2ACN. The TG and DTA curves of sample SP show a effect on the morphology of Li7P3S11.
rather smooth continuous trend throughout the temperature The PXRD patterns of sample PP (Figure S3) before
range. Most features are too small to explain. Noticeably, there annealing match well with those of the crystalline phase
992 DOI: 10.1021/acs.chemmater.7b04842
Chem. Mater. 2018, 30, 990−997
Chemistry of Materials Article

“Li3PS4·2ACN”;40 after annealing at 200 or 260 °C, the desired Li7P3S11 phase. The conversion occurs at a temperature
patterns (Figure 3) match well with those of the reported β- that is slightly higher (∼10 °C) than that of the conversion of
sample PP into β-Li3PS4, as seen from TG and DTA, which
explains the presence of both β-Li3PS4 and Li7P3S11 phases in
sample DP-200. As DP is basically PP particles coated by SP,
two processes need to occur during the conversion: (1)
decomposition of PP and formation of β-Li3PS4 with the release
of ACN and (2) solid-state reaction of SP with PP or PP
derivatives. It is inconclusive whether the two processes occur
concurrently or sequentially; the presence of β-Li3PS4 suggests
the latter mechanism is more likely. It can also be inferred that
due to the need for solid-state diffusion and reaction, the
formed crystalline Li7P3S11 may suffer from local inhomoge-
neity and nonstoichiometry. According to the SEM image
(Figure 2b), the conversion did not consume the SP coating
completely, as some amorphous coverage was still observed in
sample DP-260.
PXRD is useful in the identification of crystalline phases, but
amorphous components of the sample cannot be characterized.
Raman and 31P MAS NMR spectroscopy were employed as
complementary tools to improve our understanding of the local
structures of the samples. The Raman spectra of PP-200 and
PP-260 are similar (Figure 4). A single peak at 426 cm−1 can be

Figure 4. Raman (left) and 31P MAS NMR (right) spectra of samples
Figure 3. PXRD patterns of samples DP, PP, and SP after heat DP, PP, and SP annealed at different temperatures. The asterisk
treatment at 200 or 260 °C. The raw patterns and detailed pattern denotes a spinning sideband.
fitting can be found in Figure S2.
assigned to the local structural unit of PS43− tetrahedra,
42
Li3PS4. Peaks corresponding to Li2S were observed in both confirming that only the Li3PS4 phase is present.43 The
patterns, indicating small amounts of Li2S were present in both spectrum of sample DP-260 shows a major peak at 410 cm−1
samples. The crystallinity of sample PP is higher with a higher corresponding to the local structural unit of P2S74− and a
annealing temperature; small unknown peaks also disappear in shoulder peak at 426 cm−1 corresponding to PS43−, consistent
sample PP-260. We can unambiguously conclude from the with the crystalline structure of the Li7P3S11 phase.27 The 426
XRD and TGA/DTA data that the main phase of sample PP is cm−1 peak is stronger in sample DP-200. This is expected as the
Li3PS4·ACN, which undergoes decomposition and transforms residual β-Li3PS4 phase is present in the XRD pattern of DP-
into β-Li3PS4 phase at >190 °C. 200. The Raman spectrum of SP-260 is not presented because
For sample DP, the XRD pattern before annealing is similar of the large fluorescence background, which hinders the
to that of PP (Figure S3), suggesting that the SP coverage on determination of its local structure.
PP is indeed amorphous. After annealing, the pattern of DP- The 31P MAS NMR spectra of samples PP-260 and DP-260
260 matches well with that of the high-conductivity Li7P3S11 (Figure 4) match well with the reported spectra of 75Li2S·
phase, whereas DP-200 is a combination of Li7P3S11, β-Li3PS4, 25P2S5 and 70Li2S·30P2S5 glass ceramics.44 The two strong
and some unknown phase. It should be noted that Bragg peaks peaks at 91 and 87 ppm can be assigned to the structural units
reflect only crystalline phases, and amorphous phases could also of P2S74− and PS43− in the crystalline phases, respectively.45
be present in the sample. The completely different phases of Seino et al. analyzed the degree of crystallization in the 70Li2S·
DP-260 and PP-260 indicate the critical role of the solution 30P2S5 glass ceramics using 31P MAS NMR.46 They showed
phase (SP) in converting the powder precipitate (PP) into the that broad peaks related to P2S74− and PS43− units in the
993 DOI: 10.1021/acs.chemmater.7b04842
Chem. Mater. 2018, 30, 990−997
Chemistry of Materials Article

Figure 5. Schematic illustration of the mechanism of formation of Li7P3S11 synthesized in acetonitrile.

amorphous component convolute with sharp peaks corre- converted to the high-conductivity Li7P3S11 phase after
sponding to P2S74− and PS43− in the crystalline component in annealing. The proposed mechanism of formation of Li7P3S11
poorly crystalline samples. Minimal broadening at the base of is illustrated in Figure 5. This precipitation−solid-state
sharp crystalline peaks was observed in the spectra of samples conversion reaction mechanism may be universal to the liquid
PP-260 and DP-260, suggesting that both samples are mostly phase synthesis of thiophosphate compounds and their
crystalline. Two small peaks at 109 and 105 ppm in the derivatives, as the Li3PS4·ACN phase may be the only stable
spectrum of DP-260 can be assigned to the structural unit of precipitate. For instance, the Li7P2S8I phase has been
P2S64− in the crystalline Li4P2S6 phase. The Li4P2S6 phase is synthesized by solid-state reaction of liquid phase-synthesized
known to arise from decomposition of glassy Li4P2S7 with “Li3PS4·2ACN” and LiI.40 It follows that the morphology of
release of elemental S.45 It is also reported that prolonged these liquid-synthesized thiophosphate compounds is largely
heating of Li7P3S11 will lead to formation of the Li4P2S6 phase.47 determined by that of the Li3PS4·ACN precipitate. Therefore,
We can safely conclude from XRD, Raman, and 31P NMR the mechanism of formation of the Li3PS4·ACN precipitate is
analysis that the crystalline Li7P3S11 phase is formed by critically important, which has not been elucidated. Given the
conversion of Li3PS4 or Li3PS4·ACN and the solution phase insoluble nature of Li2S and high solubility of “Li2S·P2S5”, the
through solid-state reaction. There remains a question about formation of Li 3 PS 4 ·ACN most likely occurs through
what the solution phase is. To answer the question, we conversion of Li2S particles from outside to inside by the
conducted a similar synthesis but varied the Li2S:P2S5 molar “Li2S·P2S5” solution. Under such an assumption, the morphol-
ratio. Interestingly, the quantity of powder precipitates ogy of Li2S particles will determine the morphology of the
decreases with an increase in the amount of P2S5; e.g., at a Li3PS4·ACN precipitate. When the Li2S precursor particles are
55:45 ratio, only 0.2 g of PP was obtained from a total of 1 g of large, the conversion may never reach completion because of
precursors. The XRD pattern of PP at a 55:45 ratio is similar to the passivation of the outer Li3PS4·ACN, which can explain the
that at a 75:25 ratio (Figure S3), also matching the dried presence of Li2S in samples PP-200 and PP-260.
precursor:Li7P3S11 ratio of Yao et al.18 and that of the “Li3PS4· With annealing of the precipitates and solution phase
2ACN” phase of Rangasamy et al.,40 indicating that all the mixture, a majority of the solution phase will be “assimilated”
powder precipitates are essentially the same compound. At a during the solid-state reaction but a small amount may remain
1:1 ratio, Li2S and P2S5 dissolve quickly to form a clear on the surface of Li7P3S11 particles. Therefore, the state into
yellowish solution. Indeed, Liang et al. have shown that the 1:1 which the solution phase transforms at the annealing
Li2S/P2S5 solution can be used to coat Li2S particles.48,49 temperature will influence the transport property of DP-260.
Therefore, the solution phase is some amorphous form of The PXRD patterns of SP-200 and SP-260 have large
“Li2S·P2S5·xACN”. The term “Li2S·P2S5” is used in a vague amorphous backgrounds, consistent with the SEM observation
sense to refer to the phase with a Li2S:P2S5 molar ratio of 1:1 that the sample remains mostly amorphous. Bragg peaks in SP-
that either dissolves in or precipitates out of the solution, 200 can be attributed to P2S5, and peaks in SP-260 can be
because the solvated form of the solution phase or the structure attributed to Li4P2S6 and Li2P2S6 phases.50,51 The small
of the amorphous solids after solvent removal was not directly exothermic peak in the DTA data of sample SP can then be
determined. It should be noted that neither Li2S nor P2S5 is attributed to crystallization of the amorphous phase. The 31P
soluble in acetonitrile, so the dissolution of Li2S and P2S5 (1:1) NMR spectrum of SP-260 is quite complex. Most peaks are
is probably due to the formation of a new structural P2S62− unit broad, confirming the amorphous nature. The peak at 109 ppm
that is soluble in acetonitrile. We believe that Li2S and P2S5 (in is associated with P2S64− in the Li4P2S6 phase; the peak at 83
an x:100 − x molar ratio, where 50 < x < 75) participate in the ppm is due to glassy Li2S−P2S5.45 Two peaks of small chemical
reaction in acetonitrile as follows: shifts are unidentified. Nevertheless, none of the crystalline
Li 4 P 2S 6 , Li 2 P 2 S 6 , or glassy Li 2 S−P 2 S 5 has good ionic
x Li 2S + (100 − x)P2S5 + ACN conductivity,50−52 so the presence of the residual solution
→ 2(x − 50)Li3PS4 ·ACN(p) + 2(75 − x)Li 2S· P2S5(s) phase is expected to lower the overall ionic conductivity of
sample DP-260.
When x < 50, soluble “Li2S·P2S5” forms and some P2S5 remains The ionic conductivity of sample DP-260 was characterized
undissolved; when x > 75, the Li3PS4·ACN precipitate forms by electrochemical impedance spectroscopy over the temper-
and some Li2S remains undissolved. In the case of 70Li2S· ature range of −40 to 100 °C. The impedance plots are shown
30P2S5 under equilibrium conditions, 80 mol % Li3PS4·ACN in Figure S4, and detailed fitting information can be found in
(PP) precipitates out while 20 mol % “Li2S·P2S5” dissolves in the Supporting Information. At or above room temperature (22
the solution. °C), the impedance plots show a straight line intercepting or
After acetonitrile evaporation, the Li3PS4·ACN precipitates trending toward the x-axis at low and intermediate frequencies
are coated by the amorphous “Li2S·P2S5” phase and then whereas the bending at high frequencies is due to stray
994 DOI: 10.1021/acs.chemmater.7b04842
Chem. Mater. 2018, 30, 990−997
Chemistry of Materials Article

capacitance and stray inductance. The straight line represents a elucidated by this study and hence how sensitive the transport
contribution from the electrodes, and the intercepts give the properties can be to experimental conditions. Identifying the
impedance of the materials. It cannot be determined from these two-step reaction mechanism provides crucial clues for further
data what components contribute to the impedance; therefore, improving the quality of Li7P3S11 from the liquid phase
we denote it as total impedance (Rt). At low temperatures (≤0 synthesis. Measures that could potentially improve the
°C), a semicircle appeared whose left intercept did not go to transport property of liquid phase-synthesized Li7P3S11 should
zero. Apparently, there are at least two processes that focus on facilitating complete conversion of the two solid
significantly contribute to overall impedance. Literature phases, for instance, (1) optimizing the solvent extraction
references of low-temperature impedance measurement of process so the two phases are more uniformly mixed, (2)
Li7P3S11 solid electrolytes are scarce, so it is difficult to optimizing the annealing process for better solid-state
determine the origin of the low-frequency semicircle. We conversion, and (3) reducing the size of the Li3PS4·ACN
assume that the process at high frequency corresponds to particle to shorten diffusion length, etc.
intragrain (bulk) transport and the process at lower frequency It is evident that the intergrain process has an activation
corresponds to intergrain transport. The intergrain impedance energy much higher than that of the intragrain process.
may be due to secondary phases or simply grain boundary Although indeterminable from the impedance spectrum, the
impedance. The intragrain resistance and the intergrain intergrain resistance at 22 and 50 °C can be calculated by
resistance were obtained by fitting, and the ionic conductivities extrapolation to be 10 and 1 Ω, respectively; the intergrain
were calculated from the resistance and the sample geometry. resistance is negligible at high temperatures.


The room-temperature conductivity of sample DP-260 is 8.7 ×
10−4 S cm−1.
CONCLUSIONS
The activation energy was calculated on the basis of the
Arrhenius equation: σ = σ0/T exp(−Ea/kT). It can be seen The mechanism of formation of crystalline Li7P3S11 solid
(Figure 6) that the total ionic conductivity shows a good electrolytes synthesized in acetonitrile was revealed by
analyzing the intermediate products during the reaction. It is
found that the precursors (70:30 Li2S:P2S5 molar ratio) in
acetonitrile form Li3PS4·ACN (75:25 Li2S:P2S5) precipitates
and soluble “Li2S·P2S5” (50:50 Li2S:P2S5). Unlike the β-Li3PS4
crystalline phase that forms directly from the decomposition of
Li3PS4·ACN precipitates, crystalline Li7P3S11 forms through the
solid-state reaction of the Li3PS4·ACN precipitates and the
amorphous “Li2S·P2S5” phase from the supernatant. The
soluble species “Li2S·P2S5” appears as amorphous coverage on
the precipitate particles after acetonitrile evaporation. The
liquid phase-synthesized Li7P3S11 has a total conductivity of
0.87 mS cm−1 at room temperature. The understanding of the
formation mechanism provides clues for further optimizing the
transport properties and morphologies of liquid phase-
synthesized Li7P3S11 solid electrolytes.

Figure 6. Arrhenius plot for the total and intergrain ion conductivity of
Li7P3S11 (260 °C, DP-260).

*
ASSOCIATED CONTENT
S Supporting Information

The Supporting Information is available free of charge on the


Arrhenius behavior in the entire temperature range with an ACS Publications website at DOI: 10.1021/acs.chemma-
activation energy of 0.37 eV (36 kJ mol−1). This value is close ter.7b04842.
to that of 70Li2S·30P2S5 glass obtained by mechanical milling Additional SEM images and XRD data, impedance
and much lower than that of 70Li2S·30P2S5 glass ceramics (0.18 spectra, and detailed fitting information (PDF)
eV). Seino et al. showed that the activation energy of 70Li2S·
30P2S5 glass ceramics from mechanical milling is roughly
negatively correlated with the degree of crystallization.
However, the observed high activation energy cannot be
■ AUTHOR INFORMATION
Corresponding Authors
explained by this correlation because the XRD and NMR data
indicated that sample DP-260 is mostly crystalline. This *E-mail: dongping.lu@pnnl.gov.
observation highlights the key differences between Li7P3S11 *E-mail: jun.liu@pnnl.gov.
solid electrolytes obtained via solid-state route and liquid phase ORCID
synthesis. Glass ceramics obtained by a solid-state route arise Dongping Lu: 0000-0001-9597-8500
from crystallization of a homogeneous single-phase glass, Patrick Z. El Khoury: 0000-0002-6032-9006
whereas liquid phase synthesized samples arise from reaction Kee Sung Han: 0000-0002-3535-1818
of at least two phases. Notably, the reported activation energies Ji-Guang Zhang: 0000-0001-7343-4609
of Li7P3S11 via liquid phase synthesis, despite all having the
superionic crystal phase, have large variation, ranging from 23
Jun Liu: 0000-0001-8663-7771
to 38 kJ mol−1.18,37,39 This is not surprising considering the Notes
complexity of the formation mechanism in the liquid phase as The authors declare no competing financial interest.
995 DOI: 10.1021/acs.chemmater.7b04842
Chem. Mater. 2018, 30, 990−997
Chemistry of Materials Article

■ ACKNOWLEDGMENTS
This work was supported by the Energy Efficiency and
(18) Yao, X. Y.; Liu, D.; Wang, C. S.; Long, P.; Peng, G.; Hu, Y. S.;
Li, H.; Chen, L. Q.; Xu, X. X. High-Energy All-Solid-State Lithium
Batteries with Ultralong Cycle Life. Nano Lett. 2016, 16, 7148−7154.
Renewable Energy (EERE) Office of Vehicle Technologies of (19) Kerman, K.; Luntz, A.; Viswanathan, V.; Chiang, Y. M.; Chen, Z.
the U.S. Department of Energy (DOE) under Contract B. Review-Practical Challenges Hindering the Development of Solid
DEAC02-05CH11231 and DEAC02-98CH10886 for the State Li Ion Batteries. J. Electrochem. Soc. 2017, 164, A1731−A1744.
Advanced Battery Materials Research (BMR) Program and (20) Takada, K. Progress and prospective of solid-state lithium
the U.S. DOE EERE Water Power Technologies Office and the batteries. Acta Mater. 2013, 61, 759−770.
(21) Kamaya, N.; Homma, K.; Yamakawa, Y.; Hirayama, M.; Kanno,
U.S. Army Corps of Engineers Portland District. SEM, solid- R.; Yonemura, M.; Kamiyama, T.; Kato, Y.; Hama, S.; Kawamoto, K.;
state NMR, and Raman characterization were conducted in the Mitsui, A. A lithium superionic conductor. Nat. Mater. 2011, 10, 682−
William R. Wiley Environmental Molecular Sciences Labo- 686.
ratory (EMSL). PNNL is operated by Battelle for the DOE (22) Seino, Y.; Ota, T.; Takada, K.; Hayashi, A.; Tatsumisago, M. A
under Contract DE-AC05-76RLO1830. sulphide lithium super ion conductor is superior to liquid ion


conductors for use in rechargeable batteries. Energy Environ. Sci. 2014,
7, 627−631.
REFERENCES (23) Busche, M. R.; Weber, D. A.; Schneider, Y.; Dietrich, C.;
(1) Scrosati, B.; Garche, J. Lithium batteries: Status, prospects and Wenzel, S.; Leichtweiss, T.; Schroder, D.; Zhang, W. B.; Weigand, H.;
future. J. Power Sources 2010, 195, 2419−2430. Walter, D.; Sedlmaier, S. J.; Houtarde, D.; Nazar, L. F.; Janek, J. In Situ
(2) Armand, M.; Tarascon, J. M. Building better batteries. Nature Monitoring of Fast Li-Ion Conductor Li7P3S11 Crystallization Inside
2008, 451, 652−657. a Hot-Press Setup. Chem. Mater. 2016, 28, 6152−6165.
(3) Goodenough, J. B.; Kim, Y. Challenges for Rechargeable Li (24) Wenzel, S.; Randau, S.; Leichtweiss, T.; Weber, D. A.; Sann, J.;
Batteries. Chem. Mater. 2010, 22, 587−603. Zeier, W. G.; Janek, J. Direct Observation of the Interfacial Instability
(4) Goodenough, J. B.; Kim, Y. Challenges for rechargeable batteries. of the Fast Ionic Conductor Li10GeP2S12 at the Lithium Metal
J. Power Sources 2011, 196, 6688−6694. Anode. Chem. Mater. 2016, 28, 2400−2407.
(5) Jung, Y. S.; Oh, D. Y.; Nam, Y. J.; Park, K. H. Issues and (25) Mo, Y. F.; Ong, S. P.; Ceder, G. First Principles Study of the
Challenges for Bulk-Type All-Solid-State Rechargeable Lithium Li10GeP2S12 Lithium Super Ionic Conductor Material. Chem. Mater.
Batteries using Sulfide Solid Electrolytes. Isr. J. Chem. 2015, 55, 2012, 24, 15−17.
472−485. (26) Minami, K.; Hayashi, A.; Tatsumisago, M. Crystallization
(6) Scrosati, B.; Hassoun, J.; Sun, Y. K. Lithium-ion batteries. A look Process for Superionic Li7P3S11 Glass-Ceramic Electrolytes. J. Am.
into the future. Energy Environ. Sci. 2011, 4, 3287−3295. Ceram. Soc. 2011, 94, 1779−1783.
(7) Ashuri, M.; He, Q. R.; Shaw, L. L. Silicon as a potential anode (27) Mizuno, F.; Hayashi, A.; Tadanaga, K.; Tatsumisago, M. New,
material for Li-ion batteries: where size, geometry and structure highly ion-conductive crystals precipitated from Li2S-P2S5 glasses.
matter. Nanoscale 2016, 8, 74−103. Adv. Mater. 2005, 17, 918.
(8) Xu, W.; Wang, J. L.; Ding, F.; Chen, X. L.; Nasybulin, E.; Zhang, (28) Minami, K.; Mizuno, F.; Hayashi, A.; Tatsumisago, M. Lithium
Y. H.; Zhang, J. G. Lithium metal anodes for rechargeable batteries. ion conductivity of the Li2S-P2S5 glass-based electrolytes prepared by
Energy Environ. Sci. 2014, 7, 513−537. the melt quenching method. Solid State Ionics 2007, 178, 837−841.
(9) Cheng, X. B.; Zhang, R.; Zhao, C. Z.; Zhang, Q. Toward Safe (29) Sugiura, K.; Kubo, H.; Hashimoto, Y.; Koyama, T. Method for
Lithium Metal Anode in Rechargeable Batteries: A Review. Chem. Rev. producing sulfide solid electrolyte materials. U.S. Patent 9595735 B2,
2017, 117, 10403−10473. 2017.
(10) Manthiram, A.; Fu, Y. Z.; Chung, S. H.; Zu, C. X.; Su, Y. S. (30) Cushing, B. L.; Kolesnichenko, V. L.; O’Connor, C. J. Recent
Rechargeable Lithium-Sulfur Batteries. Chem. Rev. 2014, 114, 11751− advances in the liquid-phase syntheses of inorganic nanoparticles.
11787. Chem. Rev. 2004, 104, 3893−3946.
(11) Grande, L.; Paillard, E.; Hassoun, J.; Park, J. B.; Lee, Y. J.; Sun, (31) Han, F. D.; Yue, J.; Fan, X. L.; Gao, T.; Luo, C.; Ma, Z. H.; Suo,
L. M.; Wang, C. S. High-Performance All-Solid-State Lithium-Sulfur
Y. K.; Passerini, S.; Scrosati, B. The Lithium/Air Battery: Still an
Battery Enabled by a Mixed-Conductive Li2S Nanocomposite. Nano
Emerging System or a Practical Reality? Adv. Mater. 2015, 27, 784−
Lett. 2016, 16, 4521−4527.
800.
(32) Park, K. H.; Oh, D. Y.; Choi, Y. E.; Nam, Y. J.; Han, L. L.; Kim,
(12) Xu, K. Electrolytes and Interphases in Li-Ion Batteries and
J. Y.; Xin, H. L.; Lin, F.; Oh, S. M.; Jung, Y. S. Solution-Processable
Beyond. Chem. Rev. 2014, 114, 11503−11618. Glass LiI-Li4SnS4 Superionic Conductors for All-Solid-State Li-Ion
(13) Zhu, Y. Z.; He, X. F.; Mo, Y. F. Origin of Outstanding Stability
Batteries. Adv. Mater. 2016, 28, 1874−1883.
in the Lithium Solid Electrolyte Materials: Insights from Thermody- (33) Yubuchi, S.; Teragawa, S.; Aso, K.; Tadanaga, K.; Hayashi, A.;
namic Analyses Based on First-Principles Calculations. ACS Appl. Tatsumisago, M. Preparation of high lithium-ion conducting Li6PS5Cl
Mater. Interfaces 2015, 7, 23685−23693. solid electrolyte from ethanol solution for all-solid-state lithium
(14) Lu, D. P.; Shao, Y. Y.; Lozano, T.; Bennett, W. D.; Graff, G. L.; batteries. J. Power Sources 2015, 293, 941−945.
Polzin, B.; Zhang, J. G.; Engelhard, M. H.; Saenz, N. T.; Henderson, (34) Teragawa, S.; Aso, K.; Tadanaga, K.; Hayashi, A.; Tatsumisago,
W. A.; Bhattacharya, P.; Liu, J.; Xiao, J. Failure Mechanism for Fast- M. Preparation of Li2S-P2S5 solid electrolyte from N-methylforma-
Charged Lithium Metal Batteries with Liquid Electrolytes. Adv. Energy mide solution and application for all-solid-state lithium battery. J.
Mater. 2015, 5, 1400993. Power Sources 2014, 248, 939−942.
(15) Lv, D. P.; Zheng, J. M.; Li, Q. Y.; Xie, X.; Ferrara, S.; Nie, Z. M.; (35) Wang, Y. M.; Liu, Z. Q.; Zhu, X. L.; Tang, Y. F.; Huang, F. Q.
Mehdi, L. B.; Browning, N. D.; Zhang, J. G.; Graff, G. L.; Liu, J.; Xiao, Highly lithium-ion conductive thio-LISICON thin film processed by
J. High Energy Density Lithium-Sulfur Batteries: Challenges of Thick low-temperature solution method. J. Power Sources 2013, 224, 225−
Sulfur Cathodes. Adv. Energy Mater. 2015, 5, 1402290. 229.
(16) Nagata, H.; Chikusa, Y. A lithium sulfur battery with high power (36) Liu, Z. C.; Fu, W. J.; Payzant, E. A.; Yu, X.; Wu, Z. L.; Dudney,
density. J. Power Sources 2014, 264, 206−210. N. J.; Kiggans, J.; Hong, K. L.; Rondinone, A. J.; Liang, C. D.
(17) Kato, Y.; Hori, S.; Saito, T.; Suzuki, K.; Hirayama, M.; Mitsui, Anomalous High Ionic Conductivity of Nanoporous beta-Li3PS4. J.
A.; Yonemura, M.; Iba, H.; Kanno, R. High-power all-solid-state Am. Chem. Soc. 2013, 135, 975−978.
batteries using sulfide superionic conductors. Nat. Energy 2016, 1, (37) Xu, R. C.; Xia, X. H.; Yao, Z. J.; Wang, X. L.; Gu, C. D.; Tu, J. P.
16030. Preparation of Li7P3S11 glass-ceramic electrolyte by dissolution-

996 DOI: 10.1021/acs.chemmater.7b04842


Chem. Mater. 2018, 30, 990−997
Chemistry of Materials Article

evaporation method for all-solid-state lithium ion batteries. Electro-


chim. Acta 2016, 219, 235−240.
(38) Phuc, N. H. H.; Morikawa, K.; Mitsuhiro, T.; Muto, H.;
Matsuda, A. Synthesis of plate-like Li3PS4 solid electrolyte via liquid-
phase shaking for all-solid-state lithium batteries. Ionics 2017, 23,
2061−2067.
(39) Ito, S.; Nakakita, M.; Aihara, Y.; Uehara, T.; Machida, N. A
synthesis of crystalline Li7P3S11 solid electrolyte from 1,2-dimethoxy-
ethane solvent. J. Power Sources 2014, 271, 342−345.
(40) Rangasamy, E.; Liu, Z. C.; Gobet, M.; Pilar, K.; Sahu, G.; Zhou,
W.; Wu, H.; Greenbaum, S.; Liang, C. D. An Iodide-Based Li7P2S8I
Superionic Conductor. J. Am. Chem. Soc. 2015, 137, 1384−1387.
(41) Sakuda, A.; Hayashi, A.; Tatsumisago, M. Sulfide Solid
Electrolyte with Favorable Mechanical Property for All-Solid-State
Lithium Battery. Sci. Rep. 2013, 3, 2261 DOI: 10.1038/srep02261.
(42) Homma, K.; Yonemura, M.; Kobayashi, T.; Nagao, M.;
Hirayama, M.; Kanno, R. Crystal structure and phase transitions of
the lithium ionic conductor Li3PS4. Solid State Ionics 2011, 182, 53−
58.
(43) Tachez, M.; Malugani, J. P.; Mercier, R.; Robert, G. Ionic-
Conductivity of and Phase-Transition in Lithium Thiophosphate
Li3ps4. Solid State Ionics 1984, 14, 181−185.
(44) Murakami, M.; Shimoda, K.; Shiotani, S.; Mitsui, A.; Ohara, K.;
Onodera, Y.; Arai, H.; Uchimoto, Y.; Ogumi, Z. Dynamical Origin of
Ionic Conductivity for Li7P3S11 Metastable Crystal As Studied by Li-
6/7 and P-31 Solid-State NMR. J. Phys. Chem. C 2015, 119, 24248−
24254.
(45) Eckert, H.; Zhang, Z. M.; Kennedy, J. H. Structural
Transformation of Nonoxide Chalcogenide Glasses - the Short-
Range Order of Li2s-P2s5 Glasses Studied by Quantitative P-31 and
Li-6,7 High-Resolution Solid-State Nmr. Chem. Mater. 1990, 2, 273−
279.
(46) Seino, Y.; Nakagawa, M.; Senga, M.; Higuchi, H.; Takada, K.;
Sasaki, T. Analysis of the structure and degree of crystallisation of
70Li(2)S-30P(2)S(5) glass ceramic. J. Mater. Chem. A 2015, 3, 2756−
2761.
(47) Wei, J. J.; Kim, H.; Lee, D. C.; Hu, R. Z.; Wu, F. X.; Zhao, H. L.;
Alamgir, F. M.; Yushin, G. Influence of annealing on ionic transfer and
storage stability of Li2S-P2S5 solid electrolyte. J. Power Sources 2015,
294, 494−500.
(48) Lin, Z.; Liu, Z. C.; Dudney, N. J.; Liang, C. D. Lithium
Superionic Sulfide Cathode for All-Solid Lithium-Sulfur Batteries. ACS
Nano 2013, 7, 2829−2833.
(49) Liang, C.; Liu, Z.; Fu, W.; Lin, Z.; Dudney, N. J.; Howe, J. Y.;
Rondinone, A. J. Lithium sulfide compositions for battery electrolyte
and battery electrode coatings. U.S. Patent 8597838 B2, 2014.
(50) Dietrich, C.; Sadowski, M.; Sicolo, S.; Weber, D. A.; Sedlmaier,
S. J.; Weldert, K. S.; Indris, S.; Albe, K.; Janek, J.; Zeier, W. G. Local
Structural Investigations, Defect Formation, and Ionic Conductivity of
the Lithium Ionic Conductor Li4P2S6. Chem. Mater. 2016, 28, 8764−
8773.
(51) Dietrich, C.; Weber, D. A.; Culver, S.; Senyshyn, A.; Sedlmaier,
S. J.; Indris, S.; Janek, J.; Zeier, W. G. Synthesis, Structural
Characterization, and Lithium Ion Conductivity of the Lithium
Thiophosphate Li2P2S6. Inorg. Chem. 2017, 56, 6681−6687.
(52) Zhang, Z. M.; Kennedy, J. H. Synthesis and Characterization of
the B2s3-Li2s, the P2s5-Li2s and the B2s3-P2s5-Li2s Glass Systems.
Solid State Ionics 1990, 38, 217−224.

997 DOI: 10.1021/acs.chemmater.7b04842


Chem. Mater. 2018, 30, 990−997

You might also like