You are on page 1of 9

RESEARCH ARTICLE

www.advenergymat.de

Stepwise Dopant Selection Process for High-Nickel Layered


Oxide Cathodes
Do-Hoon Kim, Jun-Hyuk Song, Chul-Ho Jung, Donggun Eum, Byunghoon Kim,
Seong-Hyeon Hong, and Kisuk Kang*

1. Introduction
NCM-based lithium layered oxides (LiNi1–x–yCoxMnyO2) have become
prevalent cathode materials in state-of-the-art lithium-ion batteries. Higher Recently, high-nickel NCM
energy densities can be achieved in these materials by systematically (LiNi1–x–yCoxMnyO2, 1 – x – y ≥ 0.8) cathode
increasing the nickel content; however, this approach commonly results in materials have garnered considerable
research interest owing to their high spe-
inferior cycle stability. The poor cycle retention of high-nickel NCM cathodes
cific capacities (>200 mA h g−1) and have
is generally attributed to chemo-mechanical degradation (e.g., intergranular been extensively applied in state-of-the-
microcracks), vulnerability to oxygen-gas evolution, and the accompanying art lithium-ion batteries.[1,2] It has been
rocksalt phase formation via cation mixing. Herein, the feasibility of doping widely observed that a higher nickel con-
strategies is examined to mitigate these issues and effective dopants for tent in NCM cathodes leads to an increase
in the usable specific capacity of layered
high-nickel NCM cathodes are theoretically identified through a stepwise
cathodes in a given voltage window and is
pruning process based on density functional theory calculations. Specifically, practically beneficial as the nickel replaces
a sequential three-step screening process is conducted for 38 potential toxic and high-cost cobalt.[3–8] Neverthe-
dopants to scrutinize their effectiveness in mitigating chemo-mechanical less, this approach of increasing the nickel
lattice stress, oxygen evolution, and cation mixing at charged states. content is typically hindered by inferior
Using this process, promising dopant species are selected rationally and a cycle performance resulting from the
chemo-mechanical failure of high-nickel
silicon-doped LiNi0.92Co0.04Mn0.04O2 cathode is synthesized, which exhibits
NCM particles during extended cycling.
suppressed lattice expansion/contraction, fewer intergranular microcracks, Recent studies have revealed that the rapid
and reduced rocksalt formation on the surface compared with its undoped capacity fade in these materials is cor-
counterpart, leading to superior electrochemical performance. Moreover, related to intergranular cracking in sec-
a comprehensive map of dopants regarding their potential applicability is ondary particles[9–11] and the surface phase
transition[12,13] at a high state of charge.
presented, providing rational guidance for an effective doping strategy for
In particular, the large variation of the
high-nickel NCM cathodes. c-lattice parameter upon de/lithiation has
been reported as the main cause of inter-
granular cracking and the corresponding secondary particle
D.-H. Kim, J.-H. Song, C.-H. Jung, D. Eum, B. Kim, S.-H. Hong, K. Kang destruction.[14–16] The large c-lattice parameter change upon
Department of Materials Science and Engineering
Research Institute of Advanced Materials (RIAM) repetitive cycling induces anisotropic contraction and expan-
Seoul National University sion among randomly oriented primary particles, consequently
1 Gwanak-ro, Gwanak-gu, Seoul 08826, Republic of Korea leading to mechanical stress and crack propagation in sec-
E-mail: matlgen1@snu.ac.kr ondary particles.[14,15] Not surprisingly, parasitic side reactions
K. Kang continue to occur at the interface between the newly exposed
Center for Nanoparticle Research
Institute for Basic Science (IBS)
particle surface and the electrolyte, and the accumulation of the
Seoul National University by-products over cycling increases the cell impedance and grad-
1 Gwanak-ro, Gwanak-gu, Seoul 08826, Republic of Korea ually leads to degradation of the cell performance.[9,15]
K. Kang The low thermal stability of high-nickel NCM materials is
School of Chemical and Bioengineering another drawback preventing their practical adoption. A series
Institute of Engineering Research of studies have indicated that the onset temperature of oxygen
College of Engineering
Seoul National University evolution at charged states substantially decreases as the nickel
1 Gwanak-ro, Gwanak-gu, Seoul 151-742, Republic of Korea content increases in NCM compositions.[17,18] The low onset
temperature of oxygen evolution intensifies the risk of thermal
The ORCID identification number(s) for the author(s) of this article runaway of the battery system at charged states, making high-
can be found under https://doi.org/10.1002/aenm.202200136. nickel NCM materials unreliable with respect to safety.[19] In
addition, oxygen evolution is known to trigger the reconstruc-
DOI: 10.1002/aenm.202200136 tion of the surface structure of NCM particles, accompanying

Adv. Energy Mater. 2022, 12, 2200136 2200136 (1 of 9) © 2022 Wiley-VCH GmbH
16146840, 2022, 18, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202200136 by Konkuk University Sanghuh Memorial Library, Wiley Online Library on [01/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advenergymat.de

Figure 1. Schematic illustration of stepwise pruning process for selection of dopants in high-nickel NCM. The number of elements considered in each
step are displayed.

significant cation mixing[19–21] and the corresponding phase mitigate microcrack generation, enhance the thermal stability,
transition from a layered to a spinel or rock-salt structure.[22–24] and suppress the surface reconstruction involving cation disor-
These types of surface reconstruction are generally known to dering. The stepwise pruning process covers 38 elements in the
increase the charge-transfer resistance, resulting in the build- periodic table, including previously proposed dopants. Inspired
up of voltage hysteresis and premature cell failure.[22] by the dopant selection results, we experimentally synthesize
One common strategy to overcome these chronic issues has silicon-doped high-nickel NCM (LiNi0.92Co0.04Mn0.04O2) and
been to slightly modify the composition of high-nickel NCM confirm that the doped-LiNi0.92Co0.04Mn0.04O2 electrode under-
materials via elemental doping while preserving the high nickel goes smaller lattice changes during cycling and consequently
content in the layered structure.[25–30] Numerous dopants have suffers less from microcrack formation in the secondary parti-
been previously explored in both experimental and theoretical cles, as evidenced by in situ X-ray diffraction (XRD), scanning
case studies with the goals of mitigating microcrack forma- electron microscopy (SEM), and scanning transmission elec-
tion,[25,26] suppressing oxygen-gas generation,[27–29] and/or tron microscopy (STEM) analyses. Moreover, these changes in
reducing surface phase transformation.[29,30] Although some of the material behavior are demonstrated to lead to an enhanced
these attempts have resulted in notable performance enhance- electrochemical performance with prolonged capacity retention
ments,[31–34] the optimal dopant species/compositions that can and higher rate capability compared with that of the undoped
simultaneously resolve the aforementioned difficulties remain system. We believe that our strategy and findings enrich the
elusive. Moreover, there have been ambiguities and/or incon- fundamental understanding of the effects of diverse dopants
sistent results regarding the roles of dopants even for similar and are expected to provide valuable guidance for the explora-
compositional NCM electrode materials.[35–38] This issue arises tion of potential dopants for stable high-nickel NCM materials.
partly because during the doping process, contributing factors
other than intrinsic dopant effects may arise and be dominant,
which would have a substantial effect on the overall cathode 2. Results and Discussion
performance. For example, the conventional doping process
often induces a change in the primary particle/grain micro- In an effort to identify the potential dopant candidates and
structure[25,26,39] and sometimes leaves conductive carbon resi- their efficacies, we designed a three-step pruning process for
dues from using dopant precursors on the particle surface.[40,41] the dopant selection by considering the degradation mecha-
The additional heating process necessary for certain dopants nism of high-nickel NCM materials, as schematically shown in
may also alter the surface nature of the pristine NCM particles, Figure 1. For 38 dopant candidates covering Group 13–15 ele-
which can result in a significant change in the electrochemical ments and 3d to 5d metals in the periodic table, we selected
performance regardless of the presence of dopants.[42–44] Fur- the anisotropic lattice parameter variation as the first criterion
thermore, the distribution/amount of dopants in the structure/ in the pruning process, preferentially selecting candidates that
grain boundaries critically depends on the dopant species and can reduce the lattice strain during de/lithiation of the layered
the doping processes. These complexities make it challenging structure. Next, for the dopants that met the first criterion, the
to clearly understand the intrinsic effect of each dopant when effect on the oxygen-gas evolution was evaluated by estimating
they are incorporated into the NCM layered structure; further the change in the formation energy of oxygen vacancies due
clarification is indispensable in establishing rational and effec- to the presence of the dopant. In this step, we selected ele-
tive doping strategies beyond the simple trial-and-error-based ments that could increase the oxygen-vacancy formation energy
dopant selections. in the doped structure relative to that in the undoped system.
In this work, we investigate the effects of various dopants Finally, the effect of the dopants on the structural stabiliza-
through a systematic exploration of a model system of doped- tion was considered. Considering that the surface transition to
LiNiO2 based on density functional theory (DFT) calculation, a spinel or rocksalt phase involves transition-metal migration
which can serve as a map of intrinsic dopant effects for high- and disordering, the migration energy of nickel ions from the
nickel NCM materials. In this process, a three-step sequential transition-metal layer to the lithium layer (with or without
filtration is applied to identify the most suitable dopants to oxygen vacancy) was evaluated for the dopant candidates that

Adv. Energy Mater. 2022, 12, 2200136 2200136 (2 of 9) © 2022 Wiley-VCH GmbH
16146840, 2022, 18, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202200136 by Konkuk University Sanghuh Memorial Library, Wiley Online Library on [01/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advenergymat.de

passed the two prior criteria, leading to the selection of one swiftly check the intrinsic features of the dopant while under-
trial dopant for experimental verification. Although we arrived taking rigorous screening (Discussion S1 and Figure S4, Sup-
at one final dopant that optimally satisfied all of our quantita- porting Information). The c-lattice parameter variation for the
tive criteria, it should be noted that during the pruning process, doped system was calculated by taking the difference between
several promising dopant candidates lay in the marginal region the maximum (x = 0.33) and minimum (x = 0.16) c-lattice
and were excluded in the final selection. Thus, we provide the parameters for the LixNi0.92M0.08O2 (M = dopant) system. The
full calculational data for all the dopants that were examined diagram shows that all the dopants are beneficial to some
but screened out for each pruning process in the supplemen- extent in mitigating the abrupt c-lattice parameter contraction,
tary information (Figure S1, Supporting Information) for future with their efficacies varying from 2.74% for Zr to 3.88% for
reference in dopant studies. Os. Nevertheless, we expect that these values are rather over-
As the nickel content in the NCM material increases, the estimated given the relatively high doping concentration of our
material properties approach those of LixNiO2 (LNO, 0 ≤ x ≤ 1), model system (8%), considering that the typical doping level is
which is the end member of the high-nickel NCM.[45,46] There- ≈2%.[25,48,49] Thus, we applied a strict selection condition of a
fore, we assumed that the calculation for LNO would ade- c-lattice parameter change of less than 3%, even though several
quately demonstrate the phenomenon that can be expressed previous studies have reported that even a smaller remedy of
in high-nickel NCM materials (Figure S2, Supporting Informa- the c-lattice parameter collapse could help in mitigating det-
tion). For the selection process for the lattice strain, we first rimental chemo-mechanical failure.[49,50] This process led to
calculated the lattice-parameter variation that occurs in the 12 potential dopant candidates (Al, Si, V, Mn, Cu, Zr, Mo, Ag,
undoped LNO system, as a function of lithium content as a Sn, W, Re, Ir), as indicated in blue in the figure, with 14 dopants
reference. Figure S3 (Supporting Information) shows that the in a marginal region (yellow) and 12 dopants excluded (red).
c-lattice parameter gradually increases from 14.09 to 14.29 Å Further investigation of c-lattice parameter changes in these
with lithium extraction until x = 0.33. This maximum c-lattice doped systems revealed the nickel-slab distance monotonically
parameter drops to 13.56 Å with further delithiation up to decreased during the delithiation process, which is attributed
x = 0.16, resulting in c-lattice contraction of ≈5.1%.[47] In con- to the oxidation of nickel (Ni3+→ Ni4+), whereas the lithium-
trast, the maximum a-lattice parameter change was less than slab distance increased from x = 1 to x = 0.33 and significantly
0.01% (from 5.62 to 5.60 Å) for the same degree of delithia- collapsed with further delithiation to x = 0.16, consistent with
tion. These non-symmetrical lattice parameter changes imply the general understanding that the c-lattice parameter contrac-
anisotropic strain generation in the secondary particles that tion is induced by the lithium-slab contraction at high state
are typically comprised of randomly oriented primary particles. of charge[15,51] (Tables S1 and S2, Supporting Information).
Having this value as a reference, we investigated how the large Table S2 (Supporting Information) clearly shows that the
c-lattice contraction is affected by the 38 potential dopants selected dopants could mitigate the collapse of the lithium-slab
in Figure 2. We calculated high-doping concentration (8.2%) distance at the end of charge, thereby reducing the large aniso-
cases considering the size of supercells in our computations to tropic c-lattice collapse of the layered structure.

Figure 2. The c-lattice parameter contraction from Li0.33Ni0.92M0.08O2 to Li0.16Ni0.92M0.08O2 for the 38 doping candidates (M: doping candidates). The
dotted line represents the baseline used to select the first filtered candidates (within 3%). All the elements were categorized into four groups: nickel
ion as a reference element (green), candidate elements (blue), marginal elements (orange), and excluded elements (red).

Adv. Energy Mater. 2022, 12, 2200136 2200136 (3 of 9) © 2022 Wiley-VCH GmbH
16146840, 2022, 18, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202200136 by Konkuk University Sanghuh Memorial Library, Wiley Online Library on [01/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advenergymat.de

As a second step, we examined the change in the oxygen-


vacancy formation energies when the dopant candidates that
passed the first screening were incorporated into the structure,
helping to guide the selection of potential dopants that would
be beneficial for attaining structural stability at a high state of
charge. The oxygen-vacancy formation energies were calculated
for all the oxygen atoms in the unit cell (24 atoms) for each
doped system, and the oxygen vacancy with the lowest forma-
tion energy was selected as a representative for the system (cal-
culational details are provided in the Method section). Figure 3a
presents a schematic illustration of the oxygen-gas evolution
from the lattice oxygen (left) and the predicted vacancy forma-
tion energies (right) for each dopant species. For reference, the
oxygen-vacancy formation energy for the undoped system was
calculated to be −0.12 eV, indicating the spontaneous oxygen-
gas evolution at a highly charged state, which is consistent
with previous experimental and theorical works.[28,52,53] How-
ever, it was revealed that the lattice-oxygen stability is signifi-
cantly altered by the presence of dopants; the vacancy forma-
tion energies varied over a wide range from 0.16 eV to −1.24 eV
depending on the dopant. Al, Si, Mn, Cu, and Ir dopants were
observed to substantially increase the oxygen-vacancy forma-
tion energy compared with that of the undoped case (−0.12 eV).
In contrast, dopants such as V, Mo, Ag, and Re aggravated the
oxygen-gas evolution from the layered structure in the charged
state and were thus excluded in our screening process. There
were also some dopants that displayed a marginal difference in
the oxygen-vacancy formation such as Zr, Sn, and W compared
with the undoped reference, indicating that they might be ben-
eficial in mitigating the lattice strain, as determined in the first
screening, but do not notably contribute to the structural sta-
bility in the highly charged state. In this respect, we classified
the dopants into three categories: “candidates,” “marginal can-
didates” with oxygen-vacancy formation energies higher than or
similar to that of the undoped case, and “excluded elements”
with significantly lower oxygen-vacancy formation energies.
Recent studies have shown that the strength of the ionic bond
between two neighboring atoms is proportional to the effective
charge of each atom,[54,55] inferring that the different valence
states of dopants may influence the strength of the ionic bond
between the dopant and oxygen (Table S3 and Figure S5, Sup-
porting Information). Our results also indicate that the lattice-
oxygen stabilities are generally enhanced by dopants with high
valence states due to the stronger bonding with adjacent oxygen
ions. Accordingly, we selected five dopant candidates (Al, Si,
Mn, Cu, Ir) for further screening.
In the final step, we evaluated the tendency of nickel-ion
migration/disordering in the layered structure and its depend-
ency on the neighboring dopant species. Considering that the
irreversible phase transition from the layered to spinel or rocksalt
structures is initiated by cation mixing, the energetics of nickel-
ion migration were probed at a highly charged state.[56] Figure 3b
presents a schematic illustration of nickel-ion migration (left)
Figure 3. a) Oxygen-vacancy (VO) formation energy in Li0.16Ni0.92M0.08O2
(M: first filtered dopants). The dotted line with the value of −0.12 eV indi-
cates the oxygen-vacancy formation energy in Li0.16NiO2. b) Nickel-ion layer with neighboring oxygen vacancy (VO) in Li0.16Ni0.92M0.08O2 (M: third
stability at tetrahedral site in the lithium layer for Li0.16Ni0.92M0.08O2 (M: filtered dopants). The dotted lines with values of −0.70 eV indicates the
second filtered dopants). The dotted line with the value of 1.01 eV indi- nickel-ion stability at the tetrahedral site of the lithium layer with oxygen
cates the nickel-ion stability at the tetrahedral site of the lithium layer vacancies in Li0.16NiO2. All elements were categorized into 4 groups iden-
in Li0.16NiO2. c) Nickel-ion stability at the tetrahedral site of the lithium tical to the first criterion.

Adv. Energy Mater. 2022, 12, 2200136 2200136 (4 of 9) © 2022 Wiley-VCH GmbH
16146840, 2022, 18, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202200136 by Konkuk University Sanghuh Memorial Library, Wiley Online Library on [01/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advenergymat.de

and the predicted nickel stability (right) when the nickel in the the bonding nature with oxygen upon delithiation unlike the
vicinity of a dopant element migrates to a face-shared tetrahedral transition metal. The maintenance of the electrostatic repul-
site in the lithium layer. The nickel-ion stability was estimated sion between oxygen planes is expected to keep the lithium-
from the energy difference of the pristine structure and the dis- slab distance relatively stable even at high state of charge. We
ordered structure with the migrated nickel ion. As a reference, also suppose that the high oxygen-vacancy formation energy
the undoped system displayed an energy penalty of 1.01 eV when observed in the Si-doped system can be partially elucidated by
the nickel-ion disorder occurred.[57] Among the five dopant can- the modified charge distribution in the oxygen layer and the
didates, Al, Si, and Cu were observed to help inhibit the nickel corresponding metal–oxygen bond nature. It has been reported
migration, with higher energy penalties of 2.35, 2.1, and 1.92 eV, that the ionic-bond strength is proportional to the effective
respectively, whereas Mn and Ir were either less effective or det- charge of both ions, and thus, the oxygen elimination could be
rimental. We suppose that the volume change of the tetrahedral inhibited as the electron charge in the oxygen layer remains rel-
sites due to dopant incorporation would have caused the energy atively invariant.[54,55] This finding infers that the suppression
penalty for the nickel migration to the site. of the oxygen-vacancy formation in the Si-doped system can be
We further supposed that the phase transition on the sur- regarded as being the result of an enforced metal–oxygen bond
face would be more prone to occurring simultaneously with the with a high effective charge of oxygen ions.
formation of oxygen vacancies, as recent works have elucidated Inspired by theoretical investigations of the dopant effects,
that oxygen vacancies facilitate transition-metal migration into we decided to experimentally validate the effectiveness of the
the lithium layer in layered lithium oxides.[58,59] In this regard, Si dopant as a model case in high-nickel NCM materials. Sil-
we additionally considered the nickel-ion stability in the con- icon-doped LiNi0.92Co0.04Mn0.04O2 (Si-NCM9244) and undoped
text of the oxygen vacancies, as illustrated in Figure 3c. It could LiNi0.92Co0.04Mn0.04O2 (NCM9244) were synthesized using co-
be confirmed that the nickel ion was substantially stabilized in precipitation and a subsequent lithiation process, as described
the lithium layer (−0.70 eV) in the presence of oxygen vacancies in Methods. The high-resolution power diffraction (HRPD)
(Figure S6, Supporting Information), implying spontaneous analysis in Figure S8 (Supporting Information) indicated
nickel-ion migration into the lithium layer in the undoped ref- that both Si-NCM9244 and NCM9244 samples were obtained
erence in a highly charged state. However, when dopant ele- without any noticeable impurities, showing that the diffrac-
ments were present nearby, the nickel-ion migration tendency tion patterns of both materials were well matched with those
was slightly altered. The Al and Cu dopants were observed of standard R3-m layered oxides.[27] (More details on the sample
to further promote the nickel migration to the lithium layer characterization can be found in the Supporting Information
(−0.97 and −0.84 eV, respectively), whereas Si doping could help Discussion). First, we comparatively examined how the struc-
mitigate this migration (−0.50 eV) regardless of the presence of tural evolution occurs with delithiation of the two samples
oxygen vacancies. Before selecting the final dopant candidate, using in situ X-ray diffraction and post-mortem SEM analysis,
we also re-examined their effects in different dopant sites (octa- as presented in Figure 4a,b, respectively. Figure 4a depicts the
hedral vs tetrahedral sites) in the layered structure, however, it shifts of the main (003) reflection of the layered structure for
did not alter the general trend of the findings (Figure S7, Sup- the Si-NCM9244 and NCM9244 electrodes during the initial
porting Information). charge process at a current rate of 0.2C in the voltage range of
To elucidate the intrinsic role of the final candidate, the Si 2.5–4.4 V. The figure illustrates that the (003) reflection of the
dopant, we comparatively analyzed the local electronic structure NCM9244 electrode (left) gradually shifts to a low angle from
of undoped and doped systems using Bader charge analysis, 18.72 to 18.42°, indicating the expansion of the c-lattice para-
as shown in Table S5 (Supporting Information). It is generally meter, and then abruptly shifts back to 19.48° at the end of the
accepted that the lithium-slab distance variation during del- charge (≈4.4 V vs Li/Li+). This shift to a higher angle implies
ithiation is affected by both the electrostatic repulsion between the drastic contraction of the c-lattice parameter at the end of
oxygen planes and the screening effect of lithium ions.[54,60] the delithiation. The tendency of the (003) reflection shift in the
Upon delithiation, the lithium-slab distance increases with Si-NCM9244 electrode (right) is analogous to that in NCM9244
growing electrostatic repulsion between oxygen-atom planes from 18.72 to 18.42° during the initial delithiation process;
because of the weakened lithium-ion screening. However, however, the shift to a high angle at the end of the charge is
at a high state of charge, the lithium-slab distance decreases less significant (18.42 to 19.32°). This finding clearly indicates
because of the weaker repulsion between oxygen layers with that the c-lattice parameter variation in the Si-NCM9244 elec-
lower electron density that is typically caused by the stronger trode was mitigated compared with that in the undoped system.
covalent metal–oxygen bonds at higher valence state of the Accordingly, we performed cross-sectional SEM analysis for
transition metal.[61–63] Bader charge analysis revealed that when both materials to verify its outcome on the formation of micro-
Si is doped in the structure, the loss of electron density in the cracks induced after the repetitive anisotropic lattice variations.
oxygen layers is notably suppressed at the highly charged state. Figure 4b shows that although the NCM9244 particles (upper
We observed that the average charge density of six oxygen panel) had a densely packed morphology before cycling (inset),
atoms bonded to Si dopants is relatively less affected by the del- a considerable amount of microcracks and voids were exten-
ithiation and remains higher than that bonded to nickel ions sively formed from the surface to the bulk after 50 cycles. In
in the undoped system (see Table S5, Supporting Information contrast, fewer microcracks were formed in the Si-NCM9244
for details). The relative invariance of the electron charge in the particles (bottom panel) after 50 cycles with relatively intact
oxygen layer for the Si-doped system can be partly attributed to primary particles, confirming the efficacy of Si doping in miti-
the electrochemical inactivity of Si ions, which does not change gating the anisotropic lattice strain.

Adv. Energy Mater. 2022, 12, 2200136 2200136 (5 of 9) © 2022 Wiley-VCH GmbH
16146840, 2022, 18, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202200136 by Konkuk University Sanghuh Memorial Library, Wiley Online Library on [01/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advenergymat.de

Figure 4. a) Charge/discharge profiles and in situ XRD patterns of NCM9244 (left) and Si-NCM9244 (right) electrodes during charging at
a current density of 0.2 C. b) Cross-sectional SEM images of secondary particles for NCM9244 and Si-NCM9244 before (inset) and after cycling
(100 cycles). c) Atomic-resolution high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) images of NCM9244 (left)
and Si-NCM9244 (right) after 50 cycles. FFTs of regions 1–3 in NCM9244 and Si-NCM9244 are described as below. d) Capacity retention of NCM9244
and Si-NCM9244 for 100 cycles at 0.3 C. e) Discharge capacities at different C-rates in NCM9244 and Si-NCM9244.

STEM analysis further revealed the suppression of transi- 2.5–4.4 V (vs Li/Li+). The slightly lower discharge capacity
tion-metal migration at the surface of the Si-NCM9244 elec- of the Si-NCM9244 electrode may stem from the substitu-
trode compared with that of the NCM9244. The high-angle tion of the electrochemically inactive Si dopant. Figure 4d
annular dark-field (HAADF) images in Figure 4c display the shows the capacity retention of both electrodes at a current
surface of NCM9244 (left) and Si-NCM9244 (right) particles at a rate of 0.3 C. The Si-NCM9244 electrode retained a capacity
discharged state after 50 cycles. The fast-Fourier transformation of 157.15 mA h g−1 after 100 cycles, corresponding to a
(FFT) results indicate that the original layered structure was capacity retention of 75.2%, exceeding that of the NCM9244
mostly retained in the bulk; however, the surface region under- electrode under the same conditions (138.03 mA h g−1
went a phase transformation with a substantial amount of (≈66.9% retention)). In Figure 4e, the rate performances of
transition-metal migrations into the lithium layer. The diffrac- the Si-NCM9244 and NCM9244 electrodes are compared.
tion spots from the rocksalt phases (yellow, cubic region) and Notably, the Si-NCM9244 electrode could retain a higher
intermediate phases (green, intermediate region) containing capacity (149.5 mA h g−1) even when the current density
cubic and spinel phases were detected in both NCM9244 and increased to 2 C, in contrast to the significant capacity
Si-NCM9244 particles. However, the thickness of the recon- drop of the NCM9244 electrode at the same current density
structed surface regions was markedly more extensive in the (114.4 mA h g−1). This improved electrochemical performance
NCM9244 than in Si-NCM9244, indicating that transition-metal of the doped system can be attributed to enhanced structural
migration/oxygen evolution at the surface could be suppressed stability against microcrack formation and cation disordering,
by the Si doping in the high-nickel layered material. as previously discussed in the dopant screening process sec-
Finally, we comparatively examined the electrochemical tion, and the corresponding suppression of side reactions in
activity and stability of the Si-NCM9244 and NCM9244 elec- the electrochemical cell, which confirms the validity of our
trodes, as shown in Figure 4d and Figure S10 (Supporting theoretical dopant screening process.
Information). At a current rate of 0.1C, the Si-NCM9244 It is noteworthy that although we have selected one final
and NCM9244 electrodes delivered first cycle capacities of dopant candidate, Si, in our selection process, which consid-
223.8 and 228.5 mA g−1, respectively, in the voltage range of ered three major aspects, other dopants that were marginally

Adv. Energy Mater. 2022, 12, 2200136 2200136 (6 of 9) © 2022 Wiley-VCH GmbH
16146840, 2022, 18, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202200136 by Konkuk University Sanghuh Memorial Library, Wiley Online Library on [01/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advenergymat.de

screened out should not be completely ruled out for practical of the self-interaction error and lack of van der Waals (vdW) force in
applications. Certain dopants could outperform Si in some the GGA exchange functional, the vdW-density functional based on
aspects, e.g., the Zr dopant in reducing lattice strain or the Al nonlocal exchange-correlation was also adopted (optPBE, optB88, and
optB86b).[68,69] A plane-wave basis set with a kinetic energy cut-off of
dopant in suppressing oxygen gas. Considering that the three 500 eV and 2 × 2 × 1 k-points grid based on the Monkhorst–Pack scheme
major aspects would not have equal contributions to the degra- were utilized after performing the convergence test.[70] All the structures
dation of high-nickel NCM electrodes, more weighted selection were fully relaxed until the residual force in the unit cell converged to
criteria may make the final choice of dopant different than Si. within 0.02 eV Å−1. Next, 2 × 2 × 1 supercells containing 12 formula units
Furthermore, beyond our three criteria, there are other aspects of the LiNiO2 structures (space group: R3-m) were applied (Li12Ni12O24,
such as microstructure control or coatings that have been 48 atoms). In total, 38 dopants were considered as substitutes for
one nickel ion in the supercell (Li12Ni11M1O24, M = dopant). Although
known to be beneficial in high-nickel NCM electrodes. These
determining the favorable crystallographic site of dopants remains a
aspects are closely related to the surface properties (e.g., surface significant issue, it was assumed that the dopants replace nickel ions in
energy and lattice parameters at the interface), however, our the octahedral sites in the transition-metal layer to expedite the pruning
dopant selection map that revealed the intrinsic properties of process on the basis of extensive previous works.[28,30,71,72] The lithium-
dopants on crystallographic/electronic structure will also give vacancy ordering configurations with the lowest electrostatic energies
the intuition on designing the better-engineered surface. More- were enumerated using the Ewald summation method for each lithium
concentration.[73] Lattice expansion and contraction were depicted as the
over, we believe that this basic information about dopant effects
c-lattice parameter variation as a function of the lithium concentration.
can be extended to multi-dopant strategies. Taking advantage A supercell with composition Li2Ni11M1O24 was selected to reflect the
of different single dopants with their unique characteristics, high state of charge in which oxygen gas was evolved, and the oxygen-
optimal combinations of several dopant species can be incor- vacancy formation energy The oxygen-vacancy formation energy Ef(Vo)
porated into high-nickel NCM electrodes for practical purposes. was defined as
In this regard, the dopant map provided in Figures 2 and 3, and
1
Figure S1 (Supporting Information) can serve as an effective E f (Vo ) = E (Vo ) − E (pristine ) + E ( O 2 ) , (1)
2
guidebook for future endeavors.
where E(Vo) and E(pristine) are the total energies of the oxygen-
deficient and pristine structure, respectively. Because the chemical
3. Summary potential of oxygen gas is dependent on the temperature and because
DFT overestimates the oxygen binding energy, the chemical potential
Potential dopant candidates that can improve the electrochem- of oxygen was adopted, calculated to be 8.99 eV per molecule at room
ical performance of high-nickel NCM (LiNixCoyMnzO2, x ≥ 0.9) temperature.[58] The nickel-ion stability was considered as the site-energy
difference between octahedral sites in the transition-metal layer and
were comprehensively investigated by applying a systematic
tetrahedral sites in the lithium layer.[56,57] Because nickel-ion migration
theoretical stepwise pruning process. From the three criteria of was also sensitive to the local lithium environment, the lithium
mitigating anisotropic lattice collapse, reducing oxygen-gas evo- configuration for each lithium concentration was fixed and the energy
lution, and suppressing cation mixing, a total of 38 elements differences for all possible sites were estimated.
were sequentially screened, with silicon finally selected as a Material Synthesis: LiNi0.92Co0.04Mn0.04O2 (NCM9244) was prepared
model dopant for further experimental verification. Theoretical by coprecipitation and a subsequent solid-state method. The
analysis revealed that the silicon doping suppresses contraction Ni0.92Co0.04Mn0.04(OH)2 hydroxide precursor was synthesized using the
co-precipitation method with NiSO4⋅6 H2O (Daejung Chemical Co.),
of the lithium-slab distance at a high state of charge, thereby CoSO4⋅7 H2O (Daejung Chemical Co.), MnSO4⋅H2O (Daejung
alleviating the c-lattice parameter shrinkage, and reduces Chemical Co.), NaOH (Daejung Chemical Co.), and NH4OH (Daejung
oxygen–gas evolution while inhibiting nickel-ion migration. Chemical Co.) as starting materials. The homogeneously mixed solution
Experimental verification revealed that the Si-doped NCM9244 (Ni:Co:Mn = 92:4:4 in molar ratio) was fed into a batch reactor containing
exhibited significant improvement in the electrochemical a solution of NH4OH (aq) and NaOH (aq) under an N2 atmosphere.
stability and activity, which was attributed to the decrease in Next, NaOH (molar ratio of NaOH/TM = 2.0) and NH4OH as chelating
agents (molar ratio of NH4OH/TM = 1.2) were pumped separately into
secondary-particle cracking during repeated cycling and the
the reactor. The precursor powder was obtained by filtering, washing,
accompanying suppression of the side reactions. Although our vacuum drying, and precipitating at 110 °C for 12 h. The precursor was
model case presented here is limited to one final dopant for hand-mixed with LiOH·H2O (molar ratio 1:1.03) and annealed at 740 °C
experimental verifications, we believe that our map of dopant for 12 h with oxygen gas flow.
efficacy and theoretical stepwise pruning and validation proce- For Si-NCM9244, first, the hydroxide precursor was coated with
dures can be more extensively utilized as rational guidelines a silicon source using tetraethyl orthosilicate (Sigma–Aldrich) and
acetylacetone (Sigma–Aldrich). Tetraethyl orthosilicate and acetylacetone
for the exploration of effective dopants for high-nickel NCM
(1:4 mole ratio) were mixed in dehydrated ethanol and stirred for 30 min.
electrodes. Then, hydroxide powder (Si:NCM = 1:99 mole ratio) was subsequently
added and mixed for 30 min and dried at 100 °C overnight. Finally,
Si-NCM9244 was prepared by mixing the silicon-coated precursor with
4. Experimental Section LiOH·H2O (molar ratio of 1:1.03) followed by calcination at 740 °C for
12 h with oxygen flow.
First-Principles Calculations: First-principles calculations based on Structural Characterization: The crystal structure was determined using
DFT with the projector augmented wave (PAW)[64,65] method were synchrotron radiation powder X-ray diffraction (HRPD) data collected
conducted using the Vienna Ab Initio Simulation Package (VASP).[66] from the 9B HRPD beamline at the Pohang Accelerator Laboratory
The spin-polarized generalized gradient approximation (GGA) with (PAL). The FullProf program was used for the Rietveld refinement. An
the Perdew–Burke–Ernzerhof (PBE) exchange-correlation model was X-ray diffractometer (R-AXIS IV, RIGAKU) was used to monitor the phase
used as a reference for the lattice-parameter calculation.[67] Because transformation during the charge process for in situ XRD analysis. In

Adv. Energy Mater. 2022, 12, 2200136 2200136 (7 of 9) © 2022 Wiley-VCH GmbH
16146840, 2022, 18, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202200136 by Konkuk University Sanghuh Memorial Library, Wiley Online Library on [01/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advenergymat.de

detail, a modified coin cell with a 3 mm hole at its center sealed with [1] J. B. Goodenough, Y. Kim, Chem. Mater. 2010, 22, 587.
mylar film was used while the cells were charged up to 4.4 V versus Li/Li+ [2] W. Li, E. M. Erickson, A. Manthiram, Nat. Energy 2020, 5, 26.
at a current density of 0.2 C. High-resolution STEM images were obtained [3] Y.-K. Sun, S.-T. Myung, B.-C. Park, J. Prakash, I. Belharouak,
using a double aberration-corrected JEOL-ARM 200CF microscope with K. Amine, Nat. Mater. 2009, 8, 320.
an acceleration voltage of 200 kV. The cross-sectional TEM and SEM [4] A. Manthiram, J. C. Knight, S.-T. Myung, S.-M. Oh, Y.-K. Sun,
specimens were prepared using a FIB (Helios Nano Lab450, FEI) and Adv. Energy Mater. 2016, 6, 1501010.
cross-section polisher (IB-19510CP, JEOL), respectively. [5] J. Kim, H. Lee, H. Cha, M. Yoon, M. Park, J. Cho, Adv. Energy Mater.
Electrochemistry: The electrochemical properties of the half-cells were 2018, 8, 1702028.
measured using 2032 coin-type cells (Welcos, Korea). To prepare the [6] C. S. Yoon, H.-H. Ryu, G.-T. Park, J.-H. Kim, K.-H. Kim, Y.-K. Sun,
electrodes, the obtained NCM9244 powder was mixed with Super-P and
J. Mater. Chem. A 2018, 6, 4126.
polyvinylidene fluoride (PVDF) binder (weight ratio 85:8:7) in N-methyl-
[7] Y. Bi, J. Tao, Y. Wu, L. Li, Y. Xu, E. Hu, B. Wu, J. Hu, C. Wang,
2-pyrrolidone (NMP) solvent and then cast on Al foil and dried. The cell
J.-G. Zhang, Y. Qi, J. Xiao, Science 2020, 370, 1313.
was assembled with lithium-metal foil as a counter electrode, a separator
film (Celgard), and an electrolyte of EC: EMC (3:7 v/v) dissolved with [8] C. Xu, K. Märker, J. Lee, A. Mahadevegowda, P. J. Reeves, S. J. Day,
LiPF6 (1.2 M) salt. The cells were cycled using an automatic battery M. F. Groh, S. P. Emge, C. Ducati, B. Layla Mehdi, C. C. Tang,
cycler (WBCS 3000S WonATech) in the voltage range of 2.5–4.4 V vs Li/ C. P. Grey, Nat. Mater. 2021, 20, 84.
Li+. The current density of 1 C represents 200 mA g−1. [9] A. O. Kondrakov, H. Geßwein, K. Galdina, L. De Biasi, V. Meded,
E. O. Filatova, G. Schumacher, W. Wenzel, P. Hartmann,
T. Brezesinski, J. Janek, J. Phys. Chem. C 2017, 121, 24381.
[10] H.-H. Sun, A. Manthiram, Chem. Mater. 2017, 29, 8486.
Supporting Information [11] H. H. Sun, H.-H. Ryu, U.-H. Kim, J. A. Weeks, A. Heller, Y.-K. Sun,
Supporting Information is available from the Wiley Online Library or C. B. Mullins, ACS Energy Lett. 2020, 5, 1136.
from the author. [12] S.-M. Bak, K.-W. Nam, W. Chang, X. Yu, E. Hu, S. Hwang,
E. A. Stach, K.-B. Kim, K. Y. Chung, X.-Q. Yang, Chem. Mater. 2013,
25, 337.
[13] W. Li, X. Liu, Q. Xie, Y. You, M. Chi, A. Manthiram, Chem. Mater.
Acknowledgements 2020, 32, 7796.
[14] H.-H. Ryu, K.-J. Park, C. S. Yoon, Y.-K. Sun, Chem. Mater. 2018, 30, 1155.
This research was supported by the Basic Science Research Program
[15] A. O. Kondrakov, A. Schmidt, J. Xu, H. Geßwein, R. Mönig,
through the National Research Foundation of Korea (NRF) funded by
the Ministry of Education (2020R1A6A3A13070145). This work was also P. Hartmann, H. Sommer, T. Brezesinski, J. Janek, J. Phys. Chem. C
supported by the Center for Nanoparticle Research at Institute for 2017, 121, 3286.
Basic Science (IBS) (IBS-R006-A2) and the National Supercomputing [16] L. de Biasi, A. O. Kondrakov, H. Geßwein, T. Brezesinski,
Center with supercomputing resources including technical support P. Hartmann, J. Janek, J. Phys. Chem. C 2017, 121, 26163.
(KSC-2016-C3-0069). [17] H. Konishi, T. Yuasa, M. Yoshikawa, J. Power Sources 2011, 196,
6884.
[18] L. Wu, K.-W. Nam, X. Wang, Y. Zhou, J.-C. Zheng, X.-Q. Yang,
Y. Zhu, Chem. Mater. 2011, 23, 3953.
Conflict of Interest [19] S.-M. Bak, E. Hu, Y. Zhou, X. Yu, S. D. Senanayake, S.-J. Cho,
The authors declare no conflict of interest. K.-B. Kim, K. Y. Chung, X.-Q. Yang, K.-W. Nam, ACS applied materials
and interfaces 2014, 6, 22594.
[20] J. Bréger, Y. S. Meng, Y. Hinuma, S. Kumar, K. Kang, Y. Shao-Horn,
G. Ceder, C. P. Grey, Chem. Mater. 2006, 18, 4768.
Author Contributions [21] R. V. Chebiam, F. Prado, A. Manthiram, J. Electrochem. Soc. 2001,
148, A49.
D.-H.K., J.-H.S., and C.-H.J. contributed equally to this work. D.-H.K.,
J.-H.S., S.-H.H., and K.K. designed the project. D.-H.K., and J.-H.S. [22] S.-K. Jung, H. Gwon, J. Hong, K.-Y. Park, D.-H. Seo, H. Kim, J. Hyun,
conducted the DFT calculations. D.-H.K., J.-H.S., and B.K. analyzed the W. Yang, K. Kang, Adv. Energy Mater. 2014, 4, 1300787.
calculation results. C.-H.J., and D.E. carried out the synthesis, structural [23] S. -. K. Jung, H. Kim, S. H. Song, S. Lee, J. Kim, K. Kang, Adv. Funct.
characterization, and electrochemical test. B.K. provided constructive Mater. 2021, 2108790.
advice for the calculation designs and results. J.-H.S., D.-H.K., and K.K. [24] S. H. Song, M. Cho, I. Park, J. -. G. Yoo, K. -. T. Ko, J. Hong, J. Kim,
wrote the manuscript, and K.K. supervised all aspects of research. S. -. K. Jung, M. Avdeev, S. Ji, S. Lee, J. Bang, H. Kim, Adv. Energy
Mater. 2020, 10, 2000521.
[25] K.-J. Park, H.-G. Jung, L.-Y. Kuo, P. Kaghazchi, C. S. Yoon, Y.-K. Sun,
Adv. Energy Mater. 2018, 8, 1801202.
Data Availability Statement [26] U.-H. Kim, G.-T. Park, B.-K. Son, G. W. Nam, J. Liu, L.-Y. Kuo,
The data that support the findings of this study are available in the P. Kaghazchi, C. S. Yoon, Y.-K. Sun, Nat. Energy 2020, 5, 860.
supplementary material of this article. [27] U.-H. Kim, D.-W. Jun, K.-J. Park, Q. Zhang, P. Kaghazchi, D. Aurbach,
D. T. Major, G. Goobes, M. Dixit, N. Leifer, C. M. Wang, P. Yan,
D. Ahn, K.-H. Kim, C. S. Yoon, Y.-K. Sun, Energy and environmental
science 2018, 11, 1271.
Keywords [28] J. Cheng, L. Mu, C. Wang, Z. Yang, H. L. Xin, F. Lin, K. A. Persson,
chemo-mechanical degradation, density functional theory, doping, high- J. Mater. Chem. A 2020, 8, 23293.
nickel NCM cathodes, layered cathode materials [29] D. Wang, X. Li, Z. Wang, H. Guo, Y. Xu, Y. Fan, J. Ru, Electrochim.
Acta 2016, 188, 48.
Received: January 12, 2022 [30] F. Schipper, M. Dixit, D. Kovacheva, M. Talianker, O. Haik,
Revised: February 22, 2022 J. Grinblat, E. M. Erickson, C. Ghanty, D. T. Major, B. Markovsky,
Published online: March 23, 2022 D. Aurbach, J. Mater. Chem. A 2016, 4, 16073.

Adv. Energy Mater. 2022, 12, 2200136 2200136 (8 of 9) © 2022 Wiley-VCH GmbH
16146840, 2022, 18, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/aenm.202200136 by Konkuk University Sanghuh Memorial Library, Wiley Online Library on [01/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advenergymat.de

[31] Y.-C. Li, W. Xiang, Z.-G. Wu, C.-L. Xu, Y.-D. Xu, Y. Xiao, Z.-G. Yang, [50] H.-H. Ryu, G.-T. Park, C. S. Yoon, Y.-K. Sun, J. Mater. Chem. A 2019,
C.-J. Wu, G.-P. Lv, X.-D. Guo, Electrochim. Acta 2018, 291, 84. 7, 18580.
[32] H. Sun, Z. Cao, T. Wang, R. Lin, Y. Li, X. Liu, L. Zhang, F. Lin, [51] W. Li, H. Y. Asl, Q. Xie, A. Manthiram, J. Am. Chem. Soc. 2019, 141,
Y. Huang, W. Luo, Materials Today Energy 2019, 13, 145. 5097.
[33] C. S. Yoon, U.-H. Kim, G.-T. Park, S. J. Kim, K.-H. Kim, J. Kim, [52] R. Jung, M. Metzger, F. Maglia, C. Stinner, H. A. Gasteiger, J. Electrochem.
Y.-K. Sun, ACS Energy Lett. 2018, 3, 1634. Soc. 2017, 164, A1361.
[34] F. A. Susai, D. Kovacheva, A. Chakraborty, T. Kravchuk, [53] W. Hu, H. Wang, W. Luo, B. Xu, C. Ouyang, Solid State Ionics 2020,
R. Ravikumar, M. Talianker, J. Grinblat, L. Burstein, Y. Kauffmann, 347, 115257.
D. T. Major, B. Markovsky, D. Aurbach, ACS Appl. Energy Mater. [54] M. A. y de Dompablo, G. Ceder, J. Power Sources 2003, 119, 654.
2019, 2, 4521. [55] H. Chen, J. A. Dawson, J. H. Harding, J. Mater. Chem. A 2014, 2,
[35] A. Dianat, N. Seriani, M. Bobeth, G. Cuniberti, J. Mater. Chem. A 7988.
2013, 1, 9273. [56] J. Reed, G. Ceder, A. Van Der Ven, Electrochemical and Solid State
[36] P. Hou, F. Li, Y. Sun, M. Pan, X. Wang, M. Shao, X. Xu, ACS Sustain- Letters 2001, 4, A78.
able Chemistry and Engineering 2018, 6, 5653. [57] J. Reed, G. Ceder, Chem. Rev. 2004, 104, 4513.
[37] J. Li, M. Zhang, D. Zhang, Y. Yan, Z. Li, Chem. Eng. J. 2020, 402, [58] F. Kong, C. Liang, L. Wang, Y. Zheng, S. Perananthan, R. C. Longo,
126195. J. P. Ferraris, M. Kim, K. Cho, Adv. Energy Mater. 2019, 9, 1802586.
[38] L. Li, E. Han, L. Zhu, S. Qiao, C. Du, Ionics 2020, 26, 2655. [59] C. R. Fell, D. Qian, K. J. Carroll, M. Chi, J. L. Jones, Y. S. Meng,
[39] C.-H. Jung, D.-H. Kim, D. Eum, K.-H. Kim, J. Choi, J. Lee, H.-H. Kim, Chem. Mater. 2013, 25, 1621.
K. Kang, S.-H. Hong, Adv. Funct. Mater. 2021, 31, 2010095. [60] T. Ohzuku, A. Ueda, M. Nagayama, J. Electrochem. Soc. 1993, 140,
[40] N. Ravet, A. Abouimrane, M. Armand, Nat. Mater. 2003, 2, 702. 1862.
[41] S.-Y. Chung, J. T. Bloking, Y.-M. Chiang, Nat. Mater. 2003, 2, 702. [61] B. Mortemard De Boisse, M. Reynaud, J. Ma, J. Kikkawa,
[42] J. Zhao, W. Zhang, A. Huq, S. T. Misture, B. Zhang, S. Guo, L. Wu, S.-I. Nishimura, M. Casas-Cabanas, C. Delmas, M. Okubo,
Y. Zhu, Z. Chen, K. Amine, F. Pan, J. Bai, F. Wang, Adv. Energy Mater. A. Yamada, Nat. Commun. 2019, 10, 2185.
2017, 7, 1601266. [62] B. M. De Boisse, J.-H. Cheng, D. Carlier, M. Guignard, C.-J. Pan,
[43] M. Bianchini, J. Wang, R. J. Clément, B. Ouyang, P. Xiao, S. Bordere, D. Filimonov, C. Drathen, E. Suard, B.-J. Hwang,
D. Kitchaev, T. Shi, Y. Zhang, Y. Wang, H. Kim, M. Zhang, J. Bai, J. Mater. Chem. A 2015, 3, 10976.
F. Wang, W. Sun, G. Ceder, Nat. Mater. 2020, 19, 1088. [63] K. Min, K. Kim, C. Jung, S.-W. Seo, Y. Y. Song, H. S. Lee, J. Shin,
[44] W. Hua, K. Wang, M. Knapp, B. Schwarz, S. Wang, H. Liu, J. Lai, E. Cho, J. Power Sources 2016, 315, 111.
M. Müller, A. Schökel, A. Missyul, D. Ferreira Sanchez, X. Guo, [64] P. E. Blöchl, Phys. Rev. B 1994, 50, 17953.
J. R. Binder, J. Xiong, S. Indris, H. Ehrenberg, Chem. Mater. 2020, [65] G. Kresse, D. Joubert, Phys. Rev. B 1999, 59, 1758.
32, 4984. [66] G. Kresse, J. Furthmüller, Phys. Rev. B 1996, 54, 11169.
[45] H.-J. Noh, S. Youn, C. S. Yoon, Y.-K. Sun, J. Power Sources 2013, 233, [67] J. P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 1996, 77, 3865.
121. [68] J. Klimeš, D. R. Bowler, A. Michaelides, J. Phys.: Condens. Matter
[46] J. Xu, E. Hu, D. Nordlund, A. Mehta, S. N. Ehrlich, X.-Q. Yang, 2009, 22, 022201.
W. Tong, ACS applied materials and interfaces 2016, 8, 31677. [69] M. Aykol, S. Kim, C. Wolverton, J. Phys. Chem. C 2015, 119, 19053.
[47] J. C. Garcia, J. Gabriel, N. H. Paulson, J. Low, M. Stan, H. Iddir, [70] H. J. Monkhorst, J. D. Pack, Phys. Rev. B 1976, 13, 518 8.
J. Phys. Chem. C 2021. [71] Y. Shin, W. H. Kan, M. Aykol, J. K. Papp, B. D. McCloskey, G. Chen,
[48] K. Zhou, Q. Xie, B. Li, A. Manthiram, Energy Storage Mater. 2021, K. A. Persson, Nat. Commun. 2018, 9, 4597.
34, 229. [72] C. Liang, F. Kong, R. C. Longo, C. Zhang, Y. Nie, Y. Zheng, K. Cho,
[49] T. Weigel, F. Schipper, E. M. Erickson, F. A. Susai, B. Markovsky, J. Mater. Chem. A 2017, 5, 25303.
D. Aurbach, ACS Energy Lett. 2019, 4, 508. [73] P. P. Ewald, Ann. Phys. 1921, 369, 253.

Adv. Energy Mater. 2022, 12, 2200136 2200136 (9 of 9) © 2022 Wiley-VCH GmbH

You might also like