You are on page 1of 17

ll

OPEN ACCESS

Article
Direct capacity regeneration for spent Li-ion
batteries
Nobuhiro Ogihara, Katsuhiko
Nagaya, Hiroyuki
Yamaguchi, ..., Hiroki Kondo,
Tsuyoshi Sasaki, Shinobu
Okayama

ogihara@mosk.tytlabs.co.jp

Highlights
The need to recycle spent Li-ion
batteries will expand with
electrification expansion

Capacity-degraded batteries
were recovered without
dismantling by reagent injection

Potential-controlled anion
radicals realized the capacity
recovery without degradation

The same recovery effect was also


demonstrated in a large practical
battery

Efficient recycling of spent Li-ion batteries is critical for sustainability, especially


with the increasing electrification of industry. This can be achieved by reducing
costly, time-consuming, and energy-intensive processing steps. Our proposed
technology recovers battery capacity by injecting reagents, eliminating the need
for dismantling. The injection treatment of potential-controlled radical anionic
naphthalene into capacity-degraded batteries recovered capacity without
degradation. We have also succeeded in confirming the capacity-recovery effect in
large practical batteries.

Ogihara et al., Joule 8, 1–16


May 15, 2024 ª 2024 The Author(s). Published
by Elsevier Inc.
https://doi.org/10.1016/j.joule.2024.02.010
Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
OPEN ACCESS

Article
Direct capacity regeneration
for spent Li-ion batteries
Nobuhiro Ogihara,1,8,* Katsuhiko Nagaya,2 Hiroyuki Yamaguchi,2 Yasuhito Kondo,3 Yuka Yamada,3
Takahiro Horiba,4 Takeshi Baba,4,6 Nobuko Ohba,4 Shogo Komagata,3 Yoshifumi Aoki,5
Hiroki Kondo,3 Tsuyoshi Sasaki,3 and Shinobu Okayama2,7

SUMMARY CONTEXT & SCALE


With the rapid increase in lithium (Li)-ion battery applications, there Recycling spent batteries is
is growing interest in the circulation of large quantities of spent bat- important to ensure their
teries. However, existing recycling systems require not only several sustainable use. As we shift
processes for recycling itself but also remanufacturing processes, toward electrification, the number
which require increased energy consumption. Here, a recovery re- of spent batteries will increase
agent injection is proposed for regenerating spent batteries. For ca- dramatically. The current
pacity-degraded batteries with reduced electrons and carrier Li+ recycling process involves
ions, Li-naphthalene radical anions with controlled potential based dismantling the batteries to
on solvent dielectric effects are injected, selectively providing recover valuable raw materials
both electrons and carrier Li ions to the cathode and resulting in ca- and resynthesizing them.
pacity recovery without degradation with cycles. Furthermore, the However, this process is time
integration of constant-voltage treatment has successfully demon- consuming, costly, and energy
strated a high-capacity recovery effect in 4 Ah-class practical batte- intensive. It is therefore important
ries. The proposed method is expected to achieve the shortest route to reduce the number of
in battery regeneration and provide new options for circular battery processing steps involved in
systems. recycling. Our solution to this
problem is a battery capacity-
recovery technology that involves
INTRODUCTION
injecting reagents, which is the
The interest in battery recycling stems from political and environmental concerns shortest recycling route that does
regarding production and disposal,1,2 as well as the stable securing of resources not require dismantling. This
in raw materials such as cobalt and natural graphite for Li-ion batteries due to limited method reduces the
reserves or uneven distribution of production areas.3 In the recycling process in Li- environmental impact by
ion batteries, as shown in Figure 1A (and details in Figure S1), after disassembling minimizing the number of
and separating spent batteries, materials, mainly cathodes, are recovered as raw processes involved, increases the
materials through hydro- or pyrometallurgical refining: the former uses a solution value of spent batteries, and
to extract or separate metal ions from metal oxides used as cathode active materials, promotes sustainable battery
while the latter converts metal oxides to metals or metal compounds through heat utilization.
treatment.4–6 The electrode materials are resynthesized using the raw materials ob-
tained by recycling (type I in Figure 1A). Recently, a method called direct recycling
has been proposed to return the recovered electrode materials to their active states
without requiring synthesis from raw materials (type II in Figure 1A).7–9 From the
viewpoint of returning to the battery, these existing methods include not only
the recycling process but also the process of remaking the battery, which requires
the manufacturing time, energy, and cost of all these processes.

There are mainly two ways in which the capacity of Li-ion batteries can deteriorate in
long-term use: (1) damage to the active material (such as irreversible phase-transi-
tion change, particle cracking, and electrical contact loss)10–12 and (2) loss of carrier
ions.13,14 The former is particularly harmful, especially in extreme temperature

Joule 8, 1–16, May 15, 2024 ª 2024 The Author(s). Published by Elsevier Inc. 1
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
OPEN ACCESS Article

conditions,15,16 while the latter is less damaging to the active material but can still
reduce the battery’s capacity. For safety reasons, large-application batteries are
often operated in temperature-controlled systems with air or water,17,18 where the
loss of carrier ions tends to be the dominant mode of degradation. If a battery has
less damage to its active material, an energy-saving process may be recommended
after proper diagnosis and classification. Here, we propose a one-step process suit-
able for batteries with capacity degradation due to loss of carrier ions, which regen-
erates batteries by simply injecting recovered reagents for the degraded batteries
derived from the carrier ion loss, without the previously reported process described
above (type III in Figure 1A).

Typically, carrier Li+ ions are present in the layered oxide cathode during the initial
Li-ion battery production (Figure 1Bi). During the charging process, electrons are
supplied from the cathode with a stable, lower thermodynamic energy level to the
anode with an unstable, higher energy level, while carrier Li+ ions also move for
charge compensation (Figure 1Bii). During this process, carrier Li+ ions are
consumed by irreversible reductive decomposition with the electrolyte at the
graphite anode/electrolyte interface, forming a solid-electrolyte interface (SEI)
film.19 This phenomenon is responsible for the loss of carrier ions. Although the
SEI film prevents further reductive decomposition, long-term cycling gradually con-
sumes the carrier Li+ ion, resulting in a decrease in capacity (Figure 1Biii).

We considered a route to restoring capacity by providing carrier Li+ ions to the cath-
ode (Figures 1Biv and 1Bv). Though the cathode in the capacity-reduced battery has
a high potential, Li+ ions cannot be introduced as carrier Li+ ions by themself due to
a lack of a driving force (Figure 1Ci). Here, our method generates a negatively
charged state by supplying electrons to the cathode through a chemical reduction
reaction while simultaneously supplying carrier Li+ through spontaneous charge
compensation, resulting in capacity recovery (Figure 1Cii). One of the key features
of our approach regarding recovery reagents in this study is the use of alkali metal
arenide reduction reagents for radical anions that donate electrons and Li+ ions and
are often used in organic synthesis.20,21 Lithium naphthalenide (Li-Naph), a typical
substrate, is used as a strong reducing reagent, e.g., in metal-arene complex syn- 1Nobuhiro Ogihara Research Group, Frontier
thesis or two-dimensional sheet synthesis by exfoliation from graphite22 or cargo- Research Management Office, Toyota Central
genide.23 A series of lithium arenides with various potentials depending on their R&D Labs., Inc., Nagakute 480-1192, Japan
2Advanced Battery Development Div., Toyota
molecular structures were prepared (Figure S2),20,24 and this feature has recently
Motor Corporation, Toyota 471-8571, Japan
been used in Li-ion battery applications. For example, biphenylenides with 3Secondary Batteries Research-Domain,
electron-donating units have low reaction potentials—approximately 0.1 V (vs. Li/ Emerging Electrification Technology Div., Toyota
Li+—allowing lithiation to silicon anodes for Li-ion batteries.25,26 In contrast, lithium Central R&D Labs., Inc., Nagakute 480-1192,
Japan
arenides composed of polycyclic aromatic hydrocarbons have high reaction poten-
4Quantum Computing Research-Domain,
tials (approximately 0.8 V for pyrene and approximately 1.3 V for phenylene), Beyond-X Research Div., Toyota Central R&D
allowing selective lithiation to LiFePO4 and LiCoO2 cathode powders after Labs., Inc., Nagakute 480–1192, Japan
disassembly.27 5MaterialsAnalysis & Evaluation
Research-Domain, Emerging Electrification
Technology Div., Toyota Central R&D Labs., Inc.,
Arenides used for battery capacity recovery must selectively act on the cathode, as Nagakute 480–1192, Japan
shown in Figure 1Biv, without degrading the inside of the battery, especially the 6Presentaddress: R-Frontier Div., Frontier
graphite anode that reacts with the arenides leading to the destruction of the Research Center, Toyota Motor Corporation,
Susono 410-1193, Japan
layered structure,22 and for this purpose, control in the high-potential direction is 7Present address: Value Innovation Biz. Div.,
important. Although the reported use of polycyclic aromatic hydrocarbons is inter- Prime Planet Energy & Solutions, Inc., Toyota
esting in terms of high potential control, a different approach is needed for our pur- 471-8572, Japan
8Lead contact
poses, including the expansion of their use, since they cause a dramatic increase in
the raw material price relative to the widespread use of simple naphthalene (Fig- *Correspondence: ogihara@mosk.tytlabs.co.jp
ure S2). Under these circumstances, without using complex framework arenides, https://doi.org/10.1016/j.joule.2024.02.010

2 Joule 8, 1–16, May 15, 2024


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
Article OPEN ACCESS

A Type I

Type II

Coating Assembly
Type III (This work)
Resynthesis
Electrode materials
Electrodes

Batteries

Regeneration
Direct recycling (capacity)
Raw materials

Spent batteries
Pyro- or hydro- Diagnostics
metallurgy Separation Disassembly &
Classification

Spent electrode materials Spent electrodes

B L (ii) (iii) (iv) (v)

Capacity Reduced capacity Capacity recover


(Stable)
Low ← Potential → High
High ← Energy level → Low

Carrier Li+ ion Cathode


Capacity
shift

Side
reaction

Dead
Recovery agent
(Unstable)

Charging Li+ ion put into the cell


Anode

Degradation mechanism Regeneration mechanism

C D
L (ii) /L  Hٕ  /L
Negative
Li-Naph
Li+ O
O O

Tetrahydrofuran Dimethoxyethane
Electron (THF) (DME)
No driving force Li+
 Battery
Reduction Charge compensation
electrolyte components

Figure 1. Capacity recovery for lithium-ion batteries


(A) Battery cycling flow and comparison of proposed and reported processes.
(B) The concept of battery capacity degradation and its recovery are described by the movement of carrier Li+ ions (blue circles) between the potential
profiles of the NCM cathode and graphite anode.
(C) Donation mechanism of electrons and Li+ ions for capacity recovery.
(D) Overview of the reagents considered in this study for battery capacity recovery.

by injecting a common Li-Naph-based recovery reagent into batteries (Figure 1D),


whose reduction potential is controlled above 1.5 V by the dielectric constant effect
of solvation, as shown in Figure S2, we successfully demonstrated capacity recovery
without degradation not only in preliminary small cells but also in practical large
cells. This paper describes the mechanism for battery capacity-recovery reagents us-
ing calculations and basic physical properties, validates the reagent in small cells,
addresses thermodynamic approaches to improve the recovery effect, and finally,
demonstrates the effect in large cells. As can be seen from the comparison in
Figure S1, the proposed direct regeneration method, presented as a type III
approach in Figure 1A, significantly reduces the recycling processes compared
with other battery recycling systems, such as methods using hydro- or

Joule 8, 1–16, May 15, 2024 3


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
OPEN ACCESS Article

pyrometallurgical refining or direct recycling, presented as types I and II ap-


proaches, respectively, and is essential in terms of the circular economy for Li-ion
batteries.28,29

RESULTS AND DISCUSSION


Demonstration of capacity-recovery effect
The capacity-recovery studies with Li-Naph-based recovery reagent injection were
performed using simulated capacity-degraded preliminary laminated-type cells
consisting of LiNi1/3Co1/3Mn1/3O2 (NCM) cathode and graphite anode (see cell
preparation in experimental procedures and Figure S3 for details). The stability of
Li-Naph relies heavily on the type of solvent used, and it is typically made in solvents
such as tetrahydrofuran (THF) or dimethoxyethane (DME),20,30 exhibiting strong
reducing properties with potentials of approximately 0.5 V vs. Li/Li+.26,31 Here, as
a preliminary step in the recovery reagent, we prepared four solutions of the THF
and DME solutions containing 1 mol L 1 Li-Naph and their mixtures with the electro-
lyte of 1.1 mol L 1 LiPF6 in ethylene carbonate (EC)/dimethyl carbonate (DMC)/ethyl
methyl carbonate (EMC) (30:40:30 vol %) used for battery evaluation. The capacity-
recovery results after injection and subsequent rest showed that the capacity was
recovered in the mixed system with Li-Naph/DME + electrolyte, whereas it was
reduced in the non-mixed systems with Li-Naph/THF and Li-Naph/DME (Figure 2A).
In the mixed system using Li-Naph/THF + electrolyte, the capacity recovered after
injection but significantly decreased after rest. The system showing capacity recov-
ery (Li-Naph/DME + electrolyte) shows no drop in the open circuit voltage (OCV) of
the cell after injection (Figure S4), and the subsequent cycle characteristics also show
equivalent performance to that of the system without additives (Figure S5). In
contrast, the systems without capacity recovery (Li-Naph/THF, Li-Naph/DME, and
Li-Naph/THF + electrolyte) show a drop in OCV, and their cycle evaluations also
show a decrease in capacity with the number of cycles (Figure 2B). The charge-
discharge characteristics of the simulated degraded battery reconstructed with fresh
electrolyte were compared with treatment with the recovery reagent (Figure S6).
Even in the simulated capacity-degraded battery containing only new carrier Li+
ions, the capacity does not increase with repeated cycles in the absence of the recov-
ery reagent, indicating that the addition of a substance with both electrons and car-
rier Li+ ions is responsible for the reaction driving force for capacity recovery.

Potential measurements of the four recovery reagents were performed to determine


differences in the recovery effects (Figure 2C). The potential was approximately
0.5 V vs. Li/Li+ in the non-mixed systems with Li-Naph/THF and Li-Naph/DME,
similar to those reported previously,31 and an increase of 1.73 and 1.80 V vs. Li/
Li+ in the mixed systems showing the capacity-recovery effect with Li-Naph/THF +
electrolyte and Li-Naph/DME + electrolyte, respectively. To confirm the effect of
the prepared recovery reagents, the change in Li content of the charged NCM cath-
ode and graphite anode was measured using inductively coupled plasma-optical
emission spectrometry (ICP-OES) with samples immersed in the Li-Naph/THF + elec-
trolyte. The results of the Li content showed a selective increase in the NCM, while
the graphite remained unchanged (Figure 2D). The increase in Li content in NCM
varied linearly with the square root of the immersion time (Figure S7), suggesting
a one-dimensional diffusion-type mechanism,32 in which Li supply proceeds from
the electrode surface.

Based on the above considerations and the potential tendencies of each material,
as shown in Figure 2E, the highly reductive Li-Naph with low potential caused

4 Joule 8, 1–16, May 15, 2024


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
Article OPEN ACCESS

Figure 2. Confirmation of battery capacity recovery


(A) Discharge curves before and after injection of recovery reagent and after rest. Inset: the color of
injected recovery reagent.
(B) Capacity retention with cycling after recovery treatment.
(C) Measured potential change of Li-Naph vs. concentration.
(D) Effect of capacity-recovery reagent on NCM and graphite electrodes using 1 M Li-Naph/THF +
electrolyte. The gray dotted line shows the theoretical maximum Li amount at the NCM electrode.
(E) The potential relationship of the capacity-recovery reagent to cathode and anode. Graphite
reacts at approximately 0.8 V, including electrolyte decomposition, and compared to Li-Naph with
a potential of approximately 0.4 V, charge transfer occurs from the higher-energy Li-Naph to the
lower-energy graphite.

degradation by acting on the graphite anode, whereas the high-potential-control


Li-Naph + electrolyte mixture induced a recovery effect by selectively providing
electrons and carrier Li+ ions to the cathode while suppressing anode degradation.
As for the difference between THF and DME in the mixture, it is likely due to the
negative effect of the residual activity after the recovery effect. The activity of naph-
thalenide, which exists as a radical anion,20 was examined using electron spin reso-
nance (ESR). After the evaluation of the recovery tests, the electrolyte was removed

Joule 8, 1–16, May 15, 2024 5


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
OPEN ACCESS Article

from the cell, and its ESR results showed a large peak in the Li-Naph/THF + electro-
lyte (Figure S8), indicating the high activity of naphthalenide. On the other hand, a
small peak was observed for the Li-Naph/DME + electrolyte. This suggests that
the presence of residual naphthalenide after capacity recovery causes degradation
during rest and cycling.

Mechanism of potential change in recovery reagents


To reveal the mechanism of the potential increase of Li-Naph in the electrolyte
mixture related to the capacity-recovery effect, the energy of the highest occupied
molecular orbital (HOMO) of Li-Naph in solution was calculated using density func-
tional theory (DFT). The HOMO energy is related to the potential for electron-
donating properties,33,34 and its larger negative values correspond to the high-po-
tential direction.

First, to investigate the stable structure for Li-Naph in solution, the DFT calculations
were performed under a series of conditions: the type of coordination molecule, the
oxygen coordination number to Li (denoted On, where n is the oxygen coordination
number), and the dielectric constant in the polarizable continuum model using the
integral equation formalism variant (IEFPCM)35 (see Table S1 for detailed parame-
ters). Note the change in interatomic distance (Figure 3A) and charge (Figure 3B)
from O3 to O4. Li moves away from Naph and closer to the solvents (Figure 3A).
The formula charges of Li and Naph are +1 and 1, respectively, and considering
the difference between the formula charge and the calculated value as the charge
contributing to the interaction, the interaction for Li increases while that for Naph de-
creases (Figure 3B). At short Li-Naph distances below O3, Naph is stable because of
the dominance of Li+-p interaction36–38 (as shown in Figure 3C). However, at O4, it
becomes unstable due to the decrease in Li+-p interaction caused by the Li+-solva-
tion dominance at long Li-Naph distances.

Next, the contribution factors to the HOMO energy shift were quantified by machine
learning using partial least squares (PLS) regression in structural models where the
solvent oxygen coordination number is less than 3 (Figure S9). The machine learning
model with the parameters listed in Table S2 as explanatory variables for the HOMO
energy of the objective function showed a good correlation between the DFT calcu-
lated values and the machine learning predictions (Figure 3D). According to the ma-
chine learning predictions, the ordinates of the contributions of the explanatory vari-
ables estimated from the standard regression coefficients (b weights)39–41 indicate
that the inverse of the dielectric constant and the inverse of the Li-Naph distance
have a significant impact on the change in HOMO energy (Figure 3E). This suggests
that the HOMO energy associated with the potential of Li-Naph is related to the
electrostatic potential (F) of a point charge (Q) at a specific dielectric constant (ε)
and distance (r), as described by Gauss’ law:42,43

Q
F = (Equation 1)
4pεr
This equation implies a linear relationship F to the inverse of the dielectric constant
(1/ε), and indeed, the calculated HOMO energy shows a linear relationship to 1/ε in
integral equation formalism variant (IEFPCM) (Figures S10 and S11). The measured
potentials of the prepared solutions using the recovery effect study in Figure 2 also
exhibited a linear relationship to the 1/ε of the mixed solvent (Figure 3F), confirming
the above law regarding the relationship between potential and dielectric constant
in actual solutions. The primary reason for the significant potential shift is the
increased dielectric constant of the added electrolyte. Based on DFT calculations,

6 Joule 8, 1–16, May 15, 2024


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
Article OPEN ACCESS

Figure 3. Estimation of recovery reagent structure


(A and B) Dependence of Li-Naph and Li-O distances (A) and Li and Naph charges (B) on the number of coordinating oxygen atoms to Li obtained from
DFT calculations using box-and-whisker diagrams (see Table S1 for detailed values). The orange line is the median value.
(C) Schematic illustration of the stabilized structure model obtained from DFT calculations in Li-Naph and neighboring solvents for the number of
coordinating solvents.
(D) The relationship between the predicted HOMO energy obtained from the regression and the actual HOMO energy.
(E) Ordering of the contribution of each parameter from the standard regression coefficients (b weights) in PLS to the HOMO energy obtained from the
DFT calculations in PLS regression using machine learning.
(F) Relation of the measured potential in each prepared solution to the inverse of the dielectric constant for the mixed solvents. The volume ratio of the
Li-Naph solution and the mixed solution is 1:1.
(G) The structure showing the high potential (low HOMO energy) in the DFT calculation in IEFPCM for mixed solvents (Li-Naph, cis, and trans-DME
molecules). Transparent green and blue are HOMO orbitals.

Joule 8, 1–16, May 15, 2024 7


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
OPEN ACCESS Article

the solvent compositions near Li-Naph composed of cis- and trans-DME molecules
with chelate coordination, as shown in Figure 3G, exhibit a larger negative HOMO
energy corresponding to a higher potential in the solvent mixture of IEFPCM than
the solvent composition composed of THF molecules (see Table S1). This trend
is consistent with the potential relationship in the actual electrolyte mixture in
Figure 2C.

The solution exhibiting the above high potential was prepared by synthesizing Li-
Naph using ether-based solvents and then mixing it with the electrolyte. However,
the formation reaction of Li-Naph using Li metal and Naph did not proceed when
a high-dielectric-constant carbonate solvent was used instead of the ether-based
solvents. This also supports that the mixing effect on the above high potential is
mainly due to the influence of the high-dielectric-constant solvent from the sur-
roundings on the solvation center consisting of Li-Naph with ether-based solvents,
such as in the polarization continuum model, rather than uniform mixing of all
components.

Constant-voltage treatment
Based on the left panel of Figure 4A, the reaction between the Li-Naph-based recov-
ery reagent and the NCM cathode can be viewed as a reaction of Gibbs energy
(DrG). This is determined by the reaction progression (x) obtained from the electro-
chemical potential (mA / mB) in the Gibbs energy change of the system under con-
stant pressure (p) and temperature (T), which can be expressed as follows:44
 
vG
Dr G = (Equation 2)
vx p;T

DrG is related to the potential difference in the total reaction (DE) as follows.

Dr G
DE = (Equation 3)
nF
where n and F are the number of reacting electrons and the Faraday constant, respec-
tively. As shown in the right panel of Figure 4A, a large negative value in the reaction
Gibbs energy (DrG < 0) indicates a high spontaneous reaction, and a value closer to
zero (DrG / 0) indicates an equilibrium state. Therefore, when the treated cell was
left at the OCV after Li-Naph injection (OCV treatment), electrons and Li+ ions
were supplied to the cathode, as shown in Figure 4B, and DE decreased as the reac-
tion proceeded. As a result, the reaction driving force decreased (DrG / 0). In
contrast, when the treated cell was set at a constant-voltage (CV) treatment, as shown
in Figure 4C, the electrons donated to the NCM cathode from Li-Naph were supplied
to the anode via an external circuit to maintain the voltage, resulting in a large reac-
tion driving force (DrG < 0). The expectation of this method would be basically inde-
pendent of the cathode active material species if the potential difference between Li-
Naph and the cathode can be secured.

This treatment difference was found to have a significant effect on the capacity re-
covery in preliminary cells with high-density electrodes and in practical cells. The
latter is discussed in detail later in this paper. Under OCV treatment, the capacity
-ecovery effect confirmed by Li-Naph in the cells using low-density electrodes ( Fig-
ure 4Di) disappeared in the cells using high-density electrodes ( Figure 4Dii) with
high internal resistance, as shown in Figure S12. This suggests the effect of reduced
mass supply due to high-density electrodes. The capacity-recovery effect was
confirmed when the cells with high-density electrodes were treated with CV at

8 Joule 8, 1–16, May 15, 2024


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
Article OPEN ACCESS

B C

D E

Figure 4. Constant-voltage treatment for highly effective capacity recovery


(A) Relationship between reaction progress (left) and Gibbs free energy in open-circuit-voltage (OCV) and constant-voltage (CV) treatments (right).
(B and C) Movement of carrier Li+ ions (blue circles) and changes in the potential difference (DE) during capacity-recovery reactions in OCV (B) and CV
treatments (C).
(D) Comparison of discharge curves before and after injection of recovery reagent by OCV and CV treatments in batteries with electrodes of different
electrode densities.
(E) Summary of CV capacities under various conditions and recovery capacities for CV treatment.

4.0 V (Figure 4Diii), indicating that CV treatment is effective for high-energy-density


batteries with high-density electrodes.

The results of capacity-recovery studies for high-density electrode cells under


various conditions—including CV value, type of electrolyte mixture, and concentra-
tion (detailed conditions are listed in Table S3)—demonstrated a positive correlation
between the capacity at CV (Figure S13) and the recovered capacity (Figure S14),
with a larger CV capacity indicating a higher recovery effect (Figure 4E). Higher Li-
Naph concentrations led to greater recovery capacities. Although the change in

Joule 8, 1–16, May 15, 2024 9


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
OPEN ACCESS Article

electrolyte type did not affect the capacity-recovery effect, the use of imide electro-
lytes such as lithium bis(fluorosulfonyl)imide (LiFSI) or lithium bis(trifluoromethane-
sulfonyl)imide (LiTFSI) provided favorable cycle characteristics after capacity recov-
ery ([10] and [11] in Figure S15), the same as those without the addition of the
electrolyte. The recovery reagent consisting of Li-Naph/THF + electrolyte leads to
significant capacity degradation after recovery in OCV treatment (capacity retention
of approximately 55% at the 20th cycle as shown in Figure 2B), but favorable
improvement in CV treatment (capacity retention of approximately 80% at the
100th cycle as shown in [3] and [4] of Figure S15). The obvious difference in the cycle
performance between the two treatments suggests an electrochemical deactivation
effect in CV treatment for the residual active Li-Naph that leads to cycle degradation.
To investigate possible negative effects due to the presence of the recovery agent
inside the cell, charge-discharge polarization resistances (Rp) under each condition
were compared (Figure S16). The results showed that the THF solvent-based recov-
ery reagent tended to have higher resistance than the DME-based reagent, and the
DME-based reagent showed lower resistance than the cell before injection under
some conditions. One possible cell safety concern is the inclusion of low-boiling-
point ether solvents, but the boiling point, flash point, and vapor pressure of DME
are nearly identical to those of DMC, which is used as the electrolyte solvent for
Li-ion batteries. Although the low-flash-point solvent ratio is higher, there is no spe-
cial battery safety reduction. Furthermore, the ether solvents have recently been re-
ported as electrolyte solvents for Li-ion batteries to improve low-temperature
performance.45–48

Large-cell demonstration
Using capacity-degraded large 4 Ah-class square Li-ion cells consisting of a
graphite-based anode and NCM-based cathode with carbonate-based electro-
lyte,49 as shown in the inset of Figure 5A, the capacity-recovery effects were demon-
strated in the OCV and CV treatments (Figure S17). After injection of the recovery
reagent, the capacity increased with cycling in both treatments without observing
noticeable gas production (Figure 5A), with the CV treatment achieving a larger ca-
pacity-recovery effect of approximately 4% (Figure S18A). The utilization efficiency
of the added recovery reagent was higher in the large cells (approximately 75.3%
and 71.1% at CV and OCV treatments, respectively, as shown in Figure S18B), in
which penetration of the recovery reagent seems more difficult, than in the small
cells (approximately 25%). The charge-discharge profiles after the injection of the re-
covery reagent with cycles showed a gradual increase in capacity without any change
in charge-discharge polarization for both the CV (Figure 5B) and OCV treatments
(Figure S19). The differential capacitance dQ/dV plots in the charge-discharge pro-
files indicated a pair of peaks at 3.4–3.6 V (red dotted box in Figure 5C) and two pairs
of peaks at 3.6–3.8 V (blue dotted box in Figure 5C). As determined in previous re-
ports,50 the former and latter peaks correspond to the Li insertion reaction of the
graphite anode and NCM cathode, respectively. In addition, the respective peaks
in the differential capacitance dQ/dV plots shifted toward lower voltages as the ca-
pacity recovered (Figures 5D and 5E). To investigate the observed peak shift in the
differential capacitance dQ/dV plots, simulations of the elimination of the capacity
shift were performed with the respective charge curves of the cathode and anode
(Figure S20). The peak shift of the differential capacitance dQ/dV plot in the exper-
imental results was in good agreement with that in the simulation results for the
origin of capacity degradation elimination.

Fourier transform infrared spectroscopy (FTIR) analysis of NCM cathodes removed


from cells after recovery tests (Figure 5F) proved that the recovery reagent

10 Joule 8, 1–16, May 15, 2024


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
Article OPEN ACCESS

Figure 5. Demonstration in a practical battery


(A) Capacity vs. cycles for CV and OCV treatments using the recovery reagent Li-Naph/DME + electrolyte and no treatment for capacity recovery. Inset:
appearance of the 4 Ah-class lithium-ion square cells used in the demonstration.
(B) Charge-discharge curves for each cycle after CV treatment. Inset: enlarged view of the area indicated by the dotted line.
(C) Differential capacity dQ/dV plots against each cycle after CV treatment.
(D and E) Detail of the differential capacity dQ/dV plots by red (D) and blue (E) dotted lines in (C).
(F) Description of the FTIR analysis points on the cathode after CV-treatment evaluation.
(G) FTIR spectra of each location in the cathode after CV-treatment evaluation shown in (F). The sample washed with EMC was confirmed to be free of
naphthalene, thus confirming that the naphthalene observed without washing was adsorbed on the electrode.

naphthalene penetrated deep into the inside of the large cells (Figure 5G). The deep
penetration of the capacity-recovery reagent supports the observed recovery effect
and high utilization efficiency of the recovery reagent. In large Li-ion cells, the occur-
rence of reversible volume expansion and contraction, including electrode and elec-
trolyte, during charging and discharging has been reported.51–53 Therefore, the re-
sults of the capacity increase associated with cycling once the recovery reagent was
fed into the large cell suggested that the reagent was spread inside the cell using the
volume expansion and contraction during cycling as the driving force. The unique
behavior of the large cell demonstrates that the utilization of the recovery reagent
in the large cell was higher than that in the small cell.

Conclusion
Our study showcased a method for regenerating spent battery capacity through
direct injection of a reagent containing lithium arenide. We achieved this by control-
ling the reaction potential of the recovery reagent through the dielectric effect of the
solvent and implementing a CV process based on the Gibbs energy of the reaction.
This technique proved effective not only in small cells but also in practical cells.
Considering that arenides that exhibit potentials between 1.0 V and 1.5 V, which
allow capacity recovery without battery degradation, are very expensive, as shown
in Figure S2, we believe that the naphthalenide compound whose potential is
controlled by the dielectric constant is suitable for processing large volumes for
spent batteries. Moreover, the proposed method would be equally effective, under

Joule 8, 1–16, May 15, 2024 11


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
OPEN ACCESS Article

the same mechanism, not only for spent batteries but also for the initial capacity loss
caused by side reactions in manufacturing. The large variations in cell capacity loss
will also need to be combined with advanced diagnostic techniques in order to take
appropriate action. While achieving high-capacity-recovery effects may require
further exploration of reagent compositions including high concentrations that
enhance electron- and Li-ion-donating properties, cell geometries, and injection
methods, our proposal reduces energy consumption by significantly cutting down
the battery regeneration process, reduces warming-causing CO2 emissions, and
will be a promising low-process-load technology that can motivate sustainable bat-
tery circulation in the future.

EXPERIMENTAL PROCEDURES
Resource availability
Lead contact
Further information and requests for resources should be directed to and will be ful-
filled by Nobuhiro Ogihara (ogihara@mosk.tytlabs.co.jp).

Materials availability
All of the materials are available for purchase, with the exception of the large cell that
was used for the demonstration.

Data and code availability


The Python modules handle raw data parsing and feature extraction, while
JupyterLab is used for data analysis.

Li-Naph preparation
Anhydrous THF and DME (Kishida Chemical, Japan) were used as solvents. Li-Naph
was prepared using a previously described procedure.54 Naphthalene (Wako Pure
Chemical Industries, Japan) was dissolved in THF or DME to prepare a 1.0 mol L 1
solution; then, an equimolar amount of lithium metal (Honjo Metal Co., Ltd., Japan)
was added, and the solution was stirred. The color of the solution turned dark green
after adding lithium (Li-Naph/THF and Li-Naph/DME in the inset of Figure 2A). The
resulting Li-Naph solution was mixed with LiPF6-based, LiTFSI-based, and LiFSI-
based electrolytes, as shown below, to a ratio of 1:1 to generate a recovery reagent.
The color of the resulting solution changed to reddish brown (Li-Naph/THF + elec-
trolyte and Li-Naph/DME + electrolyte in the inset of Figure 2A).

Electrode and electrolyte preparation


LiNi1/3Co1/3Mn1/3O2 (NCM (Toda Kogyo Corp., Japan) cathodes were prepared by
coating a dispersion composed of NCM, carbon black (Denka, Japan) as a conducting
agent, and polyvinylidene fluoride (PVDF) as a binder (Kureha, Japan) in N-methyl-2-
pyrrolidone on aluminum foil. The compositions of the active material, conductive
carbon, and PVDF were 88:8:4 wt % and 92:5:3 wt % for the low- and high-density
electrodes, respectively. Graphite anodes were also prepared by coating a dispersion
composed of active material (Osaka Gas Chemical, Japan), carboxymethyl cellulose
as the thickening agent (CMC, Daicel, Japan), and styrene-butadiene rubber as the
binder (1 wt %, SBR, JSR, Japan) in water on a copper foil. The composition of
graphite, CMC, and SBR was 98:1:1 wt %. Prior to electrochemical cell fabrication,
both electrodes were dried at 120 C under vacuum for at least 10 h. The loading
weights of the cathode and anode were 4.4 and 3.2 mg cm 2, respectively. The elec-
trolytes were dissolved in a carbonate-based mixture of EC, DMC, and EMC (30:40:30
vol %) with a concentration of 1.1 mol L 1 LiPF6, and 1.1 mol L 1 LiTFSI and LiFSI were
used as electrolytes for mixing with Li-Naph.

12 Joule 8, 1–16, May 15, 2024


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
Article OPEN ACCESS

Cell preparation
The electrochemical properties of the respective conditions were examined using
two types of cells: a small 6 mAh class laminated type with a 10 cm2 electrode
area assembled by employing a separator filled with electrolyte in an argon-filled
glove box and a capacity-degraded large 4.6 Ah-class square type manufactured
as a practical battery.49 Prior to the recovery reagent injection test in the preliminary
laminated-type cell based on a graphite anode/NCM cathode cell (Figure S3A), a Li
anode/NCM cathode cell was fabricated and charged up to 50% of the electrode ca-
pacity (3 mAh) at 0.6 mA corresponding to a 10-h charging rate (detailed charge
curve is shown in Figure S3B). A simulated capacity-degraded cell with the NCM
cathode taken from this cell and a fresh graphite anode was then assembled with
the separator filled with the electrolyte (detailed charge and discharge curves are
shown in Figure S3C). The large cell was examined using a degraded cell with a ca-
pacity of 3.8 Ah, which corresponds to a capacity loss of approximately 20%, by con-
ducting a cycle test simulating a long-distance run49 with an initial capacity of 4.6 Ah.

Galvanostatic charge-discharge measurements


All preliminary laminated-type cells were charged to 4.1 V and discharged to 3.0 V at
1.2 mA corresponding to a 5-h charging rate after the recovery reagent injection
test. After that, the cells were charged to 4.1 V at 1.2 mA, followed by a CV charge
for 5 h, and then discharged to 3.0 V at 0.6 mA. The discharge capacity achieved in
the second cycle was used as the recovered capacity. In the cycling test, the same
cells were cycled 100 times between 3.0 and 4.1 V at 1.8 mA. AC impedance mea-
surements (Solatron 1260/1286, England) of the same cells were performed at the
open-circuit potential. The frequency was varied from 100 kHz to 100 mHz with a
perturbation amplitude of 10 mV (peak to peak). All electrochemical measurements
were performed at 25 C. The 4 Ah-class square cells were charged to 4.1 V at a con-
stant current of 2.0 A and a CV of 1 h and discharged to 3.0 V at the same constant
current and CV time. The cycling test after injection of the recovery reagent was
cycled between 3.0 and 4.1 V at 2.0 A.

Recovery agent charging process


In the OCV treatment using the laminated-type cells, the cell with confirmed
capacity was charged to 4.0 V at 1.2 mA, the cell was partially opened in an
argon-filled glove box, and 0.5 mL of recovery reagent was placed inside the
cell using a micropipette. After the recovery reagent was added, the cell was
immediately sealed in an argon-filled glove box and allowed to rest for 40 or
120 h. In the CV treatment using small cells, the same procedure used for the
OCV treatment was followed until the injection and sealing of the recovery re-
agent, followed by constant current-CV charging at 1.2 mA to the respective volt-
ages. The termination conditions were 2 h of CV or 5 mAh. For the treatment using
the 4 Ah-class square cells, after injecting and sealing approximately 16 g of recov-
ery reagent, the cells were charged at 2.0 A corresponding to a 2-h charging rate
to 3.75 V, then rested for 25 h for the OCV treatment and held voltage for 25 h for
the CV treatment. The cells were then charged at 2.0 A to 4.1 V, held at the
voltage for 1 h, and discharged to 3.0 V for 10 or 20 cycles, with an OCV rest
for 1 day after the 3.6 V charge during the cycle test.

Characterization
The quantity of Li in the cathode and anode immersed in 1.0 mol L 1 Li-Naph/
THF + electrolyte was confirmed by ICP-OES (CIROS 120EOP, Rigaku, and
SPECTRO). The radical anion activity of Li-Naph was confirmed by ESR (Elexsys
E500, Bruker). The electrolyte solution containing the recovery reagent was

Joule 8, 1–16, May 15, 2024 13


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
OPEN ACCESS Article

extracted from the cell after the test, and 100 mL of the sample solution was
placed in a flat cell for ESR measurement. The spectra of the samples were re-
corded at a microwave power of 3.17 mW (18 dB), microwave frequency of 9.6
GHz, modulation frequency of 100 kHz, modulation amplitude of 1 Gauss with a
time constant of 30 ms, and 10 scans. All FTIR spectra of the cathode after the ca-
pacity recovery test in the large cell were recorded on a Thermo Nicolet Avatar 360
FTIR spectrophotometer. Prior to FTIR measurements, the electrode samples were
taken from the electrochemical cell after evaluation and cut in an argon-filled glove
box. As a comparison, one electrode sample washed with EMC was also prepared.

Computational methods
The calculations of the structural optimization and the HOMO energies were
performed using the Gaussian 16 package (Revision C.01)55 based on the DFT
method. The geometrical optimization of the model except for the small compounds
was performed using the uB97X-D56 functional and the Def2-SV(P) basis set.57 The
single point calculation of the optimized model and all calculations for small
compounds were performed with the same functional and the Def2-TZVPPD basis
set.57,58 We also adopted the solvation effect via the polarizable continuum model
using the IEFPCM.35 In this study, THF, PC, and the same mixture of solvents
(EC:DMC:EMC = 30:40:30 vol %) as the electrolyte used in the experiment were
considered. Since the parameters for EC and mixed solvent in the IEFPCM model
were unknown, the solvents were simply assumed using the dielectric constant at
25 C from the previously reported literature.59 The molecular structures and orbitals
were visualized using Materials Studio 2021 software.60

Machine learning prediction


The contribution of the molecular model parameters to the potential shift was per-
formed by machine learning using PLS regression in Python’s scikit-learn.61 Using the
calculation results obtained by the DFT method (Table S1), the HOMO energy
related to the potential was used as the objective variable and the parameters of
the molecular model, shown in Table S2, as explanatory variables. PLS regression
was performed using K-split cross-validation after the normalization of the explana-
tory variables. The parameters for the number of components and splits in the PLS
regression were optimized by grid search. The accuracy of the obtained machine
learning model was evaluated based on the root mean square error (RMSE) and
R2 (coefficient of determination), and conditions were searched so that the smaller
RMSE and R2 were close to 1. The ordinal order of the parameters contributing to
the HOMO energy shift was determined from the standard regression coefficients
(b-weights) in PLS.39–41

SUPPLEMENTAL INFORMATION
Supplemental information can be found online at https://doi.org/10.1016/j.joule.
2024.02.010.

ACKNOWLEDGMENTS
The authors thank the members of the Secondary Batteries Research-Domain of
Toyota Central R&D Labs., Inc. for helpful comments and technical support.

AUTHOR CONTRIBUTIONS
N. Ogihara, K.N., and S.O. conceived the idea and designed the experiments. H.Y.
and Y.Y. conducted large-cell experiments. T.B. and N. Ohba performed the DFT
calculations. Y.K. conducted the potential measurements. Y.A. conducted the ESR

14 Joule 8, 1–16, May 15, 2024


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
Article OPEN ACCESS

analysis. T.H., S.K., and T.S. helped with the cell fabrication and electrochemical
characterizations. H.K. performed the dismantling treatment after the large-cell
evaluation. N. Ogihara prepared the manuscript. All authors discussed the results
and commented on the manuscript.

DECLARATION OF INTERESTS
The paper includes a patent application.

Received: September 20, 2023


Revised: November 10, 2023
Accepted: February 14, 2024
Published: March 8, 2024

REFERENCES
1. Notter, D.A., Gauch, M., Widmer, R., Wäger, P., 11. Li, T., Yuan, X.-Z., Zhang, L., Song, D., Shi, K., 20. Holy, N.L. (1974). Reactions of the radical
Stamp, A., Zah, R., and Althaus, H.J. (2010). and Bock, C. (2020). Degradation Mechanisms anions and dianions of aromatic hydrocarbons.
Contribution of Li-Ion Batteries to the and Mitigation Strategies of Nickel-Rich NMC- Chem. Rev. 74, 243–277.
Environmental Impact of Electric Vehicles. Based Lithium-Ion Batteries. Electrochem.
Environ. Sci. Technol. 44, 6550–6556. Energ. Rev. 3, 43–80. 21. Shindo, M., Koretsune, R., Yokota, W., Itoh, K.,
and Shishido, K. (2001). Practical synthesis of
2. Farjana, S.H., Huda, N., and Mahmud, M.A.P. 12. Edge, J.S., O’Kane, S., Prosser, R., Kirkaldy, ynolate anions: naphthalene-catalyzed
(2019). Life cycle assessment of cobalt N.D., Patel, A.N., Hales, A., Ghosh, A., Ai, W., reductive lithiation of a,a-dibromo esters.
extraction process. J. Sustain. Min. 18, Chen, J., Yang, J., et al. (2021). Lithium ion Tetrahedron Lett. 42, 8357–8360.
150–161. battery degradation: what you need to know.
Phys. Chem. Chem. Phys. 23, 8200–8221. 22. Jung, H., Yang, S.J., Kim, T., Kang, J.H., and
3. Baum, Z.J., Bird, R.E., Yu, X., and Ma, J. (2022). Park, C.R. (2013). Ultrafast Room-Temperature
Lithium-Ion Battery Recycling–Overview of 13. Birkl, C.R., Roberts, M.R., McTurk, E., Bruce, Reduction of Graphene Oxide to Graphene
Techniques and Trends. ACS Energy Lett. 7, P.G., and Howey, D.A. (2017). Degradation with Excellent Dispersibility by Lithium
712–719. diagnostics for lithium ion cells. J. Power Naphthalenide. Carbon 63, 165–174.
Sources 341, 373–386.
4. Harper, G., Sommerville, R., Kendrick, E., 23. Zheng, J., Zhang, H., Dong, S., Liu, Y., Nai, C.T.,
Driscoll, L., Slater, P., Stolkin, R., Walton, A., 14. Pastor-Fernández, C., Yu, T.F., Widanage, Shin, H.S., Jeong, H.Y., Liu, B., and Loh, K.P.
Christensen, P., Heidrich, O., Lambert, S., et al. W.D., and Marco, J. (2019). Critical review of (2014). High Yield Exfoliation of Two-
(2019). Recycling Lithium-Ion Batteries from non-invasive diagnosis techniques for Dimensional Chalcogenides Using Sodium
Electric Vehicles. Nature 575, 75–86. quantification of degradation modes in lithium- Naphthalenide. Nat. Commun. 5, 2995.
ion batteries. Renew. Sustain. Energy Rev. 109,
5. Emilsson, E., and Dahllöf, L. (2019). Lithium-Ion 24. Slates, R.V., and Szwarc, M. (1965). Dissociative
138–159.
Vehicle Battery Production - Status 2019 on Equilibria in the Systems Aromatic
Energy Use, CO2 Emissions, Use of Metals, 15. Sasaki, T., Nonaka, T., Oka, H., Okuda, C., Itou, Hydrocarbon[UNK],Na+ % Radical Anion[UNK] +
Products Environmental Footprint, and Y., Kondo, Y., Takeuchi, Y., Ukyo, Y., Tatsumi, Na+. J. Phys. Chem. 69, 4124–4131.
Recycling. 978-91-7883-112-8 (ISBN). Sven. K., and Muto, S. (2009). Capacity-fading
Miljöinstitutet. http://urn.kb.se/resolve? 25. Yoshida, S., Masuo, Y., Shibata, D., Haruta, M.,
mechanisms of LiNiO2-based lithium-ion Doi, T., and Inaba, M. (2015). Li Pre-doping of
urn=urn:nbn:se:ivl:diva-132. batteries: I. Analysis by electrochemical and Amorphous Silicon Electrode in Li-
6. Fan, E., Li, L., Wang, Z., Lin, J., Huang, Y., Yao, spectroscopic examination. J. Electrochem. Naphthalene Complex Solutions.
Y., Chen, R., and Wu, F. (2020). Sustainable Soc. 156, A289. Electrochemistry 83, 843–845.
Recycling Technology for Li-Ion Batteries and
16. Muto, S., Sasano, Y., Tatsumi, K., Sasaki, T., 26. Jang, J., Kang, I., Choi, J., Jeong, H., Yi, K.W.,
Beyond: Challenges and Future Prospects.
Horibuchi, K., Takeuchi, Y., and Ukyo, Y. (2009). Hong, J., and Lee, M. (2020). Molecularly
Chem. Rev. 120, 7020–7063.
Capacity-Fading Mechanisms of LiNiO2-Based Tailored Lithium–Arene Complex Enables
7. Shi, Y., Zhang, M., Meng, Y.S., and Chen, Z. Lithium-Ion Batteries: II. Diagnostic Analysis by Chemical Prelithiation of High-Capacity
(2019). Ambient-Pressure Relithiation of Electron Microscopy and Spectroscopy. Lithium-Ion Battery Anodes. Angew. Chem. Int.
Degraded LixNi0.5Co0.2Mn0.3O2 (0 < x < 1) via J. Electrochem. Soc. 156, A371. Ed. Engl. 59, 14473–14480.
Eutectic Solutions for Direct Regeneration of
Lithium-Ion Battery Cathodes. Adv. Energy 17. Tran, T.-H., Harmand, S., and Sahut, B. (2014). 27. Wu, C., Hu, J., Ye, L., Su, Z., Fang, X., Zhu, X.,
Mater. 9, 1900454. Experimental investigation on heat pipe Zhuang, L., Ai, X., Yang, H., and Qian, J. (2021).
cooling for Hybrid Electric Vehicle and Electric Direct Regeneration of Spent Li-Ion Battery
8. Wang, T., Luo, H.M., Bai, Y.C., Li, J.L., Vehicle lithium-ion battery. J. Power Sources Cathodes via Chemical Relithiation Reaction.
Belharouak, I., and Dai, S. (2020). Direct 265, 262–272. ACS Sustainable Chem. Eng. 9, 16384–16393.
Recycling of Spent NCM Cathodes through
Ionothermal Lithiation. Adv. Energy Mater. 10, 18. Chen, J., Kang, S., E, J., Huang, Z., Wei, K., 28. Curtis, T.L., Smith, L., Buchanan, H., and Heath,
2001204. Zhang, B., Zhu, H., Deng, Y., Zhang, F., and G. (2021). A Circular Economy for Lithium-Ion
Liao, G. (2019). Effects of different phase Batteries Used in Mobile and Stationary Energy
9. Xu, P., Yang, Z., Yu, X., Holoubek, J., Gao, H., Li, change material thermal management Storage: Drivers, Barriers, Enablers, and U.S.
M., Cai, G., Bloom, I., Liu, H., Chen, Y., et al. strategies on the cooling performance of the Policy Considerations. https://www.osti.gov/
(2021). Design and Optimization of the Direct power lithium ion batteries: a review. J. Power biblio/1768315.
Recycling of Spent Li-Ion Battery Cathode Sources 442, 227228.
Materials. ACS Sustainable Chem. Eng. 9, 29. Roy, J.J., Rarotra, S., Krikstolaityte, V., Zhuoran,
4543–4553. 19. Ogihara, N., Igarashi, Y., Kamakura, A., Naoi, K.W., Cindy, Y.D.I., Tan, X.Y., Carboni, M.,
K., Kusachi, Y., and Utsugi, K. (2006). Meyer, D., Yan, Q., and Srinivasan, M. (2021).
10. Palacı́n, M.R. (2018). Understanding ageing in Disordered carbon negative electrode for Green Recycling Methods to Treat Lithium-Ion
Li-ion batteries: a chemical issue. Chem. Soc. electrochemical capacitors and high-rate Batteries E-Waste: A Circular Approach to
Rev. 47, 4924–4933. batteries. Electrochim. Acta 52, 1713–1720. Sustainability. Adv. Mater. 34, e2103346.

Joule 8, 1–16, May 15, 2024 15


Please cite this article in press as: Ogihara et al., Direct capacity regeneration for spent Li-ion batteries, Joule (2024), https://doi.org/10.1016/
j.joule.2024.02.010

ll
OPEN ACCESS Article

30. Castillo, M., Metta-Magaña, A.J., and Fortier, differentiates ‘‘ground-glass’’ opacities due to 51. Lee, J.H., Lee, H.M., and Ahn, S. (2003). Battery
S. (2016). Isolation of gravimetrically COVID-19 from acute non-COVID-19 lung dimensional changes occurring during charge/
quantifiable alkali metal arenides using disease. Sci. Rep. 11, 17237. discharge cycles—thin rectangular lithium ion
18-crown-6. New J. Chem. 40, 1923–1926. and polymer cells. J. Power Sources 119–121,
41. Delli Pizzi, A., Chiarelli, A.M., Chiacchiaretta, P., 833–837.
31. Tan, K.S., and Yazami, R. (2015). Physical- d’Annibale, M., Croce, P., Rosa, C.,
Chemical and Electrochemical Studies of the Mastrodicasa, D., Trebeschi, S., Lambregts, 52. Wang, X., Sone, Y., Segami, G., Naito, H.,
Lithium Naphthalenide Anolyte. Electrochim. D.M.J., Caposiena, D., et al. (2021). MRI-based Yamada, C., and Kibe, K. (2007). Understanding
Acta 180, 629–635. clinical-radiomics model predicts tumor Volume Change in Lithium-Ion Cells during
response before treatment in locally advanced Charging and Discharging Using In Situ
32. Wright, R.B., Motloch, C.G., Belt, J.R., rectal cancer. Sci. Rep. 11, 5379. Measurements. J. Electrochem. Soc. 154,
Christophersen, J.P., Ho, C.D., Richardson, A14–A21.
R.A., Bloom, I., Jones, S.A., Battaglia, V.S., 42. Ida, N. (2015). Gauss’s Law and the Electric
Henriksen, G.L., et al. (2002). Calendar- and Potential. In Eng. Electromagn (Springer Int. 53. Oh, K.-Y., Siegel, J.B., Secondo, L., Kim, S.U.,
cycle-life studies of advanced technology Publ.), pp. 139–229. Samad, N.A., Qin, J., Anderson, D., Garikipati,
development program generation 1 lithium- K., Knobloch, A., Epureanu, B.I., et al. (2014).
ion batteries. J. Power Sources 110, 445–470. Rate dependence of swelling in lithium-ion
43. Panofsky, W.K., and Phillips, M. (2005). Classical
cells. J. Power Sources 267, 197–202.
33. Dance, I. (2006). The Correlation of Redox electricity and magnetism (Courier
Potential, HOMO Energy, and Oxidation State Corporation). 54. Ogihara, N., Ohba, N., and Kishida, Y. (2017).
in Metal Sulfide Clusters and Its Application to On/Off Switchable Electronic Conduction in
Determine the Redox Level of the FeMo-co 44. Atkins, P.W., De Paula, J., and Keeler, J. (2023). Intercalated Metal-Organic Frameworks. Sci.
Active-Site Cluster of Nitrogenase. Inorg. Atkins’ physical chemistry (Oxford University Adv. 3, e1603103.
Chem. 45, 5084–5091. Press).
55. Frisch, M.J., Trucks, G.W., Schlegel, H.B.,
34. Bender, T.P., Graham, J.F., and Duff, J.M. 45. Jia, H., Xu, Y., Burton, S.D., Gao, P., Zhang, Scuseria, G.E., Robb, M.A., Cheeseman, J.R.,
(2001). Effect of substitution on the X., Matthews, B.E., Engelhard, M.H., Zhong, Scalmani, G., Barone, V., Petersson, G.A.,
electrochemical and xerographic properties of L., Bowden, M.E., Xiao, B., et al. (2020). Nakatsuji, H., et al. (2016). Gaussian 16. Rev.
triarylamines: correlation to the Hammett Enabling Ether-Based Electrolytes for Long C 01.
parameter of the substituent and calculated Cycle Life of Lithium-Ion Batteries at High
HOMO energy level. Chem. Mater. 13, Charge Voltage. ACS Appl. Mater. Interfaces 56. Chai, J.D., and Head-Gordon, M. (2008). Long-
4105–4111. 12, 54893–54903. range corrected hybrid density functionals with
damped atom–atom dispersion corrections.
35. Tomasi, J., Mennucci, B., and Cammi, R. (2005). 46. Ren, X., Zhang, X., Shadike, Z., Zou, L., Jia, H., Phys. Chem. Chem. Phys. 10, 6615–6620.
Quantum Mechanical Continuum Solvation Cao, X., Engelhard, M.H., Matthews, B.E.,
Models. Chem. Rev. 105, 2999–3093. Wang, C., Arey, B.W., et al. (2020). Designing 57. Weigend, F., and Ahlrichs, R. (2005). Balanced
Advanced In Situ Electrode/Electrolyte basis sets of split valence, triple zeta valence
36. Vollmer, J.M., Kandalam, A.K., and Curtiss, L.A. Interphases for Wide Temperature Operation and quadruple zeta valence quality for H to Rn:
(2002). Lithium Benzene Sandwich of 4.5 V Li||LiCoO2 Batteries. Adv. Mater. 32, Design and assessment of accuracy. Phys.
Compounds: A Quantum Chemical Study. e2004898. Chem. Chem. Phys. 7, 3297–3305. https://doi.
J. Phys. Chem. A 106, 9533–9537. org/10.1039/B508541A.
47. Zhang, S.S. (2020). Reformulation of Electrolyte
37. Rao, J.S., Zipse, H., and Sastry, G.N. (2009). for Enhanced Fast-Charge Capability of Li-Ion 58. Rappoport, D., and Furche, F. (2010). Property-
Explicit Solvent Effect on Cation p Battery. J. Electrochem. Soc. 167, 60527. optimized Gaussian basis sets for molecular
Interactions: A First Principle Investigation. response calculations. J. Chem. Phys. 133,
J. Phys. Chem. B 113, 7225–7236. 134105.
48. Wang, Z., Wang, H., Qi, S., Wu, D., Huang, J.,
38. Vijay, D., and Sastry, G.N. (2008). Exploring the Li, X., Wang, C., and Ma, J. (2022). Structural 59. Hall, D.S., Self, J., and Dahn, J.R. (2015).
size dependence of cyclic and acyclic regulation chemistry of lithium ion solvation for Dielectric Constants for Quantum Chemistry
p-systems on cation–p binding. Phys. Chem. lithium batteries. Ecomat. 4, e12200. and Li-Ion Batteries: Solvent Blends of Ethylene
Chem. Phys. 10, 582–590. Carbonate and Ethyl Methyl Carbonate.
49. Nagai, H., Morita, M., and Satoh, K. (2016). J. Phys. Chem. C 119, 22322–22330.
39. Abdi, H. (2010). Partial least squares regression Development of the Li-ion Battery Cell for
and projection on latent structure regression Hybrid Vehicle. SAE Technical Paper Series. 60. Systèmes, D. (2021). Materials Studio 2021
(PLS Regression). WIREs Computational. Stat. (BIOVIA).
2, 97–106. 50. Jung, R., Metzger, M., Maglia, F., Stinner, C.,
and Gasteiger, H.A. (2017). Oxygen release 61. Pedregosa, F., Varoquaux, G., Gramfort, A.,
40. Delli Pizzi, A., Chiarelli, A.M., Chiacchiaretta, P., and its effect on the cycling stability of Michel, V., Thirion, B., Grisel, O., Blondel, M.,
Valdesi, C., Croce, P., Mastrodicasa, D., Villani, LiNixMnyCozO2 (NMC) cathode materials for Li- Prettenhofer, P., Weiss, R., and Dubourg, V.
M., Trebeschi, S., Serafini, F.L., Rosa, C., et al. ion batteries. J. Electrochem. Soc. 164, A1361– (2011). Scikit-learn: Machine learning in Python.
(2021). Radiomics-based machine learning A1377. J. Mach. Learn. Res. 12, 2825–2830.

16 Joule 8, 1–16, May 15, 2024

You might also like