You are on page 1of 7

Solid State Ionics 378 (2022) 115888

Contents lists available at ScienceDirect

Solid State Ionics


journal homepage: www.elsevier.com/locate/ssi

Electrochemical stability of a NASICON solid electrolyte from the lithium


aluminum germanium phosphate (LAGP) series
Vinicius Martins Zallocco a, c, *, Jhonys Machado Freitas b, Nerilso Bocchi b,
Ana Candida Martins Rodrigues c
a
Graduate Program in Materials Science and Engineering, Federal University of São Carlos, 13565-905 São Carlos, SP, Brazil
b
Department of Chemistry, Federal University of São Carlos, C.P. 676, 13565-905 São Carlos, SP, Brazil
c
Center for Research, Technology and Education in Vitreous Materials, Department of Materials Engineering, Federal University of São Carlos, 13565-905 São Carlos,
SP, Brazil

A R T I C L E I N F O A B S T R A C T

Keywords: Li1.5Al0.5Ge1.5(PO4)3 glass-ceramics from NASICON series was synthesized using two different methods: single
LAGP heat treatment and glass sintering with concurrent crystallization of a pressed pellet, which exhibit ionic con­
NASICON, glass-ceramic ductivity of 4.15 × 10− 4 S cm− 1. The electrochemical stability window was investigated by cyclic voltammetry,
Electrochemical window
utilizing a mixture of powder solid electrolyte, carbon, and binder as the working electrode, resulting in the Li/
Solid electrolyte
LAGP/LAGP+C/Au configuration. As a comparison, conventional set-up for cyclic voltammetry was also used,
resulting in the Li/LAGP/Au setup. An electrochemical stability window around 3.5 V was found.

1. Introduction made to study and develop new inorganic solid electrolytes and over­
come challenges that hinder the development of these batteries, for
In order to achieve the transformations needed for a more sustain­ instance high internal resistance [7,8]. Because the solid-solid contact
able way of living, the United Nations created “The 2030 Agenda for between electrodes and electrolytes exhibits high resistance, which is
Sustainable Development” which includes 17 sustainable development detrimental to the battery capacity, several studies have focused on
goals (SDGs) [1]. Renewable energy sources, electric cars, portable reducing the resistance at the interface through different strategies
electronic devices, and reducing our reliance on fossil fuel play an [9–14]. Although the interfacial contact resistance plays an important
important role in future development. In this scenario, energy storage role, the spontaneous chemical reaction between electrodes and elec­
devices are a key technology to make this future reality. trolyte may suggest chemical incompatibility [15–18], i.e., the electrode
Lithium batteries are desired because lithium is the lightest and most potentials may be outside of the electrolyte's stability window, revealing
electropositive metal available, making it a good candidate for anode another variable other than solid-solid contact to add up on the total
material yielding high-density batteries. However, the combination of Li interfacial resistance.
metal and liquid electrolyte cause dendrite growth, which leads to short- Solid-state electrolytes are considered an alternative to replace liquid
circuiting of the cell and poses a serious risk of explosion due to the electrolytes in both lithium and lithium-ion batteries because their non-
flammability of liquid electrolytes [2]. The dendrite growth was cir­ flammable nature protects against explosion if the cell is eventually
cumvented by using carbonaceous anode intercalation compounds short-circuited, thus being inherently safer than organic flammable
because, in this case, lithium is used in its ionic state. This change led to liquid electrolytes. Besides the advantage of not being flammable,
the most successful battery technology available nowadays: lithium-ion another advantage would be the wide electrochemical stability window,
battery [2]. Nevertheless, these batteries still utilize liquid flammable which has been proven by some experimental results [19–22].
electrolytes. Thus, the use of solid electrolytes has been proposed, and However, those experimental results contradict theoretical studies
all-solid-state batteries (ASSB) are now the subject of intense research based on a computational thermodynamic assessment of the phase's
[3–6]. stability as a function of the applied potential [23,24]. As stated by Zhu
To fulfill all-solid-state battery requirements, much effort has been et al. [23], the origin of the outstanding electrochemical stability

* Corresponding author.
E-mail address: vini.zallocco@gmail.com (V.M. Zallocco).

https://doi.org/10.1016/j.ssi.2022.115888
Received 23 October 2021; Received in revised form 16 February 2022; Accepted 22 February 2022
Available online 16 March 2022
0167-2738/© 2022 Elsevier B.V. All rights reserved.
V.M. Zallocco et al. Solid State Ionics 378 (2022) 115888

window observed in many experiments is originated by a kinetic po­ in batches of 15 g. After weighing, the reagents were previously mixed
larization, which requires an overpotential to detect the currents related for 3 h in a rotary mill jar with alumina balls. The homogenized material
to the sluggish rate of solid-state decomposition reactions, and the was transferred to an alumina crucible and calcinated at 700 ◦ C for 1 h to
passivation mechanism that forms a SEI (Solid Electrolyte Interface) release volatiles such as ammonia, water, and carbon dioxide from the
[23]. Thus, to investigate a more realistic electrochemical stability starting materials. The mixture was then melted at 1250 ◦ C for 30 min.
window of solid electrolytes, an improved contact area between the The melt was splat-cooled between two stainless steel plates to increase
solid electrolyte and electronic conductive material is ideal for ampli­ the glass cooling rate. The use of alumina crucible was already reported
fying the redox currents from the electrolyte decomposition [25,26], by other authors [19,37], and no extra Al was detected by ICP-OES and
since the sluggish kinetics of solid-state reaction yields tiny currents. ICP-AES in the LAGP stoichiometry. Glass transition (Tg) and crystalli­
Having said that, Han et al. [25] proposed a modified set-up for cyclic zation (Tx) temperatures were determined by Differential Scanning
voltammetry, in which the solid electrolyte is mixed with carbon pow­ Calorimetry using a DSC-Netzsch 404, at 10 ◦ C min− 1, in a platinum
der, leading to the configuration: Li/Solid Electrolyte/powder mixture crucible with a lid.
of Solid Electrolyte and Carbon/Inert Metal (e.g., Au or Pt). Using this After glass preparation, glass crystallization was achieved through
configuration, they analyzed two well-known solid electrolytes as two different procedures. Li1.5Al0.5Ge1.5(PO4)3 glass-ceramic was first
models: Li10GeP2S12 (LGPS) and Li7La3Zr2O12 (LLZO). prepared by single heat treatment for crystallization of the bulk glass. In
Several solid-state electrolytes are proposed due to their high ionic the second procedure, to improve sample homogeneity, glass-ceramics
conductivity [27–30]. Among them, NASICON (Na Superionic were obtained by sintering and concurrent crystallization of the glass
Conductor) for example, lithium‑aluminum‑titanium-phosphates powder. For the single heat treatment, the bulk glass was crystallized at
(LATP) and lithium‑aluminum‑germanium-phosphates (LAGP) have 800 ◦ C for 3 h in a vertical electric furnace having a precision of ±1 ◦ C.
been widely studied given not only their relatively high ionic conduc­ After the heating treatment, the sample was removed from the furnace
tivity but also their stability in air and water [31–33]. Thus, the and left to cool down at ambient temperature (~25 ◦ C). For the sintering
NASICON-structured Li1.5Al0.5Ge1.5(PO4)3 is seen as a promising solid- and crystallization method, the glass was grounded in agate mortar in
state electrolyte. Moreover, some studies have investigated the elec­ particles between 45 and 65 μm. Then, two different processing condi­
trochemical stability window of LAGP. Table 1 summarizes the value for tions were used. In the first case, the glass powder was isostatically cold
the cathodic and anodic limits of LAGP published in the literature by pressed at 267 MPa and was heated from room temperature to 571 ◦ C at
different authors. 5 K min− 1 for 30 min to allow sintering to occur. After that, the sample
As seen in Table 1, there is a notorious discrepancy between exper­ was heated until 800 ◦ C at 40 K min− 1 for 3 h to allow crystallization.
imental and theoretical results concerning the anodic and cathodic Then the furnace was turned down for cooling. In a second procedure,
limits for the decomposition of LAGP, as well as a disagreement between 10 wt% of dextrin was added to the glass powder, the pellet was then
the experimental results. Given that, this work aims to determine the isostatically cold pressed at 267 MPa and then heated up to 200 ◦ C (5 K
electrochemical stability window for the LAGP using a composite min− 1), held for 2 h to release volatiles from dextrin, then heated to
working electrode, where carbon and solid electrolyte powders are 571 ◦ C (5 K min− 1) for sintering and finally to 800 ◦ C (40 K min− 1) for 3
mixed to increase the contact area between the solid electrolyte and the h to allow crystallization to occur. The furnace was turned down for
electronic conductive material. The conventional cyclic voltammetry, cooling. The sintering temperature of 571 ◦ C was determined by optical
Li/LAGP/Au, was also utilized as a comparison. To optimize the ionic dilatometer (Misura HSM-ODHT 1600–8002, TA instruments), using a
conductivity of the solid electrolyte, we obtained LAGP glass-ceramic in green body of 3.2 mm in diameter and 4.5 mm height pressed at 25 MPa
different ways, by direct crystallization of the glass and by sintering the at a heating rate of 10 ◦ C min− 1.
glassy LAGP powder with and without a binder (dextrin) to obtain The obtained crystalline phase was determined by X-ray diffraction
pellets. (Rigaku Ultima IV X-ray diffractometer) between 10◦ to 80◦ 2θ angle, at
40 kV and 20 mA, using the step scan procedure with an integration time
2. Experimental of 3 s and a step size of 0.015◦ .
The ionic conductivity of the pellet was determined by complex
Li1.5Al0.5Ge1.5(PO4)3 solid electrolyte was prepared through the impedance spectroscopy, (Novocontrol Alpha-A High Performance
glass-ceramic route [37]. The precursor glass was obtained by melting Frequency Analyzer) in a two- point cell in air, at a frequency range of
stoichiometric amounts of Li2CO3 (Dinâmica, 99.00%), GeO2 (Alfa 107–101 Hz at 308 K with voltage amplitude of 300 mV. The pellets were
Aesar, 99.99%), Al2O3 (Aldrich, 99.00%) and NH4H2PO4 (Aldrich, 98%) polished to ensure two parallel surfaces. Before measurements, gold
electrodes were sputtered on both parallel sides using Quorum Q150R
ES equipment.
Table 1
The electrochemical analyses were performed using two different
Values for reduction and oxidation potentials for NASICON solid electrolyte
LAGP.
cells. The first one was similar to that proposed by Han et al. [25], in
which its components are set up as Li/LAGP/LAGP+C/Au. The com­
Cathodic Anodic Analysis type Experimental set-up Reference
posite electrode of LAGP+C/Au was prepared by mixing LAGP powder
Limit (V vs Limit (V vs or theoretical
Li/Li+) Li/Li+) approach and acetylene black (Vulcan XC 72-6P 2800, Cabot) in the mass pro­
portion of 50:50 besides 10 wt% of polyvinylidene fluoride (PVDF,
0 6* Experimental CV – Li/LAGP/SS** [19]
0.8 7* Experimental CV – Li/LAGP/Ag [20]
Aldrich) and a sufficient volume of N-methyl-2-pyrrolidone (NMP, Acros
0.9 4* Experimental CV – Li/LAGP/Au [34] Organic) to form a viscous paste. This latter was then deposited on one
1.5 6* Experimental CV – Li/LAGP/Au [35] side of the LAGP pellets and dried overnight at 120 ◦ C. Afterwards, gold
0.2 5.3 Experimental CV – three [36] was sputtered on the top of the paste to improve electrical contact. Then,
electrodes setup
a Teflon® swagelok-type cell was assembled inside a glove box
1.85 4.9 Experimental CV – composite [26]
working electrode (MBraun, model LabMaster) under an argon atmosphere, with humidity
2.70 4.27 Theoretical FPC*** [23] control and at room temperature (~25 ◦ C). Finally, metallic lithium, the
* LAGP pellet and the composite electrode were placed inside this
Represents the maximum potential applied to the experiment and no
oxidation current was detected, indicating anodic stability at least up to the swagelok-type cell to result in the Li/LAGP/LAGP+C/Au setup. The
indicated potential. CV = Cyclic Voltammetry. second cell was assembled similarly to the first one but using a semi-
**
SS – Stainless Steel. blocking electrode thus resulting in Li/LAGP/Au configuration.
***
FPC = First Principle Calculations. All cyclic voltammetries (CV) were performed using an Autolab

2
V.M. Zallocco et al. Solid State Ionics 378 (2022) 115888

PGSTAT-204 potentiostat/galvanostat (Metrohm Autolab) controlled by


the NOVA 2.1 software. Metallic lithium was used as counter and
reference electrodes. All potential values were always referred to as the
Li/Li+ reference electrode. The working electrodes were varied ac­
cording to the assemblies as described previously. A scan rate of 0.1 mV
s− 1 in various potential ranges was used in the CV measurements.

3. Results and discussion

3.1. Characterization of precursor glass and glass-ceramic

The melted glass is transparent and colorless. The DSC (Differential


Scanning Calorimetry) curve of bulk glass obtained at a heating rate of
10 K min− 1 is shown in Fig 1. As can be seen, the glass transition and
crystallization temperatures are comparable to results previously re­
Fig. 2. Retraction of LAGP glass as a function of the temperature.
ported in the literature [19,37–39], which indicates that the target glass
was successfully synthesized.
One of the methods adopted to obtain the glass-ceramic pellet
involved the sintering of glass powder and concurrent crystallization. In
this case, the sintering temperature was defined using an optical dila­
tometer test. Fig 2 exhibits the sample's retraction as a function of the
temperature. In this figure, the sintering temperature is defined by the
equipment software at the beginning of the retraction curve, i.e., at
around 5% retraction. The beginning of the retraction curve indicates
the beginning of viscous flow. For this reason, it is important to sinter
the pellet at this temperature, to minimize the crystallization effect
while sintering. The inset of Fig 2 shows a picture of the body before and
after sintering.
The next step in the glass-ceramic route is the controlled crystalli­
zation of the precursor glass, to yield a glass-ceramic with a micro­
structure aimed to optimize ionic conductivity, that is, favor grain
growth over nucleation. In fact, increased grain size reduces the inter­
facial area between grain boundaries, thus reducing the intergranular
resistance [37]. As described in the Experimental section, two different
procedures were adopted to obtain the glass-ceramic. The first one
consisted of the direct crystallization of a bulk glass sample (geometric
factor of 0.394 cm− 1) through single heat treatment (SHT). The second
procedure consisted of the sintering and crystallization of a pressed glass
powder pellet. A pellet was prepared with a sintering additive (dextrin)
aiming to improve the mechanical strength of the green body; and
another tablet was prepared without the dextrin for the sake of com­ Fig. 3. Complex impedance of LAGP bulk glass-ceramics, sintered pellet
without dextrin and with dextrin as indicated in the legend. Inset a) shows
parison. The sintered pellet without dextrin exhibited microcracks and
imperfections on the LAGP pellet sintered and crystallized without dextrin and
deformation as shown in inset a) of Fig 3.
b) shows the pellet with optimized processing conditions using dextrin.

The sintered pellet with dextrin (1.020 cm diameter and 0.079 cm


thick), without dextrin (diameter of 1.070 cm and 0.089 cm thick) and
bulk glass crystallized through SHT were then characterized by complex
impedance spectroscopy as shown in Fig 3. It is worth noting that the
maximum order of magnitude (10− 4 S cm− 1) for conductivity was ach­
ieved for this system. This value is also comparable with data published
in the literature [19,38]. As demonstrated, the sintered sample without
sintering additive, which exhibited cracks, has the lowest ionic con­
ductivity of the three samples, leading to ionic conduction equal to 1.76
× 10− 4 S cm-1. This relativity low value is expected as defects such as
pores and cracks are known to decrease the ionic conductivity of ma­
terials. On the other hand, it is worth mentioning that the sample sin­
tered with dextrin showed a higher ionic conductivity at room
temperature (4.15 × 10− 4 S cm-1), as compared to that obtained via SHT
(2.77 × 10− 4 S cm-1).
A sintered pellet was grounded to verify the obtained crystalline
phases. X-ray diffraction spectra of the glass-ceramic is shown in Fig 4.
The peaks match the NASICON LiGe2(PO4)3 - PDF: 80–1924. According
to the Rietveld refinement (Software GSAS-II), a percentage of 3.4% of
the AlPO4 (space group C 2 2 21) phase was formed in the LAGP pellet.
Fig. 1. DSC curve for Li1.5Al0.5Ge1.5(PO4)3 bulk glass at heating rate of 10
Furthermore, a CIF file of x = 0.48 was used for the refinement of
K min− 1.

3
V.M. Zallocco et al. Solid State Ionics 378 (2022) 115888

toward 0 V, the onset of a reduction process at 1.70 V is clearly observed.


This can possibly be associated with the reduction of Ge4+ to Gex+ (0 ≤
x ≤ 3) on the surface of the LAGP+C/Au electrode in contact with the
LAGP solid electrolyte. This reduction process is in agreement with the
works of Hartmann et al. [44] and Chung and Kang [45] who pointed
out that germanium undergoes a reduction when in contact with
lithium, as well as a distribution of oxidation states of germanium across
the interphase depth formed upon contact. Although the present work is
slightly different from those authors, as we are investigating a composite
material (LAGP+C) in contact with LAGP, it is expected that the LAGP of
the electrode becomes unstable as the potential values are decreased.
Therefore, Ge4+ on the surface of the LAGP+C/Au electrode possibly
starts to reduce to Ge3+ at 1.70 V initially forming an LAGP with a
modified stoichiometry [45] before decomposing in other phases. As the
potential decreases beyond 1.70 V, Ge3+ can reduce to even lower
oxidation states leading to the formation of other different germanium
Fig. 4. XRD pattern of grounded pellet of LAGP and Rietveld refinement. compounds, as well as distinct phases containing the other LAGP com­
ponents such as Li, Al and P. According to the literature [23,45,46],
possible LAGP decomposition products can be formed in the potential
Li1+xAlxGe2-x(PO4)3), yielding a theoretical density of 3.42 g cm− 3. The
range of 1.70–0 V are Li4P2O7, GeO2, AlPO4, LiPO3, Li9Al4, Ge, Li–Ge
Archimedes principle was used to calculate the experimental density of
alloys, Li3P, Li2O2 and Li2O. Applying experimental techniques to detect
four different samples, resulting in an average relative density of 80%.
these possible phases as a function of the applied potential is a challenge
Since the porosity of the sample is relatively high, it could be argued
and should be the subject of a future study. From 0 V to higher poten­
that dendrites may grow though the pores [40]. However, dendrites are
tials, the cathodic currents of the reverse scan intersect and cross over
likely to grow under high current densities, i.e., tens of mA cm-2 [41,42].
the cathodic currents of the forward scan only for the second and third
In contrast, low current densities around tens of μA cm-2 were measured
cycles. This behavior is associated to a nucleation process [47], which
in our experiments. In addition, LAGP porosity should not interfere in
occurs only after the reduction processes at ~4.26 V and ~ 2.80 V are
the cathodic and anodic limit as electrochemical window depends
revealed.
markedly on the nature of electrolyte and electrodes [43].
Regarding the anodic potential scans in Fig 5, oxidation processes are
observed at a potential range of 0.82–1.50 V in all three cycles, which
3.2. Electrochemical measurements can be related to phase oxidation such as Ge and Li–Ge alloys. The onset
of another oxidation process is clearly observed at 5.25 V for all cycles.
Fig 5 shows cyclic voltammetry profiles obtained for the LAGP+C/ This process can be attributed to the oxidation of LAGP reduction
Au electrode in the cell assembled with the Li/LAGP/LAGP+C/Au setup. products formed at lower potential values in the previous cathodic scan,
While the first cycle was performed in the potential range of 0–6 V, the such as oxides and phosphates as previously mentioned. At these higher
second and third cycles were carried out increasing the upper limit potential values, the spectrum of phases that can be formed are O2, Ge5O
potential to 7 V. Analyzing the potential scan of the first cycle from OCP (PO4)6, AlPO4, GeP2O7 and GeO2 [23,24]. Once again, the application of
experimental techniques to detect these possible phases as a function of
the applied potential is a challenge and should be the subject of a future
study. Oxidation of LAGP was also detected and will be discussed below

Fig. 5. Cyclic voltammetries obtained at a scan rate of 0.1 mV s− 1 for the first
cell: Li/LAGP/LAGP+C/Au. The inset is a picture of the process step by step,
where the composite electrode was deposited on an LAGP pellet and gold was
sputtered on the top of the composite electrode after drying overnight. (For Fig. 6. Cyclic voltammetries obtained at a scan rate of 0.1 mV s− 1 for the first
interpretation of the references to colour in this figure legend, the reader is cell: Li/LAGP/LAGP+C/Au. The inset represents the first two cycles scanned in
referred to the web version of this article.) a potential range of 0–7 V.

4
V.M. Zallocco et al. Solid State Ionics 378 (2022) 115888

based on the results shown in Fig. 6. in the region of interest, as illustrated by the inset in Fig. 7, reveals the
From the second and third cycles, it is interesting to note that the currents related to the LAGP reduction process starting at ~1.95 V vs Li/
cathodic potential scans in Fig 5 reveal two other reduction processes at Li+. Regarding the anodic potential scans, oxidation processes are also
~4.26 V and ~ 2.80 V, as well as the disappearance of that reduction observed at a potential range of 0.45–1.50 V in all three cycles, which
process observed in the first cycle after 1.70 V. The absence of the can also be related to the oxidation of Ge-containing phases, and the
reduction process at 1.70 V in the second and third cycles, added to the formation of Li–Ge alloys. In the same way as for the Li/LAGP/
fact that there is no oxidation peak related to this reduction process, LAGP+C/Au cell, an onset of oxidation processes is clearly observed at
indicates the irreversibility of the LAGP phase decomposition. As the ~5.25 V for all cycles (see inset in Fig 7). These processes are attributed
other two reduction processes seem to be associated with the oxidation's to the oxidation of the products formed at lower potential values, during
products formed at higher potential values in the previous anodic po­ the reduction of LAGP in the previous cathodic scan, such as oxides and
tential scan, a new set of cyclic voltammetries for the LAGP+C/Au phosphates described before.
electrode was performed starting from OCP toward 2 V, and then In order to detect LAGP oxidation with the cell configuration Li/
sweeping back toward 7 V, as shown in Fig. 6 and its inset for the first LAGP/Au, another set of cyclic voltammetries was performed starting
two cycles. In this figure, the third and fourth cycles were carried out from OCP toward higher potential values, as shown in Fig 8. In the first
from OCP toward 0 V, and then swept back toward 7 V. In the first two cycle, an oxidation process is clearly observed at ~5.42 V vs Li/Li+,
cycles, only an oxidation process starting at ~5.25 V is observed in the which is attributed to the oxidation of oxygen anions (O2− ) present in
anodic potential scan, which can be undoubtedly attributed to the LAGP the crystal lattice of LAGP, as predicted by other authors [23,24].
oxidation process as this scan was performed before decomposing the Regarding the cathodic scan, the onset of a reduction process at ~1.67 V
electrode material at lower potential values, i.e.: 1.70 V, as previously vs Li/Li+ is observed in the first cycle, which can be attributed to the
described. No reduction processes are evident in the subsequent LAGP decomposition, although its potential is shifted when compared
cathodic potential scans of the first two cycles. These observations with that in Fig 7. This shift occurred because the previous anodic scan
indicate the irreversibility of the LAGP oxidation process and, therefore, reached the anodic potential limit of 7 V, thus changing the surface of
can be attributed to the oxidation of oxygen anion in the LAGP crystal the electrode material and, consequently, its potential in the subsequent
lattice as mentioned before. On the other hand, the absence of reduction cathodic scan. In the second cycle, the oxidation process detected at
processes in the first two cycles together with the appearance of two ~5.25 V vs Li/Li+ corroborates previous results, where these currents
reduction processes in the third and fourth cycles, as well as the higher are attributed to the oxidation of oxygen anions present in oxides and
oxidation current observed at ~5.50 V confirm that the reduction pro­ phosphates formed at lower potential values in the previous cathodic
cesses at ~4.26 V and ~ 2.80 V are related to the oxidation products scan. Furthermore, for the second cycle, reduction processes at ~3.00 V
formed at high potential values. However, these reduction processes are vs Li/Li+ can be observed, which have already previously been observed
revealed only after the previous cathodic scan reached the potential and discussed, having been attributed to reductions of oxidation prod­
value of 0 V. ucts formed at high potential values, which are revealed only after the
As a way of comparison, Fig 7 shows cyclic voltammetric profiles previous cathodic scan reached the potential value of 0 V.
obtained for the setup Li/LAGP/Au, in which Au acts as working elec­ In both cell configurations, i.e., Li/LAGP/Au and Li/LAGP/LAGP+C/
trode in the setup assembled as Li/LAGP/Au. The cycles were carried out Au, the reduction and oxidation currents of the LAGP detected are very
in a potential range of 0–7 V, starting from OCP toward 0 V and then small (~1 μA) due to sluggish solid-state reaction kinetics. There was no
sweeping back to 7 V. At first glance, the onset of a reduction process at significant difference between the results of the two configurations. The
~0.80 V is clearly observed which was attributed to the LAGP reduction Li/LAGP/Au and Li/LAGP/LAGP+C/Au cell configurations reveal a
by Feng et al. [20]. However, the inset shows a reduction current arising LAGP reduction process starting at ~1.95 V vs Li/Li+ and ~ 1.70 V vs Li/
at ~1.95 V which is hidden in the main voltammograms. It is worth Li+, respectively. On the other hand, these configurations revealed a
noting that the Li/LAGP/Au cell exhibits, at a potential range of 0–1 V, LAGP oxidation process starting at ~5.42 V vs Li/Li+ and ~ 5.25 V vs Li/
values of a reduction current about twelve times higher than those Li+, respectively. It is interesting to note that the potential values for
registered for the Li/LAGP/LAGP+C/Au cell. However, a magnification anodic and cathodic limits are shifted about 0.16 V toward higher po­
tentials for the Li/LAGP/Au when compared to the Li/LAGP/LAGP+C/
Au cell. However, an equal electrochemical window of ~3.5 V was

Fig. 7. Cyclic voltammetries obtained at a scan rate of 0.1 mV s− 1 for the cell: Fig. 8. Cyclic voltammetries obtained at a scan rate of 0.1 mV s− 1
for the
Li/LAGP/Au. The inset is a zoom of the voltammograms between − 6 and 6 μA. second cell: Li/LAGP/Au.

5
V.M. Zallocco et al. Solid State Ionics 378 (2022) 115888

detected for both cell configurations. In contrast, oxidation processes in present in the crystal lattice of the LAGP phase oxidize yielding gas
the potential range of 0.45–1.50 V vs Li/Li+ related to Li and Li–Ge oxygen [26], and consequently, lithium ions migrate toward the lithium
alloys are better observed for the Li/LAGP/Au cell configuration. electrode (LAGP delithiation) to maintain electrical neutrality. As a
Based on the voltammetric observations described hereinbefore, a result, lithium-free phases are formed at higher potential values as
schematic illustration (Fig 9) was made to clarify the processes occur­ already pointed out previously. Finally, the experimental identification
ring at the interface of LAGP/composite electrode – (LAGP+C/Au) and of phases formed at both low and high potentials could bring insights to
to bring together critical information related to the LAGP solid elec­ the MIC formation mechanism, which is not fully understood, and
trolyte and its application on solid-state batteries. First, the image in­ should be the focus of future studies.
dicates the electrochemical stability window of the solid electrolyte A theoretical study based on a thermodynamic assessment of phases
LAGP, which is in the potential range of 1.70–5.25 V. Since this solid pointed out an electrochemical stability window of LAGP between 2.70
electrolyte has high ionic conductivity (~10− 4 S cm− 1) and low elec­ and 4.27 V [23]. The difference between the experimental and theo­
tronic conductivity (~10− 9 S cm− 1) [20] and the composite electrode retical analysis is most likely originated by kinetic polarization, and thus
LAGP+C/Au has low ionic conductivity, but high electronic conduc­ an overpotential is required in the experimental analysis to detect
tivity due to the high amount of carbon added to it, charge carriers (Li+ decomposition currents due to a sluggish rate of solid-state reactions.
and e− ) should be redistributed and concentrated at the interface under Even though the experimental results indicate an electrochemical sta­
a certain cathodic/anodic polarization. This interface is represented in bility window in the range of 1.70–5.25 V (or 1.95–5.42 V), it is
the zoom image at 1.70 V where germanium ions (Ge4+) start to be important to highlight that the LAGP decomposition is still thermody­
reduced. Only germanium ions at this interface can combine with namically feasible at this potential range, leading to slow deterioration
electrons due to the low electronic conductivity of the LAGP phase, over an extended period of time [23]. Therefore, solid-state batteries
which explains the low current values detected in the cathodic scans of operating in that experimental potential range are highly likely to in­
the voltammetric profiles; germanium ions away from the interface crease their internal resistance over time.
LAGP/LAGP+C, and therefore in the LAGP bulk maintain their oxida­
tion state (Ge4+). 4. Conclusion
As the potential decreases beyond 1.70 V, the LAGP continues
decomposing, yielding an interphase properly called MCI (mix con­ In summary, LAGP glass-ceramic pellets were successfully synthe­
ducting interphase) [44,49], as it has a substantial electronic conduc­ sized throughout sintering and crystallization methods with a better
tivity [44,45] and is unable to passivate the interface. As a result, the ionic conductivity compared to single heat treatment of the bulk glass.
MCI grows continually into the LAGP bulk, thus altering its properties. Sinter-crystallized LAGP exhibited an ionic conductivity of 4.15 × 10− 4
Composition and relative quantities of the formed MCI interphase S cm− 1, with a relative porosity of 80% and 3.4% of AlPO4 as a sec­
should vary as a potential function. For instance, the Li–Ge alloy could ondary phase. The electrochemical measurements resulted in cathodic
only be formed at potential values lower than 0.4 V [46], therefore it is and anodic potential limits of 1.70 V and 5.25 V vs. Li/Li+, respectively,
expected that this phase will not be detected in a potential interval be­ revealing an electrochemical stability window around 3.5 V, which is far
tween 0.40 and 1.70 V, although LAGP has already been decomposed at from that predicted theoretically (1.57 V). This difference can be
this range. attributed to the kinetic polarization caused by the sluggish kinetics of
On the other hand, regarding the potential values higher than the solid-state reaction that requires an overpotential to be detected.
OCP, the LAGP becomes unstable at 5.25 V. Oxygen anions (O2− )

Fig. 9. Schematic illustration of the possible processes occurring at the LAGP/composite electrode interface in the Li/LAGP/LAGP+C/Au setup as a function of the
applied potential. MCI formation indicates that the interface is not stable.

6
V.M. Zallocco et al. Solid State Ionics 378 (2022) 115888

CRediT authorship contribution statement [17] A. Sakuda, A. Hayashi, M. Tatsumisago, Interfacial observation between LiCoO2
electrode and Li2S− P2S5 solid electrolytes of all-solid-state lithium secondary
batteries using transmission electron microscopy, Chem. Mater. 22 (3) (2010)
MVZ: Conceptualization, Data Acquisition, Formal Analysis, Inves­ 949–956.
tigation, Data Curation, Original Draft.JMF: Methodology, Data Acqui­ [18] K. Takada, Progress and prospective of solid-state lithium batteries, Acta Mater. 61
sition, Formal Analysis, Software.NB: Methodology, Formal analysis, (3) (2013) 759–770.
[19] X. Xu, Z. Wen, X. Wu, X. Yang, Z. Gu, Lithium ion-conducting glass–ceramics of Li1.
Resources, Writing - Review & Editing, Supervision.ACMR: Conceptu­ 5Al0. 5Ge1. 5 (PO4) 3–xLi2O (x= 0.0–0.20) with good electrical and
alization, Formal Analysis, Resources, Writing - Review & Editing, electrochemical properties, J. Am. Ceram. Soc. 90 (9) (2007) 2802–2806.
Supervision. [20] J.K. Feng, L. Lu, M.O. Lai, Lithium storage capability of lithium ion conductor Li1.
5Al0. 5Ge1. 5 (PO4) 3, J. Alloys Compd. 501 (2) (2010) 255–258.
[21] N. Kamaya, et al., A lithium superionic conductor, Nat. Mater. 10 (9) (2011)
Declaration of Competing Interest 682–686.
[22] X. Yan, Z. Li, Z. Wen, W. Han, Li/Li7La3Zr2O12/LiFePO4 all-solid-state battery
with ultrathin nanoscale solid electrolyte, J. Phys. Chem. C 121 (3) (2017)
The authors declare that they have no known competing financial 1431–1435.
interests or personal relationships that could have appeared to influence [23] Y. Zhu, X. He, Y. Mo, Origin of outstanding stability in the lithium solid electrolyte
the work reported in this paper. materials: insights from thermodynamic analyses based on first-principles
calculations, ACS Appl. Mater. Interfaces 7 (42) (2015) 23685–23693.
[24] W.D. Richards, L.J. Miara, Y. Wang, J.C. Kim, G. Ceder, Interface stability in solid-
Acknowledgments state batteries, Chem. Mater. 28 (1) (2016) 266–273.
[25] F. Han, Y. Zhu, X. He, Y. Mo, C. Wang, Electrochemical stability of Li10GeP2S12
and Li7La3Zr2O12 solid electrolytes, Adv. Energy Mater. 6 (8) (2016) 1501590.
The authors acknowledge the São Paulo Research Foundation
[26] Y. Benabed, M. Rioux, S. Rousselot, G. Hautier, M. Dollé, Assessing the
(FAPESP) for its financial support, under CEPID Process no.2013/ electrochemical stability window of NASICON-type solid electrolytes, Frontiers in
07793-6. VNZ and NB acknowledges Brazil's National Council for Sci­ Energy Research 9 (2021) 230.
entific and Technological Development (CNPq) for scholarships granted [27] Z. Liu, et al., Anomalous high ionic conductivity of nanoporous β-Li3PS4, J. Am.
Chem. Soc. 135 (3) (2013) 975–978.
under project processes nos. 138854/2019-0 and 309900/2019-0, [28] A. Hayashi, S. Hama, F. Mizuno, K. Tadanaga, T. Minami, M. Tatsumisago,
respectively. Prof. Francisco Carlos Serbena, from State University Ponta Characterization of Li2S–P2S5 glass-ceramics as a solid electrolyte for lithium
Grossa - Brazil, is deeply acknowledged for helpful discussion on the secondary batteries, Solid State Ionics 175 (1–4) (2004) 683–686.
[29] S. Ohta, T. Kobayashi, T. Asaoka, High lithium ionic conductivity in the garnet-
Rietveld analysis. This study was financed in part by the Coordenação de type oxide Li7− X La3 (Zr2− X, NbX) O12 (X= 0–2), J. Power Sources 196 (6)
Aperfeiçoamento de Pessoal de Nível Superior - Brasil (CAPES) - Finance (2011) 3342–3345.
Code 001. [30] M.S. Michael, M.M.E. Jacob, S.R.S. Prabaharan, S. Radhakrishna, Enhanced
lithium ion transport in PEO-based solid polymer electrolytes employing a novel
class of plasticizers, Solid State Ionics 98 (3–4) (1997) 167–174.
References [31] J. Fu, Superionic conductivity of glass-ceramics in the system Li 2O-Al 2O 3-TiO 2-
P 2O 5, Solid State Ionics 96 (3–4) (1997) 195–200.
[1] U. Nations, 2030 Agenda for Sustainable Development. https://sdgs.un.org/goals, [32] M. Monchak, et al., Lithium diffusion pathway in Li1. 3Al0. 3Ti1. 7 (PO4) 3 (LATP)
2015. superionic conductor, Inorg. Chem. 55 (6) (2016) 2941–2945.
[2] J.-M. Tarascon, M. Armand, Issues and challenges facing rechargeable lithium [33] E. Zhao, F. Ma, Y. Guo, Y. Jin, Stable LATP/LAGP double-layer solid electrolyte
batteries, Nature 414 (6861) (2001) 359–367, https://doi.org/10.1038/35104644. prepared via a simple dry-pressing method for solid state lithium ion batteries, RSC
[3] F. Zheng, M. Kotobuki, S. Song, M.O. Lai, L. Lu, Review on solid electrolytes for all- Adv. 6 (95) (2016) 92579–92585.
solid-state lithium-ion batteries, J. Power Sources 389 (2018) 198–213. [34] Y. Cui, M. Rohde, T. Reichmann, M. Mahmoud, C. Ziebert, H. Seifert, Ionic
[4] R. Chen, Q. Li, X. Yu, L. Chen, H. Li, Approaching practically accessible solid-state conductivity and stability of the lithium aluminum germanium phosphate, ECS
batteries: stability issues related to solid electrolytes and interfaces, Chem. Rev. 14 Trans. 72 (Oct. 2016) 139–146, https://doi.org/10.1149/07208.0139ecst.
(120) (2019) 6820–6877. [35] Y. Liu, J. Chen, J. Gao, Preparation and chemical compatibility of lithium
[5] S. Randau, et al., Benchmarking the performance of all-solid-state lithium batteries, aluminum germanium phosphate solid electrolyte, Solid State Ionics 318 (2018)
Nat. Energy 5 (3) (2020) 259–270. 27–34.
[6] A. Manthiram, X. Yu, S. Wang, Lithium battery chemistries enabled by solid-state [36] R.B. Nuernberg, A.C.M. Rodrigues, M. Ribes, A. Pradel, Electrochemical properties
electrolytes, Nature Reviews Materials 2 (4) (2017) 1–16. of NASICON-structured glass-ceramics of the Li1+ xCrx (GeyTi1-y) 2-x (PO4) 3
[7] M. Hou, F. Liang, K. Chen, Y. Dai, D. Xue, Challenges and perspectives of system, Electrochim. Acta 283 (2018) 1835–1844.
NASICON-type solid electrolytes for all-solid-state lithium batteries, [37] A.M. Cruz, E.B. Ferreira, A.C.M. Rodrigues, Controlled crystallization and ionic
Nanotechnology 31 (13) (2020), 132003. conductivity of a nanostructured LiAlGePO4 glass–ceramic, J. Non-Cryst. Solids
[8] T. Famprikis, P. Canepa, J.A. Dawson, M.S. Islam, C. Masquelier, Fundamentals of 355 (45–47) (2009) 2295–2301.
inorganic solid-state electrolytes for batteries, Nat. Mater. 18 (12) (2019) [38] J. Fu, Fast Li+ ion conducting glass-ceramics in the system
1278–1291. Li2O–Al2O3–GeO2–P2O5, Solid State Ionics 104 (3–4) (1997) 191–194.
[9] Y.-C. Jung, S.-M. Lee, J.-H. Choi, S.S. Jang, D.-W. Kim, All solid-state lithium [39] C.J. Leo, B.V.R. Chowdari, G.V.S. Rao, J.-L. Souquet, Lithium conducting glass
batteries assembled with hybrid solid electrolytes, J. Electrochem. Soc. 162 (4) ceramic with Nasicon structure, Mater. Res. Bull. 37 (8) (2002) 1419–1430.
(2015) A704–A710. [40] Y. Ren, Y. Shen, Y. Lin, C.-W. Nan, Direct observation of lithium dendrites inside
[10] W. Zhou, S. Wang, Y. Li, S. Xin, A. Manthiram, J.B. Goodenough, Plating a garnet-type lithium-ion solid electrolyte, Electrochem. Commun. 57 (2015) 27–30.
dendrite-free lithium anode with a polymer/ceramic/polymer sandwich [41] R. Sudo, et al., Interface behavior between garnet-type lithium-conducting solid
electrolyte, J. Am. Chem. Soc. 138 (30) (Aug. 2016) 9385–9388, https://doi.org/ electrolyte and lithium metal, Solid State Ionics 262 (2014) 151–154.
10.1021/jacs.6b05341. [42] K. Ishiguro, et al., Stability of Nb-doped cubic Li7La3Zr2O12 with lithium metal,
[11] Z. Zhang, et al., An advanced construction strategy of all-solid-state lithium J. Electrochem. Soc. 160 (10) (2013) A1690.
batteries with excellent interfacial compatibility and ultralong cycle life, J. Mater. [43] K. Oldham, J. Myland, A. Bond, Electrochemical Science and Technology:
Chem. A 5 (32) (2017) 16984–16993. Fundamentals and Applications, John Wiley & Sons, 2011.
[12] S. Yu, A. Mertens, H. Tempel, R. Schierholz, H. Kungl, R.-A. Eichel, Monolithic all- [44] P. Hartmann, et al., Degradation of NASICON-type materials in contact with
phosphate solid-state lithium-ion battery with improved interfacial compatibility, Lithium metal: formation of mixed conducting interphases (MCI) on solid
ACS Appl. Mater. Interfaces 10 (26) (Jul. 2018) 22264–22277, https://doi.org/ electrolytes, J. Phys. Chem. C 117 (41) (Oct. 2013) 21064–21074, https://doi.org/
10.1021/acsami.8b05902. 10.1021/jp4051275.
[13] E. Kobayashi, L.S. Plashnitsa, T. Doi, S. Okada, J.I. Yamaki, Electrochemical [45] H. Chung, B. Kang, Mechanical and thermal failure induced by contact between a
properties of Li symmetric solid-state cell with NASICON-type solid electrolyte and Li1. 5Al0. 5Ge1. 5 (PO4) 3 solid electrolyte and Li metal in an all solid-state Li cell,
electrodes, Electrochem. Commun. (2010), https://doi.org/10.1016/j. Chem. Mater. 29 (20) (2017) 8611–8619.
elecom.2010.04.014. [46] J.S. Peña, I. Sandu, O. Joubert, F.S. Pascual, C.O. Areán, T. Brousse,
[14] J.P. Robinson, P.D. Kichambare, J.L. Deiner, R. Miller, M.A. Rottmayer, G. Electrochemical reaction between lithium and β-quartz GeO2, Electrochemical and
M. Koenig Jr., High temperature electrode-electrolyte interface formation between Solid State Letters 7 (9) (2004) A278.
LiMn1. 5Ni0. 5O4 and Li1. 4Al0. 4Ge1. 6 (PO 4) 3, J. Am. Ceram. Soc. 101 (3) [47] S.B. Emery, J.L. Hubbley, D. Roy, Voltammetric and amperometric analyses of
(2018) 1087–1094. electrochemical nucleation: electrodeposition of copper on nickel and tantalum,
[15] T. Kobayashi, A. Yamada, R. Kanno, Interfacial reactions at electrode/electrolyte J. Electroanal. Chem. 568 (2004) 121–133.
boundary in all solid-state lithium battery using inorganic solid electrolyte, thio- [49] S. Wenzel, T. Leichtweiss, D. Krüger, J. Sann, J. Janek, Interphase formation on
LISICON, Electrochim. Acta 53 (15) (2008) 5045–5050. lithium solid electrolytes—an in situ approach to study interfacial reactions by
[16] K.H. Kim, et al., Characterization of the interface between LiCoO2 and photoelectron spectroscopy, Solid State Ionics 278 (2015) 98–105, https://doi.org/
Li7La3Zr2O12 in an all-solid-state rechargeable lithium battery, J. Power Sources 10.1016/j.ssi.2015.06.001.
196 (2) (2011) 764–767.

You might also like