ArticlePDF Available

Challenges of Predicting Temperature Dependent Capacity Loss Using the Example of NMC-LMO Lithium-Ion Battery Cells

IOP Publishing
Journal of The Electrochemical Society
Authors:

Abstract and Figures

In semi-empirical aging modeling of lithium ion-batteries an Arrhenius approach is commonly applied to describe the temperature dependency of a linear capacity loss. However, this dependency can change with degradation modes which was also observed in this cyclic aging study on NMC111-LMO graphite pouch cells in a temperature range of 4 °C to 48 °C. By means of differential voltage analysis and post-mortem analysis we correlated different regimes in capacity loss to degradation modes and aging mechanisms. In the first regime, a power dependency of time was observed. A second accelerated linear regime which followed an increase in loss of active material of the positive electrode was seen for medium (∼19 °C to 25 °C) to high aging temperatures. Transition metal dissolution was suggested to cause accelerated SEI growth. An activation energy could be estimated to 0.83 eV (± 0.17 eV, 95% CI). Finally, at aging temperatures around 45 °C we propose decreased charge transfer kinetics to result in mossy dendrites on the negative electrode which cause a final knee in aging trajectory. The findings highlight the necessity of sufficient aging temperatures and testing time.
This content is subject to copyright. Terms and conditions apply.
Challenges of Predicting Temperature Dependent Capacity Loss
Using the Example of NMC-LMO Lithium-Ion Battery Cells
L. Cloos,
1,z
J. Langer,
1
M. Schifer,
2
A. Weber,
2
and Th. Wetzel
1
1
Institute of Thermal Process Engineering (TVT), Karlsruhe Institute of Technology (KIT) Kaiserstraße 12, 76131 Karlsruhe,
Germany
2
Institute for Applied Materials - Electrochemical Technologies (IAM-ET), Karlsruhe Institute of Technology (KIT)
Kaiserstraße 12, 76131 Karlsruhe, Germany
In semi-empirical aging modeling of lithium ion-batteries an Arrhenius approach is commonly applied to describe the temperature
dependency of a linear capacity loss. However, this dependency can change with degradation modes which was also observed in
this cyclic aging study on NMC111-LMO graphite pouch cells in a temperature range of 4 °C to 48 °C. By means of differential
voltage analysis and post-mortem analysis we correlated different regimes in capacity loss to degradation modes and aging
mechanisms. In the rst regime, a power dependency of time was observed. A second accelerated linear regime which followed an
increase in loss of active material of the positive electrode was seen for medium (19 °C to 25 °C) to high aging temperatures.
Transition metal dissolution was suggested to cause accelerated SEI growth. An activation energy could be estimated to 0.83 eV
0.17 eV, 95% CI). Finally, at aging temperatures around 45 °C we propose decreased charge transfer kinetics to result in mossy
dendrites on the negative electrode which cause a nal knee in aging trajectory. The ndings highlight the necessity of sufcient
aging temperatures and testing time.
© 2024 The Author(s). Published on behalf of The Electrochemical Society by IOP Publishing Limited. This is an open access
article distributed under the terms of the Creative Commons Attribution 4.0 License (CC BY, http://creativecommons.org/licenses/
by/4.0/), which permits unrestricted reuse of the work in any medium, provided the original work is properly cited. [DOI: 10.1149/
1945-7111/ad3ec3]
Manuscript submitted February 16, 2024; revised manuscript received April 8, 2024. Published April 25, 2024.
Correctly estimating the lifetime of lithium-ion batteries is
important to ensure the warranty of electric vehicles which range
from 8 to 10 years and 100 to 155 thousand miles.
1
Many models to
predict the lifetime are dependent on data tting.
2
Semi-empirical
aging models have a high applicability and simplicity when it comes
to lithium-ion battery lifetime prediction.
2
In comparison to purely
empirical models to describe capacity fade over time or charge-
throughput they include physics-informed aspects.
3
Many stress factors of cyclic aging
4
can be included in such a model.
Even when keeping the stress factors constant the capacity fade
dependency on time or charge throughput can show three different
trajectories accelerating, linear and decelerating.
5,6
To describe a
decelerating trajectory, a root-square dependency was commonly used to
account for diffusion limited steady-state growth
7
and Solid Electrolyte
Interface (SEI) growth
8
which was especially applied for calendaric
aging.
3,9
However, Attia et al.
10
have pointed out that SEI growth is not
always diffusion limited. A transition over the different potencies of time
from 1 over 0.5 to 0, which vary from reaction to diffusion and migration
limitation of the SEI, were coupled by Kolzenberg et al.
11
Cyclic aging
data can follow a power law of less than 0.5.
10,12
In other cyclic aging
studies a linear dependency of the charge throughput has been
found.
9,13,14
Kolzenberg et al.
15
correlated a transition from square-root
to linear dependency during cyclic aging to decreased self-repairing of
the SEI. These models mostly neglected the appearance of sudden
deaths.
14
This seems adequate as the sudden death usually happens after
reaching a remaining capacity of 80%
14,16
which is dened as End of
Life (EoL) by USABC test procedures.
17
An important stress factor on cyclic aging is temperature.
18,19
To
account for the temperature inuence on capacity loss during aging,
the Arrhenius relation (1)
20
was proposed to model the temperature
dependency of SEI growth,
21
including the activation energy
E
,
A
Boltzmann-constant kBand temperature T.
rA E
kT
exp 1
A
B
⎜⎟
=⋅ []
This relation was also found for the temperature dependency of
cyclic aging.
8,22
In this case, a second temperature regime that is
governed by a different aging mechanism was found.
13
Therefore, a
double exponential form is used to describe both increased SEI
growth and manganese dissolution at higher temperatures and
lithium plating at lower temperatures on the example of an NMC/
LMO blend positive electrode.
18
This dependency was also applied
to other systems.
19,23,24
For this approach, a linearization of the
aging rate is assumed.
18
However, many factors such as the
C-Rate,
23,25,26
electrode thickness or energy density
23,25,26
and the
cell chemistry
27
impact the dominant aging mechanism, which then
dictates the temperature dependency.
2527
Furthermore, the domi-
nant aging mechanism can change over the lifetime.
28,29
For non-
linear capacity fade Kucinskis et al.
25
have proposed linearization of
certain State of Health (SoH) regions, which can be visualized in a
2D aging color map.
30
However, this approach is a step-wise
solution, which does not take into account the underlying aging
mechanisms. Therefore, the question as to how the temperature
dependency of different aging trajectories correlates to the under-
lying aging mechanisms needs to be discussed. To answer this
question a cyclic aging study on a pouch cell at different aging
temperatures, Differential Voltage Analysis (DVA)
31
and post-
mortem analysis were conducted in this study.
Experimental
The investigation was performed on a 20 Ah pouch cell with
counter-tab conguration (SPB58253172P2, Enertech Internation,
Inc.). It consists of NMC111-LMO positive electrode and graphite
negative electrode which was conrmed via XRD measurements
(D8 Advance diffractometer, Bruker). The cell with its dimensions
253 mm ×172 mm ×5.8 mm and weight of approximately 465 g
has 18 positive electrode (PE) sheets and 17 negative negative (NE)
electrode sheets. The nominal voltage is 3.75 V.
Test setup, cycling and regular reference tests.The cells were
cycled with a BaSyTec XCTS (BaSyTec GmbH) in a controlled
thermal environment. More precisely, each cell was clamped
between aluminum plates with uid channels on each side of the
pouch cell. Water was used as coolant which was supplied by
thermostats and cryostats depending on the temperature level (ECO
4 S for 19 °C, 40 °C, 45 °C and 48 °C; RE1050S for 25 °C;
VARIOCOOL 2000 for 4 °C (LAUDA Dr R. WOBSER GMBH
& CO. KG)). This setup allowed individual temperature boundary
z
E-mail: lisa.cloos@kit.edu
Journal of The Electrochemical Society, 2024 171 040538
conditions between 4 °C and 48 °C. The cell names are listed
accordingly in Table I. Additional cells at the same temperature
test condition are labeled b. The temperature was controlled via
thermocouples type K (ES Electronic Sensor GmbH) placed on the
surface of each pouch cell. The entire temperature measurement path
was calibrated previous to the aging study with a setup of
OCEANUS-6 series, milliK precision thermometer and a PT-25
reference thermometer (Isothermal Technology Limited). The
equivalent aging temperature T
EA
T
19
is given in Table Ias the actual
mean surface temperature during cycling over the entire cyclic aging
study. The internal temperature of such a pouch cell could vary up to
1 K trough-plane.
32
To mitigate the inuence of the thermocouple, a
1 mm gap ller was placed between the pouch cell and the aluminum
plate (TGF-V-Si, HALA Conec GmbH & Co. KG). The aluminum
plates also ensured a constant pressure of 0.5 bar, which was applied
via springs. All cells including the described setup were placed in a
climate chamber at a constant ambient temperature of 25 °C.
Other cyclic aging stress-factors were kept constant. The cells
were cycled with 2 C between the cut-off voltages of 3 V and 4.2 V
without pauses. Regular Reference Tests (RTP) were performed at
the nominal temperature of 25 °C. Thermal and electrical equilibra-
tion was ensured with voltage and temperature before conducting the
RTP. First, a nominal capacity (
C
N) measurement at C/2 and a quasi-
Open Circuit Voltage (qOCV) at C/10 in discharge direction were
measured. The start point to these measurements was reached with a
CC-CV (C/2, C/40 cut-off) charge phase. Afterwards, 18 s pulses
were conducted at various states of charge such as 50% with a C-
Rate of 1 C in discharging direction. These RTPs were done with
varying intervals between test batch one ( a
4
C,°a
2
5C
,
°
b
2
5C
,
°
a4
5C°) and two ( b
4
C,°a19 C ,°a
0C ,°
a4
8C°). In the rst test batch, RTPs were performed every 50
full cycles for the rst 10 RTPs. Afterwards, the checkup was
performed every 100 full cycles. Test batch two was checked upon
every 100 full cycles from the beginning. Almost no difference
between the capacity fades of cell a
4
C°and b
4
C°from the
different batches can be seen in these rst RTPs (Fig. 2). Therefore,
we suggest that no impact of different RTP schedules for the test
batches one and two on the capacity fade can be observed.
Experimental cell setup.To analyze the degradation modes via
DVA for both electrodes individually, experimental cell material was
harvested from a new cell at Begin of Life (BoL) and after cycling at
45 °C until EoL. Here, the EoL is dened as the end of testing. Hence,
the SoH of the cell aged at 45 °C at EoL was 67%. The cells were
opened in a controlled inert argon atmosphere in a glove-box. The
electrodes underwent a thorough preparation process involving their
separation and triple washing with Dimethylcarbonate (DMC) to
eliminate any residual conductive salt originating from the electrolyte.
Additionally, the active material of the electrodes was delaminated
from the current collector on one side through the utilization of N-
methyl-2-pyrrolidone (NMP). Circular coins with a standardized
diameter of 18mm were punched out from the electrodes for further
experimentation.
The experimental cell conguration encapsulated within a PAT-
Cell housing (EL-CELL GmbH) was used for the electrical
characterization of the electrodes. In the context of half-cell
measurements, negative and positive electrode coins were individu-
ally assembled against lithium. For full-cell measurements, a cell
with both negative and positive electrode was constructed. The
electrode coins were isolated by an insulation sleeve (EL-CELL
GmbH), comprising a polypropylene (PP) ber/polyethylene (PE)
membrane, alongside a lithium reference. The separator was soaked
with 80 μl LP30 (1 M LiPF
6
EC/DMC) electrolyte.
Cycling of the experimental cells was executed utilizing a
BaSyTec XCTS (BaSyTec GmbH) from 3 V to 4.3 V for the
positive electrode half-cell (positive electrode vs lithium), 0.01 V
to 1 V for the negative half-cell (negative electrode vs lithium) and
3 V to 4.2 V for the full cell (negative electrode vs positive
electrode). The applied currents for cycling procedures were based
on the calculated areal capacity derived from the cells overall
capacity and the aggregate area of the active material. The cyclic
protocol included formation cycles before a cycle with C/100 was
conducted.
Differential voltage analysis and further analysis methods.To
quantify the degradation modes via DVA,
31
the approach presented
by Schmitt et al.
33
is used. They aligned half-cell DVA curves that
can be measured at one point in time to the aged pouch cell DVA
curves.
33
Afterwards, the degradation modes could be calculated
based on the alignment factors,
33
which was done accordingly in this
study. The half-cell measurements at BoL were used for this method.
The minimization was performed with the lsqnonlin solver in
MATLAB
®
. Contrary to the method described by Schmitt et al.
33
the voltage curves were used for alignment, which was also
previously done by Hu et al.
34
Another difference to Schmitt
et al.
33
is the cut-off at the sides of the voltage curves was increased
from 1% to 5% to avoid optimization to the steep slopes.
As another quantication step, the capacity loss was analyzed for
acceleration onsets and knees. A Bacon-Watts
35
model which was
adapted by Fermín-Cueto et al.
36
to identify knee-points can be used
for this purpose. Fermín-Cueto et al.
36
also proposed a double
Bacon-Watts function to predict the knee onset. Similarly, we used
this double Bacon-Watts function to identify an acceleration onset
and the nal knee. Again, the lsqnonlin solver in MATLAB
®
was
utilized.
A further analysis on the capacity loss was performed. In a rst
step, we assumed the capacity loss to follow a power function as
shown in Eq. 2with
r
pow and the exponent
α
as variables. To
visualize the evolution of the exponent
α
over the course of aging,
the power function was tted over a sliding window of 20 days with
the tfunction using the non-linear least squares method in
MATLAB
®
. In a second step, the identied changes in the aging
trajectory from the double Bacon-Watts function
36
were used to
separate the slopes in aging trajectory. Then, the power (pow)
function and if applicable a linear (lin) function (3) was tted using
the tfunction with non-linear least squares method in MATLAB
®
.
The linear function consists of the axis intercept
B
and the slope
r
.
lin
C
C
rX12
N
N,BoL
pow
=− []
α
C
C
BrX
X time or EFC 3
N
N,BoL
lin
=−
=[]
Post-mortem analysis.The aged pouch cells were opened in
the same manner as previously described for the experimental cell
setup. For each temperature levellow, medium and higha cell
Table I. Tested cells cell names and measured mean surface
temperatures during cycling (TEAT ) and regular reference test (RTP)
test batch.
Cell name TC
EAT/°Test batch
a
4
C°3.6 1
b4
C°4.5 2
a
19 C°19.0 2
a
2
5C°25.5 1
b2
5C°25.6 1
a
4
0C°40.5 2
a
4
5C°45.2 1
a
4
8C°48.1 2
Journal of The Electrochemical Society, 2024 171 040538
was opened at the respective SoH ( a
4
C
:
°85.9% SoH; b
2
5C
:
°
85.8% SoH; a
4
5C
:
°67.0% SoH). Samples were cut from the
electrodes and washed with DMC. Then, they were only exposed to
the ambient atmosphere shortly before inserting them into micro-
scopes. Both a digital light microscope (VHX 7000, Keyence
Corporation) and a scanning electron microscope (SEM) (LEO
1530 Gemini) were used. For the latter, an acceleration voltage of
2 kV to 5 kV was applied for the SEM images. Energy dispersive X-
ray spectroscopy (EDX) was performed with an acceleration voltage
of 15 kV. Inductively coupled plasma optical emission spectroscopy
(ICP-OES) was applied for the found depositions on the sample
a
4
5C.°The depositions were scraped off with a scalpel. The
sample was dissolved in acid while traces of graphite, that stuck to
the depositions, remained. The measurement was conducted with an
iCAP 7000 (ThermoFisher Scientic).
Results and Discussion
The aim is to set up a physically meaningful semi-empirical
temperature dependent aging model for the investigated NMC-
LMO/Graphite pouch cell. Therefore, the trajectories of the tem-
perature dependent capacity fade and resistance increase during
cyclic aging will be discussed in a rst step. As the temperature
dependency of the aging behavior is linked to the dominant aging
mechanism,
2527
degradation modes will be analyzed for both
experimental cell setup as well as for the pouch cells. Post-mortem
analysis is used to provide even more detailed insights into the
prevailing aging mechanisms. Lastly, correlations between capacity,
degradation modes and aging mechanisms will be discussed and a
new model will be proposed. An overview of the procedure is given
in Fig. 1.
Capacity decrease and resistance increase.The relative capa-
city loss is plotted over time in days in Fig. 2a. 100 days equal
approximately 2250 equivalent full cycles (EFC). The EFCs are
calculated with the cumulative charge throughput from both charge
and discharge of each cycle divided by the cumulative charge
throughput of the rst C/2 cycle performed at the initial RTP at
nominal conditions. The different aging temperature levels are
indicated with different colors (blue4 °C, light blue19 °C, green-
25 °C, yellow40 °C, magenta45 °C, dark magenta48 °C).
Additional cells for one temperature test condition (b) are depicted
with dashed lines. Figure 2b provides a zoom into the SoH range
from 90% to 100% of Fig. 2a. In this gure, a closer look at the cells
a
2
5C°and b
2
5C°reveal an almost indistinguishable capacity
fade, which indicates good reproducibility. The cells aged at around
4 °C show increasing deviation after approximately 15 days, which
could be due to the slight difference in average temperature over
cycling of around 1 K. This temperature increase could not be
correlated to an increase in resistance (Fig. 2c) and is therefore likely
due to the experimental setup.
An initial capacity drop can be seen for all investigated cells in
Fig. 2b. In the rst 13 days the capacity decrease trajectories are
strongly overlapping and decelerating for all temperature boundary
conditions. For the aging temperatures below 25 °C (
4
C
,
°19 C°),
the decelerating aging trajectory continues, while the medium to
higher aging temperatures (
2
5C,°
4
0C,°
4
5C°and
4
8C°) lead to
an acceleration in the aging trajectory at different EFCs. Few data
points in the capacity loss for the aging temperature of 19 °C also
hint towards an acceleration. This start of accelerated aging occurs
well above a SoH of 90%. With increasing aging temperature, the
acceleration occurs at a lower EFC. The aging trajectories then
change to a linear aging trajectory whose slope increases with
increasing aging temperature (
2
5C°<
4
0C°<
4
5C°<
4
8C.°)
This also means, that the lower aging temperatures (
4
C
,
°19 C°)
with the ongoing mostly decelerating trend show less relative
capacity fade in the long run. The least relative capacity fade at
approximately 90 days is observed for an aging temperature of
19 °C. Only for the high aging temperatures 45 C
and
4
8C°)a
nal knee can be seen below 80% SoH.
The relative 18 s pulse resistance increase is shown in Fig. 2c.
Similarly to the capacity loss, an acceleration of the resistance
increase can be seen for medium to high aging temperatures.
However, this acceleration seems to occur at a later point in time,
which can clearly be seen for the high aging temperatures. The
capacity loss only becomes sensitive to resistance increase if the
discharge ends in the at region of the voltage curve,
6
which is not
the case here. Therefore, the capacity loss acceleration is unlikely to
be due to the resistance increase. Even though resistances do not
necessarily cause a capacity knee,
6
a strong linear relation between
resistance elbows and capacity knees were seen.
37
In comparison with literature, NMC-LMO/Graphite cells usually
have a decelerating
3841
or linear capacity loss behavior.
42
Only in
one investigation for the same cell chemistry at 25 °C an accelera-
tion of the aging trajectory in the region of 95% SoH was seen.
18
Dubarry et al.
17
have discussed such an acceleration between a SoH
of 80% to 90% in a study on the same cell chemistry. They have
found an increase in Loss of Active Material of the positive
electrode (LAM
PE
).
17
Assuming similar mechanisms resulting in a
knee
6
also causing this acceleration an increase of one dominant
degradation mode
43,44
or a change in dominating degradation
mode
28,29
are possible. To get a general idea about dominating
degradation mode for the investigated cell DVA is an adequate tool.
Differential voltage analysis on half-cells and full-cells.With
half-cell measurements of positive and negative electrodes LAM can
be easily seen. LAM results in a contraction of the voltage curves.
31
Loss of Lithium Inventory (LLI)
45
causes a shift between positive
electrode and negative electrode curve, which can be seen in full-cell
conguration.
31
The cell a
4
5C°is chosen as an example for high
capacity loss. DVA graphs of the positive and negative electrode
DVAs with respective half-cell and full-cell measurement at EoL are
shown in Figs. 3a3d. The balancing of positive and negative half-
cell electrode DVA to the pouch cell DVA
31
is shown in Fig. 3a.
The NMC-LMO blend positive electrode has two minima in all
presented measurements (Fig. 3c). The one at higher potentials can
be assigned to LMO and the one at lower potentials to NMC.
41
The
positive electrode capacity contracts by about 19%, which is a
signicant amount of LAM
PE
. No clear trend can be seen of a NMC
and LMO peak shift. The negative electrode shows two very
different trends in the half-cell and full-cell measurement. The
Figure 1. Overview of the procedure in this publication to nd a semi-empirical model for temperature dependent cyclic aging.
Journal of The Electrochemical Society, 2024 171 040538
half-cell DVA of the negative electrode does not change drastically
in comparison to BoL. All graphite peaks are still visible. The
contraction is around 2%. This is not true for the negative electrode
DVA calculated from full-cell measurements at EoL. Here, the
contraction is signicant of around 47%. This means, that LAM
NE
is
minor. Any losses in active lithium were compensated by the lithium
metal counter electrode. The large difference between the half-cell
and full-cell measurement of the negative electrode at EoL indicate a
large LLI. A similar difference between the setups were seen by Sieg
et al.
46
Differential voltage analysis on pouch cells.While these half-
cell measurements only give a momentary insight into the degrada-
tion modes at BoL and EoL, a DVA at each RTP can show the
degradation mode trends over cyclic aging. In Figs. 4a4c) the
estimated LLI, LAM
PE
and LAM
NE
for the described method in
Differential Voltage Analysis and Further Analysis Methods
section is shown. The squared norm of the residuals of the method
is presented in Fig. 4e. For comparison, the capacity loss is also
plotted in Fig. 4d. In this gure, the described onsets of acceleration
and knees in capacity loss were quantied with the double
Bacon-Watts
36
model from Differential Voltage Analysis and
Further Analysis Methodssection. The onset of acceleration of
the capacity loss is marked with a dotted vertical line and the nal
knee is marked with a vertical dot dash line.
The trajectories in LLI appear like the inverse of the capacity
loss, including the acceleration. An acceleration can also be seen in
LAM
PE
. However, this acceleration starts slightly earlier. Here, both
the low aging temperatures (3 °C and 19 °C) show the least increase.
Generally, LAM
PE
is quantitatively less than LLI. As already
predicted by the half-cell measurements LAM
NE
is very low for
all investigated aging temperatures. Interestingly, LAM
NE
increases
signicantly at the same time of the capacity loss knee. It needs to be
taken into account that the squared norm of the residuals increases
with decreasing SoH.
DVA of half- and full-cells and pouch cells suggested LLI to be
the dominant degradation mode. For NMC-LMO cells, LLI was
previously found to be the dominant degradation mode at an ambient
temperature and above SoH 80%.
17,47
Two regimes that were
already identied for the capacity loss could also be identied in
quantitative DVA; being an increase in LAM
PE
followed by an
increase in LLI. Dubarry et al.
17
related an increase of LAM
PE
to the
capacity loss acceleration for an NMC111-LMO cell at room
temperature. They proposed that LLI in the rst regime leads to
grain isolation and/or pore clogging of active material and LAM.
17
Regarding these results, it seems more likely that the aging
mechanism causing an increase in LAM
PE
accelerates and then
triggers an acceleration in LLI. Furthermore, the results suggest an
increase in LAM
NE
to correlate to the nal knee in capacity loss.
Post-mortem analysis could give further evidence as to which aging
mechanisms are causing this behavior.
Post-mortem analysis.The post-mortem analysis which is
discussed in this section consisted of a visual inspection and SEM
analysis. Cells from three different aging temperature levels which
showed different aging trajectories were chosen. The low
Figure 2. Cyclic aging study at different aging temperature levels (blue4 °C, light blue19 °C, green25 °C, yellow40 °C, magenta45 °C, dark
magenta48 °C). Additional cells for one temperature test condition (b) are depicted with unlled markers. (a) Relative capacity loss over time in days;
(b) Zoom into relative capacity loss over time in days; (c) Relative 18 s pulse resistance increase at 50% state of charge over time in days. (100 days
»
2250
equivalent full cycles (EFC)).
Figure 3. (a) Balancing of half-cell negative (NE) and positive (PE) electrode to pouch cell DVA at Begin of Life (BoL); (b) Full-cell DVA with respective
negative and positive electrode DVA in full-cell conguration at End of Life (EoL) of cell a
4
5C ;°(c) Positive electrode DVA at BoL and EoL for half-cell and
full-cell conguration of cell a
4
5C ;°(d) Negative electrode DVA at BoL and EoL half-cell and full-cell conguration of cell a
4
5C ;°The arrow indicates the
shift of the negative half-cell to full-cell DVA at EoL. (blackpouch and full-cell, orangehalf-cell PE BoL, yellowhalf-cell PE BoL, greenfull-cell PE EoL,
bluehalf-cell NE BoL, light bluehalf-cell NE EoL, pinkfull-cell NE EoL).
Journal of The Electrochemical Society, 2024 171 040538
temperature level was represented by the cell aged at 4 °C, the
medium temperature level at 25 °C and the high temperature level at
45 °C respectively. Upon cell opening, the electrode sheets of cell
aged at 4 °C had a liquid lm on the sheets in the moment of
disassembly. The other cells at a higher aging temperature level
showed a ne layer on the electrodes after cell opening which
resembled a frost pattern. This could be washed off with DMC. Also,
the negative electrode sheets themselves displayed visual differences
while the positive electrode sheets were macroscopically unchanged.
The negative electrode sheets of the cell aged at 4 °C had white
depositions at the edges. The cell aged at 45 °C also had white
circular depositions in the center and close to the current collectors
(Fig. 5a).
Light microscopy of the white depositions of the cell aged at
45 °C (Fig. 5b) revealed a pillar-like structure with yellowish
discoloration in the center. SEM pictures showed a different
structure towards the center (Fig. 5c) and towards the edges
(Fig. 5d) of the white deposition. At the edges, a mossy structure
can be seen, whereas the center seems to be covered with an uneven
layer. SEM images of the white depositions on the a
4
C°cell
revealed a similar structure as described for the depositions of the
a4
5C°cell although less in height. ICP-OES measurements of the
scraped of white depositions on the
a4
5C°cell had the largest
measurable content to be Lithium. Manganese, Nickel and Cobalt
were present in small contents. SEM-EDX on the depositions
showed prominent peaks of Oxygen, Phosphorous and Fluorine.
For the low temperature aging (
4
C°) lithium plating at the edges
can be expected due to the edge effect.
48,49
Also, at high temperature
aging (
4
5C°) ICP-OES conrmed the presence of lithium in the
white pillar like depositions. Even though lithium plating at higher
temperatures is not expected
18
it can occur when inhomogeneities
are present.
50,51
Storch et al.
50
proposed an interplay between
temperature, current density and lithium concentration gradient to
cause local lithium plating in their large format cells. The local
current density is also critical to the lithium growth mechanisms.
52
The interaction between local current density, lithium concentration
and fractured SEI enables mossy dendrites even under normal charge
scenarios
53
whose shape resemble the found depositions on the
negative electrodes at an aging temperature of 45 °C (Fig. 5b). The
layer on top of the pillars, which can be seen in Fig. 5b resembles the
layer found by Paarmann et al.
54
They suggested this to be a type of
lithium plating.
54
The deposited lithium can become inactive
dead-lithium.
55
This could explain the lack of metallic sheen of
the depositions.
Apart from the observed depositions SEM also reveals differ-
ences in the electrode particles between the cells aged at different
temperatures. To illustrate these differences both negative electrode
and positive electrode SEM at BoL and EoL of the three different
temperature levels are selected as representation in Fig. 6. At BoL,
the large round particles in the positive electrode blend can be
identied as NMC and the smaller square shaped particles as
LMO.
56
When comparing the positive electrode at BoL and at
EoL of the cell a
2
5C
,
°a surface layer has formed on top of a large
NMC particle, which looks different to carbon black binder phase at
BoL (Figs. 6a and 6b). This surface layer can be seen on many
positive electrode particles for all investigated aging temperature
levels. In case of the a
4
C°cell a close up of one of these layers
which peeled off was taken (Fig. 6(d)). Here, the layer thickness can
be estimated to about 200 nm. At an aging temperature of 45 °C, not
only the surface layer can be seen but also particles seem to be
Figure 4. Quantitative analysis of degradation modes of the cyclic aging study at different aging temperature levels (blue4 °C, light blue19 °C, green25 °C,
yellow40 °C, magenta45 °C, dark magenta48 °C). Additional cells for one temperature test condition (b) are depicted with unlled markers or dashed lines.
The identied onset of capacity loss acceleration is indicated with vertical dotted lines and the identied knees with vertical dot dash lines in the respective color.
(a) Loss of lithium inventory (LLI) over time in days; (b) Loss of active material of the positive electrode (LAM
PE
) over time in days; (c) Loss of active material
of the negative electrode (LAM
NE
) over time in days; (d) Relative capacity loss over time in days, the tted aging function is shown with lines; (e) Squared norm
of the residuals of the quantitative DVA. (100 days
»
2250 equivalent full cycles (EFC)).
Journal of The Electrochemical Society, 2024 171 040538
ground to ne powder in some areas (Fig. 6c). Generally, particle
cracking can be observed in all probes but also at BoL. Therefore,
there is no clear trend.
A known aging mechanism on the positive electrode is the
formation of a Cathode Electrolyte Interface (CEI). However, CEI is
known to be couple of nm thick,
57
which is well below the estimated
layer thickness of 200 nm of the observed layer in this investigation.
Only at high voltages above 4.5 V, a higher thickness around 2 μm
layer on NMC111 particles was seen.
58
Another aging mechanism at
high voltages for NMC particles is particle pulverization,
59
which
was observed for a high aging temperature of 45 °C (Fig. 6c). In this
investigation, high cut-off voltages were avoided. The maximum
positive electrode voltage was 4.27 V vs Li/Li
+
at BoL. This is low
even when taking into account that the onset potential for aging
mechanisms such as oxygen release decreases with increasing
temperature due to lower overpotentials.
60
Darma et al.
61
argued,
that already at 4.05 V NMC was stressed mechanically in their
NMC-LMO-NCA blend material. In the following analysis, Lang
et al.
62
found microcracks in the NMC. Cracking is known to be a
critical mechanism for layer-structure positive electrodes at high
voltages.
63
Lang et al.
62
argued this to become a self-enforcing
trend. This mechanism could explain both layer formation on
positive electrode particles and positive electrode particle pulveriza-
tion at higher temperatures such as 45 °C. In an NMC-LMO blend
material Dubarry et al.
64
have seen over-lithiation of LMO due to
differences in the kinetics of the blend which should not appear
under normal cycling conditions.
Whereas layer formation on the positive electrode particles was
observed for all aging temperature levels layer formation on the
negative electrodes was noticed to various extents. In comparison to
BoL, the negative electrode of the a
2
5C°at EoL shows rounding
of the edges which can be consequence of surface layer formation
(Figs. 6e and 6f). This is also true for the negative electrodes of the
a4
5C°(Fig. 6g). For the cell aged at 4 °C the rounding of negative
electrode particles is not as visible.
A common aging mechanism of the negative electrode is the
SEI.
65
Blurring of the graphite edges is usually associated with
SEI.
54
The SEM pictures suggest that the SEI layer increases at
higher aging temperatures of 25 °C and above, which is likely to be
due to the described positive electrode aging mechanisms. At higher
temperatures such as 55 °C, Mn dissolution in LiPF
6
electrolytes
occurs.
66
This aging mechanism of metal dissolution is known for
both spinel LMO and layered oxide NMC.
67
The dissolved
manganese ions can then interact with the SEI and enhance the
formation of inactive species.
68,69
Also, in this study, ICP-OES
measurements of the depositions on the negative electrode at an
aging temperature of 45 °C revealed the presence of transition
metals. These side reactions can also cause Li salt and electrolyte
depletion
41,70
which is in line with decreased wetting of electrode
sheets upon cell disassembly.
71
Transition metal dissolution is not
only triggered by higher temperatures but also higher voltages
72
which also seems to be an issue here.
These aging mechanisms can be correlated to the features found
in capacity loss, resistance increase and DVA. The post-mortem
results suggest that over-lithiation is an issue for this blend positive
electrode. High voltage operation of a positive electrode can cause
layer formation,
58
particle cracking
63
and transition metal
dissolution.
72
Layer formation on the positive electrode did not
necessarily lead to LAM
PE
, which can be seen on the basis of an
aging temperature of 4 °C. Positive electrode surface layers are not
known to passivate.
14,73
At medium to higher aging temperatures the
processes on the positive electrode seem to accelerate, which is then
followed by an increase in LLI and decrease in capacity loss.
Transition metal dissolution is a known process to increase SEI layer
formation.
68
The insertion of manganese species in the SEI alters the
SEI
69,74
and increases the charge-transfer impedance.
75
An increase
in the pulse resistance could be observed. Decreased charge-transfer
causing plating is considered a threshold knee trajectory.
6,16
Due to
the changed SEI, we suggest an easier onset for mossy dendrites as
they grow due to a fractured SEI.
53
This knee coincides with an
increase in LAM
NE
. Another aging mechanism that can be correlated
with this nal knee is particle pulverization, which is the worst case
of particle cracking.
Consequences for temperature dependent semi-empirical aging
model.To describe the aging trajectories for the investigated cell
an assumption of a linear aging trend whose slope is often used to
describe the temperature dependency in an Arrhenius t
18
is not
applicable. Not only did different aging temperatures lead to
different aging trajectories which was already found by Kucinskis
et al.
25
but also changes in the aging trajectories were observed.
Clearly, different aging modeling are necessary. For this modeling
the nal knee below a SoH of 80% is excluded.
To analyze the changes in the capacity loss a power function (2)
was tted over a sliding window as described in "Differential
Voltage Analysis and Further Analysis Methods" section. The
exponent
α
is presented in Fig. 7a over the course of aging with
the 95% condence interval (CI) as a shaded area in the respective
color. Changes in the exponent can indicate changed limitation with
increasing SEI thickness; from reaction to diffusion to electron
Figure 5. Picture of the negative electrode sheets (a), light microscopy (b) and SEM pictures (c), (d) of a white deposition at EoL of cell a
4
5C°.
Journal of The Electrochemical Society, 2024 171 040538
migration limitation.
11
Even though the condence intervals are
wide, trends for aging temperatures below and above 25 °C can be
differentiated. While the exponent
α
at aging temperatures of 25 °C
rstly lies around 0.5 higher aging temperatures result in an
exponent of 0.7 to 1. The exponents of the medium aging
temperatures all start to increase at some point in time while the
exponents at higher aging temperatures increase from the beginning.
The increase of the exponent at low to medium aging temperatures
of 3 °C and 19 °C is unclear due to high condence intervals.
Exponents well above one can be correlated to the beginning of the
knee.
Relating these results to the investigation of SEI growth regimes
of Kolzenberg et al.
11,15
we observed the transition from self-
passivation (
α
=0.5) to non-passivation (
α
=1) SEI growth for
medium aging temperatures (19 °C, 25 °C). Either at higher aging
temperatures this transition is too fast or another mechanism is at
play. An exponent of 0.76 was previously found in an experimental
study of Gilbert et al.
72
It describes a diffusion/percolation limited
mechanism through the inner and outer SEI.
72,76
In their study for
higher cut-off voltages, this initial regime with an exponent of 0.76
was swiftly followed by an intermediate linear regime
72
which
marks a fast change in SEI growth regimes from lower to higher
growth rate. Gilbert et al.
72
argue that the linearity is an artifact of
multiple occurring processes during their cyclic aging experiments.
This linear regime is controlled by cracking of positive electrode
particles and loss of transition metal ions, which result in faster SEI
growth rate than solely diffusion controlled growth.
72
This is in line
with the ndings of this study.
Generally, the investigated cell is rather insensitive to lower
aging temperatures, which is entirely different to the results by
Waldmann et al.
18
with the same cell chemistry. We observed only
slight edge plating for aging temperatures of 3 °C. This highlights
that not only the cell chemistry but also other design factors such as
electrode thickness have an impact on the optimum aging
temperature.
25,26
This could be related to the fact that the positive
electrode chemistry and electrode material load has shown to shift
the degradation modes.
77
The analysis of the exponent
α
supports our ndings regarding
the two aging regimes. Therefore, we used the found acceleration
onsets and knees from the double Bacon-Watts model as boundaries
to t a power function (2) to the rst regime and a linear function (3)
to the second regime as described in "Differential Voltage Analysis
Figure 6. SEM pictures of positive electrode (PE) (a), (b), (c), (g), (h) and negative electrodes (NE) (c)(f) at Begin of Life (BoL) (a), (c) and End of Life (EoL)
for cells aged at temperatures of b
2
5C°(b), (d), a
4
5C°(c), (e) and a
4
C°(f)(h).
Journal of The Electrochemical Society, 2024 171 040538
and Further Analysis Methods" section. The results for the ttings
are shown with continuous lines in Fig. 4d. The values for the
parameters can be found in Table II. Due to multiple tested aging
temperatures in this investigation an Arrhenius plot can be drawn for
the slope
r
lin of the linear regime over time (Fig. 7b). The error bars
highlight the 95% CI which is higher for the aging temperature of
19 °C. This aging temperature only showed signs of acceleration. It
reveals an estimate for the activation energy of 0.83 eV 0.17 eV,
95% CI) which is higher than the measured activation energies for
cyclic aging in the high temperature regime. These range from 31.7
kJ mol
1
to 54.99 kJ mol
118,22,78
in comparison to 80.08 kJ mol
1
.
This means that the activation energy for the multiple previously
described processes is higher than conventional SEI growth. It is
also larger than the activation energy of Mn dissolution which was
found by Waldmann et al.
79
via ICP-OES with 0.33 ± 0.02 eV.
Neglecting the different regions can have minor to very severe
consequences on the aging prediction accuracy. Assuming a linear
trend for example can describe the overall trend relatively well.
However, it cannot describe the initial capacity loss especially for
high and low temperatures. Even more important is the necessary
testing time. If the testing time does not cover the accelerated regime
an extrapolation is not possible. This is illustrated in Fig. 1on the top
right on the example of an aging temperature of 25 °C. A prediction
based only on 50 days of testing would lead to a misjudgment of the
achievable days for a SoH of 90% by about 20 days (30%).
Conclusions
Temperature dependent semi-empirical aging models are often
described via an Arrhenius dependency. However, it has been shown
that this dependency can change with the slope of the aging
trajectory.
25
In this paper we set up a physically meaningful
temperature dependent aging model for a NMC111-LMO graphite
pouch cell. We did not only observe different aging trajectories for
the different aging temperatures ranging from 4 °C to 48 °C but also
an acceleration in aging trajectories over the lifetime for some aging
temperatures. With means of DVA and post-mortem analysis a
correlation between capacity loss, degradation modes and aging
mechanisms could be derived:
-Lower aging temperatures (3 °C) mostly followed a t
0.5
behavior. A
qualitatively similar behavior was observed for the LLI. The only
visible aging mechanism apart from slight edge plating was layer
formation on the positive electrode (200 nm) which could be
due to over-lithiation of the blend positive electrode.
62,64
-Medium aging temperatures (19 °C to 25 °C) rst followed a t
0.5
behavior but then accelerated into a linear regime which was also
seen for higher aging temperatures.
-High aging temperatures (40 °C to 48 °C) had a power dependency
above 0.7 in the rst regime before the acceleration into a linear
regime. An increase in LAM
PE
preceded an increase in LLI and
the accelerated capacity loss. This acceleration is a known
process at higher voltages where transition metal dissolution
leads to diffusion/percolation controlled SEI growth.
72
The
process accelerates due to further stress on the oxide particles
which leads to a fast linear growth of SEI.
72
An increased layer
formation on the negative electrode and transition metal dissolu-
tion could indeed be seen. The activation energy of these
processes could be estimated to 0.83 eV 0.17 eV, 95% CI).
Finally, we suggest the increased charge transfer resistance of the
SEI to enable mossy dendrite lithium growth which results in a
nal knee and coincides with an increase in LAM
NE
.
This study clearly shows that sufcient testing both in aging
temperatures as well as testing time is needed to depict all aging
regimes. Neglecting the second regime can lead to false estimate of
achievable days of around 30% to reach SoH 90%. This might
Figure 7. (a) Exponent
α
of the Eq. 2tted over a sliding window of 20
days of the capacity loss at different aging temperatures with the 95%
condence interval shown as shaded area (blue4 °C, light blue19 °C,
green25 °C, yellow40 °C, magenta45 °C, dark magenta48 °C); (b)
Arrhenius plot of the pre-factor
r
lin of Eq. 3tted to the linear region of
the capacity loss; The error bars mark 95% condence interval of tting
factors. (blackArrhenius t with activation Energy (
E
A)); Additional cells
for one temperature test condition (b) are depicted with unlled markers or
dashed lines.
Table II. Parameters of aging model shown in Fig. 4d.
Cell name
r
days
pow 1
/
α
/− Threshold days/
B/−
r
days
lin
1
/
a
4
C°0.0091 0.575 ——
b
4
C°0.0111 0.495 ——
a
1
9C°0.0102 0.476 65.645 0.988 0.0009
a
2
5C°0.0074 0.583 50.007 0.988 0.0013
b
2
5C°0.0080 0.560 49.382 0.986 0.0013
a
4
0C°0.0044 0.812 28.418 1.031 0.0035
a
4
5C°0.0053 0.794 16.703 1.039 0.0054
a
4
8C°0.0037 0.970 21.641 1.094 0.0077
Journal of The Electrochemical Society, 2024 171 040538
become even more challenging when thermal inhomogeneities
19
and
transients
19,80
are included.
Acknowledgments
L.C., M.S. and T.W. gratefully acknowledges the funding and
support by the German Research Foundation (DFG) within the research
training group SiMET under the project number 281041241/GRK2218.
Also, the authors would like to thank Sabrina Herberger at TVT at KIT
for support in test setup, Annette Schucker at IAM-ET at KIT for
conducting EDX and XRD measurements on the BoL cell, Elisabeth
Eiche at AGW at KIT for conducting ICP-OES measurements on the
negative electrode covering layer and Volker Zibat at LEM at KIT for
conducting SEM-EDX measurements. Many thanks also to Philipp
Dechent at University of Oxford for the valuable discussions and
thorough proof-read.
ORCID
L. Cloos https://orcid.org/0009-0006-1001-2891
M. Schifer https://orcid.org/0009-0006-4606-102X
A. Weber https://orcid.org/0000-0003-1744-3732
References
1. J. Gorzelany, (2022), By The Numbers: Comparing Electric Car Warranties https://
www.forbes.com/sites/jimgorzelany/2022/10/31/by-the-numbers-comparing-elec-
tric-car-warranties/.
2. M. S. Hosen, J. Jaguemont, J. van Mierlo, and M. Berecibar, I. Science,24, 102060
(2021).
3. M. Ecker, J. B. Gerschler, J. Vogel, S. Käbitz, F. Hust, P. Dechent, and D. U. Sauer,
J. Power Sources,215, 248 (2012).
4. T. Gewald, A. Candussio, L. Wildfeuer, D. Lehmkuhl, A. Hahn, and M. Lienkamp,
Batteries,6, 6 (2020).
5. T. Waldmann, B.-I. Hogg, and M. Wohlfahrt-Mehrens, J. Power Sources,384, 107
(2018).
6. P. M. Attia et al., J. Electrochem. Soc.,169, 60517 (2022).
7. B. E. Deal and A. S. Grove, J. Appl. Phys.,36, 3770 (1965).
8. R. Wright et al., J. Power Sources,110, 445 (2002).
9. J. Wang, J. Purewal, P. Liu, J. Hicks-Garner, S. Soukazian, E. Sherman,
A. Sorenson, L. Vu, H. Tataria, and M. W. Verbrugge, J. Power Sources,269,
937 (2014).
10. P.M.Attia,W.C.Chueh,andS.J.Harris,J. Electrochem. Soc.,167, 090535 (2020).
11. L. von Kolzenberg, A. Latz, and B. Horstmann, Chem. Sus. Chem.,13, 3901 (2020).
12. I. Bloom et al., J. Power Sources,101, 238 (2001).
13. L. Lam and P. Bauer, IEEE Trans. Power Electron.,28, 5910 (2013).
14. M. Ecker, N. Nieto, S. Käbitz, J. Schmalstieg, H. Blanke, A. Warnecke, and D.
U. Sauer, J. Power Sources,248, 839 (2014).
15. L. von Kolzenberg, A. Latz, and B. Horstmann, Batteries & Supercaps,5,
e202100216 (2022).
16. S. F. Schuster, T. Bach, E. Fleder, J. Müller, M. Brand, G. Sextl, and A. Jossen,
Journal of Energy Storage,1, 44 (2015).
17. M. Dubarry, C. Truchot, B. Y. Liaw, K. Gering, S. Sazhin, D. Jamison, and
C. Michelbacher, J. Power Sources,196, 10336 (2011).
18. T. Waldmann, M. Wilka, M. Kasper, M. Fleischhammer, and M. Wohlfahrt-
Mehrens, J. Power Sources,262, 129 (2014).
19. D. Werner, S. Paarmann, A. Wiebelt, and T. Wetzel, Batteries,6, 12 (2020).
20. S. Arrhenius, Z. Physik. Chem.,4, 226 (1889).
21. M. Broussely, S. Herreyre, P. Biensan, P. Kasztejna, K. Nechev, and R. Staniewicz,
J. Power Sources,97-98, 13 (2001).
22. B. Y. Liaw, E. Roth, R. G. Jungst, G. Nagasubramanian, H. L. Case, and D.
H. Doughty, J. Power Sources,119-121, 874 (2003).
23. X.-G. Yang and C.-Y. Wang, J. Power Sources,402, 489 (2018).
24. Y. Du, S. Shironita, E. Hosono, D. Asakura, Y. Sone, and M. Umeda, J. Power
Sources,556, 232513 (2023).
25. G. Kucinskis, M. Bozorgchenani, M. Feinauer, M. Kasper, M. Wohlfahrt-Mehrens,
and T. Waldmann, J. Power Sources,549, 232129 (2022).
26. M. Bozorgchenani, G. Kucinskis, M. Wohlfahrt-Mehrens, and T. Waldmann,
J. Electrochem. Soc.,169, 030509 (2022).
27. Y. Preger, H. M. Barkholtz, A. Fresquez, D. L. Campbell, B. W. Juba, J. Romàn-
Kustas, S. R. Ferreira, and B. Chalamala, J. Electrochem. Soc.,167, 120532 (2020).
28. M. Dubarry, C. Truchot, and B. Y. Liaw, J. Power Sources,219, 204 (2012).
29. K. Smith, A. Saxon, M. Keyser, B. Lundstrom, Z. Cao, and A. Roc, 2017 American
Control Conference (ACC) 4062 (IEEE) (2017).
30. M. Feinauer, M. Wohlfahrt-Mehrens, M. Hölzle, and T. Waldmann, J. Power
Sources,594, 233948 (2024).
31. I. Bloom, A. N. Jansen, D. P. Abraham, J. Knuth, S. A. Jones, V. S. Battaglia, and
G. L. Henriksen, J. Power Sources,139, 295 (2005).
32. T. Waldmann, G. Bisle, B.-I. Hogg, S. Stumpp, M. A. Danzer, M. Kasper,
P. Axmann, and M. Wohlfahrt-Mehrens, J. Electrochem. Soc.,162, A921 (2015).
33. J. Schmitt, M. Schindler, A. Oberbauer, and A. Jossen, J. Power Sources,532,
231296 (2022).
34. V. W. Hu and D. T. Schwartz, J. Electrochem. Soc.,169, 030539 (2022).
35. D. W. Bacon and D. G. Watts, Biometrika,58, 525 (1971).
36. P. Fermín-Cueto, E. McTurk, M. Allerhand, E. Medina-Lopez, M. F. Anjos,
J. Sylvester, and G. dos Reis, Energy and AI,1, 100006 (2020).
37. C. Strange, S. Li, R. Gilchrist, and G. dos Reis, Energies,14, 1206 (2021).
38. A. Cordoba-Arenas, S. Onori, Y. Guezennec, and G. Rizzoni, J. Power Sources,
278, 473 (2015).
39. M. Fleckenstein, O. Bohlen, and B. Bäker, WEVJ,5, 322 (2012).
40. J. Purewal, J. Wang, J. Graetz, S. Soukiazian, H. Tataria, and M. W. Verbrugge,
J. Power Sources,272, 1154 (2014).
41. D. A. Stevens, R. Y. Ying, R. Fathi, J. N. Reimers, J. E. Harlow, and J. R. Dahn,
J. Electrochem. Soc.,161, A1364 (2014).
42. G. Kovachev, C. Ellersdorfer, G. Gstrein, I. Hanzu, H. M. R. Wilkening,
T. Werling, F. Schauwecker, and W. Sinz, E. Transportation,6, 100087 (2020).
43. X.-G. Yang, Y. Leng, G. Zhang, S. Ge, and C.-Y. Wang, J. Power Sources,360,28
(2017).
44. D. Anseán, M. Dubarry, A. Devie, B. Y. Liaw, V. M. García, J. C. Viera, and
M. González, J. Power Sources,356, 36 (2017).
45. M. Dubarry and B. Y. Liaw, J. Power Sources,194, 541 (2009).
46. J. Sieg, M. Storch, J. Fath, A. Nuhic, J. Bandlow, B. Spier, and D. U. Sauer, Journal
of Energy Storage,30, 101582 (2020).
47. A. J. Smith, P. Svens, M. Varini, G. Lindbergh, and R. W. Lindström,
J. Electrochem. Soc.,168, 110530 (2021).
48. M. Tang, P. Albertus, and J. Newman, J. Electrochem. Soc.,156, A390 (2009).
49. F. Grimsmann, T. Gerbert, F. Brauchle, A. Gruhle, J. Parisi, and M. Knipper,
Journal of Energy Storage,15, 17 (2018).
50. M. Storch, J. P. Fath, J. Sieg, D. Vrankovic, C. Krupp, B. Spier, and R. Riedel,
Journal of Energy Storage,41, 102887 (2021).
51. Y. Zhu et al., Nat. Commun.,10, 2067 (2019).
52. P. Bai, J. Guo, M. Wang, A. Kushima, L. Su, J. Li, F. R. Brushett, and M.
Z. Bazant, Joule,2, 2434 (2018).
53. C.-J. Ko, C.-H. Chen, and K.-C. Chen, J. Power Sources,563, 232779 (2023).
54. S. Paarmann, K. Schuld, and T. Wetzel, Energy Tech.,10, 2200384 (2022).
55. I. Yoshimatsu, T. Hirai, and J. Yamaki, J. Electrochem. Soc.,135, 2422 (1988).
56. N. M. Jobst, G. Gabrielli, P. Axmann, M. Hölzle, and M. Wohlfahrt-Mehrens,
J. Electrochem. Soc.,168, 070550 (2021).
57. Z. Zhang et al., Matter,4, 302 (2021).
58. M. Varini, J. Y. Ko, P. Svens, U. Mattinen, M. Klett, H. Ekström, and
G. Lindbergh, Journal of Energy Storage,31, 101616 (2020).
59. W. Li, X. Liu, Q. Xie, Y. You, M. Chi, and A. Manthiram, Chem. Mater.,32, 7796
(2020).
60. R. Jung, P. Strobl, F. Maglia, C. Stinner, and H. A. Gasteiger, J. Electrochem. Soc.,
165, A2869 (2018).
61. M. S. D. Darma, M. Lang, K. Kleiner, L. Mereacre, V. Liebau, F. Fauth,
T. Bergfeldt, and H. Ehrenberg, J. Power Sources,327, 714 (2016).
62. M. Lang, M. S. D. Darma, K. Kleiner, L. Riekehr, L. Mereacre, M. Ávila Pérez,
V. Liebau, and H. Ehrenberg, J. Power Sources,326, 397 (2016).
63. P. Yan, J. Zheng, M. Gu, J. Xiao, J.-G. Zhang, and C.-M. Wang, Nat. Commun.,8,
14101 (2017).
64. M. Dubarry, C. Truchot, A. Devie, B. Y. Liaw, K. Gering, S. Sazhin, D. Jamison,
and C. Michelbacher, J. Electrochem. Soc.,162, A1787 (2015).
65. E. Peled, J. Electrochem. Soc.,126, 2047 (1979).
66. A. Blyr, C. Sigala, G. Amatucci, D. Guyomard, Y. Chabre, and J.-M. Tarascon,
J. Electrochem. Soc.,145, 194 (1998).
67. S. K. Martha et al., J. Power Sources,189, 288 (2009).
68. H. Tsunekawa, A. S. Tanimoto, R. Marubayashi, M. Fujita, K. Kifune, and
M. Sano, J. Electrochem. Soc.,149, A1326 (2002).
69. G. Amatucci, A. Du Pasquier, A. Blyr, T. Zheng, and J.-M. Tarascon, Electrochim.
Acta,45, 255 (1999).
70. P. Yang, J. Zheng, S. Kuppan, Q. Li, D. Lv, J. Xiao, G. Chen, J.-G. Zhang, and C.-
M. Wang, Chem. Mater.,27, 7447 (2015).
71. M. Ecker, P. Shaei Sabet, and D. U. Sauer, Appl. Energy,206, 934 (2017).
72. J. A. Gilbert, I. A. Shkrob, and D. P. Abraham, J. Electrochem. Soc.,164, A389
(2017).
73. K. Edström, T. Gustafsson, and J. O. Thomas, Electrochim. Acta,50, 397 (2004).
74. C. Zhan, J. Lu, A. Jeremy Kropf, T. Wu, A. N. Jansen, Y.-K. Sun, X. Qiu, and
K. Amine, Nat. Commun.,4, 2437 (2013).
75. K. Amine, J. Liu, S. Kang, I. Belharouak, Y. Hyung, D. Vissers, and G. Henriksen,
J. Power Sources,129, 14 (2004).
76. A. Nitzan and M. A. Ratner, J. Phys. Chem.,98, 1765 (1994).
77. B.-R. Chen, C. M. Walker, S. Kim, M. R. Kunz, T. R. Tanim, and E. J. Dufek,
Joule,6, 2776 (2022).
78. I. Baghdadi, O. Briat, J.-Y. Delétage, P. Gyan, and J.-M. Vinassa, J. Power Sources,
325, 273 (2016).
79. T. Waldmann, N. Ghanbari, M. Kasper, and M. Wohlfahrt-Mehrens,
J. Electrochem. Soc.,162, A1500 (2015).
80. L. Cloos, O. Queisser, A. Chahbaz, S. Paarmann, D. U. Sauer, and T. Wetzel,
Batteries & Supercaps,7, e202300445 (2024).
Journal of The Electrochemical Society, 2024 171 040538
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Cyclic aging tests of lithium‐ion batteries are very time‐consuming. Therefore, it is necessary to reduce the testing time by tightening the testing conditions. However, the acceleration with this approach is limited without altering the aging mechanisms. In this paper, we investigate whether and how thermal transients accelerate the aging. The tests are performed on NMC/graphite pouch cells by applying temperatures in a range of 5 °C to 45 °C to the cell surface. The results show, that an accelerated capacity loss can be achieved in comparison to the reference cell at a steady‐state temperature of 25 °C. However, capacity difference analysis (CDA) prognoses a covering layer for the transient cells, which is confirmed upon post‐mortem analysis. We suspect the origin to lie in the dynamics of temperature fields and current distribution during temperature changes when charging. More specifically, areas of higher temperature in the cell lead to high local current densities and plating. Subsequently, high temperatures promote the reaction of the plated lithium with electrolyte. The results show that thermal transients are a critical condition for lifetime and safety and should be treated with caution as they can occur during real life operation.
Article
Full-text available
Low C-rate charge and discharge experiments, plus complementary differential voltage or differential capacity analysis, are among the most common battery characterization methods. Here, we adapt the multi-species, multi-reaction (MSMR) half-cell thermodynamic model to low C-rate cycling of whole-cell Li-ion batteries. MSMR models for the anode and cathode are coupled through whole-cell charge balances and cell-cycling voltage constraint equations, forming the basis for model-based estimation of MSMR half-cell parameters from whole-cell experimental data. Emergent properties of the whole-cell, like slippage of the anode and cathode lithiation windows, are also computed as cells cycle and degrade. A sequential least-square optimization scheme is used for parameter estimation from low-C cycling data of Samsung 18650 NMC∣C cells. Low-error fits of the open-circuit cell voltage (e.g., under 5 mV mean absolute error for charge or discharge curves) and differential voltage curves for fresh and aged cells are achieved. We explore the features (and limitations) of using literature reference values for the MSMR half-cell thermodynamic parameters (reducing our whole-cell formulation to a 1-degree-of-freedom fit) and demonstrate the benefits of expanding the degrees of freedom by letting the MSMR parameters be tailored to the cell under test, within a constrained neighborhood of the half-cell reference values. Bootstrap analysis is performed on each dataset to show the robustness of our fitting to experimental noise and data sampling over the course of 600 cell cycles. The results show which specific MSMR insertion reactions are most responsible for capacity loss in each half-cell and the collective interactions that lead to whole-cell slippage and changes in useable capacity. Open-source software is made available to easily extend this model-based analysis to other labs and battery chemistries.
Article
Full-text available
Lithium‐ion batteries (LIB) are widely used as electrochemical energy storage devices due to their advantages in energy and power density as well as their reliability. One research focus is the cyclic lifetime of LIB, where the influence of thermal conditions during cycling is not yet completely understood. In previous publications, we presented investigations on the cyclic ageing behaviour of LIB under various thermal boundary conditions, including defined temperature gradients and temperature transients. Thereupon, in the present work, the results of cell openings and an extensive post‐mortem study with analyses of scanning electron micrographs (SEM), electrode thickness, X‐ray diffraction (XRD) and inductively coupled plasma optical emission spectroscopy (ICP‐OES) are presented, evaluated and correlated with the thermal conditions. The results reveal significant variations on the electrode and atomic scale for the different thermal boundary conditions during cycling. The electrodes exposed to a temperature gradient exhibit an inhomogeneous distribution of ageing behaviour that directly correlates with the local temperature. Cycling with superimposed transient temperatures reveals fundamentally different ageing effects. With this knowledge, critical temperature conditions can be avoided in applications to prolong cyclic lifetime. This article is protected by copyright. All rights reserved.
Article
Full-text available
Lithium-ion batteries can last many years but sometimes exhibit rapid, nonlinear degradation that severely limits battery lifetime. Here, we review prior work on “knees” in lithium-ion battery aging trajectories. We first review definitions for knees and three classes of “internal state trajectories” (termed snowball, hidden, and threshold trajectories) that can cause a knee. We then discuss six knee “pathways”, including lithium plating, electrode saturation, resistance growth, electrolyte and additive depletion, percolation-limited connectivity, and mechanical deformation, some of which have internal state trajectories with signals that are electrochemically undetectable. We also identify key design and usage sensitivities for knees. Finally, we discuss challenges and opportunities for knee modeling and prediction. Our findings illustrate the complexity and subtlety of lithium-ion battery degradation and can aid both academic and industrial efforts to improve battery lifetime.
Article
A comprehensive understanding of lithium-ion battery (LiB) lifespan is the key to designing durable batteries and optimizing use protocols. Although battery lifetime prediction methods are flourishing, diagnosis of the root causes of aging and degradation have not yet been well developed nor studied for a broad mixture of designs and use cases. Here, we create a machine-learning (ML)-based framework that distinguishes aging modes using multiple electrochemical signatures recorded cycle-by-cycle. The predominant aging behaviors include a combination of loss of active materials in cathode (LAMPE) and a loss of Li inventory (LLI) in Li plating or solid electrolyte interphase (SEI) formation, manifested from 44 batteries representing two cathode chemistries, two electrode loadings, and five charging rates. The aging mode classification accuracy is 86% using features within the first 50 cycles and increases to 88% beyond 225 cycles. The same features can quantify the percentage of end-of-life LAMPE with only 4.3% of error.
Article
We present an extensive analysis of Li-ion battery ageing via Arrhenius plots. The V-shaped Arrhenius plots show minima at an optimum temperature corresponding to the longest cycle-life. The V-shape of the Arrhenius plots signifies the crossover between the two dominating ageing mechanisms – solid electrolyte interphase (SEI) growth in the high temperature range and lithium deposition in the low temperature range. Subjects of our investigations are commercial 5 Ah high energy 21,700-type cells with LiNi0.90Co0.05Al0.05O2 + LiNiO2 (NCA + LNO) cathode and Si/graphite anode (∼3% Si) and 0.1 Ah lab-made pouch cells with LiNi1/3Mn1/3Co1/3O2 (NMC111) cathode and a graphite anode. The results of the Arrhenius plots are analysed in the context of C-rate, cell ageing, and electrode properties. We find that the crossover between the dominating ageing mechanism and hence the optimum operating temperature of the Li-ion cells depend on C-rate, anode coating thickness/particles sizes, the state of health, and the shape of the capacity fade curve. Considering the change of the dominant ageing mechanism can help designing battery systems with longer service life. Additionally, we show a lifetime estimation for temperature dependent cycling of batteries emphasizing the merit of Arrhenius plots in the context of battery cell ageing.
Article
The shape of the open-circuit potential (OCP) curve of silicon–graphite blend electrodes changes during cycle aging due to the faster degradation of the silicon in comparison to the graphite. In this study, the impact of these changes on the open-circuit voltage (OCV) curve of full-cells is investigated. Reconstructing the OCV curve of aged cells by shifting and linearly scaling pristine half-cell OCP curves is an established diagnostic method of determining the degradation modes occurring in lithium-ion cells. We reconstruct the full-cell OCV curves of cycle-aged commercial cells with silicon–graphite anodes using both pristine and aged silicon–graphite OCP curves. Lower estimates are obtained for the loss of anode active material and higher estimates for the loss of both cathode active material and lithium inventory, when aging-induced changes in the shape of the silicon–graphite OCP are considered. Aging-induced changes in the shape of silicon–graphite OCP curves are integrated in the diagnostic method by using a blend electrode OCP model. This not only improves the validity of the determined degradation modes, but also enables the non-destructive estimation of the anode capacity fraction provided by silicon, based on full-cell OCV measurements.