Uploaded by Duke Shih

acsenergylett.1c01573

advertisement
Downloaded via 72.36.94.221 on February 26, 2022 at 17:50:53 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Improving Contact Impedance via
Electrochemical Pulses Applied to Lithium−
Solid Electrolyte Interface in Solid-State
Batteries
Anand Parejiya, Ruhul Amin,* Marm B. Dixit, Rachid Essehli, Charl J. Jafta, David L. Wood, III,
and Ilias Belharouak*
Cite This: ACS Energy Lett. 2021, 6, 3669−3675
ACCESS
Metrics & More
Read Online
Article Recommendations
sı Supporting Information
*
ABSTRACT: Stabilizing interfaces in solid-state batteries (SSBs) is
crucial for development of high energy density batteries. In this work, we
report a facile electrochemical protocol to improve the interfacial
impedance and contact at the interface of Li | Li6.25Al0.25La3Zr2O12
(LALZO). Application of short duration, high-voltage pulses to poorly
formed interfaces leads to lower contact impedance. It is found that the
local high current density that results from these pulses at the vicinity of
the interfacial pores can lead to a better contact between Li and LALZO
because of local Joule heating, as supported by theoretical simulations.
The pulse technique, which has also been applied to a Li |
Li6.4La3Zr1.4Ta0.6O12 (LLZTO) | LiNi0.6Mn0.2Co0.2O2 (NMC622) cell,
results in remarkable reduction of the charge-transfer resistance. Ex situ
characterizations, which include X-ray photoelectron spectroscopy and
scanning electron microscopy techniques, reveal that there is no detrimental effects of the pulse on cathode and solid
electrolyte bulks and interfaces. This electrochemical pulse technique sheds light on a facile, nondestructive method that has
the potential to significantly improve the interfacial contacts in a solid-state battery configuration.
spatial kinetics as well as low stack pressure.20,23 Nonuniform
interfacial contact arising from the voids leads to the formation
of degradation pathways due to high contact resistance which
results in flux/stress hot-spots, Li-filament growth, and
cracking.20−22,24
A few studies have introduced anode interlayers to lower the
contact resistance at the Li | SE interface and improve the
Coulombic efficiency of Li electrodeposition/dissolution.25−28
Recently, stack pressure was also identified as a parameter that
can enable high rate and stable Li cycling. High stack pressure
enables intimate contact by promoting creep-assisted transport
of Li metal to the interface.23 However, theoretical and
experimental results claim that higher stack pressure can lead
T
he energy density and safety advantages of an all-solidstate battery (SSB) can be enabled by implementation
of Li metal anode (−3.04 V vs SHE, 3860 mAh g−1)
and elimination of flammable solvents present in lithium-ion
batteries.1 Commercial application of SSB is facing several
challenges that can be broadly divided into two main
categories: (1) interfacial resistances and stability of battery
components and (2) electrode processing and cell manufacturing.2,3 Indeed, the interfacial resistance, chemical stability of Li
against solid electrolytes (SE), processability, and mechanical
robustness of SE are among the key manufacturing and
integration challenges.4−12 Electrochemical stability, interphase
formation, dendrite formation and propagation are the other
challenges for battery operation and management.9,13−16
Moreover, engineering and maintaining an intimate contact
at the electrode | electrolyte interface is a key requirement for
high-rate performance SSBs.17 However, the generation of
voids at the SE | Li-metal interface can lead to chemomechanical degradation which impacts battery lifetime as
well as safety.18−22 Interfacial voids are generated because of
several factors including improper assembly, heterogeneous
© 2021 The Authors. Published by
American Chemical Society
Received: July 27, 2021
Accepted: September 20, 2021
Published: September 23, 2021
3669
https://doi.org/10.1021/acsenergylett.1c01573
ACS Energy Lett. 2021, 6, 3669−3675
Letter
http://pubs.acs.org/journal/aelccp
ACS Energy Letters
http://pubs.acs.org/journal/aelccp
Letter
Figure 1. (a) Schematic diagram depicting the effect of voltage pulse on stabilizing the Li | LALZO interface. (b) EIS spectra for symmetric
Li | LALZO | Li cell obtained before and after applying a voltage pulse. Here, the voltage pulse refers to application of high current (10 mA
cm−2) with cutoff voltages set at 10 V and −10 V for 20 cycles. (c) Polarization curves for symmetric Li | LALZO | Li cell showing initial
cycles, voltage pulse, and cycling after voltage pulse. The cell is run at room temperature.
galvanostatic cycling at 40 μA cm−2, the average cell
polarization was 0.12 V (Figure 1c). An increase of the
average polarization after 30 min of plating and stripping cycles
was found to be around 3 mV. The increase of polarization on
cycling can arise from interfacial reaction or void formation at
the interface.34 Because LALZO is relatively stable against Li
metal, void formation is likely occurring on cycling under the
applied current density. After five cycles, 20 voltage pulses
were applied on the cell by the application of high current (10
mA cm−2) with cutoff voltages set at 10 V and −10 V. In this
case, the cutoff voltages have been reached immediately on the
application of the high current pulse within 100 ms. EIS
spectra after the pulse show significant reduction in interfacial
resistance (Figure 1b). As seen from the EIS spectra, only the
interfacial resistance of the cell decreases while the bulk
resistance remains constant. No change in bulk resistance
indicates the absence of dendrite formation or propagation in
the bulk of the solid electrolyte.35 The decrease in interfacial
resistance suggests improvement in interfacial contact at the Li
metal | SE interface. Interfacial resistance after 20 pulses was
1.90 kΩ·cm2, which was 26% lower compared to that of the
pristine cell (Figure 1b). Presence of voids at the Li | SE
interface diminishes the electrochemically active area because
of the absence of ionic and electronic pathways leading to
regions with higher local current densities (current focusing).
Application of the pulse causes high lithium deposition at the
pore edges leading to the improvement of the contact
impedance (Figure 1a). Subsequent galvanostatic cycling
performed under the initial current density of 40 μA cm−2
showed that the average cell polarization decreased from 0.12
to 0.09 V (Figure 1c). Improved EIS spectra as well as lower
to cell failures due to facile lithium metal propagation through
solid electrolyte pores.29,30 It should be noted that most
strategies aimed at stabilizing the lithium anode interface
involve additional processing steps, introduction of new
materials, or altered operating conditions, which eventually
increases system cost and complexity.
Overall, physical contact loss at the Li metal | SE interface is
a major technology challenge in the operation of SSBs. Despite
the abundance of explanations for the physical contact loss
supported by experiments and theory, the solution strategies
reported in the literature are scarce.29,31 Therefore, preventive
measures or solution strategies are required to ensure long
cycle life and high-rate performance of SSBs. This work reports
an electrochemical strategy for improving the contact
resistance of Li metal | SE interface in SSBs. Results showcase
large improvement in interfacial resistance (15−58%) within a
short time pulse of (0.1−0.5 s) at nominal pressures (∼1 kPa).
The proposed electrochemical stabilization pathway can
function as a formation step for SSB cells as well as a
control/management protocol for extending cycle life of SSBs.
A symmetric Li | LALZO | Li cell was assembled inside a
glovebox using the Swagelok configuration. Detailed experimental procedures are provided in the Supporting Information. To replicate inhomogeneous Li morphology and void
formation occurring during long-term cycling of the SSB, the
cell was conditioned such that the interface was nonoptimal.
Imperfect physical contact at the interface was validated by
electrochemical impedance spectroscopy (EIS) measurement
at room temperature, which showed an interfacial resistance of
2.58 kΩ·cm2. In comparison, typical impedance values for
pristine cells with good contact between LALZO and Li are
usually in the range of 10−100s Ω·cm2.23,32,33 During
3670
https://doi.org/10.1021/acsenergylett.1c01573
ACS Energy Lett. 2021, 6, 3669−3675
ACS Energy Letters
http://pubs.acs.org/journal/aelccp
Letter
Figure 2. (a) EIS spectra for symmetric Li | LALZO | Li cell obtained before and after multiple voltage pulses. (c) Polarization curves for
symmetric Li | LALZO | Li cell showing initial cycles and cycling profiles with multiple voltage pulses. Here, the voltage pulse refers to
application of high current (10 mA cm−2) with cutoff voltages set at 10 V and −10 V for 20 cycles. The cell is run at room temperature.
Figure 3. (a) Schematic diagram showing current flow in the vicinity of a pore at the Li | LALZO interface. (b) Temperature change in
lithium in the vicinity of a pore as a function of the pore size and the shell size considered due to Joule heating. (c) COMSOL simulation
showing current density increase at the vicinity of the pore region.
Local current density in the vicinity of the pore is expected
to be higher because of the absence of an electronic pathway
through the voids (Figures 3a,c and S1). This redistribution of
the current can lead to a local rise of the temperature of the
lithium metal due to Joule heating upon application of higher
current pulses. We carried out theoretical calculations and
computational fluid dynamics (CFD) simulations to assess the
impact of interfacial pores on current density and local heating.
A cylindrical geometry in the vicinity of the pore is considered
where the diameter of the pore is a variable and the cylinder
diameter around it is specified as a function of a given shell
thickness (Figure 3a). The height of the cylinder in all cases is
considered equal to the diameter of the cylinder. CFD
simulations showcase that the current density in the vicinity
of the pore increases 4-fold compared to the applied current
density. High local current density can lead to preferential
deposition in these regions during the plating cycle, which can
lead to the filling of the pores and improvement of the
interface. The calculations for local temperature increase
assume that the electrical power associated with the local
current density in the vicinity of the pore (J2R, where J is the
local electric current density and R is the area-specific
resistance) converts to heat. This heat is assumed to dissipate
in the Li metal within the cylindrical element considered, and
the local temperature rise is estimated using standard heat
polarization unveil the effectiveness of the voltage pulse
protocol in improving the interfacial resistance.
Thereafter, the effect of multiple pulse cycles on cell
performance was investigated. Individual pulse cycles refer to
application of 20 voltage pulses (10 mA cm−2 with cutoff
voltages set at 10 V and −10 V) as described above. The
interfacial resistance of the pristine cell was 2.89 kΩ·cm2 which
decreased to 2.30 Ω·cm2 after the first voltage pulse cycle,
showcasing 20% improvement in interfacial contact (Figure
2a). After a period of 20 h of galvanostatic cycling at the
nominal current density (40 μA cm−2), 20 identical pulses
were applied multiple times followed by 20 h of galvanostatic
cycling after each pulse (Figure 2b). The interfacial resistance
kept on decreasing without any change in bulk resistance,
indicating an improvement of the contact at the anode | SE
interface. The unchanged bulk resistance indicates that there is
an absence of dendrite formation with the application of the
voltage pulse. After 5 pulse cycles, interface resistance
decreased to 1.78 kΩ·cm2 compared to the initial resistance
of 2.89 kΩ·cm2 (−38.4%). While further optimization of pulse
duration, sequence, and magnitude is required to maximize the
interfacial contact, this work highlights the applicability of this
strategy to improve contact impedance of the Li | LALZO
interface.
3671
https://doi.org/10.1021/acsenergylett.1c01573
ACS Energy Lett. 2021, 6, 3669−3675
ACS Energy Letters
http://pubs.acs.org/journal/aelccp
Letter
Figure 4. (a) Series and charge-transfer resistance for symmetric Li | LALZO | Li cell obtained before and after applying a voltage pulse. (b)
Polarization curves for symmetric Li | LALZO | Li cell showing initial cycles, voltage pulse, and cycling after voltage pulse. (c) Series and
charge-transfer resistance for symmetric Li | LALZO | Li cell obtained before and after multiple voltage pulses as a function of cumulative
charge passed through the cell. (d) Polarization curves for symmetric Li | LALZO | Li cell showing initial cycles and cycling profiles with
multiple voltage pulses. Here, the voltage pulse refers to application of high current (50 mA cm−2) with cutoff voltages set at 10 V and −10 V
for 20 cycles. The cells are run at 60 °C.
However, it is necessary to investigate the impact of voltage
pulse on a cell with intimate interfacial contact. Thus, a similar
symmetric Li | LALZO | Li cell was assembled, conditioned,
and operated at 60 °C to improve cell impedance. EIS spectra
reveal a very low interfacial resistances of 135 Ω·cm2
comparable to those reported for Li | LLZO | Li cells with
intimate interfacial contact and high critical current density
(Figure 4a).23,32,33 In this case, the first few galvanostatic
plating and striping cycles were performed at 80 μA cm−2,
which resulted in 0.02 V average cell polarization (Figure 4b).
As mentioned above, a cell polarization profile reveals
information on the effectiveness of the electrodeposition of
Li ions at the electrode. Indeed, cell polarization increased by
10 μV only after 30 min of cycling and showed very flat, stable
profiles, which suggests that the electrodeposition and
dissolution of Li metal were highly reversible with high
Coulombic efficiency.
A voltage pulse cycle was applied (50 mA cm−2 with 10 V
and −10 V cut off voltage, 20 pulses) on the symmetric cell.
After a pulse cycle, no change in the EIS spectra and cell
polarization were observed (Figures 4a,b and S2). Galvanostatic cycling at the nominal current density after the pulse
application revealed no degradation of the cell (Figure 4b).
Another cell was tested with multiple pulses (Figures 4c,d and
S3) which showed similar response with no significant change
in EIS or cell polarization even after 5 pulse cycles (Figure
capacity formulation. Results show that the local temperature
rise is within the range of 10−300 °C depending upon the
cylinder and pore geometry (Figure 3b). Larger pores result in
typically higher-temperature changes because of the larger
effective current density in the vicinity of these pores. In
addition, increasing the cylinder diameter reduces the local
heating because of an increase in the amount of material for
heat dissipation. Higher local temperature of Li essentially
implies an increased homologous temperature resulting in
higher creep. Improved creep flow of Li can aid in reduction of
the interfacial resistance.23 While the assumptions of heat
generation and dissipation are limiting, these results do provide
an estimate for local temperature rise through Joule heating.
Such a temperature increase can significantly affect lithium
flow behavior and facilitate stabilization of the interface.
Combined, the preferential deposition and improved creep
flow lead to electrochemical stabilization of the anode | SE
interface. It should be noted that the local heating of Li metal
was leveraged in previous studies to heal dendrites in
conventional Li-ion batteries wherein Li symmetric cells were
cycled at an excess of 9 mA cm−2 to remove dendrites.36 In
those studies, local heating and subsequent fast surface
diffusion helped in healing the failed cells. The interface
stabilization achieved in this work functions under similar
mechanism for improved Li flow and surface diffusion through
local heating under high-current pulses.
3672
https://doi.org/10.1021/acsenergylett.1c01573
ACS Energy Lett. 2021, 6, 3669−3675
ACS Energy Letters
http://pubs.acs.org/journal/aelccp
Letter
Figure 5. (a) EIS spectra of a full cell before and after voltage pulsing cycle at room temperature. Here, the voltage pulse refers to application
of high current (40 mA cm−2) with cutoff voltages set at 10 V and −10 V for 10 cycles. (b) Galvanostatic charge−discharge profile of full cell
at 70 °C after pulsing cycle. SEM micrographs of (c) NMC622 cathode secondary particle and (d) cathode−SE interface and anode−SE
interface after pulsing cycle. Scale bar in panel c is 2 μm, and for panel d it is 50 μm. Comparison of XPS spectra of (e) pristine cathode (red)
and cathode after pulsing cycle (black) for C 1s, (f) LLZTO before and after application of pulse for C 1s. For pulsed SE, XPS profiles for
anode and cathode side are visualized.
formation/propagation in the solid electrolyte. Reduction of
charge-transfer resistances indicate the formation of intimate
contact and absence of insulating decomposition products at
the cathode and anode | solid electrolyte interfaces. The same
cell was placed inside an environmental chamber at 70 °C for
12 h before cycling. Galvanostatic charge−discharge was
carried out between 3 and 4.3 V at the C/20 rate at 70 °C.
The charge profile plateaus at 3.6 V, which is typical for
NMC622 cathode, and the first charge and discharge capacities
of ∼206 mAh/g and ∼155 mAh/g were obtained (Figure 5b).
The cell was able to operate well after the application of 20
pulse cycles (200 individual pulses) with no indication of
detrimental impacts of the pulse. To further assess the chemomechanical impact of pulse application, another identical cell
was prepared, and a similar pulse protocol was followed on the
cell at room temperature. Post-mortem analysis was carried out
using SEM (Figures 5c,d and S5) and XPS (Figures 5e,f, S6,
and S7) on the cell components. SEM of cathode and solid
electrolyte was performed to assess mechanical degradation.
The secondary particles show no cracking due to possible
stresses generated via electrochemical pulse, as seen from
Figure 5c. Previous studies have reported that interphases in
solid electrolyte show up on SEM as regions of disparate
contrast.21 However, no interphase was apparent from the
SEM analysis of SE at both the anode and cathode interfaces
(Figure 5d). Further evidence of absence of interphase
formation was obtained by ex situ XPS studies carried out on
pristine and pulsed full cell components. Typical decomposition products at the cathode | SE interphase can be broadly
classified as organic (C-containing species) and inorganic (Libased insulating compounds). C 1s, O 1s, F 1s, and Li 1s
spectra are visualized for the pristine and pulsed cathode as
well as the LLZTO material at the cathode and Li interfaces. C
1s spectra for the cathode contain the nominally expected
surface species C−C (284.5 eV), C−X (281.9 eV), C−O (286
eV), O−C=O (289.7 eV), and C−Fx (292.6 eV), which are
consistent across both the pristine and the electrochemically
4c,d). To better compare the EIS profiles, an equivalent circuit
fitting was carried out to estimate the series and charge-transfer
resistances for individual spectra. These values are plotted
against the cumulative charge passed through the cell
components and show no obvious changes upon the pulse
application (Figure 4c). In addition, the long-term stability of
the cell after application of the pulse was evaluated at room
temperature. No sign of filament formation or growth had
been noticed after 700 h of cycling of the Li | LALZO | Li cell
upon passing 25 mAh cm−2 charge, indicating an absence of
long-term detrimental impact of the pulse application (Figure
S4a). These experiments suggest that while the voltage pulse
technique can address the interface issue for a nonideal
interface, it shows no impact on a cell with intimate anode | SE
interface contact. These results are important to ensure that
the technique can be used eventually to address interfacial
resistance issues in defective cells without altering the cells
with intimate contacts under standard cell pressures.
A Li | LLZTO | NMC622-composite cell was assembled
inside a glovebox using the Swagelok configuration. Detailed
experimental procedures are provided in the Supporting
Information. Room-temperature EIS spectra of the asassembled cell showed two semicircles with two types of
diffusion tails (Figure 5a). The two semicircles can be assigned
to two interfacial regions in the cell: anode−SE and cathode−
SE. The two distinct diffusion tails can be attributed to two
types of diffusion in solid electrolyte and cathode. The total
cell resistance at room temperature was 1.2 kΩ·cm2. A voltage
pulse cycle (10 pulses) was applied on the cell at the same time
by application of high current (40 mA cm−2) with cutoff
voltages set at 10 V and −10 V. A total of 20 pulsing cycles
were carried out with the cell allowed to equilibrate at OCV for
10 min after every pulse cycle (Figure S4b,c). After this pulse
cycle was completed, total cell resistance was reduced to 0.5
kΩ·cm2, which is a 58% reduction in the total resistance from
the as-assembled cell. As seen from the EIS spectra, the bulk
resistance remains constant, indicating the absence of dendrite
3673
https://doi.org/10.1021/acsenergylett.1c01573
ACS Energy Lett. 2021, 6, 3669−3675
ACS Energy Letters
http://pubs.acs.org/journal/aelccp
Marm B. Dixit − Electrification & Energy Infrastructure
Division, Oak Ridge National Laboratory, Oak Ridge,
Tennessee 37831, United States; orcid.org/0000-00029599-9288
Rachid Essehli − Electrification & Energy Infrastructure
Division, Oak Ridge National Laboratory, Oak Ridge,
Tennessee 37831, United States
Charl J. Jafta − Electrification & Energy Infrastructure
Division, Oak Ridge National Laboratory, Oak Ridge,
Tennessee 37831, United States; orcid.org/0000-00029773-6799
David L. Wood, III − Bredesen Center for Interdisciplinary
Research and Graduate Education, University of Tennessee,
Knoxville, Tennessee 37996, United States; orcid.org/
0000-0002-2471-4214
cycled cathode material (Figure 5e).37,38 F 1s spectra show the
presence of C−F, LFSI−, and LiF species for the pristine and
cycled cathode samples (Figure S6). Similar profiles are also
observed for Ni 1s, Ni 2p, O 1s, and Li 1s spectra (Figure S6).
Similar results are also observed for the LLZTO, wherein the
spectra for pristine LLZTO and the electrochemically cycled
LLZTO (anode and cathode side) show similar profiles
(Figures 5f and S7). Overall, the XPS results strongly support
the assertion that the application of pulse does not lead to the
formation of decomposition products. These results also
suggest that there was no material degradation upon high
current pulse application. Therefore, the high current pulse
strategy is an effective method to improve interfacial contacts
in a solid-state battery.
Several other strategies have been explored in the literature
to improve interfacial contact at Li anode | SE. Most of these
strategies come with a penalty of additional processing steps,
materials, or operating conditions that limit the technoeconomic feasibility of SSBs. Herein, we introduced an electrochemical protocol applied to Li | LALZO | Li symmetric and
NMC | LLZTO | Li cells that show clear improvement of
interfacial contact upon application of very short duration high
current pulses. This electrochemical pulse method can be
leveraged as an integration as well as a management step to
ensure seamless contact between Li-metal anode, cathode, and
SE. Ex situ characterizations demonstrate that there was no
detrimental chemo-mechanical impact on both bulks and
interfaces of the cell components upon the application of the
high current pulse protocol. The latter is found to help ensure
stable interfaces on cycling which is key to achieving highperforming SSBs. This new finding can have applications in
direct cell integration and degradation management for SSBs.
■
Complete contact information is available at:
https://pubs.acs.org/10.1021/acsenergylett.1c01573
Notes
The authors declare no competing financial interest.
■
ACKNOWLEDGMENTS
This research at Oak Ridge National Laboratory, managed by
UT Battelle, LLC, for the US Department of Energy (DOE)
under contract DE-AC05-00OR22725, was sponsored by
Laboratory Directed Research and Development (LDRD)
Program at Oak Ridge National Laboratory, and the Office of
Energy Efficiency and Renewable Energy (EERE) Vehicle
Technologies Office (VTO) (Deputy Director: David Howell)
Applied Battery Research subprogram (Program Manager:
Peter Faguy).
■
ASSOCIATED CONTENT
* Supporting Information
sı
REFERENCES
(1) Cheng, X.-B.; Zhang, R.; Zhao, C.-Z.; Zhang, Q. Toward Safe
Lithium Metal Anode in Rechargeable Batteries: A Review. Chem. Rev.
2017, 117 (15), 10403−10473.
(2) Schnell, J.; Günther, T.; Knoche, T.; Vieider, C.; Köhler, L.; Just,
A.; Keller, M.; Passerini, S.; Reinhart, G. All-Solid-State Lithium-Ion
and Lithium Metal Batteries − Paving the Way to Large-Scale
Production. J. Power Sources 2018, 382, 160−175.
(3) Kerman, K.; Luntz, A.; Viswanathan, V.; Chiang, Y.-M.; Chen, Z.
ReviewPractical Challenges Hindering the Development of Solid
State Li Ion Batteries. J. Electrochem. Soc. 2017, 164 (7), A1731−
A1744.
(4) Balaish, M.; Gonzalez-Rosillo, J. C.; Kim, K. J.; Zhu, Y.; Hood, Z.
D.; Rupp, J. L. M. Processing Thin but Robust Electrolytes for SolidState Batteries. Nat. Energy 2021, 6, 227.
(5) Xia, S.; Wu, X.; Zhang, Z.; Cui, Y.; Liu, W. Practical Challenges
and Future Perspectives of All-Solid-State Lithium-Metal Batteries.
Chem. 2019, 5 (4), 753−785.
(6) Wang, L.; Li, J.; Lu, G.; Li, W.; Tao, Q.; Shi, C.; Jin, H.; Chen,
G.; Wang, S. Fundamentals of Electrolytes for Solid-State Batteries:
Challenges and Perspectives. Front. Mater. 2020, 7 (July), 1−5.
(7) Chen, A.; Qu, C.; Shi, Y.; Shi, F. Manufacturing Strategies for
Solid Electrolyte in Batteries. Front. Energy Res. 2020, 8, 571440.
(8) Tian, Y.; Shi, T.; Richards, W. D.; Li, J.; Kim, J. C.; Bo, S. H.;
Ceder, G. Compatibility Issues between Electrodes and Electrolytes in
Solid-State Batteries. Energy Environ. Sci. 2017, 10 (5), 1150−1166.
(9) Zhu, Y.; He, X.; Mo, Y. First Principles Study on Electrochemical
and Chemical Stability of Solid Electrolyte-Electrode Interfaces in AllSolid-State Li-Ion Batteries. J. Mater. Chem. A 2016, 4 (9), 3253−
3266.
(10) Riphaus, N.; Stiaszny, B.; Beyer, H.; Indris, S.; Gasteiger, H. A.;
Sedlmaier, S. J. Editors’ ChoiceUnderstanding Chemical Stability
The Supporting Information is available free of charge at
https://pubs.acs.org/doi/10.1021/acsenergylett.1c01573.
■
Letter
Experimental methods and theoretical modeling framework and addition characterization results (PDF)
AUTHOR INFORMATION
Corresponding Authors
Ruhul Amin − Electrification & Energy Infrastructure
Division, Oak Ridge National Laboratory, Oak Ridge,
Tennessee 37831, United States; orcid.org/0000-00020054-3510; Email: aminr@ornl.gov
Ilias Belharouak − Electrification & Energy Infrastructure
Division, Oak Ridge National Laboratory, Oak Ridge,
Tennessee 37831, United States; Bredesen Center for
Interdisciplinary Research and Graduate Education,
University of Tennessee, Knoxville, Tennessee 37996, United
States; orcid.org/0000-0002-3985-0278; Phone: +1
(682) 407 3459; Email: belharouaki@ornl.gov; Fax: +1
(865) 576 7342
Authors
Anand Parejiya − Electrification & Energy Infrastructure
Division, Oak Ridge National Laboratory, Oak Ridge,
Tennessee 37831, United States; Bredesen Center for
Interdisciplinary Research and Graduate Education,
University of Tennessee, Knoxville, Tennessee 37996, United
States
3674
https://doi.org/10.1021/acsenergylett.1c01573
ACS Energy Lett. 2021, 6, 3669−3675
ACS Energy Letters
http://pubs.acs.org/journal/aelccp
Letter
lyte and Li-Metal Anode by a Germanium Layer. Adv. Mater. 2017, 29
(22), 1606042.
(28) Zhou, W.; Wang, S.; Li, Y.; Xin, S.; Manthiram, A.;
Goodenough, J. B. Plating a Dendrite-Free Lithium Anode with a
Polymer/Ceramic/Polymer Sandwich Electrolyte. J. Am. Chem. Soc.
2016, 138 (30), 9385−9388.
(29) Tu, Q.; Barroso-Luque, L.; Shi, T.; Ceder, G. Electrodeposition
and Mechanical Stability at Lithium-Solid Electrolyte Interface during
Plating in Solid-State Batteries Electrodeposition and Mechanical
Stability at Lithium-Solid Electrolyte Interface during Plating in SolidState Batteries. Cell Reports Phys. Sci. 2020, 1 (7), 100106.
(30) Doux, J. M.; Nguyen, H.; Tan, D. H. S.; Banerjee, A.; Wang, X.;
Wu, E. A.; Jo, C.; Yang, H.; Meng, Y. S. Stack Pressure Considerations
for Room-Temperature All-Solid-State Lithium Metal Batteries. Adv.
Energy Mater. 2020, 10 (1), 1903253.
(31) Koerver, R.; Zhang, W.; De Biasi, L.; Schweidler, S.;
Kondrakov, A. O.; Kolling, S.; Brezesinski, T.; Hartmann, P.; Zeier,
W. G.; Janek, J. Chemo-Mechanical Expansion of Lithium Electrode
Materials-on the Route to Mechanically Optimized All-Solid-State
Batteries. Energy Environ. Sci. 2018, 11 (8), 2142−2158.
(32) Sharafi, A.; Kazyak, E.; Davis, A. L.; Yu, S.; Thompson, T.;
Siegel, D. J.; Dasgupta, N. P.; Sakamoto, J. Surface Chemistry
Mechanism of Ultra-Low Interfacial Resistance in the Solid-State
Electrolyte Li7La3Zr2O12. Chem. Mater. 2017, 29 (18), 7961−7968.
(33) Connell, J. G.; Fuchs, T.; Hartmann, H.; Krauskopf, T.; Zhu,
Y.; Sann, J.; Garcia-Mendez, R.; Sakamoto, J.; Tepavcevic, S.; Janek, J.
Kinetic versus Thermodynamic Stability of LLZO in Contact with
Lithium Metal. Chem. Mater. 2020, 32 (23), 10207−10215.
(34) Chen, K. H.; Wood, K. N.; Kazyak, E.; Lepage, W. S.; Davis, A.
L.; Sanchez, A. J.; Dasgupta, N. P. Dead Lithium: Mass Transport
Effects on Voltage, Capacity, and Failure of Lithium Metal Anodes. J.
Mater. Chem. A 2017, 5 (23), 11671−11681.
(35) Parejiya, A.; Amin, R.; Essehli, R.; Wood, D. L.; Belharouak, I.
Electrochemical Healing of Dendrites in Garnet-Based Solid Electrolytes. ACS Energy Lett. 2020, 5 (11), 3368−3373.
(36) Li, L.; Basu, S.; Wang, Y.; Chen, Z.; Hundekar, P.; Wang, B.;
Shi, J.; Shi, Y.; Narayanan, S.; Koratkar, N. Self-Heating-Induced
Healing of Lithium Dendrites. Science (Washington, DC, U. S.) 2018,
359 (6383), 1513−1516.
(37) Jafta, C. J.; Sun, X.-G.; Veith, G. M.; Jensen, G. V.; Mahurin, S.
M.; Paranthaman, M. P.; Dai, S.; Bridges, C. A. Probing Microstructure and Electrolyte Concentration Dependent Cell Chemistry
via Operando Small Angle Neutron Scattering. Energy Environ. Sci.
2019, 12 (6), 1866−1877.
(38) Jafta, C. J.; Sun, X.-G.; Lyu, H.; Chen, H.; Thapaliya, B. P.;
Heller, W. T.; Cuneo, M. J.; Mayes, R. T.; Paranthaman, P.; Dai, S.;
Bridges, C. A.; Chen, H.; Thapaliya, B. P.; Dai, S.; Heller, W. T.;
Cuneo, M. J. Insight into the Solid Electrolyte Interphase Formation in
Bis(Fluorosulfonyl)Imide Based Ionic Liquid Electrolytes; 2021.
DOI: 10.1002/adfm.202008708.
Issues between Different Solid Electrolytes in All-Solid-State Batteries.
J. Electrochem. Soc. 2019, 166 (6), A975−A983.
(11) Chen, R.; Li, Q.; Yu, X.; Chen, L.; Li, H. Approaching
Practically Accessible Solid-State Batteries: Stability Issues Related to
Solid Electrolytes and Interfaces. Chem. Rev. 2020, 120 (14), 6820−
6877.
(12) Lou, S.; Zhang, F.; Fu, C.; Chen, M.; Ma, Y.; Yin, G.; Wang, J.
Interface Issues and Challenges in All-Solid-State Batteries: Lithium,
Sodium, and Beyond. Adv. Mater. 2021, 33, 2000721.
(13) Wang, S.; Xu, H.; Li, W.; Dolocan, A.; Manthiram, A. Interfacial
Chemistry in Solid-State Batteries: Formation of Interphase and Its
Consequences. J. Am. Chem. Soc. 2018, 140 (1), 250−257.
(14) Banerjee, A.; Wang, X.; Fang, C.; Wu, E. A.; Meng, Y. S.
Interfaces and Interphases in All-Solid-State Batteries with Inorganic
Solid Electrolytes. Chem. Rev. 2020, 120 (14), 6878−6933.
(15) Cao, D.; Sun, X.; Li, Q.; Natan, A.; Xiang, P.; Zhu, H. Lithium
Dendrite in All-Solid-State Batteries: Growth Mechanisms, Suppression Strategies, and Characterizations. Matter 2020, 3 (1), 57−94.
(16) Xiao, Y.; Wang, Y.; Bo, S. H.; Kim, J. C.; Miara, L. J.; Ceder, G.
Understanding Interface Stability in Solid-State Batteries. Nat. Rev.
Mater. 2020, 5 (2), 105−126.
(17) Li, F.; Li, J.; Zhu, F.; Liu, T.; Xu, B.; Kim, T. H.; Kramer, M. J.;
Ma, C.; Zhou, L.; Nan, C. W. Atomically Intimate Contact between
Solid Electrolytes and Electrodes for Li Batteries. Matter 2019, 1 (4),
1001−1016.
(18) Tu, Q.; Barroso-Luque, L.; Shi, T.; Ceder, G. Electrodeposition
and Mechanical Stability at Lithium-Solid Electrolyte Interface during
Plating in Solid-State Batteries Electrodeposition and Mechanical
Stability at Lithium-Solid Electrolyte Interface during Plating in SolidState Batteries. Cell Reports Phys. Sci. 2020, 1 (7), 100106.
(19) Kasemchainan, J.; Zekoll, S.; Spencer Jolly, D.; Ning, Z.;
Hartley, G. O.; Marrow, J.; Bruce, P. G. Critical Stripping Current
Leads to Dendrite Formation on Plating in Lithium Anode Solid
Electrolyte Cells. Nat. Mater. 2019, 18 (10), 1105−1111.
(20) Dixit, M. B.; Verma, A.; Zaman, W.; Zhong, X.; Kenesei, P.;
Park, J. S.; Almer, J.; Mukherjee, P. P.; Hatzell, K. B. Synchrotron
Imaging of Pore Formation in Li Metal Solid-State Batteries Aided by
Machine Learning. ACS Appl. Energy Mater. 2020, 3 (10), 9534−
9542.
(21) Lewis, J. A.; Cortes, F. J. Q.; Liu, Y.; Miers, J. C.; Verma, A.;
Vishnugopi, B. S.; Tippens, J.; Prakash, D.; Marchese, T. S.; Han, S.
Y.; Lee, C.; Shetty, P. P.; Lee, H. W.; Shevchenko, P.; De Carlo, F.;
Saldana, C.; Mukherjee, P. P.; McDowell, M. T. Linking Void and
Interphase Evolution to Electrochemistry in Solid-State Batteries
Using Operando X-Ray Tomography. Nat. Mater. 2021, 20, 503.
(22) Dixit, M. B.; Singh, N.; Horwath, J. P.; Shevchenko, P. D.;
Jones, M.; Stach, E. A.; Arthur, T. S.; Hatzell, K. B. In Situ
Investigation of Chemomechanical Effects in Thiophosphate Solid
Electrolytes. Matter 2020, 3 (6), 2138−2159.
(23) Wang, M. J.; Choudhury, R.; Sakamoto, J. Characterizing the
Li-Solid-Electrolyte Interface Dynamics as a Function of Stack
Pressure and Current Density. Joule 2019, 3 (9), 2165−2178.
(24) Deng, T.; Ji, X.; Zhao, Y.; Cao, L.; Li, S.; Hwang, S.; Luo, C.;
Wang, P.; Jia, H.; Fan, X.; Lu, X.; Su, D.; Sun, X.; Wang, C.; Zhang, J.
G. Tuning the Anode−Electrolyte Interface Chemistry for GarnetBased Solid-State Li Metal Batteries. Adv. Mater. 2020, 32 (23),
2000030.
(25) Han, X.; Gong, Y.; Fu, K.; He, X.; Hitz, G. T.; Dai, J.; Pearse,
A.; Liu, B.; Wang, H.; Rubloff, G.; Mo, Y.; Thangadurai, V.;
Wachsman, E. D.; Hu, L. Negating Interfacial Impedance in GarnetBased Solid-State Li Metal Batteries. Nat. Mater. 2017, 16 (5), 572−
579.
(26) Kazyak, E.; Chen, K. H.; Wood, K. N.; Davis, A. L.; Thompson,
T.; Bielinski, A. R.; Sanchez, A. J.; Wang, X.; Wang, C.; Sakamoto, J.;
Dasgupta, N. P. Atomic Layer Deposition of the Solid Electrolyte
Garnet Li7La3Zr2O12. Chem. Mater. 2017, 29 (8), 3785−3792.
(27) Luo, W.; Gong, Y.; Zhu, Y.; Li, Y.; Yao, Y.; Zhang, Y.; Fu, K.;
Pastel, G.; Lin, C.-F.; Mo, Y.; Wachsman, E. D.; Hu, L. Reducing
Interfacial Resistance between Garnet-Structured Solid-State Electro3675
https://doi.org/10.1021/acsenergylett.1c01573
ACS Energy Lett. 2021, 6, 3669−3675
Download