You are on page 1of 19

ll

OPEN ACCESS

Article
Quantification of electrochemical-mechanical
coupling in lithium-ion batteries
Chunhao Yuan, Youngwon
Hahn, Wenquan Lu, Victor
Oancea, Jun Xu

jun.xu@uncc.edu

Highlights
An electrochemical-mechanical
coupling model is developed for
lithium-ion batteries

Electrochemical-mechanical
loading experiments are designed

Constraint leads to decreased


electrode porosity and higher
charge voltage

Reaction force is affected by


mechanical response and
electrochemical (de)lithiation

Developing a method to understand the electrochemical-mechanical coupling


mechanisms in solid-liquid electrode-electrolyte and solid-solid active materials in
lithium-ion batteries is consequential for battery system design. Yuan et al. design
mechanically constrained charge/discharge characterizations with synergetic
efforts from multiphysics modeling to unravel coupling mechanisms in lithium-ion
batteries.

Yuan et al., Cell Reports Physical Science 3,


101158
December 21, 2022 ª 2022 The Author(s).
https://doi.org/10.1016/j.xcrp.2022.101158
ll
OPEN ACCESS

Article
Quantification of electrochemical-mechanical
coupling in lithium-ion batteries
Chunhao Yuan,1,2 Youngwon Hahn,3 Wenquan Lu,4 Victor Oancea,3 and Jun Xu1,2,5,6,*

SUMMARY
Lithium-ion battery safety and durability by nature are dependent
on electrochemical and mechanical coupling. Interdisciplinary ef-
forts are required to understand and quantify coupling behaviors.
Here we design and conduct mechanically constrained charge and
discharge characterizations with efforts supported by multiphysics
modeling to unravel the coupling mechanisms of solid-liquid elec-
trode-electrolyte and solid-solid active materials in lithium-ion bat-
teries. We demonstrate that a lithium-ion battery cell under mechan-
ical constraint exhibits a higher voltage during charging and a
shorter charging time because of increased electrolyte resistance
and decreased diffusivity caused by decreased electrode porosity.
The reaction force response of the cell is a combined result of
the cell structural response mechanically and lithium-ion intercala-
tion/de-intercalation-induced volume variation electrochemically.
Under mechanical constraint, cell capacity is significantly reduced
in fast-charge scenarios; however, it can be recovered by a con-
stant-voltage charge protocol. The results highlight the promise of
multiphysics approaches to unravel the electrochemical-mechanical
coupling mechanisms to direct battery system design and manage-
ment.

INTRODUCTION
To satisfy the increasingly stringent requirement for environmental and energy sus-
tainability, lithium-ion batteries (LIBs) are widely applied because of their relatively
high energy density and long lifespan compared with their counterparts.1–3 As an
energy storage device, a LIB undergoes electrochemical reactions; for example,
lithium (Li)-ion intercalation/de-intercalation in the electrodes, to serve as the power
source.4,5 Such an electrochemical process is deeply coupled with mechanics via
phase transformation of active particles accompanied by volume expansion/ 1Department of Mechanical Engineering and
contraction.6–8 This lithiation/delithiation-induced stress, in turn, also changes over- Engineering Science, The University of North
potential, leading to changed electrochemical behavior of the cell.9,10 The heat gen- Carolina at Charlotte, Charlotte, NC 28223, USA
2Vehicle Energy & Safety Laboratory (VESL), The
eration during this process may also influence the conductivity and diffusivity of the
University of North Carolina at Charlotte,
electrolyte, leading to varied ion transportation behaviors.11–13 Charlotte, NC 28223, USA
3Dassault Systèmes SIMULIA Corp., 1301 Atwood
Pioneering research efforts toward these multiphysics problems were initiated from Ave, Suite 101W, Johnston, RI 02919, USA
4ChemicalSciences and Engineering Division,
a decoupled approach; i.e., focusing on a single physics domain (mechanical, elec-
Argonne National Laboratory, Lemont, IL 60439,
trochemical, or thermal field). For instance, mechanical characterizations have re- USA
vealed the strain rate dependency of the mechanical properties of the cell and 5School of Data Science, The University of North
electrode materials.14–19 Salt concentration, solvent composition, and ambient tem- Carolina at Charlotte, Charlotte, NC 28223, USA
perature were proven to quantitatively influence the transport number, conductivity, 6Lead contact
and diffusivity of the electrolyte via electrochemical characterizations.11–13 Acceler- *Correspondence: jun.xu@uncc.edu
ating rate calorimetry was used to track the temporal evolution of temperature https://doi.org/10.1016/j.xcrp.2022.101158

Cell Reports Physical Science 3, 101158, December 21, 2022 ª 2022 The Author(s). 1
This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
ll
OPEN ACCESS
Article

during heating in an adiabatic chamber.20,21 Based on the experimental character-


izations, mechanical models,16,22,23 electrochemical models,24,25 and thermal
models26,27 were developed to reveal the internal phenomena for a specific physics
domain, such as stress/strain fields, voltage and current profiles, and temperature
evolutions.

More studies tried to touch on LIB multiphysics behavior in a coupled way for consid-
eration of two physics domains to investigate the underlying interactions among
different physics and scales; i.e., coupling of mechanical-electrochemical and
thermal-electrochemical fields. Anode swelling was induced by the electrochemical
process and was directly related to lattice expansion at different states of charges
(SOCs),28 and cathode volume variation was caused by Li insertion/extraction in
active particles, which is determined by the material system.29 The mechanical prop-
erties of the electrode materials have been shown to be state of health (SOH) and
SOC dependent.19,30,31 External mechanical loading was also found to have a
nontrivial influence over the open-circuit voltage of the cell.31,32 In situ experiments
observed significant stress generated on the surface layer of the Li metal anode
during Li plating/stripping.33 In terms of electrical-thermal physics, the nonuniform
temperature distribution in the cell was caused by uneven distribution of current
density34 and charging rates.35 Ambient temperature greatly influenced the thermal
behavior of the cell.35,36

More recently, electro-chemo-mechanical coupling problems were targeted and


investigated in a simplified, weakly coupled manner. Usually, a single mechanical
loading domain serves as the boundary condition for the following electrochemical
and thermal behaviors. Multiphysics computational models were developed to
describe the process from mechanical abuse to internal short circuit (ISC) and, finally,
temperature-induced thermal runaway, where the mechanical model provides the
deformed geometry for electrochemical and thermal models so that these models
are sequentially one way coupled.37–43 Some studies have focused on electro-
chemo-mechanical coupling in LIBs from a strong-coupled perspective at the parti-
cle or electrode level. The concentration variation of particles causes swelling and
shrinkage, also evidenced by the interpretation of electrode thickness.44 The
contact between particles can cause a high nonuniform mechanical stress gradient
during charging, which further influences Li-ion diffusion behavior45 and causes
nonuniform intercalation/de-intercalation currents.10 The charging-induced interfa-
cial failure between particles and surrounding materials increases electrochemical
resistance and degrades the cell’s performance.46,47

In real engineering and application scenarios for LIBs, the cells generally have
mechanical constraints to form battery modules/packs and/or are mechanically con-
nected to rigid frames (e.g., vehicle chassis). Thus, LIBs always suffer from external
mechanical and electrochemical loading during charge/discharge in addition to
multiphysics interactions because of intrinsic behavior. Such complicated scenarios
call for a quantified understanding of electro-chemo-mechanical coupling behaviors
of LIBs at the cell level, which has not yet been investigated thoroughly.

To fill this blank, we investigate the nature of mechanical-electrochemical loading


through designed experiments and developed multiphysics models. Cell charge/-
discharge experiments under external mechanical constraints are conducted in
various scenarios (i.e., galvanostatic charging, constant-voltage charging, and
cycling) to observe the coupled voltage and reaction force responses. Then a
geometrically detailed three-dimensional (3D) multiphysics model (i.e., active

2 Cell Reports Physical Science 3, 101158, December 21, 2022


ll
OPEN ACCESS
Article

Figure 1. Typical voltage and mechanical force responses of the cell under mechanical constraint
at C rates of 0.1 C, 0.5 C, and 1 C
(A) Illustration of the experimental setup and procedures and the interaction between
electrochemical charge and force variation.
(B) Experimental voltage and mechanical force responses of the cell during galvanostatic charging/
discharging with external mechanical constraint. t 0 is the nominal charge time for different C rates;
t 0 = 36,000 s, 7,200 s, and 3,600 s for 0.1 C, 0.5 C, and 1 C, respectively.

material coatings, current collectors, and separator) is developed to quantitatively


investigate in detail the profiles of relevant parameters. Through synergetic efforts
from experimental measurements and numerical modeling, we perform analyses
regarding the mechanical constraint influence on electrochemical behavior, reac-
tion force response because of active materials swelling, cycling effects, and con-
stant-voltage charge effect to deepen the understanding of coupled multiphysics/
multiscale behaviors in LIBs.

RESULTS AND DISCUSSION


Charging and discharging experiments
Charge/discharge experiments of the LIBs under external mechanical constraints
were carried out to investigate the effect of electrochemical-mechanical-coupled
loading on battery performance (more details in Experimental procedures). The
LIBs used in this study are 4,100-mAh pouch cells composed of a LiCoO2 (LCO) cath-
ode, graphite anode, and polyethylene (PE)/Al2O3 separator. In the first step, the
compression platen pre-compressed the cell to 1,000 N (i.e., 0.23 MPa) to elimi-
nate any adverse effect of cell surface roughness on measurement of mechanical
force between the cell and the platen. In the second step, the platen was kept in
the fixed position as in the previous step, and charge/discharge tests between 3 V
and 4.4 V were performed on the cell (Figure 1A). Various charge rates are consid-
ered: 0.1 C, 0.5 C, and 1 C.

Galvanostatic charge/discharge experiments on a cell without any external mechan-


ical constraints serve as baseline experiments, providing a preliminary overview of
the behavior of LIBs in a mechanical stress-free status (i.e., only voltage and temper-
ature responses are available in Figures S1 and S2). The cell temperature Tcell

Cell Reports Physical Science 3, 101158, December 21, 2022 3


ll
OPEN ACCESS
Article

increases during charge/discharge and reaches its maximum value at the end of
discharge. The maximum Tcell is larger at higher charge rates (C rates); specifically,
25.7 C, 32.9 C, and 41.2 C for 0.1 C, 0.5 C, and 1 C, respectively. More details
about the baseline experiment results can be found in the supplemental information.

To provide electrochemical-mechanical-coupled boundary conditions to the cell, we


apply an external mechanical constraint to the cell sample during the charge/
discharge cycle (Figure 1A). To eliminate the influence of different time scales and
simultaneously present an intuitive comparison of various C rates, the normalized
time t/t0 is adopted here (Figure 1B); the charge/discharge time t is normalized by
the nominal charge time at each C rate (i.e., t0 = 36,000 s, 7,200 s, and 3,600 s for
0.1 C, 0.5 C, and 1C, respectively). We observe that the mechanical reaction force
F rises as the LIB is continuously charged, and F reaches the peak value at the end
of charging; upon discharge, F begins to decrease until the end of discharge at all
C rates (C rate = 0.1 C, 0.5 C, and 1 C; Figure 1B). The peak force is greater than
3,000 N for 0.1 C, indicating a strongly developed stress response caused by
charging. The synchronous responses of cell voltage Ecell and mechanical force F
intuitively demonstrate the electro-chemo-mechanical-coupled phenomena. We
analyze this coupling behavior from the mechanisms of Ecell and F responses during
charge/discharge with the assistance of the validated multiphysics model
(Figures S2 and S3). Because the externally applied current mainly governs the elec-
trochemical reaction kinetics in the cell, the charge transfer reaction is faster at a
higher C rate. The electrochemical reaction at a low C rate is close to a quasi-static
process, which can be well described by the model. Higher C rates pose greater
challenges for the model because of the more obvious influence of a faster reaction
rate, more dynamic ion transport and diffusion processes, and larger polarity at the
electrode-electrolyte interface, which should be the limitation of the Newman
battery model (this phenomenon is also common in the available literature). More
efforts to improve the capability of the model considering the fast charging scenario
are needed in our future research.

Voltage response under electrochemo-mechanical loading


During the experiment, after an 1,000 N pre-load, the battery is mechanically con-
strained by the fixed platen and charged/discharged at various C rates. Li-ion inter-
calation-induced anode swelling dominates the battery volume variation during
charging, and F rises because of the constraints of the mechanical fixtures
(Figure 2A). During discharge, cell shrinkage mitigates the mechanical response pro-
duced during the charge process. In terms of electrochemical response, cells with
mechanical constraint show a higher Ecell during charging and a shorter charge
time than ones without constraint at all C rates (Figure 2A), demonstrating that the
external mechanical constraint has a nontrivial effect on the electrochemical
response and the capacity of the battery.

The primitive response of battery swelling under mechanical constraint is geometric


deformation to accommodate the volume variation. The governing parameters of
the deformation are cell porosities, which can be further decomposed into the elec-
trolyte volume fraction εe in cathode εe c and anode εe a . These three vital variables,
at the same time, influence the cell’s electrochemical performance. The profiles of
εe , εe c , and εe a can be monitored and extracted from the validated multiphysics
model. Without any constraint, εe c = 0.189 and εe a = 0.274 remain unchanged dur-
ing charging/discharging (Figures 2B and 2C), serving as the baseline values. When
the external mechanical loading/constraint is imposed, the electrodes are com-
pressed into a denser state, and εe begins to decrease upon pre-loading; i.e., εe c

4 Cell Reports Physical Science 3, 101158, December 21, 2022


ll
OPEN ACCESS
Article

Figure 2. Mechanical constraint effect on battery performance


(A) Comparison of test voltage responses between batteries with and without (w/ and w/o, respectively) constraint.
(B) Cathode porosity values of the battery under constraint in simulation.
(C) Anode porosity values of the battery under constraint in simulation.
(D) Comparison of cathode-side overpotential values between the simulation with or without (w/ and w/o, respectively) stress effect on overpotential.
(E) Comparison of anode-side overpotential values between the simulation with or without stress effect on overpotential.
The time t is normalized by the nominal charge time t 0 ; t 0 = 36,000 s, 7,200 s, and 3,600 s for 0.1 C, 0.5 C, and 1 C, respectively.

and εe a are reduced to 0.148 and 0.25 after 1,000 N pre-loading, respectively.
During charging, the Li ions intercalate into the active particles of the graphite
anode, and the volume of the particles expands, causing swelling of the anode. At
the same time, the cathode particles shrink because of the loss of Li ions. The focus

Cell Reports Physical Science 3, 101158, December 21, 2022 5


ll
OPEN ACCESS
Article

of this study is on out-of-plane deformation (i.e., thickness direction), and in-plane


deformation is not discussed here. The volume expansion of the graphite particles
reaches 13.1%, an order of magnitude larger than the volume contraction of
LCO cathode particles’ 1.9% during charging.29 Thus, the overall volume variation
of the cell is dominated by anode swelling and shrinkage during charging and dis-
charging (with the assumption of similar volumes for the cathode and anode),
respectively. As a result, the electrodes are further squeezed because of the external
mechanical constraint because the cell exhibits expansion during charging. εe con-
tinues to decrease and minimize the value at the end of charging; i.e., εe c is further
reduced to 0.111, 0.107, and 0.102, and εe a is reduced to 0.226, 0.224, and 0.223
for 1 C, 0.5 C, and 0.1 C, respectively (Figures 2B and 2C). To fairly compare the re-
sults at different C rates, the time t here is normalized by the nominal charging time
t0; i.e., t0 = 36,000 s, 7,200 s, and 3,600 s for 0.1 C, 0.5 C, and 1 C, respectively. A
smaller C rate results in a lower εe because of the fact that more capacity is charged
into the cell at a smaller C rate (based on our charging protocol; Experimental pro-
cedures) to cause a larger expansion so that the anode/cathode is more severely
compressed. According to Equations 10 and 11, the decreased εe directly results
in reduced effective electrolyte conductivity keff e and diffusivity Deeff , causing
increased resistance in the electrolyte phase. The cell voltage Ecell can be calculated
by

Ecell = Fs;pos  Fs;neg = Epos  Eneg + DFl (Equation 1)

where Fs;pos and Fs;neg are the solid-phase potentials; Epos and Eneg are the electrode
potential in the cathode and anode, respectively; DFl = icell ,Rsol is the potential drop
over the electrolyte phase between the cathode and anode; icell is the current density
of the cell; and Rsol is the electrolyte resistance. During charging, the current flows
from the cathode to anode in the cell; thus, DFl is positive, and the increased elec-
trolyte resistance results in larger DFl , which finally leads to a higher cell voltage Ecell .
The higher Ecell of the cell with constraint results in achieving the cutoff voltage 4.4V
earlier than a cell without constraint (Figure 2A), which shortens the charge time and
reduces battery capacity in the galvanostatic charge scenario. The fully coupled
electrochemical-mechanical model developed in this study employs a detailed 3D
geometry, which solves the deformation in the x, y, and z directions for all battery
components. Thus, electrode porosity is calculated based on deformation of the
cathode/anode in all directions (Figure S4). The pressure effect on electrode
porosity and cell voltage responses initiates when the external constraint is imposed
on the cell during charging, which is further promoted by the increased initial pre-
compression force (Figure S5). The effect of external pressure on contact resistance
in the cell is significant during the battery manufacturing process,48 but under the
scenario of lithiation-induced deformation, such an effect is dependent on the bat-
tery type and is only significant when contact loss between components happens
because of particle contraction. In this study, commercial LIBs with an LCO cathode
and graphite anode are investigated; we focus on the first charge-discharge cycle of
the fresh cell with good contact between components, and then the pressure effect
on contact resistance is trivial in this case. But the pressure effect on contact resis-
tance should be taken into account when the cycling scenario or high-energy-density
material like Si is included.

Stress produced by mechanical constraint can also influence the electrochemical


overpotential, reflected in the last term  Ush =F (Equation 3 in Experimental proced-
ures). The extracted overpotential values in the cathode/anode show little difference
with/without consideration of the stress effect in the model (Figures 2D and 2E),
revealing that the stress effect on overpotential caused by external mechanical

6 Cell Reports Physical Science 3, 101158, December 21, 2022


ll
OPEN ACCESS
Article

constraint is not significant in this scenario. The stress level caused by charging un-
der constraint is low, 1 MPa, so  Ush =F z -3 3 105 V, which is several orders of
magnitude less than the overpotential value of approximately G0.03 V, indicating
little stress effect on overpotential under the external mechanical scenario.

Mechanical force response under electrochemo-mechanical loading


The mechanical force responses of the cell during charging/discharging under me-
chanical constraints are dominated by the complex electrochemical-mechanical
phenomena in the battery. To obtain a basic understanding of the force response
under pure mechanical loading, we compressed a fully discharged cell to 1,000 N
and then fixed the compression platen to monitor the mechanical force of the cell
without charging/discharging. We observed that F drops from 1,000 N to 815 N
within 200 s and then gradually decreases to the steady stage of 650 N at
18,000 s (Figure 3A). The first important reason is due to the mechanical behavior
of the component materials of batteries. The binder and separator are polymers
with significant stress relaxation behavior under mechanical loading,49,50 and the
active materials show obvious plastic properties,18 which contribute to the force
drop of the whole cell. Another responsible factor is the electrolyte flow in the cell
because of the stress to accommodate possible deformations. In a complementary
experiment, we observed that F decreases to 797 and 538 N in 200 s and 18,000 s,
respectively, for the open-case cell (where the electrolyte can flow out of the cell).
These two representative values are 2.20% and 17.23% smaller than their counter-
parts in the closed-case cell. Therefore, the decreasing mechanical force of the
cell under pure mechanical stress is caused by its structural response, including (1)
the plastic and stress relaxation behavior of the component materials and (2) redis-
tribution of the liquid electrolyte.

Then, we charged/discharged the cells under the fixed compression platen after
1,000 N pre-loading to study the mechanical force response under electrochemical
loading. Upon charging, immediately after pre-compression, F decreases from
1,000 N to 830 N, 886 N, 914 N, and 927 N in 300 s, 75 s, 40 s, and 25 s for 0.1
C, 0.5 C, 1 C, and 2 C, respectively. We may observe that, after the initial force
drop because of the abovementioned structural response, F increases as charging
proceeds, demonstrating that the swelling-induced force increase exceeds the
structural response-caused force decrease. A larger C rate gains a faster force in-
crease (e.g., F begins to increase at 25 s for 2 C and 300 s for 0.1 C), which implies
that a higher C rate causes faster volume expansion because of the larger amount of
intercalated Li ions in a shorter time. F increases nonlinearly, and there is an inflec-
tion stage at mid-charge for all C rates (Figure 3A). The mechanical force increase
during charging is dominated by anode swelling, and anode particle expansion is
caused by Li intercalation. Therefore, the mechanical force is intrinsically related
to the Li concentration in the particle cs. Therefore, the nonlinear evolution of cs (Fig-
ure 3B) dominates the nonlinear increase of F (Figure 3A).

The mechanical force reaches its peak value FP at the end of charging and decreases
nonlinearly to the ending force FE at the end of discharging. FP decreases with
increasing C rate, whereas FE exhibits the opposite trend; FP = 3,460, 3,190,
2,900, and 1,590 N and FE = 280, 350, 450, and 760 N for 0.1 C, 0.5 C, 1 C, and 2
C, respectively (Figure 3C). The charge capacity CC of the cell is linearly related to
the charge time, and CC decreases with increasing C rate: CC =3,995.85,
3,457.78, 2,947.44, and 1,115.66 mAh for 0.1 C, 0.5 C, 1 C, and 2 C, respectively
(Figure 3D), because of the galvanostatic charging protocol. Larger CC indicates
more Li ions intercalated into the anode particles. Thus, the volume expansion of

Cell Reports Physical Science 3, 101158, December 21, 2022 7


ll
OPEN ACCESS
Article

Figure 3. Mechanism of mechanical force response of the cell under mechanical constraint during
charging/discharging with 1,000 N pre-loading
(A) Mechanical force of the cell under mechanical constraint at 0.1 C, 0.5 C, 1 C, and 2 C charging/
discharging and mechanical force of the cell under long-time mechanical constraint without or with
an open battery case.
(B) Li concentration in anode particles at various C rates. Here the concentration is extracted from
the center position of the anode layer through the thickness direction.
(C) Peak mechanical force at the end of charging and ending mechanical force at the end of
discharge under various C rates.
(D) Galvanostatic charge capacity and discharge capacity of the cell under various C rates with a
lowest voltage of 3 V and highest voltage of 4.4 V.
(E) Schematic of the structural response of the cell subject to external mechanical constraint.
The time t in (A) and (B) is normalized by the nominal charging time t 0 ; t 0 = 36,000 s, 7,200 s, 3,600 s,
and 1,800 s for 0.1 C, 0.5 C, 1 C, and 2 C, respectively. The time for curves of the closed case and
open case without charging/discharging is normalized by 36,000 s.

8 Cell Reports Physical Science 3, 101158, December 21, 2022


ll
OPEN ACCESS
Article

the anode and cell are larger at a smaller C rate, causing a higher peak force. At the
end of discharge, the difference between CC and discharge capacity CD becomes
larger at higher C rates (specifically, dC = CC  CD = 14.41, 59.56, 140.65, and
326.98 mAh for 0.1 C, 0.5 C, 1 C, and 2 C, respectively). The remaining capacity is
larger, and, thus, the cell shrinkage during discharging is less than the swelling dur-
ing charging at a higher C rate, which contributes to the larger FE at the end of
discharge for higher discharge C rates. The total charge-discharge time is shorter
for a higher C rate so that the time-dependent structural response for the reduction
of F is also less significant at a higher C rate.

The mechanical force of the cell under mechanical constraint is the coupled result of
the structural response from the mechanical perspective and the Li-ion intercalation/
de-intercalation-induced volume variation from the electrochemical aspect. The
structural response includes the flow of the liquid electrolyte, stress relaxation of
the component materials (mainly the binder and separator), and the porosity varia-
tion (Figure 3E). In this study, the charge/discharge-induced volume variation of
the cell is mainly dominated by electrode particles with a larger swelling ratio, i.e.,
graphite particles in the anode.

Cycling effects
Similarly, there are two types of cycling experiments: galvanostatic charge/discharge with
and without external mechanical constraint. 1 C and 2 C rates with 5 cycles are selected
here. For cells without constraint, in the first cycle, Tcell increases during each cycle and
reaches its peak value at the end of discharge (Figure 4A). Tcell starts to decrease at the
charging beginning from the second cycle, possibly because the heat dissipation by con-
vection Qd is larger than the heat accumulation generated by charging Qg. Higher Tcell
leads to a larger temperature difference with the ambient temperature, resulting in a
stronger thermal convection effect. The highest temperature TH and lowest temperature
TL remain almost unchanged among all cycles for all C rate scenarios. TH = 38.1 C and
TL = 27.8 C for 1 C are lower than TH = 55.3 C and TL = 45.1 C for 2 C (Figure 4A), demon-
strating that cells at a higher C rate produce a larger Qg. Under external mechanical
constraint, the mechanical force of the cell increases during charging and begins to
decrease upon discharging in each cycle (Figure 4B). The external mechanical constraint
imposed on the battery during galvanostatic charging generates a larger cell voltage Ecell
(mainly because of the porosity change, as explained in Figure 2). Thus, the total duration
ttot of five cycles is reduced for constrained cells compared with non-constrained cells.
More specifically, ttot = 27,725 s is reduced by 9.2% to 25,178 s at 1 C, and ttot =
10,066 s is reduced by 56.9% to 4,335 s at 2C (Figure 4C). The significantly reduced
charging time indicates greatly decreased capacities of the cells (Figure 4D), implying
that the mechanical constraint significantly affects the battery capacity, particularly in
fast charging scenarios.

The peak force FP at the end of charging and ending force FE at the end of discharge
of each cycle show opposite trends for cells at 1 C and 2 C. FP and FE decrease with
increasing cycling at 1 C, whereas they increase with cycling at 2 C (Figures 4E and
4F). After the first cycle, CD remains nearly the same as Cc at 1 C. Therefore, the effect
of structural response results in a gradual decrease of FP and FE, as explained in Fig-
ure 3E, as driven by the flow of electrolyte and relaxation effects in binders and sep-
arators. For 2 C, CD is smaller than Cc so that the remaining capacity is accumulated
after each cycle, which results in more pronounced swelling of the cell volume. The
average remaining capacity is 96.21 mAh/cycle, corresponding to a linearly esti-
mated force increase of 50.8 N/cycle (based on the fact that F increases from
1,000 N to 1,590 N with a charge capacity of 1,115.6 mAh at the first cycle of 2

Cell Reports Physical Science 3, 101158, December 21, 2022 9


ll
OPEN ACCESS
Article

Figure 4. Voltage, temperature, and mechanical force responses of the battery considering cycling effects
(A) Voltage and temperature curves of the battery without constraint in five cycles at 1 C and 2 C.
(B) Voltage and mechanical force curves of the battery with constraint in five cycles at 1 C and 2 C.
(C) Comparison of the voltage curves between without and with constraint in five cycles at 1 C and 2 C.
(D) Charge/discharge capacity versus cycling number at 1 C/2 C without or with constraint.
(E) Peak mechanical force versus cycling number of 1 C and 2 C.
(F) Ending mechanical force at discharge end versus cycling number of 1 C and 2 C.

C; Figure 4B). For the pure structural response (Figure 3A), F decreases from 763 N at
836 s (2 C first cycling end) to 702 N at 4,335 s (2 C fifth cycling end); thus, the
average force decrease is 15.25 N/cycle less than the force increase of 50.8 N/cycle.
Such capacity-induced swelling exceeds the effect of structural response and causes
the increasing FP and FE. These results demonstrate that the external mechanical
constraint should be fully considered because it can significantly alter the electro-
chemical-mechanical responses, especially in a fast charging scenario.

Constant-voltage charge effects


A general strategy is to use a constant-voltage charge protocol to achieve maximized
cell capacity. This strategy may also serve as a possible solution to improve the
reduced capacity of the cells under external mechanical constraints. Here we use a con-
stant-current (CC) charge-CC discharge scenario for baseline results (Figure 5). For the
cells without constraint, Tcell increases during CC charging to the first peak

10 Cell Reports Physical Science 3, 101158, December 21, 2022


ll
OPEN ACCESS
Article

Figure 5. Voltage, temperature, and mechanical force responses of the battery considering constant-voltage charging effects
(A) Comparison of cell voltage and temperature responses without constraint between CC charging-CC discharging and CC charging-CV charging-CC
discharging at 1 C and 2 C. CC, constant current; CV, constant voltage.
(B) Comparison of cell voltage and force responses with constraint between CC charging-CC discharging and CC charging-CV charging-CC
discharging at 1 C and 2 C.
(C) Comparison of the cell voltage between cells without and with constraint under CC charging-CV charging-CC discharging at 1 C and 2 C.
(D) Li concentration profiles of the anode particles with constraint during charging.
(E) Charge/discharge capacity C at CC charging, CV charging, and CC discharging stages.

temperature TP1, decreases during constant-voltage (CV) charging to almost ambient


temperature, and then increases again during CC discharging to the second peak tem-
perature TP2 (also the highest temperature) at the end of discharge (Figure 5A). This
phenomenon indicates that the heat generated by CV charging is trivial compared
with the heat dissipated through the thermal convection effect. TP1 and TP2 are
34.8 C and 42.5 C for 1 C and 44.4 C and 57.3 C for 2 C, respectively. For 1 C and
2 C, TP1 is lower than TP2, which demonstrates more heat generated by CC discharging
than by CC charging. We observe an 3 C temperature difference between separate
experiments for CC and CC-CV charging at 1 C in (Figure 5A), which is caused by
reasonable room temperature variation, but the voltage/force responses are rarely
affected by such a temperature difference, and the unlocked electrochemical-mechan-
ical coupling mechanism still holds. When the temperature effect is non-negligible, the
thermal experiment should be carefully designed in a temperature chamber.

Cell Reports Physical Science 3, 101158, December 21, 2022 11


ll
OPEN ACCESS
Article

For the cells with external mechanical constraint, after 1,000-N pre-compression, F
increases upon charging to FP1 at the end of CC charging, and then increases with
a gradually decreasing slope to FP2 at the end of CV charging, and finally, F de-
creases during CC discharge (Figure 5B). As explained above, the mechanical force
is largely dominated by the Li concentration in the active particles cs. Thus, it is
straightforward to observe that the cs shares the same evolution profiles as F, espe-
cially the gradually increasing trend during CV charging (Figure 5D). FP1 and FP2 are
2,810 N and 3,324 N for 1 C and 1,590 N and 3,705 N for 2 C, respectively. The force
increase (FP1-FP2) becomes more evident at 133% at 2C compared with 18% at 1 C.

We may also observe that mechanical constraint reduces the CC charge duration but
increases the CV charge time (Figure 5C). The reduced CC charge period is due to
the larger cell voltage caused by porosity change and reaching 4.4 V early. Because
the mechanical constraint reduces the CC charge capacity CCC-Charging, there is more
capacity available for CV charging; thus, the CV charging period tCV-Charging is longer
for constrained cells (i.e., tCV-Charging increases from 5,075 s to 5,665 s for 1 C and
from 4,939 s to 6,493 s for 2 C).

Generally speaking, two charging protocols are widely used: CC charging-CC dis-
charging (type 1), and CC charging-CV charging-CC discharging (type 2). For type
1, compared with the non-constrained case, CCC-Charging and CCC-Discharging decrease
under mechanical constraint by 8.29% and 9.74% at 1 C and 52.8% and 64.23% at 2
C, respectively (Figure 5E). Again, the capacity at a high C rate is more affected by
the mechanical constraint. For type 2, the CV portion of charging under mechanical
constraint has a larger CCV-Charging than its non-constrained counterpart; i.e., CCV-
Charging increases from 1,022 to 1,260 mAh for 1 C and from 1,744 to 2,864 mAh
for 2 C. At 2 C with constraint, CCV-Charging (2,864 mAh) is more than two times larger
than CCC-Charging (1,313 mAh) and results in a total CC (4,177 mAh) comparable with
the non-constrained case. The total CD is 3,478 mAh, almost 440% of the available
CD (788 mAh) for type 1 (Figure 5E). This demonstrates that the CV charge protocol
can significantly compensate for the capacity reduced by mechanical constraint,
especially at a high C rate.

The electrochemical-mechanical behaviors of LIBs are unique multiphysics/multi-


scale problems that create difficulties in gaining a fundamental understanding of
coupling behaviors in batteries. To gain a quantitative understanding of the electro-
chemo-mechanical coupling in LIBs, we designed an experiment with coupled
loading conditions and established a computational model to explain and under-
stand the origin of the exhibited coupling behaviors. We discovered that externally
mechanically constrained cells exhibit a higher voltage during charging and, thus,
reach the upper cutoff voltage earlier because of the increased electrolyte resistance
caused by the reduced electrode porosity of the swollen active particles. The
mechanical force behavior of the constrained cells during charging/discharging is
the ultimate coupled result of the structural response (i.e., stress relaxation, plastic
behavior, electrolyte flow) from a mechanical perspective and the Li-ion intercala-
tion/de-intercalation-induced volume change from an electrochemical perspective.
The mechanical constraint greatly reduces the available cell capacity during CC
charging-CC discharging, especially under a fast-charging cycling scenario. CV
charging can rescue the cell capacity loss affected by the external constraint.

The results in this study provide a comprehensive and quantitative understanding of


the electrochemical-mechanical coupled behaviors of LIBs, highlight the origin of
the coupling behaviors in LIBs, and shed light on engineering applications for

12 Cell Reports Physical Science 3, 101158, December 21, 2022


ll
OPEN ACCESS
Article

design, assembly, installation, and management of LIBs. To accurately predict the


mechanical behavior of the cell, especially under long-time cycling, the stress relax-
ation effect should be included, which is not considered in the current model and will
be improved in our future research.

EXPERIMENTAL PROCEDURES
Resource availability
Lead contact
Further information and requests for resources should be directed to the lead con-
tact, J.X. (jun.xu@uncc.edu).

Materials availability
This study generated no new materials.

Data and code availability


The datasets used in this study are available from the corresponding author upon
reasonable request.

Charging/discharging without constraint


The LIBs used in this study are composed of an LCO cathode, graphite anode, PE/
Al2O3 separator, and 1:1 ethylene carbonate (EC):diethyl carbonate (DEC) electro-
lyte with a typical capacity of 4,100 mAh and internal resistance of 0.155 U. The up-
per and lower cutoff voltages are 4.4 V and 3 V, respectively. The length, width, and
thickness of the cell are 76 mm, 57.21 mm, and 4.343 mm, respectively. All cells were
pre-discharged at 0.1 C to reach the initial state 3 V before the following experi-
ments. Three charge/discharge scenarios were considered in the study: (1) CC
charging-CC discharging, (2) CC charging-CV charging-CC discharging, and (3)
CC charging-CC discharging in multiple cycles. The charge/discharge tests were
conducted by LANHE CT3001K at different C rates (0.1 C, 0.5 C, 1 C, and 2 C),
and the battery temperature was monitored simultaneously by HIOKI LR8431-30
at three different locations (Figures 6A and 6B). The air temperature was recorded
in a synchronized manner as well. The normal C rate of the cell in this study is 1 C.
For CV charging, the cutoff current value was set as 82 mA (1/50 C). The voltage mea-
surement frequency (i.e., data collection frequency of the cycler, LANHE CT3001K) is
5 Hz, which is usually not high enough to capture the rapid voltage response at the
end of discharge in the experiment, especially at high C rates (like 1 C and 2 C). The
cell voltage at the end of discharge end may be below the cutoff voltage 3V, which
can be improved by the cycler itself with a high measurement frequency. All tests
were conducted at room temperature around 24 C.

Charging/discharging with constraint


Charge/discharge tests of the LIBs under external mechanical constraints were car-
ried out to investigate the effect of electrochemical-mechanical-coupled loading on
battery performance. First, a customized rectangular compression platen pre-com-
pressed the battery to 1,000 N via Instron 5966 to eliminate any negative effect of
surface roughness on the measurement of mechanical force. Then the platen was
kept in the fixed position with 1,000 N pre-loading, and simultaneously the
charge/discharge test was performed on the cell (Figures 6A and 6C). All charge/
discharge scenarios mentioned above were considered here to investigate the me-
chanical constraint effect. The mechanical force between the compression platen
and cell and cell voltage were recorded. Because the compression platen occupied
the battery surface to impose mechanical constraint, the temperature measurement
was not included here.

Cell Reports Physical Science 3, 101158, December 21, 2022 13


ll
OPEN ACCESS
Article

Figure 6. Schematic of the experimental setup


(A) Control system of the equipment.
(B) Battery charging-discharging, with temperature response measured at three positions.
(C) Pre-compression of the battery to 1,000 N and then keeping the compression platen fixed and simultaneously charging-discharging the battery.

Mechanical characterization
Mechanical compression experiments were conducted via Instron 5966 to obtain the
basic constitutive properties of the battery component materials, including the cathode
and anode, as well as the force-displacement (F-d) response of the cell. The electrodes,
disassembled from the cell, were cut into square samples with a side length of 15 mm.
Then 20 layers were stacked and compressed, based on the Instron 2345 platform, at a
loading rate of 2 mm/min. The constitutive stress-strain curves of the cathode and
anode can be calculated from the F-d results obtained in the compression tests, which
serve as the input data of the mechanical model. Calculation details can be found in Fig-
ure S6. Out-of-plane compression of the battery was also performed at a 2 mm/min
loading rate, and the F-d curve (Figure S7) was used to validate the mechanical model.

Numerical simulation
The detailed 3D computational model was established based on the ABAQUS plat-
form to unravel the internal mechanism of electrochemical-mechanical coupling
behavior. Each layer of separator, electrode active material, and current collector,
as well as the tabs, is considered in the model (Figure 7). There are 38 layers of sepa-
rator, 19 layers of anode, and 18 layers and 2 half-layers of cathode (the first half layer
is at the top, and the second half layer is at the bottom). To describe the experi-
ment’s phenomena, coupled electrochemical-mechanical governing equations are
included.

Electrochemical model
The model calculates the solid/liquid-phase potential and Li-ion concentration dur-
ing the charge/discharge process. The electrochemical kinetics at the electrode-
electrolyte interface are governed by the Butler-Volmer equation:25,46,51–54
    
aa Fh ac Fh
IBV = I0 exp  exp  (Equation 2)
RT RT

14 Cell Reports Physical Science 3, 101158, December 21, 2022


ll
OPEN ACCESS
Article

Figure 7. Computational model of the pouch cell with a detailed layer of separator, electrode
current collector, and active material
The 3D geometry of the cell follows a stacking structure, including the cathode (active material and
current collector, 18 layers and 2 half-layers), separator (38 layers), and anode (active material and
current collector, 19 layers), with 10 meshing elements for each layer. The cathode and anode
current collector layers are connected by the positive tab and negative tab, respectively.

  Ush
h = Fs  Fl  Eref cssurf ; T  (Equation 3)
F
where IBV is the current density; I0 is the exchange current density; h is the overpo-
tential; aa and ac are the anodic and cathodic charge-transfer coefficients, respec-
tively; F is Faraday’s constant; R is the gas constant; T is the temperature; Fs is the
solid-phase potential; Fl is the liquid-phase potential; Eref is the open cell potential;
U is the partial molar volume; and sh = tr(s)/3= (s11+s22+s33)/3 is the hydrostatic
stress.

In the solid phase, the solid-phase potential is described by


 
s VFs  as IBV
0 =  V,  keff (Equation 4)

s is the effective electrical conductivity in the solid phase, and as is the active
where keff
surface area per unit electrode volume.

In the liquid phase, Li-ion transport and liquid-phase potential are governed by25,51

0 =  V,Ie + as IBV (Equation 5)


 
vce t+ as IBV
εe =  V,  Deeff Vce + Ie + (Equation 6)
vt F F
   
2RT dlnfG
Ie =  ke VFe 
eff
1+ ð1  t + ÞVlnce (Equation 7)
F dlnce
where ce is the Li-ion concentration; Ie and Fe are the current and electrical potential
in the liquid phase, respectively; t + is the transference number; dlnf G
dlnce is part of the
molar activity coefficient, De is the effective diffusivity in electrolyte; and keff
eff
e is
the effective electrical conductivity in the liquid phase.

At the particle level, Li diffusion is governed by Fick’s law:25,51

Cell Reports Physical Science 3, 101158, December 21, 2022 15


ll
OPEN ACCESS
Article
 
vcs 1 v vcs
=  2 r 2 Dseff Ds (Equation 8)
vt r vr vr
where cs is the Li concentration in the solid phase, and Ds is the diffusion coefficient in
IBV
the solid phase. The boundary conditions are as follows: vcvt = 0jr = 0 , Ds vr = F jr = rp .
s vcs

The effective coefficients are calculated using Bruggeman relationship:46,55,56


εs
keff
s = ks = ε1s + a ks (Equation 9)
ts
εe
e =
keff ke = ε1e + a ke (Equation 10)
te
Deeff = ε1e + a De (Equation 11)
where εs is the volume fraction of the solid phase, εe is the volume fraction of elec-
trolyte (i.e., the porosity), t i = ε
i
a
ði = s; eÞ is tortuosity, and a =0.5 is the Brugge-
man exponent.

Mechanical model
The mechanical model solves the stress and deformation fields. The force equilib-
rium is represented by weak form:
Z Z Z  
vdv
t , dvdS + f ,dvdV = s: dV (Equation 12)
S V vx
where t is force per unit area, f is body force per unit volume, dv is the virtual velocity
field, and s is the Cauchy stress matrix. The deformation gradient matrix F is written as:

F = V,R (Equation 13)


where V is the stretch matrix to completely define the stretch in three orthogonal di-
rections, and R is the rotation to define the rigid body rotation. Li-ion intercalation
into the active particles in the electrode will cause a volume change, which is
described by volumetric expansion as a function of concentration:46

DVp    
DQ = = Q csave  Q csave;initial (Equation 14)
Vp0

where DQ is the volumetric strain; Vp0 and DVp are the initial volume and volume
change, respectively; and Qðcsave Þ and Qðcsave;initial Þ are the current and initial normalized
volume Vp =Vp0 , respectively. The liquid-phase electrolyte is not physically modeled,
but the cathode, anode, and separator are saturated in the electrolyte (i.e., the pores
of electrode/separator are filled with electrolyte), so the liquid electrolyte volume frac-
tion change can be obtained from the volume fraction change of solid-phase compo-
nents based on the total volume fraction of the solid phase and liquid phase being un-
changed. Because of particle swelling, the volume fraction of the solid/liquid phase will
change accordingly, following Dεs = εs DQ and Dεe =  Dεs . The specific active sur-
face area as is also affected by Das = 2as DQ=3. The simplified linear relationship be-
tween volumetric strain and normalized concentration c~s = csave =csmax is defined in the
model. Because the linear swelling ratio of the graphite anode in the electrode thick-
ness direction is 15%,28 which is one-third of the volumetric swelling ratio, the anode
swelling coefficient in the model is defined as 39% after trial and error. The focus of
this study is on the mechanical constraint effect, and the temperature is not considered
in the model. All parameters are summarized in Table S1.

SUPPLEMENTAL INFORMATION
Supplemental information can be found online at https://doi.org/10.1016/j.xcrp.
2022.101158.

16 Cell Reports Physical Science 3, 101158, December 21, 2022


ll
OPEN ACCESS
Article

ACKNOWLEDGMENTS
C.Y. gratefully acknowledges the Graduate School Summer Fellowship at UNC Char-
lotte. W.L. gratefully acknowledges support from the U.S. Department of Energy
(DOE), Vehicle Technologies Office. Argonne is operated for the DOE Office of Sci-
ence by UChicago Argonne, LLC, under contract DE-AC02-06CH11357. J.X. grate-
fully acknowledges support from Dassault Systèmes.

AUTHOR CONTRIBUTIONS
Conceptualization, J.X.; supervision, J.X.; methodology, C.Y., Y.H., V.O., and J.X.;
formal analysis, C.Y., Y.H., V.O., and W.L.; investigation, C.Y., Y.H., V.O., and J.X.;
writing – original draft, C.Y.; writing – review & editing, Y.H., W.L., V.O., and J.X.

DECLARATION OF INTERESTS
Y.H. and V.O., developers/employees of Dassault Systèmes SIMULIA Corp., are au-
thors of the manuscript.

Received: September 1, 2022


Revised: October 2, 2022
Accepted: November 2, 2022
Published: November 30, 2022

REFERENCES
1. Liu, B., Jia, Y., Yuan, C., Wang, L., Gao, X., Yin, 8. Wei, P., Zhou, J., Pang, X., Liu, H., Deng, K., 15. Hu, W., Wang, L., Wang, G., Huan, Y., and Xu, J.
S., and Xu, J. (2020). Safety issues and Wang, G., Wu, Y., and Chen, B. (2014). Effects (2018). Mechanical characterization and
mechanisms of lithium-ion battery cell upon of dislocation mechanics on diffusion-induced modeling for anodes and cathodes in lithium-
mechanical abusive loading: a review. Energy stresses within a spherical insertion particle ion batteries. Interv. Neurol. 7, 265–270. https://
Storage Mater. 24, 85–112. https://doi.org/10. electrode. J. Mater. Chem. 2, 1128–1136. doi.org/10.1016/j.jpowsour.2018.05.007.
1016/j.ensm.2019.06.036. https://doi.org/10.1039/c3ta13925e.
16. Wang, L., Yin, S., and Xu, J. (2019). A detailed
2. Jia, Y., Li, J., Yuan, C., Gao, X., Yao, W., Lee, M., 9. Li, H., Song, Y., Lu, B., and Zhang, J. (2018). computational model for cylindrical lithium-ion
and Xu, J. (2021). Data-driven safety risk Effects of stress dependent electrochemical batteries under mechanical loading: from cell
prediction of lithium-ion battery. Adv. Energy reaction on voltage hysteresis of lithium ion deformation to short-circuit onset. J. Power
Mater. 11, 2003868. https://doi.org/10.1002/ batteries. Appl. Math. Mech. 39, 1453–1464. Sources 413, 284–292. https://doi.org/10.1016/
aenm.202003868. https://doi.org/10.1007/s10483-018-2373-8. j.jpowsour.2018.12.059.

3. Deng, J., Bae, C., Denlinger, A., and Miller, T. 10. Wu, B., and Lu, W. (2017). A battery model that 17. Khattra, N.S., Karlsson, A.M., Santare, M.H.,
(2020). Electric vehicles batteries: requirements fully couples mechanics and electrochemistry Walsh, P., and Busby, F.C. (2012). Effect of time-
and challenges. Joule 4, 511–515. https://doi. at both particle and electrode levels by dependent material properties on the
org/10.1016/j.joule.2020.01.013. incorporation of particle interaction. J. Power mechanical behavior of PFSA membranes
Sources 360, 360–372. https://doi.org/10.1016/ subjected to humidity cycling. J. Power
4. Han, X., Meng, Q., Wan, X., Sun, B., Zhang, Y., j.jpowsour.2017.05.115. Sources 214, 365–376. https://doi.org/10.1016/
Shen, B., Gao, J., Ma, Y., Zuo, P., Lou, S., and j.jpowsour.2012.04.065.
Yin, G. (2021). Intercalation pseudocapacitive 11. Ding, M.S., Xu, K., Zhang, S.S., Amine, K.,
electrochemistry of Nb-based oxides for fast Henriksen, G.L., and Jow, T.R. (2001). Change 18. Wang, L., Duan, X., Liu, B., Li, Q.M., Yin, S., and
charging of lithium-ion batteries. Nano Energy of conductivity with salt content, solvent Xu, J. (2020). Deformation and failure behaviors
81, 105635. https://doi.org/10.1016/j.nanoen. composition, and temperature for electrolytes of anode in lithium-ion batteries: model and
2020.105635. of LiPF6 in ethylene carbonate-ethyl methyl mechanism. J. Power Sources 448, 227468.
carbonate. J. Electrochem. Soc. 148, A1196– https://doi.org/10.1016/j.jpowsour.2019.
5. Hua, W., Schwarz, B., Azmi, R., Müller, M., Dewi A1204. https://doi.org/10.1149/1.1403730. 227468.
Darma, M.S., Knapp, M., Senyshyn, A., Heere,
M., Missyul, A., Simonelli, L., et al. (2020). 12. Valøen, L.O., and Reimers, J.N. (2005). Transport 19. Sonwane, A., Yuan, C., and Xu, J. (2021).
Lithium-ion (de)intercalation mechanism in properties of LiPF6-based Li-ion battery Coupling effect of state-of-charge and strain
core-shell layered Li(Ni, Co, Mn)O2 cathode electrolytes. J. Electrochem. Soc. 152, A882– rate on the mechanical behavior of electrodes
materials. Nano Energy 78, 105231. https://doi. A891. https://doi.org/10.1149/1.1872737. of 21700 lithium-ion battery. J. Electrochem.
org/10.1016/j.nanoen.2020.105231. Energy Convers. Storage 18, 020905. https://
13. Nyman, A., Behm, M., and Lindbergh, G. doi.org/10.1115/1.4049042.
6. Gao, X., Jia, Y., Zhang, W., Yuan, C., and Xu, J. (2008). Electrochemical characterisation and
(2022). Mechanics-driven anode material modelling of the mass transport phenomena in 20. Zheng, S., Wang, L., Feng, X., and He, X. (2018).
failure in battery safety and capacity LiPF6–EC–EMC electrolyte. Electrochim. Acta Probing the heat sources during thermal
deterioration issues. Appl. Mech. Rev. 74. 53, 6356–6365. https://doi.org/10.1016/j. runaway process by thermal analysis of
https://doi.org/10.1115/1.4054566/1141026. electacta.2008.04.023. different battery chemistries. J. Power Sources
378, 527–536. https://doi.org/10.1016/j.
7. McDowell, M.T., Xia, S., and Zhu, T. (2016). The 14. Wang, L., Yin, S., Yu, Z., Wang, Y., Yu, T., Zhao, jpowsour.2017.12.050.
mechanics of large-volume-change J., Xie, Z., Li, Y., and Xu, J. (2018). Unlocking the
transformations in high-capacity battery significant role of shell material for lithium-ion 21. Ren, D., Liu, X., Feng, X., Lu, L., Ouyang, M., Li,
materials. Extreme Mech. Lett. 9, 480–494. battery safety. Mater. Des. 160, 601–610. J., and He, X. (2018). Model-based thermal
https://doi.org/10.1016/j.eml.2016.03.004. https://doi.org/10.1016/j.matdes.2018.10.002. runaway prediction of lithium-ion batteries

Cell Reports Physical Science 3, 101158, December 21, 2022 17


ll
OPEN ACCESS
Article

from kinetics analysis of cell components. Appl. thermal model of pouch battery during normal 46. Xu, R., Yang, Y., Yin, F., Liu, P., Cloetens, P., Liu,
Energy 228, 633–644. https://doi.org/10.1016/j. discharge and internal short circuit process. Y., Lin, F., and Zhao, K. (2019). Heterogeneous
apenergy.2018.06.126. Appl. Therm. Eng. 120, 506–516. https://doi. damage in Li-ion batteries: experimental
org/10.1016/j.applthermaleng.2017.03.135. analysis and theoretical modeling. J. Mech.
22. Greve, L., and Fehrenbach, C. (2012). Phys. Solids 129, 160–183. https://doi.org/10.
Mechanical testing and macro-mechanical 35. Ping, P., Wang, Q., Chung, Y., and Wen, J. 1016/j.jmps.2019.05.003.
finite element simulation of the deformation, (2017). Modelling electro-thermal response of
fracture, and short circuit initiation of cylindrical lithium-ion batteries from normal to abuse 47. Gao, X., and Xu, J. (2021). Multiscale modeling
Lithium ion battery cells. J. Power Sources 214, conditions. Appl. Energy 205, 1327–1344. of electro-chemo-mechanical degradation in
377–385. https://doi.org/10.1016/j.jpowsour. https://doi.org/10.1016/j.apenergy.2017. Si/C core–shell anode for the lithium-ion
2012.04.055. 08.073. battery of high energy density. J. Electrochem.
23. Zhu, J., Li, W., Wierzbicki, T., Xia, Y., and Energy Convers. Storage 18. https://doi.org/
36. Chiew, J., Chin, C.S., Toh, W.D., Gao, Z., Jia, J.,
10.1115/1.4048704.
Harding, J. (2019). Deformation and failure of and Zhang, C.Z. (2019). A pseudo three-
lithium-ion batteries treated as a discrete dimensional electrochemical-thermal model of
layered structure. Int. J. Plast. 121, 293–311. 48. Li, R., Li, W., Singh, A., Ren, D., Hou, Z., and
a cylindrical LiFePO4/graphite battery. Appl.
https://doi.org/10.1016/j.ijplas.2019.06.011. Therm. Eng. 147, 450–463. https://doi.org/10. Ouyang, M. (2022). Effect of external pressure
1016/j.applthermaleng.2018.10.108. and internal stress on battery performance and
24. Doyle, M., Fuller, T.F., and Newman, J. (1993). lifespan. Energy Storage Mater. 52, 395–429.
Modeling of galvanostatic charge and 37. Deng, J., Bae, C., Marcicki, J., Masias, A., and https://doi.org/10.1016/j.ensm.2022.07.034.
discharge of the lithium/polymer/insertion cell. Miller, T. (2018). Safety modelling and testing
J. Electrochem. Soc. 140, 1526–1533. https:// of lithium-ion batteries in electrified vehicles. 49. Chen, Z., Christensen, L., and Dahn, J.R. (2003).
doi.org/10.1149/1.2221597. Nat. Energy 3, 261–266. https://doi.org/10. A study of the mechanical and electrical
1038/s41560-018-0122-3. properties of a polymer/carbon black binder
25. Fuller, T.F., Doyle, M., and Newman, J. (1994). system used in battery electrodes. J. Appl.
Simulation and optimization of the dual lithium 38. Zhang, C., Santhanagopalan, S., Sprague, Polym. Sci. 90, 1891–1899. https://doi.org/10.
ion insertion cell. J. Electrochem. Soc. 141, M.A., and Pesaran, A.A. (2015). Coupled 1002/app.12863.
1–10. https://doi.org/10.1149/1.2054684. mechanical-electrical-thermal modeling for
short-circuit prediction in a lithium-ion cell 50. Wang, Y., Xing, Y., and Li, Q.M. (2022). A hyper-
26. Taheri, P., Yazdanpour, M., and Bahrami, M. under mechanical abuse. J. Power Sources 290, viscoelastic model for battery separators based
(2013). Transient three-dimensional thermal 102–113. https://doi.org/10.1016/j.jpowsour. on inverse-stress-solution parametrical
model for batteries with thin electrodes. 2015.04.162. calibration method. Int. J. Mech. Sci. 225,
J. Power Sources 243, 280–289. https://doi.org/
107361. https://doi.org/10.1016/j.ijmecsci.
10.1016/j.jpowsour.2013.05.175. 39. Yuan, C., Gao, X., Wong, H.K., Feng, B., and Xu, 2022.107361.
J. (2019). A multiphysics computational
27. Coman, P.T., Rayman, S., and White, R.E.
framework for cylindrical battery behavior upon 51. Ecker, M., Käbitz, S., Laresgoiti, I., and Sauer,
(2016). A lumped model of venting during
mechanical loading based on LS-DYNA. D.U. (2015). Parameterization of a physico-
thermal runaway in a cylindrical Lithium Cobalt
J. Electrochem. Soc. 166, A1160–A1169. chemical model of a lithium-ion battery.
Oxide lithium-ion cell. J. Power Sources 307,
https://doi.org/10.1149/2.1071906jes. J. Electrochem. Soc. 162, A1849–A1857.
56–62. https://doi.org/10.1016/j.jpowsour.
2015.12.088. 40. Liu, B., Zhao, H., Yu, H., Li, J., and Xu, J. (2017). https://doi.org/10.1149/2.0541509jes.
Multiphysics computational framework for
28. Zhang, N., and Tang, H. (2012). Dissecting 52. Carmona, E.A., Wang, M.J., Song, Y.,
cylindrical lithium-ion batteries under
anode swelling in commercial lithium-ion Sakamoto, J., and Albertus, P. (2021). The
mechanical abusive loading. Electrochim. Acta
batteries. J. Power Sources 218, 52–55. https:// effect of mechanical state on the equilibrium
256, 172–184. https://doi.org/10.1016/j.
doi.org/10.1016/j.jpowsour.2012.06.071. potential of alkali metal/ceramic single-ion
electacta.2017.10.045.
conductor systems. Adv. Energy Mater. 11,
29. Zhang, F., Huang, Q.-A., Tang, Z., Li, A., Shao,
41. Duan, X., Wang, H., Jia, Y., Wang, L., Liu, B., 2101355. https://doi.org/10.1002/aenm.
Q., Zhang, L., Li, X., and Zhang, J. (2020). A
and Xu, J. (2022). A multiphysics understanding 202101355.
review of mechanics-related material damages
in all-solid-state batteries: mechanisms, of internal short circuit mechanisms in lithium-
ion batteries upon mechanical stress abuse. 53. Song, X., Lu, Y., Wang, F., Zhao, X., and Chen, H.
performance impacts and mitigation
Energy Storage Mater. 45, 667–679. https:// (2020). A coupled electro-chemo-mechanical
strategies. Nano Energy 70, 104545. https://
doi.org/10.1016/j.ensm.2021.12.018. model for all-solid-state thin film Li-ion batteries:
doi.org/10.1016/j.nanoen.2020.104545.
the effects of bending on battery performances.
30. Wang, L., Jia, Y., and Xu, J. (2021). Mechanistic 42. Marcicki, J., Zhu, M., Bartlett, A., Yang, X.G., J. Power Sources 452, 227803. https://doi.org/
understanding of the electrochemo- Chen, Y., Miller, T., L’Eplattenier, P., and 10.1016/j.jpowsour.2020.227803.
dependent mechanical behaviors of battery Caldichoury, I. (2017). A simulation framework
anodes. J. Power Sources 510, 230428. https:// for battery cell impact safety modeling using 54. Ganser, M., Hildebrand, F.E., Klinsmann, M.,
doi.org/10.1016/j.jpowsour.2021.230428. LS-DYNA. J. Electrochem. Soc. 164, A6440– Hanauer, M., Kamlah, M., and McMeeking,
A6448. https://doi.org/10.1149/2.0661701jes. R.M. (2019). An extended formulation of butler-
31. Xu, J., Jia, Y., Liu, B., Zhao, H., Yu, H., Li, J., and volmer electrochemical reaction kinetics
Yin, S. (2018). Coupling effect of state-of-health 43. Li, H., Liu, B., Zhou, D., and Zhang, C. (2020). including the influence of mechanics.
and state-of-charge on the mechanical integrity Coupled mechanical–electrochemical–thermal
J. Electrochem. Soc. 166, H167–H176. https://
of lithium-ion batteries. Exp. Mech. 58, 633–643. study on the short-circuit mechanism of doi.org/10.1149/2.1111904jes.
https://doi.org/10.1007/s11340-018-0380-9. lithium-ion batteries under mechanical abuse.
J. Electrochem. Soc. 167, 120501. https://doi.
55. Doyle, M., Newman, J., Gozdz, A.S., Schmutz,
32. Xu, J., Liu, B., and Hu, D. (2016). State of charge org/10.1149/1945-7111/aba96f.
C.N., and Tarascon, J. (1996). Comparison of
dependent mechanical integrity behavior of
44. Gao, X., Lu, W., and Xu, J. (2020). Modeling modeling predictions with experimental data
18650 lithium-ion batteries. Sci. Rep. 6, 21829. from plastic lithium ion cells. J. Electrochem.
https://doi.org/10.1038/srep21829. framework for multiphysics-multiscale behavior
of Si–C composite anode. J. Power Sources Soc. 143, 1890–1903. https://doi.org/10.1149/
33. Cho, J.H., Xiao, X., Guo, K., Liu, Y., Gao, H., and 449, 227501. https://doi.org/10.1016/j. 1.1836921.
Sheldon, B.W. (2020). Stress evolution in lithium jpowsour.2019.227501.
metal electrodes. Energy Storage Mater. 24, 56. Stephenson, D.E., Hartman, E.M., Harb, J.N.,
281–290. https://doi.org/10.1016/j.ensm.2019. 45. Gao, X., He, P., Ren, J., and Xu, J. (2019). and Wheeler, D.R. (2007). Modeling of particle-
08.008. Modeling of contact stress among compound particle interactions in porous cathodes for
particles in high energy lithium-ion battery. lithium-ion batteries. J. Electrochem. Soc. 154,
34. Chen, M., Bai, F., Song, W., Lv, J., Lin, S., Feng, Energy Storage Mater. 18, 23–33. https://doi. A1146–A1155. https://doi.org/10.1149/1.
Z., Li, Y., and Ding, Y. (2017). A multilayer electro- org/10.1016/j.ensm.2019.02.007. 2783772.

18 Cell Reports Physical Science 3, 101158, December 21, 2022

You might also like