SlideShare a Scribd company logo
1 of 13
Download to read offline
Journal of The Electrochemical
Society
OPEN ACCESS
Increased Disorder at Graphite Particle Edges Revealed by Multi-length
Scale Characterization of Anodes from Fast-Charged Lithium-Ion Cells
To cite this article: Saran Pidaparthy et al 2021 J. Electrochem. Soc. 168 100509
View the article online for updates and enhancements.
This content was downloaded from IP address 186.211.111.10 on 07/12/2021 at 13:16
Increased Disorder at Graphite Particle Edges Revealed by Multi-
length Scale Characterization of Anodes from Fast-Charged
Lithium-Ion Cells
Saran Pidaparthy,1,2
Marco-Tulio F. Rodrigues,1,* Jian-Min Zuo,2,3
and Daniel
P. Abraham1,*,z
1
Chemical Sciences and Engineering Division, Argonne National Laboratory, Lemont, Illinois, 60439, United States of
America
2
Department of Materials Science and Engineering, University of Illinois at Urbana-Champaign, Urbana, Illinois, 61801,
United States of America
3
Materials Research Laboratory, University of Illinois at Urbana-Champaign, Urbana, Illinois, 61801, United States of
America
Fast charging of lithium-ion cells increases voltage polarization of the electrodes and creates conditions that are favorable for Li-
deposition at the graphite anode. Repeated fast charging induces changes in the capacity-voltage profiles and increases the
probability of lithium-plating on the electrode. This higher probability results from structural, morphological and chemical
modifications that are revealed by multi-length scale characterization of graphite anodes extracted from discharged lithium-ion
cells, previously charged at rates up to 6 C. The distinct differences between anodes with lithium-plating and as-prepared electrodes
are clearly seen in analytical electron microscopy data. Scanning electrode microscopy (SEM) images show that the fast-charged
anode is significantly thicker, apparently because of the electrolyte reduction/hydrolysis products that accumulate in electrode
pores. High resolution electron microscopy (HREM) images reveal wavy graphite fringes near the particle edges. Analysis of
scanning electron nanodiffraction (SEND) data reveal higher d-spacings and greater lattice rotations, indicating disorder in the
graphite near the particle edges that extend about 20 nm into the bulk. The extent of this disorder is greater near larger internal
pores, highlighting nanoscale heterogeneities within particles. As graphite lithiation occurs primarily through edge planes, this
permanent disorder would hinder Li+
intercalation kinetics and favor Li0
plating during repeated cycling.
© 2021 The Author(s). Published on behalf of The Electrochemical Society by IOP Publishing Limited. This is an open access
article distributed under the terms of the Creative Commons Attribution 4.0 License (CC BY, http://creativecommons.org/licenses/
by/4.0/), which permits unrestricted reuse of the work in any medium, provided the original work is properly cited. [DOI: 10.1149/
1945-7111/ac2a7f]
Manuscript submitted August 25, 2021; revised manuscript received September 17, 2021. Published October 8, 2021.
Supplementary material for this article is available online
Advances in lithium-ion battery (LIB) technology have fueled the
rise of the electric vehicle (EV) industry.1
Most high-energy LIBs
contain a positive electrode (cathode) comprised of a layered oxide
with Ni, Co and Mn, a negative electrode (anode) comprised of
graphite, and an electrolyte with LiPF6 salt in a mixture of organic
solvents. These cell components are being continually optimized so
that LIB technology can deliver high energy densities safely and
reliably during a battery’s 10–15-year lifetime. State-of-the-art LIBs
can be fully charged in about 1 h with minimal degradation in cell
performance. However, full charging over shorter durations causes
irreversible damage to the battery via losses in the inventory of
mobile Li+
ions (i.e., capacity fade) and resistance increases (i.e.,
power fade) in the cells.2,3
The U.S. Department of Energy (DOE) is
funding research to identify the underlying mechanisms that limit
fast-charge performance of LIB cells and to develop technological
solutions that would enable full battery charge in less than 10 min.4–7
Solutions to these limitations will allay range anxiety concerns and
accelerate customer acceptance of EVs.
During cell charging, lithium ions are extracted from the
transition metal oxide cathode and intercalated between atomic
planes of the graphite particles. Under fast charge conditions, there is
salt enrichment in the oxide cathode and salt depletion in the
graphite anode, leading to non-uniform distribution of Li+
ions
across the electrode thickness.8,9
These conditions lower Li+
ion
conductivity in the electrolyte within the electrode pores and
increase impedance, causing polarization in both electrodes.
Cathode polarization causes the cell to reach its upper cut-off
voltage (UCV) before Li+
ions are fully extracted from the oxide,
while anode polarization causes the cell to dip below the Li-plating
potential creating conditions that are favorable for lithium nucleation
and growth on the graphite electrode.10–12
The situation is further
complicated by non-uniformities in local reactivity resulting from
factors that include uneven mechanical compression of the separator
that affects its porosity/tortuosity, and irregular distribution of active
particles that affects distribution of local currents within the
electrodes.13–16
These inhomogeneities cause large lithium concen-
tration gradients in the electrodes, both across the surface and along
its thickness, promoting non-uniform aging of the cell materials. The
high charge currents can also enhance structural stresses in the
electrode matrix causing the coating to peel off the current collector
and within the active material particles causing their fracture and
disconnection from the electron conduction network.17
In our recent series of articles we have used operando X-ray
techniques to quantify phase evolution in the graphite and lithium
concentration gradients that develop in the electrode during fast
charging.18,19
In addition, we have described techniques to identify
lithium deposition on the graphite electrode and detailed electro-
chemical charging protocols that can mitigate this plating.20–22
We
observed that the likelihood of Li-plating on the graphite anode
increases during repeated fast-charge cycling.10
We note here that
commercial graphite typically contains a mixture of the hexagonal
(2H) and rhombohedral (3 R) phases: the stacking sequence is
ABAB in the 2H structure and ABCABC in the 3 R structure.23
The 2H phase is more thermodynamically stable than the 3 R phase
at room temperature and pressure. The Li intercalation mechanism is
reported to be similar for both phases: during Li+
insertion, the
adjacent graphene sheets glide relative to each other to lower the
overall energy of the system and attain AAAA layer stacking.24
Any
material feature that hinders the movement of graphene sheets will
prevent Li+
intercalation. During fast charging, Li+
intercalation
into the graphite competes with Li0
nucleation on the particles.
Greater layer-to-layer stacking disorder in the graphite in the form of
added strain, local misfits, and misorientation angles in the stack,
would make Li+
intercalation more difficult and expedite Li0
z
E-mail: abraham@anl.gov
*Electrochemical Society Member.
Journal of The Electrochemical Society, 2021 168 100509
plating. Hence, an in-depth understanding of changes in the graphite
particles subjected to fast charging is crucial for the development of
materials that enable intercalation and mitigate deterioration of cell
performance during the electrochemical cycling.
In this work, we examine the morphological, chemical, and
structural changes to the graphite materials and electrodes, using a
suite of electrochemical and physicochemical techniques. The
characterization was performed across multiple length-scales on
as-prepared electrodes and on numerous electrodes harvested from
cycled cells; only representative data from the various techniques
are shown to highlight key changes that result from the cycling. The
data in this manuscript are from an as-prepared graphite (Gr-AP)
electrode that was never exposed to the electrolyte, graphite
electrode with no plating (Gr-NP) harvested from a cell after a
few low-rate cycles and graphite electrode with lithium plating (Gr-
LP) harvested from a fast-charged cell. For purposes of clarity and
contrast, we show only the Gr-AP and Gr-LP data in the main text;
data from the Gr-NP sample are included in the Supporting
Information (SI). Note that all Tables and Figures in the SI are
marked with the designator S, as in Fig. S1 (available online at
stacks.iop.org/JES/168/100509/mmedia).
We start by describing electrochemical tests, conducted on full
cells with LiNi0.5Co0.2Mn0.3O2 (NCM523) cathodes and graphite
(Gr) anodes, which include charging rates up to 6 C (here, 1 C is the
current needed to fully discharge the cell in 1 h). Following this, we
compare electrochemical behavior of the Gr-AP electrode with that
of the Gr-LP electrode. Then, after presenting high-resolution X-ray
diffraction (XRD) data from the materials, we show microscopy and
spectroscopy results obtained on electrode cross-sections fabricated
via ion milling and on sections prepared with a focused ion beam
(FIB). At the bulk-scale, we use scanning electron microscopy
(SEM) and energy dispersive X-ray spectroscopy (EDS) to probe the
morphological and chemical changes in cross-sections of the
graphite anode. At the nano- to atomic-scale, we rely on both
conventional transmission electron microscopy (TEM) and scanning
TEM (STEM) to reveal local structural disorder and chemical
variation within individual graphite particle sections through a
combination of high-resolution imaging, EDS, and scanning electron
nanodiffraction (SEND). Our investigations reveal that anodes from
the fast-charged cells have lower capacity, present a sloping voltage
profile, display an increase in thickness and show greater lattice
disorder in localized areas near the graphite particle edges.
Experimental Section
Electrode materials, cell assembly and electrochemical
testing.—In the NCM523//Gr cells, the positive electrode has a
71 μm thick coating, with 90 wt% Li(Ni0.5Co0.2Mn0.3)O2 (Toda),
5 wt% C45 carbon (Timcal) and 5 wt% PVDF binder (KF-9300,
Kureha), on a 20 μm thick Al current collector.10
The negative
electrode has a 70 μm thick coating, with ∼92 wt% SLC1506T
graphite (Superior Graphite), 2 wt% C45 carbon and 6 wt% PVDF
binder, on a 10 μm thick Cu current collector. The electrolyte (aka
Gen 2, purchased from Tomiyama) is a solution of 1.2 M lithium
hexafluorophosphate (LiPF6) salt dissolved in a 3:7 w/w solvent
mixture of ethylene carbonate (EC) and ethyl methyl carbonate
(EMC). Celgard 2320 is the separator between the electrodes. Before
cell assembly, the electrodes and separator are dried in vacuum
ovens set at 120 °C and 70 °C, respectively. The 3-electrode
measurements are conducted in custom-made stainless-steel cells,
which are assembled and tested inside an Ar-atmosphere glove box.
The reference electrode (RE) probe is placed between two separa-
tors, sandwiched between 20.3 cm2
disks of the oxide-positive and
graphite-negative electrodes (Fig. 1): the RE consists of a copper
wire (25 μm diameter) with a thin layer of Li metal plated on the
exposed tip.9
Electrochemical cycling is typically conducted at 30 °C with a
Maccor Model 2400 cycler. The Gr-NP electrode is from a cell that
experienced two C/10 and five C/2 cycles between 3 and 4.2 V at 30
°C. The Gr-LP electrode was harvested from a cell that experienced
multiple sequences of an “escalating C-rate” protocol (Fig. 1). To
accelerate cell aging, three of the sequences were at 30 °C, one at 40
°C and one at 50 °C. Including the initial cycles and periodic
reference performance tests (C/25 rate), the cell experienced a total
of 50 cycles; of these, 25 charge segments were at rates of 2 C and
above.
After cycling, cells were discharged to (and held for 10 h at) 3 V
before disassembly in an argon-atmosphere glove box. Various
diagnostic tests were conducted on the harvested cell materials to
determine the effect of cycling on the graphite material. For the
electrochemical tests, 2032-type coin cells were assembled with 1.6
cm2
area harvested-electrode samples, lithium metal counter elec-
trode, fresh Celgard 2320 separator, and fresh Gen2 electrolyte.
Cross-section polishing.—Cross-sections of the graphite elec-
trode samples were prepared via argon ion milling using the Gatan
Model 685.O Precision Etching Coating System II (PECS II). The
electrode was gently handled with metal tweezers and cut with sharp
metal scissors to leave at least one flat-side of ∼5 mm; all metal
tools were cleaned using an ethanol and acetone rinse prior to
handling the electrode. The cut electrode was mounted onto the
PECS II sample blade (Gatan, Product No. 693.08500) using fast-
drying silver paint (Ted Pella, Product No. 16040–30) and adjusted
with a micrometer to ensure ∼200 μm overhang of the flat-side of
the electrode from the top of the sample blade. Once the electrode
was secure on the blade, the electrode/blade polishing sample was
placed onto the cross-section specimen mount (Gatan, Product No.
Figure 1. In this escalating C-rate experiment, the NCM523//Gr cell is
charged with increasing currents to a fixed capacity of 87 mAh g−1
. Each
charge is followed by 1 h rest and C/5 discharge to 3 V and a hold. The cell
voltage changes during the cycles are shown in (a), while the graphite-
electrode potential changes are shown in (b). Lithium plating on the graphite
particles becomes likely when the electrode potential dips below 0 V vs
Li/Li+
; this Li-plating condition (LPC) is indicated by the dotted line in (b).
In the above experiment, Li-plating on the graphite electrode becomes likely
when the charge rate exceeds ∼2 C.
Journal of The Electrochemical Society, 2021 168 100509
685.08540) and loaded into the PECS II instrument. After the sample
was loaded, the instrument was set to “Etch by Time” with a “Single
Modulation” etching condition, a Gun Tilt angle of 0° for both ion
guns, and a rotation rate of 0.5 RPM. The ion milling was performed
in two stages: first, a high-voltage milling at 8 kV (with current
readings in the range 20–40 μA) that was run for 4 h to create a
rough cross-section; second, a low-voltage milling at 4 kV (with
current readings of 7–15 μA) that ran for 30 min to apply a low-kV
fine polishing.
Synchrotron powder diffraction.—Powder diffraction data were
collected at beamline 11-BM of the Advanced Photon Source,
Argonne National Laboratory. Coatings from the as-prepared
electrodes and cycled electrodes, scraped of the copper current
collectors, were loaded into sealed capillaries for examination. An
X-ray wavelength of 0.41283 Å was used to collect the diffraction
data, over an angular range from –6.0° to 28.0° with a step size of
0.001° and a time per step of 0.1 s.
Scanning electron microscopy and energy-dispersive X-ray
spectroscopy.—Both imaging and elemental examinations were
performed using a JEOL 7000 F analytical SEM. The electrode
samples were prepared in the following ways: (1) for top-down
imaging, a small portion of the electrode is mounted onto a flat
sample holder via carbon tape; (2) for cross-sectional imaging, the
samples prepared by the PECS II ion milling are loaded onto a
standard cross-section specimen mount designed for the SEM. For
the SEM imaging, the instrument was set to 7 kV and operated under
secondary electron imaging mode with a medium probe current of
∼0.3 nA. For EDS acquisition, the sample was aligned to an optimal
working distance of 10 mm and the probe current was adjusted to a
setting of ∼15 nA. The mapping data were acquired using the
ThermoFisher NSS 2.0 software using its “Spectral Imaging” mode.
The datasets were summed for 150 frames, each acquired for 10 s
per frame with a 50 μs dwell time. The mapping data are reported as
counts per element and the quantification was performed by the
software automatically using a built-in database and algorithm based
on the standard phi-rho-z method. The line profile data for the cross-
section samples are acquired using the “X-ray Linescans” mode with
a single scan with 50 points from the bottom to the top of the anode
with a dwell time of 10 s per point. Line profile data are further
binned for every 5 μm step.
Focused-ion beam milling.—Samples for nano- to atomic-scale
electron imaging and diffraction were prepared with a ThermoFisher
Scientific Scios 2 FIB-SEM. The electrodes were cut with sharp
metal scissors and mounted onto a flat sample holder via carbon tape
and loaded into the instrument. The cross-section samples were
selected from areas near the electrode surface and prepared via
standard focused-ion beam lift-out procedure provided by
ThermoFisher Scientific onboard user guidance (included with the
instrument software). The samples were mounted onto one of four
posts of a PELCO FIB lift-out TEM copper grid and thinned to a
final thickness around 100–200 nm.
Transmission electron microscopy.—Transmission electron mi-
croscopy (TEM) was performed on a Hitachi 9500 TEM equipped
with a LaB6 emitter and operated at 300 kV. Image acquisition was
done using the K2 direct electron detector operated in counting
mode.
Scanning transmission electron microscopy, energy dispersive
X-ray spectroscopy, and scanning electron nanodiffraction.—
Nano- and atomic-scale imaging, chemical analysis, diffraction
data were acquired using the ThermoFisher Scientific Titan
Themis Z STEM equipped with a Schottky field emission gun
operated at 300 kV accelerating voltage. The electrode sample was
loaded into a standard double-tilt holder and, during operation, tilted
close to the 100
〈 〉 zone axis or the 210
〈 〉 zone axis (with respect to
the Graphite-2H phase), as both orientations allowed us to track the
graphite interplanar spacing in both imaging and diffraction modes.
Imaging was performed at 15–30 pA screen current with a
convergence angle of 18.0 mrad and acquired using a high-angle
annular dark field (HAADF) detector. The EDS acquisition was
performed using the above mentioned imaging conditions via the
Velox software (ThermoFisher Scientific) with an acquisition time
of ∼1 h using a continuous scan with a frame time of ∼1 s. SEND
data were collected in microprobe STEM (μP STEM) mode with a
convergence angle of 0.46 mrad. The SEND probe size is 1.7–-
1.8 nm at full-width half maximum. The diffraction datasets were
acquired at a camera length of 145 mm over a real-space scan area of
Figure 2. (a) Capacity-potential profiles of the graphite electrode from a NCM532//Gr cell with a reference electrode. The lithiation profiles are at the indicated
rates; the 0.2 C rate delithiation profile shown is after lithiation at the 2 C rate followed by a 1 h rest. (b) Profiles for as-prepared (black) and harvested electrodes
(red); the latter are from a cell subjected to multiple “escalating C-rate” cycling. The as-prepared electrode data are at C/100 rate, whereas the harvested electrode
plots are at both C/100 (solid) and C/10 (dashed) rates. In both (a) and (b), the specific capacities are obtained by dividing the cell capacity (mAh) by the graphite
weight (gGr
−1
) in the anode.
Journal of The Electrochemical Society, 2021 168 100509
150 nm × 150 nm using 75 × 75 total steps with drift correction
performed after each line acquisition.
Results and Discussion
Electrochemical cycling.—Capacity-potential profiles of the
graphite electrode from a NCM523//Gr cell are shown in Fig. 2a.
The graphite is lithiated and delithiated during cell charge and
discharge, respectively. At low lithiation rates (<C/5), the inter-
calation progresses in distinctive stages seen as plateaus in the plots;
the Li+
ions fill spaces between the carbon sheets forming distinct
LixC6 phases (x = 0.22, 0.33, 0.5, 1). At higher lithiation rates
(⩾1 C), the plateaus are not distinct as intercalation occurs under far-
from-equilibrium conditions. Multiple mixed and strained phases are
probably formed so that the profile appears continuous, rather than
staged; note that LixC6 phases of various compositions have been
reported in the research literature.25–29
During the C/5 delithiation,
staging plateaus are always observed, regardless of the behavior
during the preceding lithiation cycle (Fig. 2a). We observed this
difference between the lithiation and delithiation behavior during
operando XRD studies; therein, we attributed the blurring of phase
boundaries during lithiation to lattice strain and the formation of
distinct LixC6 phases during delithiation to abrupt separation from
oversaturated solid solutions.18
As indicated earlier, the cells in this study were subjected to
multiple applications of the escalating C-rate protocol. Capacity-
voltage profiles from a representative cell, before and after the
protocol, are shown in Fig. S1. The electrochemical cycling causes a
significant loss in cell capacity. The Li+
ions are lost to the solid
electrolyte interphase (SEI) formed during electrolyte reduction on
the graphite electrode and to the dead lithium that accumulates in the
electrode because of repeated excursions below the Li-plating
potential (Fig. 1b).22,30
Figure 2b compares plots from a lithium-plated graphite elec-
trode (Gr-LP) harvested from the cell in Fig. S1, with plots from an
as-prepared graphite (Gr-AP) electrode. The lithiation profile of the
Gr-AP electrode (at C/100 rate) displays distinctive plateaus around
210, 117 and 83 mV vs Li/Li+
. These plateaus are also very distinct
in the C/10 rate lithiation data (not shown), though they occur at
slightly lower potentials because of polarization arising from
electrode impedance. In contrast, the plateaus are not distinct in
the lithiation profile of the Gr-LP electrode, even at a C/100 rate.
The capacity of this aged-harvested electrode (326 mAh g−1
) is also
smaller than that of the as-prepared electrode (357 mAh g−1
). The C/
10 lithiation data of this Gr-LP electrode resembles the 2 C graphite-
lithiation data in Fig. 2a; the sloped profile, the barely visible staging
plateaus and the lower capacity all suggest changes in the graphite
particles that alter lithium insertion into the material. Interestingly,
the staging plateaus are visible both in the C/100 and C/10
delithiation profiles of the harvested material, which is consistent
with the distinct differences between the lithium insertion and
extraction processes mentioned earlier.
The theoretical capacity of graphite is 372 mAh g−1
but the Gr-
AP electrode yields only 357 mAh g−1
even at very low rates. The
underlying reason for this smaller capacity is the inherent turbos-
tratic stacking disorder in the commercial graphite, which limits Li+
ion intercalation into the material.31,32
Fast-charging appears to
create additional disorder, further decreasing capacity. Moreover, the
increased structure-disorder increases the number of electronically
and geometrically non-equivalent sites for Li-intercalation causing
the sloping lithiation profile of the Gr-LP even at extremely low
rates.33
In contrast, most of the sites in Gr-AP are thermodynami-
cally equivalent for Li+
intercalation, thereby producing the
observed staging behavior at low lithiation rates.
Ensemble phase analysis via powder X-ray diffraction.—We
evaluated the bulk-scale structure and ensemble lattice spacing of the
graphite anode coatings using synchrotron X-ray diffraction.
Diffraction patterns obtained on the Gr-AP and Gr-LP electrodes
are shown in Figs 3a and 3b, respectively. Figure 3a shows that the
as-prepared electrode contains two distinct graphite phases: the 2H
phase (marked as #) and the 3 R phase (marked as *). The 3 R phase
is not observed in the Gr-LP electrode pattern (Fig. 3b); we postulate
that this phase transforms into the more stable 2H form during
repeated Li+
insertion into and extraction from the graphite, which
requires gliding of the graphene planes as mentioned in the
Introduction.
We calculated the graphite interplanar d-spacing from the strong
diffraction peak at ∼1.87 Å–1
, which corresponds to the (002)
reflection for the Graphite-2H phase or the (003) reflection for the
Graphite-3R phase. For both Gr-AP and Gr-LP electrode samples we
obtained d 3.356 Å,
interplanar
〈 〉 ≅ which indicates that the latter
sample is fully delithiated and that the amount of residual Li+
ions, within the graphite of the fully discharged cell, is negligible.
However, diffraction pattern of the Gr-LP sample (Fig. 3b) shows
new peaks at ∼2.70 Å–1
and ∼4.40 Å–1
, which arise from lithium
fluoride (LiF) (marked as ♢). The presence of LiF was confirmed by
selected area electron diffraction (SAED) data on a FIB cross-
section prepared from the top surface of the Gr-LP sample (Fig. S2).
LiF is known to be a component of the graphite SEI and forms
through decomposition reactions of the LiPF6 salt.34,35
There was no
evidence of other known SEI phases in the Gr-LP pattern, which
indicates that these compounds are either fully amorphous or lack
the long-range order needed to coherently scatter the X-ray beam.
SEM imaging and compositional analysis.—Figure 4 reveals the
dramatic change to the graphite anode surface after being subjected
to cycling under the fast-charge, lithium-plating condition. Figure 4a
shows a photograph of the Gr-AP sample (cut to a 14.2 mm
diameter); the EDS data in Figs. 4b and S3 indicate that the coating
contains only C and F, reflecting the graphite, C45 and PVDF
Figure 3. Synchrotron powder diffraction data from the following graphite
electrodes: (a) as-prepared (Gr-AP) and (b) cycled with lithium plating (Gr-
LP). The Bragg reflections are indexed according to JCPDS cards 01–071-
–3739 for Graphite-2H, 01–075–2078 for Graphite-3R, and 01–071–3743 for
LiF. The phases are marked using the following symbols: # for Graphite-2H,
* for Graphite-3R, and ♢ for LiF.
Journal of The Electrochemical Society, 2021 168 100509
materials used for electrode preparation. Figure 4c is an SEM image
of the Gr-AP sample surface at 2k × magnification to illustrate that
the electrode is comprised of densely packed, potato-shaped graphite
particles up to 10 μm in size.
SEM and EDS data from the Gr-NP electrode are displayed in
Fig. S4. The graphite particles appear unchanged by the low-rate
cycles; the elemental analyses reveal the presence of C, O, F and P,
which arise from electrolyte decomposition products and residues on
the electrode. By contrast, Fig. 4d shows a photograph of the Gr-LP
sample (cut to 14.2 mm diameter) with an uneven white-coloration
across the surface; the average elemental composition (in wt%) of
the electrode surface (Fig. 4e) is 21% O, 49% F, 27% P and 3% C.
Figure 4f shows the surface morphology of the Gr-LP sample at 2k
× magnification; in this image, there is complete coverage of the
graphite particles by products formed by reactions of the plated
lithium with the electrolyte (also see Fig. S5).
To further determine extent of changes, we examined the
electrodes in cross-section to study morphological and elemental
characteristics throughout their depth (see Figs. 5, S6, S7 and S8).
Details of the coating thickness calculations are provided in Section
4 of the SI and Fig. S9. Figure 5a shows that the Gr-AP sample has
an average coating thickness of 76.6 μm. Figure 5b shows a FIB
cross-section of a typical graphite particle, revealing the presence of
internal pores. The elemental composition across the electrode
thickness was measured via EDS line-scans. For the Gr-AP sample,
the carbon signal, which arises predominantly from the graphite,
measures an average of 97 wt% across the depth of the coating layer
above the copper current collector, while the fluorine signal, which
arises from PVDF binder, measures an average of 3 wt% (Figs. 5c
S6). Cross-sections of the Gr-NP electrode (Fig. S7) reveal an
average coating thickness of 78.1 μm, which is about 2% greater
than that of the Gr-AP electrode. Irreversible increases in graphite
electrode thickness have been noted previously in electrochemical
dilatometry studies; these typically range from 2%–5% after a few
cycles.36–38
EDS scans reveal carbon, oxygen, fluorine and phos-
phorous signals, distributed across the electrode. These elements are
from the electrolyte decomposition products that accumulate in the
electrode.
By contrast, the Gr-LP electrode has an average coating thickness
of 91.3 μm, which does not include the ∼15–20 μm thick layer
resulting from lithium plating on the anode surface (see Fig. 5d).
This coating thickness is 19.2% and 16.9% greater than that of the
Gr-AP and Gr-NP electrodes, respectively. We attribute this
electrode thickness increase to SEI buildup in the electrode pores
and rearrangement of graphite particles during the electrochemical
cycling. The process is exacerbated by Li plating and subsequent
electrolyte reduction, which increases the amount of SEI. This SEI
buildup is also seen on individual graphite particles, on the exterior
surfaces and in the internal pores (Fig. 5e). Moreover, EDS line-
scans show that the elemental composition is non-uniform across the
electrode thickness (Figs. 5f and S8). The topmost portion, which
has no graphite, contains mainly oxygen, fluorine and phosphorus
indicating that the products of lithium plating are mainly inorganic—
for example, LiF and lithium (oxy)fluoro-phosphate compounds.
Below this portion, in the electrode coating, carbon is the dominant
element but we also observe an abundance of fluorine and
phosphorus, possibly (oxy)fluoro-phosphate compounds resulting
from hydrolysis reactions of the electrolyte salt.39,40
STEM-EDS compositional analysis.—Using STEM-EDS, we
further show that the (oxy)fluoro-phosphate compounds only coat
the surface of the graphite particle and do not penetrate the bulk.
Figure 4. Comparison of the anode surface morphology and composition. The as-prepared graphite (Gr-AP) anode surface details are elucidated via: (a)
photograph at 1x magnification, (b) EDS elemental composition, in weight percent, and (c) SEM image at 2000 × magnification. Similarly, the Gr-LP anode
surface details are elucidated via: (d) photograph at 1x magnification, (e) EDS elemental composition in weight percent, and (f) SEM image at 2000 ×
magnification. Scale bars in (a) and (d): 5 mm.
Journal of The Electrochemical Society, 2021 168 100509
Figure 6a shows a survey of the Gr-LP particle section (prepared by
FIB) with EDS elemental analysis performed in the marked region,
which is magnified in Fig. 6b.
We reveal compositional heterogeneity by comparing the local
EDS spectra from graphite edges (marked as Area #1 and Area #3)
to the graphite center (marked as Area #2) shown in Fig. 6c(i). The
spectra from Area #1 and Area #3 show signals for O, F, and P in
addition to the signal from C. However, Area #2 only shows C
signature, which indicates a lack of (oxy)fluoro-phosphate com-
pounds in the graphite interior. Figure 6c(ii) shows the spectra
corresponding to the whole area in Fig. 6b. The elemental maps
(Fig. 6d) further confirm the spatial heterogeneity between carbon
and the (oxy)fluoro-phosphates; the C K signal appears only in the
graphite bulk, while the O, F, and P signals appear on the graphite
band edges. Moreover, the O, F, and P signals appear in the same
Figure 5. SEM imaging and elemental mapping of the electrode cross-sections. The as-prepared graphite (Gr-AP) sample details are elucidated via (a) SEM
image of the electrode cross-section, (b) SEM image of an individual particle section, and (c) EDS elemental data acquired across the thickness from bottom of
the current collector to the top of the coating surface—see white arrow in (a). Similarly, the lithium-plated graphite electrode (Gr-LP) sample details are
elucidated via (d) SEM image of the electrode cross-section, (e) SEM image of an individual particle section, and (f) EDS elemental data acquired across the
thickness depth from bottom of the current collector to the top surface—see white arrow in (d). Scale bars: (a), (d) are 25 μm; (b), (e) are 2 μm.
Journal of The Electrochemical Society, 2021 168 100509
locations across the EDS maps. Additional EDS maps for different
regions of the Gr-LP electrodes are provided in Figs. S10 and S11.
The above observations indicate that the SEI only accumulates on
the graphite particle edges and does not penetrate the bulk material.
Electrolyte solvent co-intercalation would cause excessive strain
and/or damage to the host graphite. We do not see such damage in
the bulk graphite and, therefore, do not anticipate extensive
electrolyte penetration. We also observe that smaller pores within
the graphite yield weaker electrolyte residue signatures compared to
the larger pores. The electrolyte is known to penetrate the intra-
particle pores during electrochemical cycling.41
Larger pores pro-
vide easier access to the electrolyte; hence, more SEI would be
expected to accumulate in the larger pores than in the smaller pores.
Revealing localized disorder via high-resolution TEM.—High-
resolution TEM (HRTEM) data reveal the nature of the turbostratic
disorder that manifests in graphite particles at the lattice-level
(Fig. 7). Figure 7a, spanning a 50 nm × 50 nm field-of-view, shows
that the Gr-AP sample has straight and uniform graphite lattice
fringes corresponding to the constituent graphene planes. From this
image, we can qualitatively assess the degree of uniformity by
studying Region 1 (Fig. 7b) and Region 2 (Fig. 7c). The lattice-
fringe spacing, measured directly from the images of both regions, is
∼0.33 nm, which is consistent with the interplanar d-spacing of
graphite determined by XRD.
Fast-Fourier Transform (FFT) analysis is useful for quantifying
geometric and periodic patterns in electron microscopy images.42,43
The localized Fast-Fourier Transform (FFT) of the Region 1 and
Region 2 images are shown in Figs. 7d and 7e, respectively;
information to explain features observed in the HRTEM images is
contained in Section 6 and Fig. S12 of the SI. Briefly, the FFT of a
real-space intensity image (with distances x in nm) yields a
representative image in the corresponding spatial frequency domain
(with distances x
1/ in nm–1
).42,43
By applying FFT to the image in
Fig. 7b, the periodic stripe pattern of the graphite lattice (with
measured spacing 0.33 nm
∼ ) is mapped to points in Fig. 7d
corresponding to the frequency of the lattice spacing (i.e.,
1 0.33 nm 1
∼ / − from the center). Moreover, because the intensity
variation of the fringes lies along the direction normal to the fringes
in Fig. 7b (specified as direction 1 on the shown axis), the
corresponding spots in the frequency domain of Fig. 7d lie along
this same direction. The higher frequency spots in Fig. 7d that lie
along the same line (direction 1) correspond to additional harmonics
(i.e., n 1 0.33 nm 1
× ∼ / − where n is an integer) of the real-space
lattice; observation of such harmonics implies that the lattice is
relatively straight and uniform (i.e., ordered) within Region 1. In the
specified direction 2 (normal to direction 1), spots are occasionally
observed within the lattice fringes (Fig. 7b); these spots likely arise
from atomic columns within the graphite and the distance between
them correspond to the spacing between such columns. Similar
observations can be made for the Region 2 image (Fig. 7c) and its
corresponding FFT (Fig. 7e), thereby illustrating spatial uniformity
of the Gr-AP lattice.
Figure 6. STEM-EDS of a graphite particle cross-section from the Gr-LP sample. HAADF-STEM images of (a) graphite particle section and (b) a close-up of
the area where the EDS maps were acquired. (c) EDS spectra represented by the following regions: (i) from selected areas 1 (top-edge of graphite edge), 2 (center
of graphite band), and 3 (bottom-edge of graphite band) as marked in (b); (ii) whole-area EDS spectrum. (d) EDS maps of (i) C, (ii) O, (iii) F, and (iv) P signals.
Scale bars: (a) 2 μm, (b) 100 nm, (d) (i)–(iv) 50 nm.
Journal of The Electrochemical Society, 2021 168 100509
By contrast, HRTEM images of the Gr-LP sample (Fig. 7f), show
significant nonuniformity. This nonuniformity is illustrated by
example from the magnified Region 1 (Fig. 7g) and Region 2
(Fig. 7h); localized FFTs of Regions 1 and 2 are shown in Figs. 7i
and 7j, respectively. Region 1, which is taken near the surface of an
internal pore within a graphite particle, is different from both
Regions 1 and 2 of the Gr-AP sample. From Fig. 7g, we find that
the Gr-LP lattice of Region 1 appears wavy; based on Fig. 7f, we
estimate that this waviness extends about 20 nm into the graphite
bulk. The FFT of Region 1 (Fig. 7i) confirms lattice disorder in the
following ways: (i) arcs are found instead of spots along direction 1;
(ii) the high frequency harmonic spot (marked by the white arrow) is
relatively faint compared to those in the Gr-AP sample. Furthermore,
the interplanar d-spacing from Fig. 7g measures up to ∼0.37 nm,
indicating that the lattice in this region does not return to its original
delithiated state. This nanoscopic increase in interplanar spacing is
not captured in the powder diffraction data (Fig. 3) because the latter
technique provides an average from all portions of the sample. On
the other hand, Region 2, which is in the bulk of the Gr-LP sample,
shows straight, uniform lattice fringes (Fig. 7h) similar to those
observed in the Gr-AP sample (Figs. 7b, 7c). The corresponding FFT
of Region 2 (Fig. 7j) has spots along direction 1 and a high-
frequency harmonic, which suggests greater ordering within the
graphite bulk. Taken together, the HRTEM images and FFT results
indicate there is greater disorder in the graphite lattice near the
Figure 7. Direct electron imaging of the graphite lattice via HRTEM. The as-prepared graphite (Gr-AP) electrode lattice details are elucidated via (a) TEM
imaging with highlighted Regions 1 and 2; real-space magnification of (b) Region 1 and (c) Region 2; localized Fast-Fourier Transforms of (d) Region 1 and (e)
Region 2. Correspondingly, the Gr-LP lattice details are elucidated via (f) TEM imaging with highlighted Regions 1 and 2; real-space magnification of (g)
Region 1 and (h) Region 2; and localized Fast-Fourier Transforms of (i) Region 1 and (j) Region 2. Scale bars: (a), (f) are 10 nm; (b), (c), (g), (h) are 2 nm; (d),
(e), (i), (j) are 5 nm–1
.
Journal of The Electrochemical Society, 2021 168 100509
surface of its internal pores and that this disorder can extend up to
∼20 nm into the particle bulk.
Quantification of nanoscopic turbostratic disorder via SEND.—
In Fig. 8 and 9, we quantify the nanoscale heterogeneity of the
graphite by measuring the interplanar d-spacing and local lattice
rotation using SEND.44
In this technique, a ∼2 nm probe (at full-
width half-maximum) is rastered across a 150 nm × 150 nm real-
space area generating an electron diffraction pattern for each probe
position. SEND provides position-resolved structural information
regarding turbostratic disorder in the materials. Details of the SEND
data acquisition are briefly described in the experimental section;
data analysis information is provided in Section 7 of the SI, which
includes Fig. S13. In the SEND experiments, we quantify both the
graphite interplanar spacing and the lattice rotation by tracking
changes to the Ginterplanar vector (with respect to the Graphite-2H
phase) for each real-space sample coordinate of the SEND scan. The
quantification is done by determining (1) changes to the length of the
vector G d
1
interplanar interplanar
∣ ∣ = to measure position-to-position
nanoscopic interplanar d-spacing, which is then represented as a
dinterplanar map to highlight local d-spacing changes; (2) changes to
the in-plane orientation θ of each vector G ,
interplanar which is then
represented as a map of θ
∣∇ ∣ to identify locations where rotational
changes in the lattice occur.
Figure 8a(i) contains HAADF-STEM image from a Gr-AP
particle cross-section, with marked Region 1 (magnified in
Fig. 8a(ii)) and Region 2 (magnified in Fig. 8a(iii)) where SEND
data were acquired. First, we discuss the nanoscale structure of
Region 1 in Fig. 8b. The data indicate that the lattice rotation shows
negligible variation ( 0.09 nm 1
∼ ° − ) across the sample (Fig. 8b(i)) and
that the interplanar d-spacing map across Region 1 is relatively
uniform (Fig. 8b(ii)). Figure 8b(iii) is a plot of the overall d-spacing
from the map as a histogram; the data show a distribution of d-
spacings in the sample with a mean of 0.334 nm and a standard
deviation of 0.006 nm. Figure 8c illustrates that Region 2 is
comparable to Region 1 in terms of both lattice rotation
(Fig. 8c(i)) and interplanar spacings (Figs. 8c(ii) and (iii)).
However, we find a slightly elevated value for rotational variation
in the particle bulk of Fig. 8c(i). This observation points to an
important characteristic of graphite, which is that the graphene
planes tend to group as bands of 10–20 nm thickness (Fig. S14).
Within each graphite band, the magnitude of the rotational variation
is negligible; between band-to-band, the in-plane lattice angle can
exhibit sudden jumps. Overall, though, the results for the Gr-AP
sample are consistent with the HRTEM observations, which also
indicated a uniform lattice. Furthermore, SEND data from the Gr-NP
electrode (Fig. S15) are very similar to those from the Gr-AP
electrode, which indicate that the low-rate cycles have negligible
impact on the graphite particle structure.
By contrast, a non-uniform lattice and an increase in lattice
disorder at the particle edges are observed in particles from the Gr-
LP electrode. Figure 9a(i) contains a HAADF-STEM image with
areas marked Region 1 (magnified in Fig. 9a(ii)) and Region 2
(magnified in Fig. 9a(iii)) where SEND data were collected. In
Fig. 9b(i), we show position-to-position changes of the in-plane
rotation angle of Ginterplanar in Region 1, which is near a relatively
thin pore region. The graphite bulk in Region 1 has a uniform lattice
rotation ( 0.14 nm 1
∼ ° − ), while the graphite edge has greater disorder
(with rotational changes 2 nm 1
> ° − ). Figure 9b(ii) shows the corre-
sponding interplanar d-spacing reconstruction map; the d-spacing is
elevated ( 0.37 nm
> ) only within ∼2 nm of the particle edge while
the bulk has d-spacings similar to those in the Gr-AP sample. The d-
spacing distribution (Fig. 9b(iii)) shows a mean around 0.337 nm
and a standard deviation of 0.008 nm, only slightly larger than those
for the Gr-AP sample.
On the other hand, Region 2 (acquired near the edge of a larger
graphite pore) of the Gr-LP sample (Fig. 9c) shows greater lattice
disorder. The average rotation gradient in this region is 0.30 nm,
̃ °/
while the particle edges have values 4 nm
> °/ (Fig. 9c(i)). Moreover,
the frequency of lattice rotation (i.e., the “wavy” nature of the
lattice) occurs up to ∼20 nm into the graphite bulk from the pore
edge. A similar behavior is observed in Fig. 9c(ii), which shows
elevated d-spacings near the pore edge and extending into the
particle bulk. Figure 9c(iii) shows that the mean d-spacing of Region
2 is 0.35 nm, which is larger than that calculated from the X-ray
data. The d-spacing distribution also shows a tail that extends from
∼0.36 nm to 0.40 nm, which indicates that the graphite edges in
Region 2 have sustained greater damage than those in Region 1. This
result also points to the spatial heterogeneity in lattice disorder
within each graphite particle—the edges with greater exposure to the
electrolyte will experience more Li+
ion (de)intercalation reactions
and apparently greater damage, especially during high-rate charging.
Damage localized at the graphite particle edges is not entirely
unexpected.45
Prior studies have pointed out that Li+
ion intercala-
tion into graphite occurs primarily through the edge planes.46,47
At
high charge rates, significant build-up of Li+
ions is expected at the
edges planes; the concentration gradients between the surface and
bulk sites could induce stresses that are significant enough to deform
the graphite lattice near the particle edges.
Mitigating damage at the graphite particle edges is crucial to
prolonging the life of LIB cells subjected to fast charge. One
solution would be to raise the lower cutoff voltage (LCV) for cell
cycling (e.g. from 2.8 to 3.3 V) to prevent complete delithiation of
the graphite. The residual dilute LixC6 phases have higher d-spacings
than the pristine graphite,29
which would facilitate Li+
ion inter-
calation and (partially) mitigate the stresses induced during the
process. Alternative solutions could include the use of graphite with
higher d-spacings in the anode, as has been suggested for both Li+
and Na+
ion cells.48,49
However, such “expanded” graphite materials
typically have lower electronic conductivity, which would limit the
overall rate of intercalation. In addition, these materials have higher
surface areas and a proclivity to trap more lithium in the SEI formed
during electrolyte reduction reactions. Other solutions proposed to
facilitate rapid cycling include the use of functionalized-graphite that
have higher d-spacings at the particle edges,50
graphite with particle
edges aligned perpendicular to the current collector,51
graphite with
carbon coatings that have larger interlamellar spacing,52
and anodes
containing a mix of graphite and hard carbons.7
These efforts, along
with research on cathode and electrolyte materials, electrode
architectures and cycling protocols, are important steps towards
making fast-charging of LIB cells, a commercial reality.
Conclusions
Fast charging of LIB cells can degrade their energy and power
performance characteristics. Here we show representative data from
as-prepared graphite electrodes (Gr-AP) and from electrodes ex-
tracted from (discharged) cells that were previously charged at rates
up to 6 C (Gr-LP). Electrochemical measurements show that the
harvested electrodes have lower capacities than the Gr-AP electrode
and display a sloping lithiation profile, rather than a staged profile,
even at low rates. Analysis of X-ray diffraction data from the Gr-AP
and Gr-LP electrodes yield similar interplanar d-spacings, indicating
that the graphite bulk is not altered by the electrochemical cycling.
Similar conclusions about the graphite particle bulk can be drawn
from the analysis of HREM and HAADF-STEM images and SEND
data. However, these nanoscale data clearly indicate damage at the
graphite edges, which extend up to ∼20 nm into the particle bulk for
the Gr-LP sample. This damage is seen as arcs (rather than spots) in
the FFT data of HREM images and as increased rotations and higher
d-spacings in the SEND data; both datasets suggest increased
turbostratic disorder and buckling of the graphene planes at particle
edges. In addition to changes in graphite-edge structure, the Gr-LP
electrode shows a significant thickness increase from the build-up of
SEI and particle rearrangements that occur during the cycling. This
SEI contains a mix of compounds, including LiF (seen in the XRD
Journal of The Electrochemical Society, 2021 168 100509
patterns) and (oxy)fluoro-phosphates suggested by the EDS data.
Larger quantities of this SEI are present adjacent to the larger
internal pores of the graphite, which likely have easier access to the
electrolyte than the smaller pores. Furthermore, the SEI is localized
Figure 8. Quantitative lattice rotation and interplanar d-spacing analysis of the as-prepared graphite (Gr-AP) sample. (a) HAADF-STEM images (i) at low-
magnification, showing Regions 1 and 2 where SEND acquisition was performed; (ii), (iii) magnified images of Regions 1 and 2, respectively. Diffractive image
reconstructions and analysis of (b) Region 1 and (c) Region 2 containing the following: (i) in-plane rotation gradient map, (ii) interplanar spacing map, (iii)
interplanar spacing distribution. Scale bars: (a) (i) is 2 μm; (a) (ii), (a) (iii), (b) (i), (b) (ii), (c) (i), (c) (ii) are 50 nm.
Journal of The Electrochemical Society, 2021 168 100509
at the particle edges and does not penetrate the graphite bulk, suggesting that only Li+
ions (no electrolyte components) intercalate
into the particles (effective desolvation of ions).
Figure 9. Quantitative lattice rotation and interplanar d-spacing analysis of the fast-charged graphite (Gr-LP) electrode. (a) HAADF-STEM images (i) at low-
magnification showing Regions 1 and 2 where SEND data were acquired; (ii), (iii) magnified images of Regions 1 and Region 2. Diffractive image
reconstructions and analysis of (b) Region 1 and (c) Region 2 containing the following: (i) in-plane rotation gradient map, (ii) interplanar spacing map, (iii)
interplanar spacing distribution. Scale bars: (a) (i) is 2 μm; (a) (ii), (a) (iii), (b) (i), (b) (ii), (c) (i), (c) (ii) are 50 nm.
Journal of The Electrochemical Society, 2021 168 100509
The implication of our study is that even a small number of high-
rate cycles appears to be sufficient to induce significant and
permanent disorder in graphite domains that are closest to the
electrolyte, be it at particle surfaces or the inner pores. Such disorder
could affect the lithiation kinetics, preventing the graphite particles
from accepting charge and favoring the occurrence of Li plating.
Moreover, we show that these structural changes are highly non-
uniform across the various pores within a same particle, which could
either be a cause or consequence of reaction heterogeneities
commonly observed at high rates. Awareness and mitigation of the
structural evolution at the graphite/electrolyte interfaces is crucial to
the development of LIBs that can endure repeated fast-charging and
yet reliably deliver high-performance over the 10 + year lifespan of
electric vehicles.
Acknowledgments
SP acknowledges support from the U.S. Department of Energy
(DOE) Graduate Student Research (SCGSR) program. The SCGSR
program is administered by the Oak Ridge Institute for Science and
Education for DOE under contract number DE‐SC0014664. This
work was carried out in part at the Materials Research Laboratory
Central Research Facilities, University of Illinois. DA and MTFR
acknowledge support from DOE’s Vehicle Technologies Office
(VTO). The electrodes used in this article were fabricated at
Argonne’s Cell Analysis, Modeling and Prototyping (CAMP)
Facility, which is fully supported by the VTO. We are grateful to
our many colleagues, especially Stephen Trask and Andrew Jansen
at Argonne. Use of the Advanced Photon Source (APS) at Argonne
National Laboratory was supported by the U.S. Department of
Energy, Office of Science, Office of Basic Energy Sciences, under
Contract No. DE-AC02–06CH11357. We are grateful to Saul
Lapidus at the APS for help with the high-resolution powder
diffraction experiments. The submitted manuscript has been created
by UChicago Argonne, LLC, Operator of Argonne National
Laboratory (“Argonne”). Argonne, a U.S. Department of Energy
Office of Science laboratory, is operated under Contract No. DE-
AC02–06CH11357. The U.S. Government retains for itself, and
others acting on its behalf, a paid-up nonexclusive, irrevocable
worldwide license in said article to reproduce, prepare derivative
works, distribute copies to the public, and perform publicly and
display publicly, by or on behalf of the Government.
ORCID
Saran Pidaparthy https://orcid.org/0000-0002-0783-3094
Marco-Tulio F. Rodrigues https://orcid.org/0000-0003-0833-6556
Jian-Min Zuo https://orcid.org/0000-0002-5151-3370
Daniel P. Abraham https://orcid.org/0000-0003-0402-9620
References
1. A. Masias, J. Marcicki, and W. A. Paxton, ACS Energy Lett., 6, 621 (2021).
2. A. Raj, M.-T. F. Rodrigues, and D. P. Abraham, J. Electrochem. Soc., 167, 120517
(2020).
3. A. Tomaszewska et al., eTransportation, 1, 100011 (2019).
4. S. Ahmed et al., J. Power Sources, 367, 250 (2017).
5. A. M. Colclasure et al., Electrochim. Acta, 337, 135854 (2020).
6. X.-G. Yang, T. Liu, Y. Gao, S. Ge, Y. Leng, D. Wang, and C.-Y. Wang, Joule, 3,
3002 (2019).
7. K.-H. Chen, V. Goel, M. J. Namkoong, M. Wied, S. Müller, V. Wood, J. Sakamoto,
K. Thornton, and N. P. Dasgupta, Adv. Energy Mater., 11, 2003336 (2021).
8. D. W. Dees, M.-T. F. Rodrigues, K. Kalaga, S. E. Trask, I. A. Shkrob, D.
P. Abraham, and A. N. Jansen, J. Electrochem. Soc., 167, 120528 (2020).
9. M.-T. F. Rodrigues, K. Kalaga, S. E. Trask, D. W. Dees, I. A. Shkrob, and D.
P. Abraham, J. Electrochem. Soc., 166, A996 (2019).
10. I. A. Shkrob, M.-T. F. Rodrigues, D. W. Dees, and D. P. Abraham, J. Electrochem.
Soc., 166, A3305 (2019).
11. M.-T. F. Rodrigues, I. A. Shkrob, A. M. Colclasure, and D. P. Abraham,
J. Electrochem. Soc., 167, 130508 (2020).
12. T. Waldmann, B.-I. Hogg, and M. Wohlfahrt-Mehrens, J. Power Sources, 384, 107
(2018).
13. L. Somerville, J. Bareño, S. Trask, P. Jennings, A. McGordon, C. Lyness, and
I. Bloom, J. Power Sources, 335, 189 (2016).
14. J. Cannarella and C. B. Arnold, J. Electrochem. Soc., 162, A1365 (2015).
15. J. S. Okasinski, I. A. Shkrob, A. Chuang, M.-T. F. Rodrigues, A. Raj, D. W. Dees,
and D. P. Abraham, Phys. Chem. Chem. Phys., 22, 21977 (2020).
16. S. Müller, J. Eller, M. Ebner, C. Burns, J. Dahn, and V. Wood, J. Electrochem. Soc.,
165, A339 (2018).
17. A. Mukhopadhyay and B. W. Sheldon, Prog. Mater Sci., 63, 58 (2014).
18. A. Raj, I. A. Shkrob, J. S. Okasinski, M.-T. Fonseca Rodrigues, A. C. Chuang,
X. Huang, and D. P. Abraham, J. Power Sources, 484, 229247 (2021).
19. K. P. C. Yao, J. S. Okasinski, K. Kalaga, I. A. Shkrob, and D. P. Abraham, Energy
Environ. Sci., 12, 656 (2019).
20. M. T. F. Rodrigues, S.-B. Son, A. M. Colclasure, I. A. Shkrob, S. E. Trask, I.
D. Bloom, and D. P. Abraham, ACS Appl. Energy Mater., 4, 1063 (2021).
21. M.-T. Fonseca Rodrigues, V. A. Maroni, D. J. Gosztola, K. P. C. Yao, K. Kalaga, I.
A. Shkrob, and D. P. Abraham, ACS Appl. Energy Mater., 2, 873 (2019).
22. P. P. Paul et al., , Adv. Energy Mater., 11, 2100372 (2021).
23. H. Shi, J. Barker, M. Y. Saïdi, R. Koksbang, and L. Morris, J. Power Sources, 68,
291 (1997).
24. A. Van der Ven, Z. Deng, S. Banerjee, and S. P. Ong, Chem. Rev., 120, 6977
(2020).
25. H. Fujimoto, H. Kiuchi, S. Takagi, K. Shimoda, K. Okazaki, Z. Ogumi, and T. Abe,
J. Electrochem. Soc., 168, 40509 (2021).
26. H. He, C. Huang, C.-W. Luo, J.-J. Liu, and Z.-S. Chao, Electrochim. Acta, 92, 148
(2013).
27. S. Schweidler, L. de Biasi, A. Schiele, P. Hartmann, T. Brezesinski, and J. Janek,
J. Phys. Chem. C, 122, 8829 (2018).
28. J. Wilhelm, S. Seidlmayer, S. Erhard, M. Hofmann, R. Gilles, and A. Jossen,
J. Electrochem. Soc., 165, A1846 (2018).
29. E. Hazrati, G. A. de Wijs, and G. Brocks, Phys. Rev. B, 90, 155448 (2014).
30. S.-B. Son, D. Robertson, Z. Yang, Y. Tsai, S. Lopykinski, and I. Bloom,
J. Electrochem. Soc., 167, 140506 (2020).
31. R. E. Franklin, Acta Crystallogr., 4, 253 (1951).
32. J. R. Dahn, A. K. Sleigh, H. Shi, J. N. Reimers, Q. Zhong, and B. M. Way,
Electrochim. Acta, 38, 1179 (1993).
33. J. R. Dahn, R. Fong, and M. J. Spoon, Phys. Rev. B, 42, 6424 (1990).
34. P. Verma, P. Maire, and P. Novák, Electrochim. Acta, 55, 6332 (2010).
35. A. M. Andersson and K. Edström, J. Electrochem. Soc., 148, A1100 (2001).
36. D. Sauerteig, S. Ivanov, H. Reinshagen, and A. Bund, J. Power Sources, 342, 939
(2017).
37. B. Rieger, S. Schlueter, S. V. Erhard, J. Schmalz, G. Reinhart, and A. Jossen,
J. Energy Storage, 6, 213 (2016).
38. A. Y. R. Prado, M.-T. F. Rodrigues, S. E. Trask, L. Shaw, and D. P. Abraham,
J. Electrochem. Soc., 167, 160551 (2020).
39. R. Wagner, M. Korth, B. Streipert, J. Kasnatscheew, D. R. Gallus, S. Brox,
M. Amereller, I. Cekic-Laskovic, and M. Winter, ACS Appl. Mater. Interfaces, 8,
30871 (2016).
40. M. Stich, M. Göttlinger, M. Kurniawan, U. Schmidt, and A. Bund, J. Phys. Chem.
C, 122, 8836 (2018).
41. D. J. Miller, C. Proff, J. G. Wen, D. P. Abraham, and J. Bareño, Adv. Energy
Mater., 3, 1098 (2013).
42. A. Marion, , ed. A. Marion (Springer, US, Boston, MA) p. 255 (1991).
43. R. C. Gonzalez and R. E. Richard E. Woods, Digital image processing 2nd ed.
(Prentice Hall, Upper Saddle River, N.J) (2002).
44. J. M. Zuo and J. C. H. Spence, Advanced Transmission Electron Microscopy
(Springer, New York, NY) (2017).
45. V. A. Sethuraman, L. J. Hardwick, V. Srinivasan, and R. Kostecki, J. Power
Sources, 195, 3655 (2010).
46. R. J. Rice and R. L. McCreery, Anal. Chem., 61, 1637 (1989).
47. K. Persson, V. A. Sethuraman, L. J. Hardwick, Y. Hinuma, Y. S. Meng, A. van der
Ven, V. Srinivasan, R. Kostecki, and G. Ceder, J. Phys. Chem. Lett., 1, 1176 (2010).
48. T.-H. Kim, E. K. Jeon, Y. Ko, B. Y. Jang, B.-S. Kim, and H.-K. Song, J. Mater.
Chem. A, 2, 7600 (2014).
49. Y. Wen, K. He, Y. Zhu, F. Han, Y. Xu, I. Matsuda, Y. Ishii, J. Cumings, and
C. Wang, Nat. Commun., 5, 4033 (2014).
50. J.-S. Park, M.-H. Lee, I.-Y. Jeon, H.-S. Park, J.-B. Baek, and H.-K. Song, ACS
Nano, 6, 10770 (2012).
51. J. Billaud, F. Bouville, T. Magrini, C. Villevieille, and A. R. Studart, Nat. Energy,
1, 16097 (2016).
52. W. Cai, C. Yan, Y.-X. Yao, L. Xu, R. Xu, L.-L. Jiang, J.-Q. Huang, and Q. Zhang,
Small Struct., 1, 2000010 (2020).
Journal of The Electrochemical Society, 2021 168 100509

More Related Content

What's hot

Synthesis characterisation_and_evaluation_of_ir_o2_based_binary_metal_oxide_...
Synthesis  characterisation_and_evaluation_of_ir_o2_based_binary_metal_oxide_...Synthesis  characterisation_and_evaluation_of_ir_o2_based_binary_metal_oxide_...
Synthesis characterisation_and_evaluation_of_ir_o2_based_binary_metal_oxide_...madlovescience
 
Walker Electrochemical Paper
Walker Electrochemical PaperWalker Electrochemical Paper
Walker Electrochemical PaperPatrick Walker
 
33nd battery conference poster zhou TB JC
33nd battery conference poster zhou TB JC33nd battery conference poster zhou TB JC
33nd battery conference poster zhou TB JCJigang Zhou
 
Synthesis and-optimisation-of-ir o-2-electrocatalysts-by-adams-fusion-method-...
Synthesis and-optimisation-of-ir o-2-electrocatalysts-by-adams-fusion-method-...Synthesis and-optimisation-of-ir o-2-electrocatalysts-by-adams-fusion-method-...
Synthesis and-optimisation-of-ir o-2-electrocatalysts-by-adams-fusion-method-...Science Padayatchi
 
Depositacion electroforetica dentro de campos electricos modulados
Depositacion electroforetica dentro de campos electricos moduladosDepositacion electroforetica dentro de campos electricos modulados
Depositacion electroforetica dentro de campos electricos moduladosMario ML
 
FYP Report-Xing Dan
FYP Report-Xing DanFYP Report-Xing Dan
FYP Report-Xing Dan#Xing Dan#
 
Lithium Batteries and Supercapacitors
Lithium Batteries and SupercapacitorsLithium Batteries and Supercapacitors
Lithium Batteries and SupercapacitorsBing Hsieh
 
F041033947
F041033947F041033947
F041033947IOSR-JEN
 
Image reversal resist photolithography of silicon based platinum and silver m...
Image reversal resist photolithography of silicon based platinum and silver m...Image reversal resist photolithography of silicon based platinum and silver m...
Image reversal resist photolithography of silicon based platinum and silver m...Journal Papers
 
Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...
Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...
Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...materials87
 
Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...
Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...
Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...sudesh789
 
2014 Journal of Power Sources 247 (2014) 572-578
2014 Journal of Power Sources 247 (2014) 572-5782014 Journal of Power Sources 247 (2014) 572-578
2014 Journal of Power Sources 247 (2014) 572-578Alexis B. B
 
Proton‐functionalized graphitic carbon nitride for efficient metal‐free disin...
Proton‐functionalized graphitic carbon nitride for efficient metal‐free disin...Proton‐functionalized graphitic carbon nitride for efficient metal‐free disin...
Proton‐functionalized graphitic carbon nitride for efficient metal‐free disin...Journal Papers
 

What's hot (20)

J0436469
J0436469J0436469
J0436469
 
C0421318
C0421318C0421318
C0421318
 
Paper
PaperPaper
Paper
 
Synthesis characterisation_and_evaluation_of_ir_o2_based_binary_metal_oxide_...
Synthesis  characterisation_and_evaluation_of_ir_o2_based_binary_metal_oxide_...Synthesis  characterisation_and_evaluation_of_ir_o2_based_binary_metal_oxide_...
Synthesis characterisation_and_evaluation_of_ir_o2_based_binary_metal_oxide_...
 
Walker Electrochemical Paper
Walker Electrochemical PaperWalker Electrochemical Paper
Walker Electrochemical Paper
 
33nd battery conference poster zhou TB JC
33nd battery conference poster zhou TB JC33nd battery conference poster zhou TB JC
33nd battery conference poster zhou TB JC
 
Synthesis and-optimisation-of-ir o-2-electrocatalysts-by-adams-fusion-method-...
Synthesis and-optimisation-of-ir o-2-electrocatalysts-by-adams-fusion-method-...Synthesis and-optimisation-of-ir o-2-electrocatalysts-by-adams-fusion-method-...
Synthesis and-optimisation-of-ir o-2-electrocatalysts-by-adams-fusion-method-...
 
Depositacion electroforetica dentro de campos electricos modulados
Depositacion electroforetica dentro de campos electricos moduladosDepositacion electroforetica dentro de campos electricos modulados
Depositacion electroforetica dentro de campos electricos modulados
 
FYP Report-Xing Dan
FYP Report-Xing DanFYP Report-Xing Dan
FYP Report-Xing Dan
 
Lithium Batteries and Supercapacitors
Lithium Batteries and SupercapacitorsLithium Batteries and Supercapacitors
Lithium Batteries and Supercapacitors
 
F041033947
F041033947F041033947
F041033947
 
Image reversal resist photolithography of silicon based platinum and silver m...
Image reversal resist photolithography of silicon based platinum and silver m...Image reversal resist photolithography of silicon based platinum and silver m...
Image reversal resist photolithography of silicon based platinum and silver m...
 
Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...
Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...
Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...
 
Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...
Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...
Synthesis and optimisation of ir o2 electrocatalysts by adams fusion method f...
 
01
0101
01
 
1-s2.0-S1369800114006258-main
1-s2.0-S1369800114006258-main1-s2.0-S1369800114006258-main
1-s2.0-S1369800114006258-main
 
Mitra. paper
Mitra. paperMitra. paper
Mitra. paper
 
2014 Journal of Power Sources 247 (2014) 572-578
2014 Journal of Power Sources 247 (2014) 572-5782014 Journal of Power Sources 247 (2014) 572-578
2014 Journal of Power Sources 247 (2014) 572-578
 
Proton‐functionalized graphitic carbon nitride for efficient metal‐free disin...
Proton‐functionalized graphitic carbon nitride for efficient metal‐free disin...Proton‐functionalized graphitic carbon nitride for efficient metal‐free disin...
Proton‐functionalized graphitic carbon nitride for efficient metal‐free disin...
 
93 pooja
93 pooja93 pooja
93 pooja
 

Similar to Pidaparthy 2021 j._electrochem._soc._168_100509

A. Rock - Final Report - 2015
A. Rock - Final Report - 2015A. Rock - Final Report - 2015
A. Rock - Final Report - 2015Anthony Rock
 
Lithium Iron Phosphate: Olivine Material for High Power Li-Ion Batteries - Cr...
Lithium Iron Phosphate: Olivine Material for High Power Li-Ion Batteries - Cr...Lithium Iron Phosphate: Olivine Material for High Power Li-Ion Batteries - Cr...
Lithium Iron Phosphate: Olivine Material for High Power Li-Ion Batteries - Cr...CrimsonPublishersRDMS
 
Rechargeable Li-ion batteries based on Olivine-structured (LiFePO4) catho...
Rechargeable Li-ion batteries based on Olivine-structured     (LiFePO4) catho...Rechargeable Li-ion batteries based on Olivine-structured     (LiFePO4) catho...
Rechargeable Li-ion batteries based on Olivine-structured (LiFePO4) catho...Arun Kumar
 
Research plan 2
Research plan 2Research plan 2
Research plan 2Toru Hara
 
Electrochemical Investigation of Electrolyte & Anodic Materials for Sodium Io...
Electrochemical Investigation of Electrolyte & Anodic Materials for Sodium Io...Electrochemical Investigation of Electrolyte & Anodic Materials for Sodium Io...
Electrochemical Investigation of Electrolyte & Anodic Materials for Sodium Io...CrimsonPublishersRDMS
 
Synopsis : Bismuth Sodium Titanate ( A lead free Ferroelectric Material
Synopsis : Bismuth Sodium Titanate ( A lead free Ferroelectric MaterialSynopsis : Bismuth Sodium Titanate ( A lead free Ferroelectric Material
Synopsis : Bismuth Sodium Titanate ( A lead free Ferroelectric MaterialNational Tsing Hua University
 
The Most Complete Interpretation of Anode Materials Standards for Lithium-ion...
The Most Complete Interpretation of Anode Materials Standards for Lithium-ion...The Most Complete Interpretation of Anode Materials Standards for Lithium-ion...
The Most Complete Interpretation of Anode Materials Standards for Lithium-ion...etekware
 
Interpretation of Anode Materials Standards for Lithium-ion Batteries.pdf
Interpretation of Anode Materials Standards for Lithium-ion Batteries.pdfInterpretation of Anode Materials Standards for Lithium-ion Batteries.pdf
Interpretation of Anode Materials Standards for Lithium-ion Batteries.pdfETEK1
 
Dendritic Electroless Deposits of Lead From Lead Acetate Solution
Dendritic Electroless Deposits of Lead From Lead Acetate SolutionDendritic Electroless Deposits of Lead From Lead Acetate Solution
Dendritic Electroless Deposits of Lead From Lead Acetate Solutioninventionjournals
 
Hot hole transfer from Ag nanoparticles to multiferroic YMn2O5 nanowires enab...
Hot hole transfer from Ag nanoparticles to multiferroic YMn2O5 nanowires enab...Hot hole transfer from Ag nanoparticles to multiferroic YMn2O5 nanowires enab...
Hot hole transfer from Ag nanoparticles to multiferroic YMn2O5 nanowires enab...Pawan Kumar
 
Nanomaterial Applications in Lithium Batteries - Christian DiCenso
Nanomaterial Applications in Lithium Batteries - Christian DiCensoNanomaterial Applications in Lithium Batteries - Christian DiCenso
Nanomaterial Applications in Lithium Batteries - Christian DiCensoChristian DiCenso
 
Electrode material for battery in automobile
Electrode material for battery in automobileElectrode material for battery in automobile
Electrode material for battery in automobileayushkamalecell
 
An introduction to electrocoagulation
An introduction to electrocoagulationAn introduction to electrocoagulation
An introduction to electrocoagulationChristos Charisiadis
 
Understanding of thermal stability of lithium ion batteries
Understanding of thermal stability of lithium ion batteriesUnderstanding of thermal stability of lithium ion batteries
Understanding of thermal stability of lithium ion batteriesKhue Luu
 
Lithium-Seawater Battery for Undersea Sensors and Vehicles
Lithium-Seawater Battery for Undersea Sensors and VehiclesLithium-Seawater Battery for Undersea Sensors and Vehicles
Lithium-Seawater Battery for Undersea Sensors and Vehicleschrisrobschu
 

Similar to Pidaparthy 2021 j._electrochem._soc._168_100509 (20)

A. Rock - Final Report - 2015
A. Rock - Final Report - 2015A. Rock - Final Report - 2015
A. Rock - Final Report - 2015
 
Batteries ppt
Batteries pptBatteries ppt
Batteries ppt
 
Lithium Iron Phosphate: Olivine Material for High Power Li-Ion Batteries - Cr...
Lithium Iron Phosphate: Olivine Material for High Power Li-Ion Batteries - Cr...Lithium Iron Phosphate: Olivine Material for High Power Li-Ion Batteries - Cr...
Lithium Iron Phosphate: Olivine Material for High Power Li-Ion Batteries - Cr...
 
Rechargeable Li-ion batteries based on Olivine-structured (LiFePO4) catho...
Rechargeable Li-ion batteries based on Olivine-structured     (LiFePO4) catho...Rechargeable Li-ion batteries based on Olivine-structured     (LiFePO4) catho...
Rechargeable Li-ion batteries based on Olivine-structured (LiFePO4) catho...
 
Research plan 2
Research plan 2Research plan 2
Research plan 2
 
Lithium and Lithium-ion Batteries: Challenges and Prospects
Lithium and Lithium-ion Batteries: Challenges and ProspectsLithium and Lithium-ion Batteries: Challenges and Prospects
Lithium and Lithium-ion Batteries: Challenges and Prospects
 
detailed study on Nanobatterys
detailed study on Nanobatterys detailed study on Nanobatterys
detailed study on Nanobatterys
 
Electrochemical Investigation of Electrolyte & Anodic Materials for Sodium Io...
Electrochemical Investigation of Electrolyte & Anodic Materials for Sodium Io...Electrochemical Investigation of Electrolyte & Anodic Materials for Sodium Io...
Electrochemical Investigation of Electrolyte & Anodic Materials for Sodium Io...
 
Synopsis : Bismuth Sodium Titanate ( A lead free Ferroelectric Material
Synopsis : Bismuth Sodium Titanate ( A lead free Ferroelectric MaterialSynopsis : Bismuth Sodium Titanate ( A lead free Ferroelectric Material
Synopsis : Bismuth Sodium Titanate ( A lead free Ferroelectric Material
 
The Most Complete Interpretation of Anode Materials Standards for Lithium-ion...
The Most Complete Interpretation of Anode Materials Standards for Lithium-ion...The Most Complete Interpretation of Anode Materials Standards for Lithium-ion...
The Most Complete Interpretation of Anode Materials Standards for Lithium-ion...
 
Interpretation of Anode Materials Standards for Lithium-ion Batteries.pdf
Interpretation of Anode Materials Standards for Lithium-ion Batteries.pdfInterpretation of Anode Materials Standards for Lithium-ion Batteries.pdf
Interpretation of Anode Materials Standards for Lithium-ion Batteries.pdf
 
Dendritic Electroless Deposits of Lead From Lead Acetate Solution
Dendritic Electroless Deposits of Lead From Lead Acetate SolutionDendritic Electroless Deposits of Lead From Lead Acetate Solution
Dendritic Electroless Deposits of Lead From Lead Acetate Solution
 
Hot hole transfer from Ag nanoparticles to multiferroic YMn2O5 nanowires enab...
Hot hole transfer from Ag nanoparticles to multiferroic YMn2O5 nanowires enab...Hot hole transfer from Ag nanoparticles to multiferroic YMn2O5 nanowires enab...
Hot hole transfer from Ag nanoparticles to multiferroic YMn2O5 nanowires enab...
 
Nanomaterial Applications in Lithium Batteries - Christian DiCenso
Nanomaterial Applications in Lithium Batteries - Christian DiCensoNanomaterial Applications in Lithium Batteries - Christian DiCenso
Nanomaterial Applications in Lithium Batteries - Christian DiCenso
 
Vishwakarma-JPS2015
Vishwakarma-JPS2015Vishwakarma-JPS2015
Vishwakarma-JPS2015
 
Electrode material for battery in automobile
Electrode material for battery in automobileElectrode material for battery in automobile
Electrode material for battery in automobile
 
Rowena brugge imperial college
Rowena brugge   imperial collegeRowena brugge   imperial college
Rowena brugge imperial college
 
An introduction to electrocoagulation
An introduction to electrocoagulationAn introduction to electrocoagulation
An introduction to electrocoagulation
 
Understanding of thermal stability of lithium ion batteries
Understanding of thermal stability of lithium ion batteriesUnderstanding of thermal stability of lithium ion batteries
Understanding of thermal stability of lithium ion batteries
 
Lithium-Seawater Battery for Undersea Sensors and Vehicles
Lithium-Seawater Battery for Undersea Sensors and VehiclesLithium-Seawater Battery for Undersea Sensors and Vehicles
Lithium-Seawater Battery for Undersea Sensors and Vehicles
 

More from Ary Assuncao

A closed loop ammonium salt system for recovery of high-purity lead tetroxide...
A closed loop ammonium salt system for recovery of high-purity lead tetroxide...A closed loop ammonium salt system for recovery of high-purity lead tetroxide...
A closed loop ammonium salt system for recovery of high-purity lead tetroxide...Ary Assuncao
 
Graphene data sheet update-jan2020 (1)
Graphene data sheet update-jan2020 (1)Graphene data sheet update-jan2020 (1)
Graphene data sheet update-jan2020 (1)Ary Assuncao
 
Graphene data sheet update-jan2020 (2)
Graphene data sheet update-jan2020 (2)Graphene data sheet update-jan2020 (2)
Graphene data sheet update-jan2020 (2)Ary Assuncao
 
2020 application of electrochemical impedance spectroscopy to commercial li...
2020   application of electrochemical impedance spectroscopy to commercial li...2020   application of electrochemical impedance spectroscopy to commercial li...
2020 application of electrochemical impedance spectroscopy to commercial li...Ary Assuncao
 
Graphene data sheet update-jan2020
Graphene data sheet update-jan2020Graphene data sheet update-jan2020
Graphene data sheet update-jan2020Ary Assuncao
 

More from Ary Assuncao (6)

A closed loop ammonium salt system for recovery of high-purity lead tetroxide...
A closed loop ammonium salt system for recovery of high-purity lead tetroxide...A closed loop ammonium salt system for recovery of high-purity lead tetroxide...
A closed loop ammonium salt system for recovery of high-purity lead tetroxide...
 
Graphene data sheet update-jan2020 (1)
Graphene data sheet update-jan2020 (1)Graphene data sheet update-jan2020 (1)
Graphene data sheet update-jan2020 (1)
 
Graphene data sheet update-jan2020 (2)
Graphene data sheet update-jan2020 (2)Graphene data sheet update-jan2020 (2)
Graphene data sheet update-jan2020 (2)
 
2020 application of electrochemical impedance spectroscopy to commercial li...
2020   application of electrochemical impedance spectroscopy to commercial li...2020   application of electrochemical impedance spectroscopy to commercial li...
2020 application of electrochemical impedance spectroscopy to commercial li...
 
Pavlov1981
Pavlov1981Pavlov1981
Pavlov1981
 
Graphene data sheet update-jan2020
Graphene data sheet update-jan2020Graphene data sheet update-jan2020
Graphene data sheet update-jan2020
 

Recently uploaded

Call Girls Varanasi Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Varanasi Just Call 9907093804 Top Class Call Girl Service AvailableCall Girls Varanasi Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Varanasi Just Call 9907093804 Top Class Call Girl Service AvailableDipal Arora
 
Call Girls Kochi Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Kochi Just Call 9907093804 Top Class Call Girl Service AvailableCall Girls Kochi Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Kochi Just Call 9907093804 Top Class Call Girl Service AvailableDipal Arora
 
VIP Call Girls Tirunelveli Aaradhya 8250192130 Independent Escort Service Tir...
VIP Call Girls Tirunelveli Aaradhya 8250192130 Independent Escort Service Tir...VIP Call Girls Tirunelveli Aaradhya 8250192130 Independent Escort Service Tir...
VIP Call Girls Tirunelveli Aaradhya 8250192130 Independent Escort Service Tir...narwatsonia7
 
Call Girls Darjeeling Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Darjeeling Just Call 9907093804 Top Class Call Girl Service AvailableCall Girls Darjeeling Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Darjeeling Just Call 9907093804 Top Class Call Girl Service AvailableDipal Arora
 
Call Girl Coimbatore Prisha☎️ 8250192130 Independent Escort Service Coimbatore
Call Girl Coimbatore Prisha☎️  8250192130 Independent Escort Service CoimbatoreCall Girl Coimbatore Prisha☎️  8250192130 Independent Escort Service Coimbatore
Call Girl Coimbatore Prisha☎️ 8250192130 Independent Escort Service Coimbatorenarwatsonia7
 
Vip Call Girls Anna Salai Chennai 👉 8250192130 ❣️💯 Top Class Girls Available
Vip Call Girls Anna Salai Chennai 👉 8250192130 ❣️💯 Top Class Girls AvailableVip Call Girls Anna Salai Chennai 👉 8250192130 ❣️💯 Top Class Girls Available
Vip Call Girls Anna Salai Chennai 👉 8250192130 ❣️💯 Top Class Girls AvailableNehru place Escorts
 
Bangalore Call Girls Hebbal Kempapura Number 7001035870 Meetin With Bangalor...
Bangalore Call Girls Hebbal Kempapura Number 7001035870  Meetin With Bangalor...Bangalore Call Girls Hebbal Kempapura Number 7001035870  Meetin With Bangalor...
Bangalore Call Girls Hebbal Kempapura Number 7001035870 Meetin With Bangalor...narwatsonia7
 
All Time Service Available Call Girls Marine Drive 📳 9820252231 For 18+ VIP C...
All Time Service Available Call Girls Marine Drive 📳 9820252231 For 18+ VIP C...All Time Service Available Call Girls Marine Drive 📳 9820252231 For 18+ VIP C...
All Time Service Available Call Girls Marine Drive 📳 9820252231 For 18+ VIP C...Arohi Goyal
 
Call Girls Aurangabad Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Aurangabad Just Call 9907093804 Top Class Call Girl Service AvailableCall Girls Aurangabad Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Aurangabad Just Call 9907093804 Top Class Call Girl Service AvailableDipal Arora
 
Call Girls Colaba Mumbai ❤️ 9920874524 👈 Cash on Delivery
Call Girls Colaba Mumbai ❤️ 9920874524 👈 Cash on DeliveryCall Girls Colaba Mumbai ❤️ 9920874524 👈 Cash on Delivery
Call Girls Colaba Mumbai ❤️ 9920874524 👈 Cash on Deliverynehamumbai
 
Call Girls Service Jaipur Grishma WhatsApp ❤8445551418 VIP Call Girls Jaipur
Call Girls Service Jaipur Grishma WhatsApp ❤8445551418 VIP Call Girls JaipurCall Girls Service Jaipur Grishma WhatsApp ❤8445551418 VIP Call Girls Jaipur
Call Girls Service Jaipur Grishma WhatsApp ❤8445551418 VIP Call Girls Jaipurparulsinha
 
Bangalore Call Girl Whatsapp Number 100% Complete Your Sexual Needs
Bangalore Call Girl Whatsapp Number 100% Complete Your Sexual NeedsBangalore Call Girl Whatsapp Number 100% Complete Your Sexual Needs
Bangalore Call Girl Whatsapp Number 100% Complete Your Sexual NeedsGfnyt
 
Call Girls Siliguri Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Siliguri Just Call 9907093804 Top Class Call Girl Service AvailableCall Girls Siliguri Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Siliguri Just Call 9907093804 Top Class Call Girl Service AvailableDipal Arora
 
Call Girls Cuttack Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Cuttack Just Call 9907093804 Top Class Call Girl Service AvailableCall Girls Cuttack Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Cuttack Just Call 9907093804 Top Class Call Girl Service AvailableDipal Arora
 
Kesar Bagh Call Girl Price 9548273370 , Lucknow Call Girls Service
Kesar Bagh Call Girl Price 9548273370 , Lucknow Call Girls ServiceKesar Bagh Call Girl Price 9548273370 , Lucknow Call Girls Service
Kesar Bagh Call Girl Price 9548273370 , Lucknow Call Girls Servicemakika9823
 
Call Girls Horamavu WhatsApp Number 7001035870 Meeting With Bangalore Escorts
Call Girls Horamavu WhatsApp Number 7001035870 Meeting With Bangalore EscortsCall Girls Horamavu WhatsApp Number 7001035870 Meeting With Bangalore Escorts
Call Girls Horamavu WhatsApp Number 7001035870 Meeting With Bangalore Escortsvidya singh
 
Call Girls Service Surat Samaira ❤️🍑 8250192130 👄 Independent Escort Service ...
Call Girls Service Surat Samaira ❤️🍑 8250192130 👄 Independent Escort Service ...Call Girls Service Surat Samaira ❤️🍑 8250192130 👄 Independent Escort Service ...
Call Girls Service Surat Samaira ❤️🍑 8250192130 👄 Independent Escort Service ...CALL GIRLS
 
Lucknow Call girls - 8800925952 - 24x7 service with hotel room
Lucknow Call girls - 8800925952 - 24x7 service with hotel roomLucknow Call girls - 8800925952 - 24x7 service with hotel room
Lucknow Call girls - 8800925952 - 24x7 service with hotel roomdiscovermytutordmt
 
Night 7k to 12k Navi Mumbai Call Girl Photo 👉 BOOK NOW 9833363713 👈 ♀️ night ...
Night 7k to 12k Navi Mumbai Call Girl Photo 👉 BOOK NOW 9833363713 👈 ♀️ night ...Night 7k to 12k Navi Mumbai Call Girl Photo 👉 BOOK NOW 9833363713 👈 ♀️ night ...
Night 7k to 12k Navi Mumbai Call Girl Photo 👉 BOOK NOW 9833363713 👈 ♀️ night ...aartirawatdelhi
 
The Most Attractive Hyderabad Call Girls Kothapet 𖠋 6297143586 𖠋 Will You Mis...
The Most Attractive Hyderabad Call Girls Kothapet 𖠋 6297143586 𖠋 Will You Mis...The Most Attractive Hyderabad Call Girls Kothapet 𖠋 6297143586 𖠋 Will You Mis...
The Most Attractive Hyderabad Call Girls Kothapet 𖠋 6297143586 𖠋 Will You Mis...chandars293
 

Recently uploaded (20)

Call Girls Varanasi Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Varanasi Just Call 9907093804 Top Class Call Girl Service AvailableCall Girls Varanasi Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Varanasi Just Call 9907093804 Top Class Call Girl Service Available
 
Call Girls Kochi Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Kochi Just Call 9907093804 Top Class Call Girl Service AvailableCall Girls Kochi Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Kochi Just Call 9907093804 Top Class Call Girl Service Available
 
VIP Call Girls Tirunelveli Aaradhya 8250192130 Independent Escort Service Tir...
VIP Call Girls Tirunelveli Aaradhya 8250192130 Independent Escort Service Tir...VIP Call Girls Tirunelveli Aaradhya 8250192130 Independent Escort Service Tir...
VIP Call Girls Tirunelveli Aaradhya 8250192130 Independent Escort Service Tir...
 
Call Girls Darjeeling Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Darjeeling Just Call 9907093804 Top Class Call Girl Service AvailableCall Girls Darjeeling Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Darjeeling Just Call 9907093804 Top Class Call Girl Service Available
 
Call Girl Coimbatore Prisha☎️ 8250192130 Independent Escort Service Coimbatore
Call Girl Coimbatore Prisha☎️  8250192130 Independent Escort Service CoimbatoreCall Girl Coimbatore Prisha☎️  8250192130 Independent Escort Service Coimbatore
Call Girl Coimbatore Prisha☎️ 8250192130 Independent Escort Service Coimbatore
 
Vip Call Girls Anna Salai Chennai 👉 8250192130 ❣️💯 Top Class Girls Available
Vip Call Girls Anna Salai Chennai 👉 8250192130 ❣️💯 Top Class Girls AvailableVip Call Girls Anna Salai Chennai 👉 8250192130 ❣️💯 Top Class Girls Available
Vip Call Girls Anna Salai Chennai 👉 8250192130 ❣️💯 Top Class Girls Available
 
Bangalore Call Girls Hebbal Kempapura Number 7001035870 Meetin With Bangalor...
Bangalore Call Girls Hebbal Kempapura Number 7001035870  Meetin With Bangalor...Bangalore Call Girls Hebbal Kempapura Number 7001035870  Meetin With Bangalor...
Bangalore Call Girls Hebbal Kempapura Number 7001035870 Meetin With Bangalor...
 
All Time Service Available Call Girls Marine Drive 📳 9820252231 For 18+ VIP C...
All Time Service Available Call Girls Marine Drive 📳 9820252231 For 18+ VIP C...All Time Service Available Call Girls Marine Drive 📳 9820252231 For 18+ VIP C...
All Time Service Available Call Girls Marine Drive 📳 9820252231 For 18+ VIP C...
 
Call Girls Aurangabad Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Aurangabad Just Call 9907093804 Top Class Call Girl Service AvailableCall Girls Aurangabad Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Aurangabad Just Call 9907093804 Top Class Call Girl Service Available
 
Call Girls Colaba Mumbai ❤️ 9920874524 👈 Cash on Delivery
Call Girls Colaba Mumbai ❤️ 9920874524 👈 Cash on DeliveryCall Girls Colaba Mumbai ❤️ 9920874524 👈 Cash on Delivery
Call Girls Colaba Mumbai ❤️ 9920874524 👈 Cash on Delivery
 
Call Girls Service Jaipur Grishma WhatsApp ❤8445551418 VIP Call Girls Jaipur
Call Girls Service Jaipur Grishma WhatsApp ❤8445551418 VIP Call Girls JaipurCall Girls Service Jaipur Grishma WhatsApp ❤8445551418 VIP Call Girls Jaipur
Call Girls Service Jaipur Grishma WhatsApp ❤8445551418 VIP Call Girls Jaipur
 
Bangalore Call Girl Whatsapp Number 100% Complete Your Sexual Needs
Bangalore Call Girl Whatsapp Number 100% Complete Your Sexual NeedsBangalore Call Girl Whatsapp Number 100% Complete Your Sexual Needs
Bangalore Call Girl Whatsapp Number 100% Complete Your Sexual Needs
 
Call Girls Siliguri Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Siliguri Just Call 9907093804 Top Class Call Girl Service AvailableCall Girls Siliguri Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Siliguri Just Call 9907093804 Top Class Call Girl Service Available
 
Call Girls Cuttack Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Cuttack Just Call 9907093804 Top Class Call Girl Service AvailableCall Girls Cuttack Just Call 9907093804 Top Class Call Girl Service Available
Call Girls Cuttack Just Call 9907093804 Top Class Call Girl Service Available
 
Kesar Bagh Call Girl Price 9548273370 , Lucknow Call Girls Service
Kesar Bagh Call Girl Price 9548273370 , Lucknow Call Girls ServiceKesar Bagh Call Girl Price 9548273370 , Lucknow Call Girls Service
Kesar Bagh Call Girl Price 9548273370 , Lucknow Call Girls Service
 
Call Girls Horamavu WhatsApp Number 7001035870 Meeting With Bangalore Escorts
Call Girls Horamavu WhatsApp Number 7001035870 Meeting With Bangalore EscortsCall Girls Horamavu WhatsApp Number 7001035870 Meeting With Bangalore Escorts
Call Girls Horamavu WhatsApp Number 7001035870 Meeting With Bangalore Escorts
 
Call Girls Service Surat Samaira ❤️🍑 8250192130 👄 Independent Escort Service ...
Call Girls Service Surat Samaira ❤️🍑 8250192130 👄 Independent Escort Service ...Call Girls Service Surat Samaira ❤️🍑 8250192130 👄 Independent Escort Service ...
Call Girls Service Surat Samaira ❤️🍑 8250192130 👄 Independent Escort Service ...
 
Lucknow Call girls - 8800925952 - 24x7 service with hotel room
Lucknow Call girls - 8800925952 - 24x7 service with hotel roomLucknow Call girls - 8800925952 - 24x7 service with hotel room
Lucknow Call girls - 8800925952 - 24x7 service with hotel room
 
Night 7k to 12k Navi Mumbai Call Girl Photo 👉 BOOK NOW 9833363713 👈 ♀️ night ...
Night 7k to 12k Navi Mumbai Call Girl Photo 👉 BOOK NOW 9833363713 👈 ♀️ night ...Night 7k to 12k Navi Mumbai Call Girl Photo 👉 BOOK NOW 9833363713 👈 ♀️ night ...
Night 7k to 12k Navi Mumbai Call Girl Photo 👉 BOOK NOW 9833363713 👈 ♀️ night ...
 
The Most Attractive Hyderabad Call Girls Kothapet 𖠋 6297143586 𖠋 Will You Mis...
The Most Attractive Hyderabad Call Girls Kothapet 𖠋 6297143586 𖠋 Will You Mis...The Most Attractive Hyderabad Call Girls Kothapet 𖠋 6297143586 𖠋 Will You Mis...
The Most Attractive Hyderabad Call Girls Kothapet 𖠋 6297143586 𖠋 Will You Mis...
 

Pidaparthy 2021 j._electrochem._soc._168_100509

  • 1. Journal of The Electrochemical Society OPEN ACCESS Increased Disorder at Graphite Particle Edges Revealed by Multi-length Scale Characterization of Anodes from Fast-Charged Lithium-Ion Cells To cite this article: Saran Pidaparthy et al 2021 J. Electrochem. Soc. 168 100509 View the article online for updates and enhancements. This content was downloaded from IP address 186.211.111.10 on 07/12/2021 at 13:16
  • 2. Increased Disorder at Graphite Particle Edges Revealed by Multi- length Scale Characterization of Anodes from Fast-Charged Lithium-Ion Cells Saran Pidaparthy,1,2 Marco-Tulio F. Rodrigues,1,* Jian-Min Zuo,2,3 and Daniel P. Abraham1,*,z 1 Chemical Sciences and Engineering Division, Argonne National Laboratory, Lemont, Illinois, 60439, United States of America 2 Department of Materials Science and Engineering, University of Illinois at Urbana-Champaign, Urbana, Illinois, 61801, United States of America 3 Materials Research Laboratory, University of Illinois at Urbana-Champaign, Urbana, Illinois, 61801, United States of America Fast charging of lithium-ion cells increases voltage polarization of the electrodes and creates conditions that are favorable for Li- deposition at the graphite anode. Repeated fast charging induces changes in the capacity-voltage profiles and increases the probability of lithium-plating on the electrode. This higher probability results from structural, morphological and chemical modifications that are revealed by multi-length scale characterization of graphite anodes extracted from discharged lithium-ion cells, previously charged at rates up to 6 C. The distinct differences between anodes with lithium-plating and as-prepared electrodes are clearly seen in analytical electron microscopy data. Scanning electrode microscopy (SEM) images show that the fast-charged anode is significantly thicker, apparently because of the electrolyte reduction/hydrolysis products that accumulate in electrode pores. High resolution electron microscopy (HREM) images reveal wavy graphite fringes near the particle edges. Analysis of scanning electron nanodiffraction (SEND) data reveal higher d-spacings and greater lattice rotations, indicating disorder in the graphite near the particle edges that extend about 20 nm into the bulk. The extent of this disorder is greater near larger internal pores, highlighting nanoscale heterogeneities within particles. As graphite lithiation occurs primarily through edge planes, this permanent disorder would hinder Li+ intercalation kinetics and favor Li0 plating during repeated cycling. © 2021 The Author(s). Published on behalf of The Electrochemical Society by IOP Publishing Limited. This is an open access article distributed under the terms of the Creative Commons Attribution 4.0 License (CC BY, http://creativecommons.org/licenses/ by/4.0/), which permits unrestricted reuse of the work in any medium, provided the original work is properly cited. [DOI: 10.1149/ 1945-7111/ac2a7f] Manuscript submitted August 25, 2021; revised manuscript received September 17, 2021. Published October 8, 2021. Supplementary material for this article is available online Advances in lithium-ion battery (LIB) technology have fueled the rise of the electric vehicle (EV) industry.1 Most high-energy LIBs contain a positive electrode (cathode) comprised of a layered oxide with Ni, Co and Mn, a negative electrode (anode) comprised of graphite, and an electrolyte with LiPF6 salt in a mixture of organic solvents. These cell components are being continually optimized so that LIB technology can deliver high energy densities safely and reliably during a battery’s 10–15-year lifetime. State-of-the-art LIBs can be fully charged in about 1 h with minimal degradation in cell performance. However, full charging over shorter durations causes irreversible damage to the battery via losses in the inventory of mobile Li+ ions (i.e., capacity fade) and resistance increases (i.e., power fade) in the cells.2,3 The U.S. Department of Energy (DOE) is funding research to identify the underlying mechanisms that limit fast-charge performance of LIB cells and to develop technological solutions that would enable full battery charge in less than 10 min.4–7 Solutions to these limitations will allay range anxiety concerns and accelerate customer acceptance of EVs. During cell charging, lithium ions are extracted from the transition metal oxide cathode and intercalated between atomic planes of the graphite particles. Under fast charge conditions, there is salt enrichment in the oxide cathode and salt depletion in the graphite anode, leading to non-uniform distribution of Li+ ions across the electrode thickness.8,9 These conditions lower Li+ ion conductivity in the electrolyte within the electrode pores and increase impedance, causing polarization in both electrodes. Cathode polarization causes the cell to reach its upper cut-off voltage (UCV) before Li+ ions are fully extracted from the oxide, while anode polarization causes the cell to dip below the Li-plating potential creating conditions that are favorable for lithium nucleation and growth on the graphite electrode.10–12 The situation is further complicated by non-uniformities in local reactivity resulting from factors that include uneven mechanical compression of the separator that affects its porosity/tortuosity, and irregular distribution of active particles that affects distribution of local currents within the electrodes.13–16 These inhomogeneities cause large lithium concen- tration gradients in the electrodes, both across the surface and along its thickness, promoting non-uniform aging of the cell materials. The high charge currents can also enhance structural stresses in the electrode matrix causing the coating to peel off the current collector and within the active material particles causing their fracture and disconnection from the electron conduction network.17 In our recent series of articles we have used operando X-ray techniques to quantify phase evolution in the graphite and lithium concentration gradients that develop in the electrode during fast charging.18,19 In addition, we have described techniques to identify lithium deposition on the graphite electrode and detailed electro- chemical charging protocols that can mitigate this plating.20–22 We observed that the likelihood of Li-plating on the graphite anode increases during repeated fast-charge cycling.10 We note here that commercial graphite typically contains a mixture of the hexagonal (2H) and rhombohedral (3 R) phases: the stacking sequence is ABAB in the 2H structure and ABCABC in the 3 R structure.23 The 2H phase is more thermodynamically stable than the 3 R phase at room temperature and pressure. The Li intercalation mechanism is reported to be similar for both phases: during Li+ insertion, the adjacent graphene sheets glide relative to each other to lower the overall energy of the system and attain AAAA layer stacking.24 Any material feature that hinders the movement of graphene sheets will prevent Li+ intercalation. During fast charging, Li+ intercalation into the graphite competes with Li0 nucleation on the particles. Greater layer-to-layer stacking disorder in the graphite in the form of added strain, local misfits, and misorientation angles in the stack, would make Li+ intercalation more difficult and expedite Li0 z E-mail: abraham@anl.gov *Electrochemical Society Member. Journal of The Electrochemical Society, 2021 168 100509
  • 3. plating. Hence, an in-depth understanding of changes in the graphite particles subjected to fast charging is crucial for the development of materials that enable intercalation and mitigate deterioration of cell performance during the electrochemical cycling. In this work, we examine the morphological, chemical, and structural changes to the graphite materials and electrodes, using a suite of electrochemical and physicochemical techniques. The characterization was performed across multiple length-scales on as-prepared electrodes and on numerous electrodes harvested from cycled cells; only representative data from the various techniques are shown to highlight key changes that result from the cycling. The data in this manuscript are from an as-prepared graphite (Gr-AP) electrode that was never exposed to the electrolyte, graphite electrode with no plating (Gr-NP) harvested from a cell after a few low-rate cycles and graphite electrode with lithium plating (Gr- LP) harvested from a fast-charged cell. For purposes of clarity and contrast, we show only the Gr-AP and Gr-LP data in the main text; data from the Gr-NP sample are included in the Supporting Information (SI). Note that all Tables and Figures in the SI are marked with the designator S, as in Fig. S1 (available online at stacks.iop.org/JES/168/100509/mmedia). We start by describing electrochemical tests, conducted on full cells with LiNi0.5Co0.2Mn0.3O2 (NCM523) cathodes and graphite (Gr) anodes, which include charging rates up to 6 C (here, 1 C is the current needed to fully discharge the cell in 1 h). Following this, we compare electrochemical behavior of the Gr-AP electrode with that of the Gr-LP electrode. Then, after presenting high-resolution X-ray diffraction (XRD) data from the materials, we show microscopy and spectroscopy results obtained on electrode cross-sections fabricated via ion milling and on sections prepared with a focused ion beam (FIB). At the bulk-scale, we use scanning electron microscopy (SEM) and energy dispersive X-ray spectroscopy (EDS) to probe the morphological and chemical changes in cross-sections of the graphite anode. At the nano- to atomic-scale, we rely on both conventional transmission electron microscopy (TEM) and scanning TEM (STEM) to reveal local structural disorder and chemical variation within individual graphite particle sections through a combination of high-resolution imaging, EDS, and scanning electron nanodiffraction (SEND). Our investigations reveal that anodes from the fast-charged cells have lower capacity, present a sloping voltage profile, display an increase in thickness and show greater lattice disorder in localized areas near the graphite particle edges. Experimental Section Electrode materials, cell assembly and electrochemical testing.—In the NCM523//Gr cells, the positive electrode has a 71 μm thick coating, with 90 wt% Li(Ni0.5Co0.2Mn0.3)O2 (Toda), 5 wt% C45 carbon (Timcal) and 5 wt% PVDF binder (KF-9300, Kureha), on a 20 μm thick Al current collector.10 The negative electrode has a 70 μm thick coating, with ∼92 wt% SLC1506T graphite (Superior Graphite), 2 wt% C45 carbon and 6 wt% PVDF binder, on a 10 μm thick Cu current collector. The electrolyte (aka Gen 2, purchased from Tomiyama) is a solution of 1.2 M lithium hexafluorophosphate (LiPF6) salt dissolved in a 3:7 w/w solvent mixture of ethylene carbonate (EC) and ethyl methyl carbonate (EMC). Celgard 2320 is the separator between the electrodes. Before cell assembly, the electrodes and separator are dried in vacuum ovens set at 120 °C and 70 °C, respectively. The 3-electrode measurements are conducted in custom-made stainless-steel cells, which are assembled and tested inside an Ar-atmosphere glove box. The reference electrode (RE) probe is placed between two separa- tors, sandwiched between 20.3 cm2 disks of the oxide-positive and graphite-negative electrodes (Fig. 1): the RE consists of a copper wire (25 μm diameter) with a thin layer of Li metal plated on the exposed tip.9 Electrochemical cycling is typically conducted at 30 °C with a Maccor Model 2400 cycler. The Gr-NP electrode is from a cell that experienced two C/10 and five C/2 cycles between 3 and 4.2 V at 30 °C. The Gr-LP electrode was harvested from a cell that experienced multiple sequences of an “escalating C-rate” protocol (Fig. 1). To accelerate cell aging, three of the sequences were at 30 °C, one at 40 °C and one at 50 °C. Including the initial cycles and periodic reference performance tests (C/25 rate), the cell experienced a total of 50 cycles; of these, 25 charge segments were at rates of 2 C and above. After cycling, cells were discharged to (and held for 10 h at) 3 V before disassembly in an argon-atmosphere glove box. Various diagnostic tests were conducted on the harvested cell materials to determine the effect of cycling on the graphite material. For the electrochemical tests, 2032-type coin cells were assembled with 1.6 cm2 area harvested-electrode samples, lithium metal counter elec- trode, fresh Celgard 2320 separator, and fresh Gen2 electrolyte. Cross-section polishing.—Cross-sections of the graphite elec- trode samples were prepared via argon ion milling using the Gatan Model 685.O Precision Etching Coating System II (PECS II). The electrode was gently handled with metal tweezers and cut with sharp metal scissors to leave at least one flat-side of ∼5 mm; all metal tools were cleaned using an ethanol and acetone rinse prior to handling the electrode. The cut electrode was mounted onto the PECS II sample blade (Gatan, Product No. 693.08500) using fast- drying silver paint (Ted Pella, Product No. 16040–30) and adjusted with a micrometer to ensure ∼200 μm overhang of the flat-side of the electrode from the top of the sample blade. Once the electrode was secure on the blade, the electrode/blade polishing sample was placed onto the cross-section specimen mount (Gatan, Product No. Figure 1. In this escalating C-rate experiment, the NCM523//Gr cell is charged with increasing currents to a fixed capacity of 87 mAh g−1 . Each charge is followed by 1 h rest and C/5 discharge to 3 V and a hold. The cell voltage changes during the cycles are shown in (a), while the graphite- electrode potential changes are shown in (b). Lithium plating on the graphite particles becomes likely when the electrode potential dips below 0 V vs Li/Li+ ; this Li-plating condition (LPC) is indicated by the dotted line in (b). In the above experiment, Li-plating on the graphite electrode becomes likely when the charge rate exceeds ∼2 C. Journal of The Electrochemical Society, 2021 168 100509
  • 4. 685.08540) and loaded into the PECS II instrument. After the sample was loaded, the instrument was set to “Etch by Time” with a “Single Modulation” etching condition, a Gun Tilt angle of 0° for both ion guns, and a rotation rate of 0.5 RPM. The ion milling was performed in two stages: first, a high-voltage milling at 8 kV (with current readings in the range 20–40 μA) that was run for 4 h to create a rough cross-section; second, a low-voltage milling at 4 kV (with current readings of 7–15 μA) that ran for 30 min to apply a low-kV fine polishing. Synchrotron powder diffraction.—Powder diffraction data were collected at beamline 11-BM of the Advanced Photon Source, Argonne National Laboratory. Coatings from the as-prepared electrodes and cycled electrodes, scraped of the copper current collectors, were loaded into sealed capillaries for examination. An X-ray wavelength of 0.41283 Å was used to collect the diffraction data, over an angular range from –6.0° to 28.0° with a step size of 0.001° and a time per step of 0.1 s. Scanning electron microscopy and energy-dispersive X-ray spectroscopy.—Both imaging and elemental examinations were performed using a JEOL 7000 F analytical SEM. The electrode samples were prepared in the following ways: (1) for top-down imaging, a small portion of the electrode is mounted onto a flat sample holder via carbon tape; (2) for cross-sectional imaging, the samples prepared by the PECS II ion milling are loaded onto a standard cross-section specimen mount designed for the SEM. For the SEM imaging, the instrument was set to 7 kV and operated under secondary electron imaging mode with a medium probe current of ∼0.3 nA. For EDS acquisition, the sample was aligned to an optimal working distance of 10 mm and the probe current was adjusted to a setting of ∼15 nA. The mapping data were acquired using the ThermoFisher NSS 2.0 software using its “Spectral Imaging” mode. The datasets were summed for 150 frames, each acquired for 10 s per frame with a 50 μs dwell time. The mapping data are reported as counts per element and the quantification was performed by the software automatically using a built-in database and algorithm based on the standard phi-rho-z method. The line profile data for the cross- section samples are acquired using the “X-ray Linescans” mode with a single scan with 50 points from the bottom to the top of the anode with a dwell time of 10 s per point. Line profile data are further binned for every 5 μm step. Focused-ion beam milling.—Samples for nano- to atomic-scale electron imaging and diffraction were prepared with a ThermoFisher Scientific Scios 2 FIB-SEM. The electrodes were cut with sharp metal scissors and mounted onto a flat sample holder via carbon tape and loaded into the instrument. The cross-section samples were selected from areas near the electrode surface and prepared via standard focused-ion beam lift-out procedure provided by ThermoFisher Scientific onboard user guidance (included with the instrument software). The samples were mounted onto one of four posts of a PELCO FIB lift-out TEM copper grid and thinned to a final thickness around 100–200 nm. Transmission electron microscopy.—Transmission electron mi- croscopy (TEM) was performed on a Hitachi 9500 TEM equipped with a LaB6 emitter and operated at 300 kV. Image acquisition was done using the K2 direct electron detector operated in counting mode. Scanning transmission electron microscopy, energy dispersive X-ray spectroscopy, and scanning electron nanodiffraction.— Nano- and atomic-scale imaging, chemical analysis, diffraction data were acquired using the ThermoFisher Scientific Titan Themis Z STEM equipped with a Schottky field emission gun operated at 300 kV accelerating voltage. The electrode sample was loaded into a standard double-tilt holder and, during operation, tilted close to the 100 〈 〉 zone axis or the 210 〈 〉 zone axis (with respect to the Graphite-2H phase), as both orientations allowed us to track the graphite interplanar spacing in both imaging and diffraction modes. Imaging was performed at 15–30 pA screen current with a convergence angle of 18.0 mrad and acquired using a high-angle annular dark field (HAADF) detector. The EDS acquisition was performed using the above mentioned imaging conditions via the Velox software (ThermoFisher Scientific) with an acquisition time of ∼1 h using a continuous scan with a frame time of ∼1 s. SEND data were collected in microprobe STEM (μP STEM) mode with a convergence angle of 0.46 mrad. The SEND probe size is 1.7–- 1.8 nm at full-width half maximum. The diffraction datasets were acquired at a camera length of 145 mm over a real-space scan area of Figure 2. (a) Capacity-potential profiles of the graphite electrode from a NCM532//Gr cell with a reference electrode. The lithiation profiles are at the indicated rates; the 0.2 C rate delithiation profile shown is after lithiation at the 2 C rate followed by a 1 h rest. (b) Profiles for as-prepared (black) and harvested electrodes (red); the latter are from a cell subjected to multiple “escalating C-rate” cycling. The as-prepared electrode data are at C/100 rate, whereas the harvested electrode plots are at both C/100 (solid) and C/10 (dashed) rates. In both (a) and (b), the specific capacities are obtained by dividing the cell capacity (mAh) by the graphite weight (gGr −1 ) in the anode. Journal of The Electrochemical Society, 2021 168 100509
  • 5. 150 nm × 150 nm using 75 × 75 total steps with drift correction performed after each line acquisition. Results and Discussion Electrochemical cycling.—Capacity-potential profiles of the graphite electrode from a NCM523//Gr cell are shown in Fig. 2a. The graphite is lithiated and delithiated during cell charge and discharge, respectively. At low lithiation rates (<C/5), the inter- calation progresses in distinctive stages seen as plateaus in the plots; the Li+ ions fill spaces between the carbon sheets forming distinct LixC6 phases (x = 0.22, 0.33, 0.5, 1). At higher lithiation rates (⩾1 C), the plateaus are not distinct as intercalation occurs under far- from-equilibrium conditions. Multiple mixed and strained phases are probably formed so that the profile appears continuous, rather than staged; note that LixC6 phases of various compositions have been reported in the research literature.25–29 During the C/5 delithiation, staging plateaus are always observed, regardless of the behavior during the preceding lithiation cycle (Fig. 2a). We observed this difference between the lithiation and delithiation behavior during operando XRD studies; therein, we attributed the blurring of phase boundaries during lithiation to lattice strain and the formation of distinct LixC6 phases during delithiation to abrupt separation from oversaturated solid solutions.18 As indicated earlier, the cells in this study were subjected to multiple applications of the escalating C-rate protocol. Capacity- voltage profiles from a representative cell, before and after the protocol, are shown in Fig. S1. The electrochemical cycling causes a significant loss in cell capacity. The Li+ ions are lost to the solid electrolyte interphase (SEI) formed during electrolyte reduction on the graphite electrode and to the dead lithium that accumulates in the electrode because of repeated excursions below the Li-plating potential (Fig. 1b).22,30 Figure 2b compares plots from a lithium-plated graphite elec- trode (Gr-LP) harvested from the cell in Fig. S1, with plots from an as-prepared graphite (Gr-AP) electrode. The lithiation profile of the Gr-AP electrode (at C/100 rate) displays distinctive plateaus around 210, 117 and 83 mV vs Li/Li+ . These plateaus are also very distinct in the C/10 rate lithiation data (not shown), though they occur at slightly lower potentials because of polarization arising from electrode impedance. In contrast, the plateaus are not distinct in the lithiation profile of the Gr-LP electrode, even at a C/100 rate. The capacity of this aged-harvested electrode (326 mAh g−1 ) is also smaller than that of the as-prepared electrode (357 mAh g−1 ). The C/ 10 lithiation data of this Gr-LP electrode resembles the 2 C graphite- lithiation data in Fig. 2a; the sloped profile, the barely visible staging plateaus and the lower capacity all suggest changes in the graphite particles that alter lithium insertion into the material. Interestingly, the staging plateaus are visible both in the C/100 and C/10 delithiation profiles of the harvested material, which is consistent with the distinct differences between the lithium insertion and extraction processes mentioned earlier. The theoretical capacity of graphite is 372 mAh g−1 but the Gr- AP electrode yields only 357 mAh g−1 even at very low rates. The underlying reason for this smaller capacity is the inherent turbos- tratic stacking disorder in the commercial graphite, which limits Li+ ion intercalation into the material.31,32 Fast-charging appears to create additional disorder, further decreasing capacity. Moreover, the increased structure-disorder increases the number of electronically and geometrically non-equivalent sites for Li-intercalation causing the sloping lithiation profile of the Gr-LP even at extremely low rates.33 In contrast, most of the sites in Gr-AP are thermodynami- cally equivalent for Li+ intercalation, thereby producing the observed staging behavior at low lithiation rates. Ensemble phase analysis via powder X-ray diffraction.—We evaluated the bulk-scale structure and ensemble lattice spacing of the graphite anode coatings using synchrotron X-ray diffraction. Diffraction patterns obtained on the Gr-AP and Gr-LP electrodes are shown in Figs 3a and 3b, respectively. Figure 3a shows that the as-prepared electrode contains two distinct graphite phases: the 2H phase (marked as #) and the 3 R phase (marked as *). The 3 R phase is not observed in the Gr-LP electrode pattern (Fig. 3b); we postulate that this phase transforms into the more stable 2H form during repeated Li+ insertion into and extraction from the graphite, which requires gliding of the graphene planes as mentioned in the Introduction. We calculated the graphite interplanar d-spacing from the strong diffraction peak at ∼1.87 Å–1 , which corresponds to the (002) reflection for the Graphite-2H phase or the (003) reflection for the Graphite-3R phase. For both Gr-AP and Gr-LP electrode samples we obtained d 3.356 Å, interplanar 〈 〉 ≅ which indicates that the latter sample is fully delithiated and that the amount of residual Li+ ions, within the graphite of the fully discharged cell, is negligible. However, diffraction pattern of the Gr-LP sample (Fig. 3b) shows new peaks at ∼2.70 Å–1 and ∼4.40 Å–1 , which arise from lithium fluoride (LiF) (marked as ♢). The presence of LiF was confirmed by selected area electron diffraction (SAED) data on a FIB cross- section prepared from the top surface of the Gr-LP sample (Fig. S2). LiF is known to be a component of the graphite SEI and forms through decomposition reactions of the LiPF6 salt.34,35 There was no evidence of other known SEI phases in the Gr-LP pattern, which indicates that these compounds are either fully amorphous or lack the long-range order needed to coherently scatter the X-ray beam. SEM imaging and compositional analysis.—Figure 4 reveals the dramatic change to the graphite anode surface after being subjected to cycling under the fast-charge, lithium-plating condition. Figure 4a shows a photograph of the Gr-AP sample (cut to a 14.2 mm diameter); the EDS data in Figs. 4b and S3 indicate that the coating contains only C and F, reflecting the graphite, C45 and PVDF Figure 3. Synchrotron powder diffraction data from the following graphite electrodes: (a) as-prepared (Gr-AP) and (b) cycled with lithium plating (Gr- LP). The Bragg reflections are indexed according to JCPDS cards 01–071- –3739 for Graphite-2H, 01–075–2078 for Graphite-3R, and 01–071–3743 for LiF. The phases are marked using the following symbols: # for Graphite-2H, * for Graphite-3R, and ♢ for LiF. Journal of The Electrochemical Society, 2021 168 100509
  • 6. materials used for electrode preparation. Figure 4c is an SEM image of the Gr-AP sample surface at 2k × magnification to illustrate that the electrode is comprised of densely packed, potato-shaped graphite particles up to 10 μm in size. SEM and EDS data from the Gr-NP electrode are displayed in Fig. S4. The graphite particles appear unchanged by the low-rate cycles; the elemental analyses reveal the presence of C, O, F and P, which arise from electrolyte decomposition products and residues on the electrode. By contrast, Fig. 4d shows a photograph of the Gr-LP sample (cut to 14.2 mm diameter) with an uneven white-coloration across the surface; the average elemental composition (in wt%) of the electrode surface (Fig. 4e) is 21% O, 49% F, 27% P and 3% C. Figure 4f shows the surface morphology of the Gr-LP sample at 2k × magnification; in this image, there is complete coverage of the graphite particles by products formed by reactions of the plated lithium with the electrolyte (also see Fig. S5). To further determine extent of changes, we examined the electrodes in cross-section to study morphological and elemental characteristics throughout their depth (see Figs. 5, S6, S7 and S8). Details of the coating thickness calculations are provided in Section 4 of the SI and Fig. S9. Figure 5a shows that the Gr-AP sample has an average coating thickness of 76.6 μm. Figure 5b shows a FIB cross-section of a typical graphite particle, revealing the presence of internal pores. The elemental composition across the electrode thickness was measured via EDS line-scans. For the Gr-AP sample, the carbon signal, which arises predominantly from the graphite, measures an average of 97 wt% across the depth of the coating layer above the copper current collector, while the fluorine signal, which arises from PVDF binder, measures an average of 3 wt% (Figs. 5c S6). Cross-sections of the Gr-NP electrode (Fig. S7) reveal an average coating thickness of 78.1 μm, which is about 2% greater than that of the Gr-AP electrode. Irreversible increases in graphite electrode thickness have been noted previously in electrochemical dilatometry studies; these typically range from 2%–5% after a few cycles.36–38 EDS scans reveal carbon, oxygen, fluorine and phos- phorous signals, distributed across the electrode. These elements are from the electrolyte decomposition products that accumulate in the electrode. By contrast, the Gr-LP electrode has an average coating thickness of 91.3 μm, which does not include the ∼15–20 μm thick layer resulting from lithium plating on the anode surface (see Fig. 5d). This coating thickness is 19.2% and 16.9% greater than that of the Gr-AP and Gr-NP electrodes, respectively. We attribute this electrode thickness increase to SEI buildup in the electrode pores and rearrangement of graphite particles during the electrochemical cycling. The process is exacerbated by Li plating and subsequent electrolyte reduction, which increases the amount of SEI. This SEI buildup is also seen on individual graphite particles, on the exterior surfaces and in the internal pores (Fig. 5e). Moreover, EDS line- scans show that the elemental composition is non-uniform across the electrode thickness (Figs. 5f and S8). The topmost portion, which has no graphite, contains mainly oxygen, fluorine and phosphorus indicating that the products of lithium plating are mainly inorganic— for example, LiF and lithium (oxy)fluoro-phosphate compounds. Below this portion, in the electrode coating, carbon is the dominant element but we also observe an abundance of fluorine and phosphorus, possibly (oxy)fluoro-phosphate compounds resulting from hydrolysis reactions of the electrolyte salt.39,40 STEM-EDS compositional analysis.—Using STEM-EDS, we further show that the (oxy)fluoro-phosphate compounds only coat the surface of the graphite particle and do not penetrate the bulk. Figure 4. Comparison of the anode surface morphology and composition. The as-prepared graphite (Gr-AP) anode surface details are elucidated via: (a) photograph at 1x magnification, (b) EDS elemental composition, in weight percent, and (c) SEM image at 2000 × magnification. Similarly, the Gr-LP anode surface details are elucidated via: (d) photograph at 1x magnification, (e) EDS elemental composition in weight percent, and (f) SEM image at 2000 × magnification. Scale bars in (a) and (d): 5 mm. Journal of The Electrochemical Society, 2021 168 100509
  • 7. Figure 6a shows a survey of the Gr-LP particle section (prepared by FIB) with EDS elemental analysis performed in the marked region, which is magnified in Fig. 6b. We reveal compositional heterogeneity by comparing the local EDS spectra from graphite edges (marked as Area #1 and Area #3) to the graphite center (marked as Area #2) shown in Fig. 6c(i). The spectra from Area #1 and Area #3 show signals for O, F, and P in addition to the signal from C. However, Area #2 only shows C signature, which indicates a lack of (oxy)fluoro-phosphate com- pounds in the graphite interior. Figure 6c(ii) shows the spectra corresponding to the whole area in Fig. 6b. The elemental maps (Fig. 6d) further confirm the spatial heterogeneity between carbon and the (oxy)fluoro-phosphates; the C K signal appears only in the graphite bulk, while the O, F, and P signals appear on the graphite band edges. Moreover, the O, F, and P signals appear in the same Figure 5. SEM imaging and elemental mapping of the electrode cross-sections. The as-prepared graphite (Gr-AP) sample details are elucidated via (a) SEM image of the electrode cross-section, (b) SEM image of an individual particle section, and (c) EDS elemental data acquired across the thickness from bottom of the current collector to the top of the coating surface—see white arrow in (a). Similarly, the lithium-plated graphite electrode (Gr-LP) sample details are elucidated via (d) SEM image of the electrode cross-section, (e) SEM image of an individual particle section, and (f) EDS elemental data acquired across the thickness depth from bottom of the current collector to the top surface—see white arrow in (d). Scale bars: (a), (d) are 25 μm; (b), (e) are 2 μm. Journal of The Electrochemical Society, 2021 168 100509
  • 8. locations across the EDS maps. Additional EDS maps for different regions of the Gr-LP electrodes are provided in Figs. S10 and S11. The above observations indicate that the SEI only accumulates on the graphite particle edges and does not penetrate the bulk material. Electrolyte solvent co-intercalation would cause excessive strain and/or damage to the host graphite. We do not see such damage in the bulk graphite and, therefore, do not anticipate extensive electrolyte penetration. We also observe that smaller pores within the graphite yield weaker electrolyte residue signatures compared to the larger pores. The electrolyte is known to penetrate the intra- particle pores during electrochemical cycling.41 Larger pores pro- vide easier access to the electrolyte; hence, more SEI would be expected to accumulate in the larger pores than in the smaller pores. Revealing localized disorder via high-resolution TEM.—High- resolution TEM (HRTEM) data reveal the nature of the turbostratic disorder that manifests in graphite particles at the lattice-level (Fig. 7). Figure 7a, spanning a 50 nm × 50 nm field-of-view, shows that the Gr-AP sample has straight and uniform graphite lattice fringes corresponding to the constituent graphene planes. From this image, we can qualitatively assess the degree of uniformity by studying Region 1 (Fig. 7b) and Region 2 (Fig. 7c). The lattice- fringe spacing, measured directly from the images of both regions, is ∼0.33 nm, which is consistent with the interplanar d-spacing of graphite determined by XRD. Fast-Fourier Transform (FFT) analysis is useful for quantifying geometric and periodic patterns in electron microscopy images.42,43 The localized Fast-Fourier Transform (FFT) of the Region 1 and Region 2 images are shown in Figs. 7d and 7e, respectively; information to explain features observed in the HRTEM images is contained in Section 6 and Fig. S12 of the SI. Briefly, the FFT of a real-space intensity image (with distances x in nm) yields a representative image in the corresponding spatial frequency domain (with distances x 1/ in nm–1 ).42,43 By applying FFT to the image in Fig. 7b, the periodic stripe pattern of the graphite lattice (with measured spacing 0.33 nm ∼ ) is mapped to points in Fig. 7d corresponding to the frequency of the lattice spacing (i.e., 1 0.33 nm 1 ∼ / − from the center). Moreover, because the intensity variation of the fringes lies along the direction normal to the fringes in Fig. 7b (specified as direction 1 on the shown axis), the corresponding spots in the frequency domain of Fig. 7d lie along this same direction. The higher frequency spots in Fig. 7d that lie along the same line (direction 1) correspond to additional harmonics (i.e., n 1 0.33 nm 1 × ∼ / − where n is an integer) of the real-space lattice; observation of such harmonics implies that the lattice is relatively straight and uniform (i.e., ordered) within Region 1. In the specified direction 2 (normal to direction 1), spots are occasionally observed within the lattice fringes (Fig. 7b); these spots likely arise from atomic columns within the graphite and the distance between them correspond to the spacing between such columns. Similar observations can be made for the Region 2 image (Fig. 7c) and its corresponding FFT (Fig. 7e), thereby illustrating spatial uniformity of the Gr-AP lattice. Figure 6. STEM-EDS of a graphite particle cross-section from the Gr-LP sample. HAADF-STEM images of (a) graphite particle section and (b) a close-up of the area where the EDS maps were acquired. (c) EDS spectra represented by the following regions: (i) from selected areas 1 (top-edge of graphite edge), 2 (center of graphite band), and 3 (bottom-edge of graphite band) as marked in (b); (ii) whole-area EDS spectrum. (d) EDS maps of (i) C, (ii) O, (iii) F, and (iv) P signals. Scale bars: (a) 2 μm, (b) 100 nm, (d) (i)–(iv) 50 nm. Journal of The Electrochemical Society, 2021 168 100509
  • 9. By contrast, HRTEM images of the Gr-LP sample (Fig. 7f), show significant nonuniformity. This nonuniformity is illustrated by example from the magnified Region 1 (Fig. 7g) and Region 2 (Fig. 7h); localized FFTs of Regions 1 and 2 are shown in Figs. 7i and 7j, respectively. Region 1, which is taken near the surface of an internal pore within a graphite particle, is different from both Regions 1 and 2 of the Gr-AP sample. From Fig. 7g, we find that the Gr-LP lattice of Region 1 appears wavy; based on Fig. 7f, we estimate that this waviness extends about 20 nm into the graphite bulk. The FFT of Region 1 (Fig. 7i) confirms lattice disorder in the following ways: (i) arcs are found instead of spots along direction 1; (ii) the high frequency harmonic spot (marked by the white arrow) is relatively faint compared to those in the Gr-AP sample. Furthermore, the interplanar d-spacing from Fig. 7g measures up to ∼0.37 nm, indicating that the lattice in this region does not return to its original delithiated state. This nanoscopic increase in interplanar spacing is not captured in the powder diffraction data (Fig. 3) because the latter technique provides an average from all portions of the sample. On the other hand, Region 2, which is in the bulk of the Gr-LP sample, shows straight, uniform lattice fringes (Fig. 7h) similar to those observed in the Gr-AP sample (Figs. 7b, 7c). The corresponding FFT of Region 2 (Fig. 7j) has spots along direction 1 and a high- frequency harmonic, which suggests greater ordering within the graphite bulk. Taken together, the HRTEM images and FFT results indicate there is greater disorder in the graphite lattice near the Figure 7. Direct electron imaging of the graphite lattice via HRTEM. The as-prepared graphite (Gr-AP) electrode lattice details are elucidated via (a) TEM imaging with highlighted Regions 1 and 2; real-space magnification of (b) Region 1 and (c) Region 2; localized Fast-Fourier Transforms of (d) Region 1 and (e) Region 2. Correspondingly, the Gr-LP lattice details are elucidated via (f) TEM imaging with highlighted Regions 1 and 2; real-space magnification of (g) Region 1 and (h) Region 2; and localized Fast-Fourier Transforms of (i) Region 1 and (j) Region 2. Scale bars: (a), (f) are 10 nm; (b), (c), (g), (h) are 2 nm; (d), (e), (i), (j) are 5 nm–1 . Journal of The Electrochemical Society, 2021 168 100509
  • 10. surface of its internal pores and that this disorder can extend up to ∼20 nm into the particle bulk. Quantification of nanoscopic turbostratic disorder via SEND.— In Fig. 8 and 9, we quantify the nanoscale heterogeneity of the graphite by measuring the interplanar d-spacing and local lattice rotation using SEND.44 In this technique, a ∼2 nm probe (at full- width half-maximum) is rastered across a 150 nm × 150 nm real- space area generating an electron diffraction pattern for each probe position. SEND provides position-resolved structural information regarding turbostratic disorder in the materials. Details of the SEND data acquisition are briefly described in the experimental section; data analysis information is provided in Section 7 of the SI, which includes Fig. S13. In the SEND experiments, we quantify both the graphite interplanar spacing and the lattice rotation by tracking changes to the Ginterplanar vector (with respect to the Graphite-2H phase) for each real-space sample coordinate of the SEND scan. The quantification is done by determining (1) changes to the length of the vector G d 1 interplanar interplanar ∣ ∣ = to measure position-to-position nanoscopic interplanar d-spacing, which is then represented as a dinterplanar map to highlight local d-spacing changes; (2) changes to the in-plane orientation θ of each vector G , interplanar which is then represented as a map of θ ∣∇ ∣ to identify locations where rotational changes in the lattice occur. Figure 8a(i) contains HAADF-STEM image from a Gr-AP particle cross-section, with marked Region 1 (magnified in Fig. 8a(ii)) and Region 2 (magnified in Fig. 8a(iii)) where SEND data were acquired. First, we discuss the nanoscale structure of Region 1 in Fig. 8b. The data indicate that the lattice rotation shows negligible variation ( 0.09 nm 1 ∼ ° − ) across the sample (Fig. 8b(i)) and that the interplanar d-spacing map across Region 1 is relatively uniform (Fig. 8b(ii)). Figure 8b(iii) is a plot of the overall d-spacing from the map as a histogram; the data show a distribution of d- spacings in the sample with a mean of 0.334 nm and a standard deviation of 0.006 nm. Figure 8c illustrates that Region 2 is comparable to Region 1 in terms of both lattice rotation (Fig. 8c(i)) and interplanar spacings (Figs. 8c(ii) and (iii)). However, we find a slightly elevated value for rotational variation in the particle bulk of Fig. 8c(i). This observation points to an important characteristic of graphite, which is that the graphene planes tend to group as bands of 10–20 nm thickness (Fig. S14). Within each graphite band, the magnitude of the rotational variation is negligible; between band-to-band, the in-plane lattice angle can exhibit sudden jumps. Overall, though, the results for the Gr-AP sample are consistent with the HRTEM observations, which also indicated a uniform lattice. Furthermore, SEND data from the Gr-NP electrode (Fig. S15) are very similar to those from the Gr-AP electrode, which indicate that the low-rate cycles have negligible impact on the graphite particle structure. By contrast, a non-uniform lattice and an increase in lattice disorder at the particle edges are observed in particles from the Gr- LP electrode. Figure 9a(i) contains a HAADF-STEM image with areas marked Region 1 (magnified in Fig. 9a(ii)) and Region 2 (magnified in Fig. 9a(iii)) where SEND data were collected. In Fig. 9b(i), we show position-to-position changes of the in-plane rotation angle of Ginterplanar in Region 1, which is near a relatively thin pore region. The graphite bulk in Region 1 has a uniform lattice rotation ( 0.14 nm 1 ∼ ° − ), while the graphite edge has greater disorder (with rotational changes 2 nm 1 > ° − ). Figure 9b(ii) shows the corre- sponding interplanar d-spacing reconstruction map; the d-spacing is elevated ( 0.37 nm > ) only within ∼2 nm of the particle edge while the bulk has d-spacings similar to those in the Gr-AP sample. The d- spacing distribution (Fig. 9b(iii)) shows a mean around 0.337 nm and a standard deviation of 0.008 nm, only slightly larger than those for the Gr-AP sample. On the other hand, Region 2 (acquired near the edge of a larger graphite pore) of the Gr-LP sample (Fig. 9c) shows greater lattice disorder. The average rotation gradient in this region is 0.30 nm, ̃ °/ while the particle edges have values 4 nm > °/ (Fig. 9c(i)). Moreover, the frequency of lattice rotation (i.e., the “wavy” nature of the lattice) occurs up to ∼20 nm into the graphite bulk from the pore edge. A similar behavior is observed in Fig. 9c(ii), which shows elevated d-spacings near the pore edge and extending into the particle bulk. Figure 9c(iii) shows that the mean d-spacing of Region 2 is 0.35 nm, which is larger than that calculated from the X-ray data. The d-spacing distribution also shows a tail that extends from ∼0.36 nm to 0.40 nm, which indicates that the graphite edges in Region 2 have sustained greater damage than those in Region 1. This result also points to the spatial heterogeneity in lattice disorder within each graphite particle—the edges with greater exposure to the electrolyte will experience more Li+ ion (de)intercalation reactions and apparently greater damage, especially during high-rate charging. Damage localized at the graphite particle edges is not entirely unexpected.45 Prior studies have pointed out that Li+ ion intercala- tion into graphite occurs primarily through the edge planes.46,47 At high charge rates, significant build-up of Li+ ions is expected at the edges planes; the concentration gradients between the surface and bulk sites could induce stresses that are significant enough to deform the graphite lattice near the particle edges. Mitigating damage at the graphite particle edges is crucial to prolonging the life of LIB cells subjected to fast charge. One solution would be to raise the lower cutoff voltage (LCV) for cell cycling (e.g. from 2.8 to 3.3 V) to prevent complete delithiation of the graphite. The residual dilute LixC6 phases have higher d-spacings than the pristine graphite,29 which would facilitate Li+ ion inter- calation and (partially) mitigate the stresses induced during the process. Alternative solutions could include the use of graphite with higher d-spacings in the anode, as has been suggested for both Li+ and Na+ ion cells.48,49 However, such “expanded” graphite materials typically have lower electronic conductivity, which would limit the overall rate of intercalation. In addition, these materials have higher surface areas and a proclivity to trap more lithium in the SEI formed during electrolyte reduction reactions. Other solutions proposed to facilitate rapid cycling include the use of functionalized-graphite that have higher d-spacings at the particle edges,50 graphite with particle edges aligned perpendicular to the current collector,51 graphite with carbon coatings that have larger interlamellar spacing,52 and anodes containing a mix of graphite and hard carbons.7 These efforts, along with research on cathode and electrolyte materials, electrode architectures and cycling protocols, are important steps towards making fast-charging of LIB cells, a commercial reality. Conclusions Fast charging of LIB cells can degrade their energy and power performance characteristics. Here we show representative data from as-prepared graphite electrodes (Gr-AP) and from electrodes ex- tracted from (discharged) cells that were previously charged at rates up to 6 C (Gr-LP). Electrochemical measurements show that the harvested electrodes have lower capacities than the Gr-AP electrode and display a sloping lithiation profile, rather than a staged profile, even at low rates. Analysis of X-ray diffraction data from the Gr-AP and Gr-LP electrodes yield similar interplanar d-spacings, indicating that the graphite bulk is not altered by the electrochemical cycling. Similar conclusions about the graphite particle bulk can be drawn from the analysis of HREM and HAADF-STEM images and SEND data. However, these nanoscale data clearly indicate damage at the graphite edges, which extend up to ∼20 nm into the particle bulk for the Gr-LP sample. This damage is seen as arcs (rather than spots) in the FFT data of HREM images and as increased rotations and higher d-spacings in the SEND data; both datasets suggest increased turbostratic disorder and buckling of the graphene planes at particle edges. In addition to changes in graphite-edge structure, the Gr-LP electrode shows a significant thickness increase from the build-up of SEI and particle rearrangements that occur during the cycling. This SEI contains a mix of compounds, including LiF (seen in the XRD Journal of The Electrochemical Society, 2021 168 100509
  • 11. patterns) and (oxy)fluoro-phosphates suggested by the EDS data. Larger quantities of this SEI are present adjacent to the larger internal pores of the graphite, which likely have easier access to the electrolyte than the smaller pores. Furthermore, the SEI is localized Figure 8. Quantitative lattice rotation and interplanar d-spacing analysis of the as-prepared graphite (Gr-AP) sample. (a) HAADF-STEM images (i) at low- magnification, showing Regions 1 and 2 where SEND acquisition was performed; (ii), (iii) magnified images of Regions 1 and 2, respectively. Diffractive image reconstructions and analysis of (b) Region 1 and (c) Region 2 containing the following: (i) in-plane rotation gradient map, (ii) interplanar spacing map, (iii) interplanar spacing distribution. Scale bars: (a) (i) is 2 μm; (a) (ii), (a) (iii), (b) (i), (b) (ii), (c) (i), (c) (ii) are 50 nm. Journal of The Electrochemical Society, 2021 168 100509
  • 12. at the particle edges and does not penetrate the graphite bulk, suggesting that only Li+ ions (no electrolyte components) intercalate into the particles (effective desolvation of ions). Figure 9. Quantitative lattice rotation and interplanar d-spacing analysis of the fast-charged graphite (Gr-LP) electrode. (a) HAADF-STEM images (i) at low- magnification showing Regions 1 and 2 where SEND data were acquired; (ii), (iii) magnified images of Regions 1 and Region 2. Diffractive image reconstructions and analysis of (b) Region 1 and (c) Region 2 containing the following: (i) in-plane rotation gradient map, (ii) interplanar spacing map, (iii) interplanar spacing distribution. Scale bars: (a) (i) is 2 μm; (a) (ii), (a) (iii), (b) (i), (b) (ii), (c) (i), (c) (ii) are 50 nm. Journal of The Electrochemical Society, 2021 168 100509
  • 13. The implication of our study is that even a small number of high- rate cycles appears to be sufficient to induce significant and permanent disorder in graphite domains that are closest to the electrolyte, be it at particle surfaces or the inner pores. Such disorder could affect the lithiation kinetics, preventing the graphite particles from accepting charge and favoring the occurrence of Li plating. Moreover, we show that these structural changes are highly non- uniform across the various pores within a same particle, which could either be a cause or consequence of reaction heterogeneities commonly observed at high rates. Awareness and mitigation of the structural evolution at the graphite/electrolyte interfaces is crucial to the development of LIBs that can endure repeated fast-charging and yet reliably deliver high-performance over the 10 + year lifespan of electric vehicles. Acknowledgments SP acknowledges support from the U.S. Department of Energy (DOE) Graduate Student Research (SCGSR) program. The SCGSR program is administered by the Oak Ridge Institute for Science and Education for DOE under contract number DE‐SC0014664. This work was carried out in part at the Materials Research Laboratory Central Research Facilities, University of Illinois. DA and MTFR acknowledge support from DOE’s Vehicle Technologies Office (VTO). The electrodes used in this article were fabricated at Argonne’s Cell Analysis, Modeling and Prototyping (CAMP) Facility, which is fully supported by the VTO. We are grateful to our many colleagues, especially Stephen Trask and Andrew Jansen at Argonne. Use of the Advanced Photon Source (APS) at Argonne National Laboratory was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, under Contract No. DE-AC02–06CH11357. We are grateful to Saul Lapidus at the APS for help with the high-resolution powder diffraction experiments. The submitted manuscript has been created by UChicago Argonne, LLC, Operator of Argonne National Laboratory (“Argonne”). Argonne, a U.S. Department of Energy Office of Science laboratory, is operated under Contract No. DE- AC02–06CH11357. The U.S. Government retains for itself, and others acting on its behalf, a paid-up nonexclusive, irrevocable worldwide license in said article to reproduce, prepare derivative works, distribute copies to the public, and perform publicly and display publicly, by or on behalf of the Government. ORCID Saran Pidaparthy https://orcid.org/0000-0002-0783-3094 Marco-Tulio F. Rodrigues https://orcid.org/0000-0003-0833-6556 Jian-Min Zuo https://orcid.org/0000-0002-5151-3370 Daniel P. Abraham https://orcid.org/0000-0003-0402-9620 References 1. A. Masias, J. Marcicki, and W. A. Paxton, ACS Energy Lett., 6, 621 (2021). 2. A. Raj, M.-T. F. Rodrigues, and D. P. Abraham, J. Electrochem. Soc., 167, 120517 (2020). 3. A. Tomaszewska et al., eTransportation, 1, 100011 (2019). 4. S. Ahmed et al., J. Power Sources, 367, 250 (2017). 5. A. M. Colclasure et al., Electrochim. Acta, 337, 135854 (2020). 6. X.-G. Yang, T. Liu, Y. Gao, S. Ge, Y. Leng, D. Wang, and C.-Y. Wang, Joule, 3, 3002 (2019). 7. K.-H. Chen, V. Goel, M. J. Namkoong, M. Wied, S. Müller, V. Wood, J. Sakamoto, K. Thornton, and N. P. Dasgupta, Adv. Energy Mater., 11, 2003336 (2021). 8. D. W. Dees, M.-T. F. Rodrigues, K. Kalaga, S. E. Trask, I. A. Shkrob, D. P. Abraham, and A. N. Jansen, J. Electrochem. Soc., 167, 120528 (2020). 9. M.-T. F. Rodrigues, K. Kalaga, S. E. Trask, D. W. Dees, I. A. Shkrob, and D. P. Abraham, J. Electrochem. Soc., 166, A996 (2019). 10. I. A. Shkrob, M.-T. F. Rodrigues, D. W. Dees, and D. P. Abraham, J. Electrochem. Soc., 166, A3305 (2019). 11. M.-T. F. Rodrigues, I. A. Shkrob, A. M. Colclasure, and D. P. Abraham, J. Electrochem. Soc., 167, 130508 (2020). 12. T. Waldmann, B.-I. Hogg, and M. Wohlfahrt-Mehrens, J. Power Sources, 384, 107 (2018). 13. L. Somerville, J. Bareño, S. Trask, P. Jennings, A. McGordon, C. Lyness, and I. Bloom, J. Power Sources, 335, 189 (2016). 14. J. Cannarella and C. B. Arnold, J. Electrochem. Soc., 162, A1365 (2015). 15. J. S. Okasinski, I. A. Shkrob, A. Chuang, M.-T. F. Rodrigues, A. Raj, D. W. Dees, and D. P. Abraham, Phys. Chem. Chem. Phys., 22, 21977 (2020). 16. S. Müller, J. Eller, M. Ebner, C. Burns, J. Dahn, and V. Wood, J. Electrochem. Soc., 165, A339 (2018). 17. A. Mukhopadhyay and B. W. Sheldon, Prog. Mater Sci., 63, 58 (2014). 18. A. Raj, I. A. Shkrob, J. S. Okasinski, M.-T. Fonseca Rodrigues, A. C. Chuang, X. Huang, and D. P. Abraham, J. Power Sources, 484, 229247 (2021). 19. K. P. C. Yao, J. S. Okasinski, K. Kalaga, I. A. Shkrob, and D. P. Abraham, Energy Environ. Sci., 12, 656 (2019). 20. M. T. F. Rodrigues, S.-B. Son, A. M. Colclasure, I. A. Shkrob, S. E. Trask, I. D. Bloom, and D. P. Abraham, ACS Appl. Energy Mater., 4, 1063 (2021). 21. M.-T. Fonseca Rodrigues, V. A. Maroni, D. J. Gosztola, K. P. C. Yao, K. Kalaga, I. A. Shkrob, and D. P. Abraham, ACS Appl. Energy Mater., 2, 873 (2019). 22. P. P. Paul et al., , Adv. Energy Mater., 11, 2100372 (2021). 23. H. Shi, J. Barker, M. Y. Saïdi, R. Koksbang, and L. Morris, J. Power Sources, 68, 291 (1997). 24. A. Van der Ven, Z. Deng, S. Banerjee, and S. P. Ong, Chem. Rev., 120, 6977 (2020). 25. H. Fujimoto, H. Kiuchi, S. Takagi, K. Shimoda, K. Okazaki, Z. Ogumi, and T. Abe, J. Electrochem. Soc., 168, 40509 (2021). 26. H. He, C. Huang, C.-W. Luo, J.-J. Liu, and Z.-S. Chao, Electrochim. Acta, 92, 148 (2013). 27. S. Schweidler, L. de Biasi, A. Schiele, P. Hartmann, T. Brezesinski, and J. Janek, J. Phys. Chem. C, 122, 8829 (2018). 28. J. Wilhelm, S. Seidlmayer, S. Erhard, M. Hofmann, R. Gilles, and A. Jossen, J. Electrochem. Soc., 165, A1846 (2018). 29. E. Hazrati, G. A. de Wijs, and G. Brocks, Phys. Rev. B, 90, 155448 (2014). 30. S.-B. Son, D. Robertson, Z. Yang, Y. Tsai, S. Lopykinski, and I. Bloom, J. Electrochem. Soc., 167, 140506 (2020). 31. R. E. Franklin, Acta Crystallogr., 4, 253 (1951). 32. J. R. Dahn, A. K. Sleigh, H. Shi, J. N. Reimers, Q. Zhong, and B. M. Way, Electrochim. Acta, 38, 1179 (1993). 33. J. R. Dahn, R. Fong, and M. J. Spoon, Phys. Rev. B, 42, 6424 (1990). 34. P. Verma, P. Maire, and P. Novák, Electrochim. Acta, 55, 6332 (2010). 35. A. M. Andersson and K. Edström, J. Electrochem. Soc., 148, A1100 (2001). 36. D. Sauerteig, S. Ivanov, H. Reinshagen, and A. Bund, J. Power Sources, 342, 939 (2017). 37. B. Rieger, S. Schlueter, S. V. Erhard, J. Schmalz, G. Reinhart, and A. Jossen, J. Energy Storage, 6, 213 (2016). 38. A. Y. R. Prado, M.-T. F. Rodrigues, S. E. Trask, L. Shaw, and D. P. Abraham, J. Electrochem. Soc., 167, 160551 (2020). 39. R. Wagner, M. Korth, B. Streipert, J. Kasnatscheew, D. R. Gallus, S. Brox, M. Amereller, I. Cekic-Laskovic, and M. Winter, ACS Appl. Mater. Interfaces, 8, 30871 (2016). 40. M. Stich, M. Göttlinger, M. Kurniawan, U. Schmidt, and A. Bund, J. Phys. Chem. C, 122, 8836 (2018). 41. D. J. Miller, C. Proff, J. G. Wen, D. P. Abraham, and J. Bareño, Adv. Energy Mater., 3, 1098 (2013). 42. A. Marion, , ed. A. Marion (Springer, US, Boston, MA) p. 255 (1991). 43. R. C. Gonzalez and R. E. Richard E. Woods, Digital image processing 2nd ed. (Prentice Hall, Upper Saddle River, N.J) (2002). 44. J. M. Zuo and J. C. H. Spence, Advanced Transmission Electron Microscopy (Springer, New York, NY) (2017). 45. V. A. Sethuraman, L. J. Hardwick, V. Srinivasan, and R. Kostecki, J. Power Sources, 195, 3655 (2010). 46. R. J. Rice and R. L. McCreery, Anal. Chem., 61, 1637 (1989). 47. K. Persson, V. A. Sethuraman, L. J. Hardwick, Y. Hinuma, Y. S. Meng, A. van der Ven, V. Srinivasan, R. Kostecki, and G. Ceder, J. Phys. Chem. Lett., 1, 1176 (2010). 48. T.-H. Kim, E. K. Jeon, Y. Ko, B. Y. Jang, B.-S. Kim, and H.-K. Song, J. Mater. Chem. A, 2, 7600 (2014). 49. Y. Wen, K. He, Y. Zhu, F. Han, Y. Xu, I. Matsuda, Y. Ishii, J. Cumings, and C. Wang, Nat. Commun., 5, 4033 (2014). 50. J.-S. Park, M.-H. Lee, I.-Y. Jeon, H.-S. Park, J.-B. Baek, and H.-K. Song, ACS Nano, 6, 10770 (2012). 51. J. Billaud, F. Bouville, T. Magrini, C. Villevieille, and A. R. Studart, Nat. Energy, 1, 16097 (2016). 52. W. Cai, C. Yan, Y.-X. Yao, L. Xu, R. Xu, L.-L. Jiang, J.-Q. Huang, and Q. Zhang, Small Struct., 1, 2000010 (2020). Journal of The Electrochemical Society, 2021 168 100509