You are on page 1of 19

MRS Advances © 2019 Materials Research Society

DOI: 10.1557/adv.2019.443
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

On our Limited Understanding of Electrodeposition

Aashutosh Mistry1 and Venkat Srinivasan*,2

Argonne National Laboratory, Lemont, Illinois 60439, United States

ABSTRACT

The energy density of electrodeposition reactions makes them attractive for energy storage.
Although its scientific inquiries nearly date back to the inception of electrochemistry, its
behavior at microscopic dimensions (relevant to battery application) is mysteriously
uncontrollable. We examine experimental reports of singular spatiotemporal evolutions with a
hope to identify universality in deposition patterns. We conclude that a macroscopic mass
transport instability cannot account for various growth morphologies and alludes to poorly
understood materials interplay at smaller scales. We summarize representative
characteristics of electrodeposition to encourage mechanistic investigations.

Introduction

Electrodeposition represents the formation of a solid phase from an ionic solution (i.e.,
electrolyte) upon passage of electricity. It is analogous to solidification as a metallurgical
process. It is an interesting proposition given its unique characteristics, e.g., an electric
current is easier to modulate as compared to a heat flux (amount of solid formed is
proportional to the current in electrodeposition and the heat flux in solidification).
It also seems lucrative from the standpoint of an energy storing reaction given
apparent simplicity of reaction step(s) as compared to intercalation in solid hosts (the
state-of-the-art battery chemistry). The formed solid contributes to the energy in its
entirety. Thus electrodeposition reactions can provide higher energy and greater power.
Consequently, electrodeposition is being pursued for futuristic energy storage
technologies1. Such a scheme represents a paradigm shift from bulk energy storage to an
interfacial mode. In situ, in operando and ex situ studies, however, reveal idiosyncrasies
ranging from nonuniform deposition 2 to semi-reversibility3 – in general, a certain degree
of uncertainty and unpredictability.

1
Email: amistry@anl.gov (AM) and vsrinivasan@anl.gov (VS)
1
ORCID: 0000 – 0002 – 4359 – 4975
2
ORCID: 0000 – 0002 – 1248 – 5952

2843
The conventional interpretation assigns the uncontrollability to
electrodeposition instability, wherein a growth front develops well-defined
inhomogeneities for certain operating conditions. Such instability is analogous to the
morphological instability observed for solidification 4 at fast cooling rates. Corresponding
mechanistic arguments define (for constant current deposition) limiting current as a
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

measure of growth instability. When the deposition is carried out faster than the limiting
current, a planar growth front cannot be maintained and dendritic patterns emerge 5,6.
Certain electrodeposition morphologies, especially for copper deposition (a usual choice
for theoretical validation), conform to such a view. On the other hand, innumerable
electroplating experiments disclose nonuniformity7–12 which is in contrast to the typical
fractal nature of fast unstable growth.
Physically, electrodeposition refers to the formation of a solid phase. Herein the
geometrical evolution and reaction distribution are strongly interdependent. The
usefulness of electrodeposition relates to the ability to maintain desired growth geometry
during its time evolution. Hence, a scientific understanding of associated
uncontrollability is essential to its application.
Such electrodeposition nonuniformities remain, at best, partially understood.
We revisit representative instances of electrodeposition to outline the variants of material
response. If underlying interactions were universal across different occurrences, resultant
growth patterns should exhibit morphological similarities. We discuss the contemporary
mechanistic interpretation to identify possible commonalities. We identify the unresolved
material interactions that symptomize the irregularities in electrodeposition.

Observations, present understanding and discrepancies

Electrodeposition is an intrinsically transient interaction, and as a result, an explicit


probing necessitates continuous tracking. Additionally, the study of nonuniformities
demands spatially resolved measurements. Such constraints on simultaneously resolving
space and time complicate in situ or in operando observations of electrodeposition, and
one usually resorts to incomplete information. For example, optical cells 2 involve large
interelectrode spacings to facilitate observations that promote electrolyte mass transport
limitation; tomography cells9 examine slow dynamics given the beam exposure timings
for spatial sampling, and experiments resolving fast transients record macroscopic
descriptors such as voltage and impedance 13. A usual compromise is to complement in
situ macroscopic data with ex situ imaging7, which presupposes equivalent dynamics
across cells that are imaged at different electrochemical states to correlate time evolution
of electrodeposition geometry (also, surface contamination is possible when exposed to
non-electrolytic environments). Such attention to measurement signatures is essential to
consistently interpret dynamical regimes.
As alluded earlier, the general understanding of electrodeposition is largely
attributed to measurements with metallic electrodes in optical cells 2,14. Since the
electrodepositing interface locally consumes the ion being deposited (often a cation), if
fast electrodeposition is carried out, local ionic concentration depletes in a finite time
(referred to as the Sand’s time). The limiting current is defined as the smallest current
that could deplete the concentration of the reacting ion close to the electrodepositing
interface (limiting current and Sand’s time are correlated; for higher currents, Sand’s
time is smaller implying an earlier onset of cation starvation). Beyond the limiting
current density, small perturbations of the electrodepositing interface grow in an
uncontrolled fashion at times larger than the Sand’s time14,15. Once local ionic
concentration depletes, a slight perturbation on a flat interface seeks partially depleted
regions and these (peak) locations grow selectively. Such a preferential growth manifests

2844
as uncontrolled dendritic deposits14. Note that before Sand’s time, the electrodepositing
interface is stable wherein the spatial fluctuations do not grow in an unbounded fashion.
Expressed differently, irregular electrodeposition front is caused by a mismatch
between growth (quantified as reaction rate) and supply of reacting species to the
interface (defined by diffusive transport in the electrolyte). If reactants cannot reach
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

certain locations on the electrodepositing interface, locally further growth is inhibited.


Thus, nonuniformity of the deposition scales with local concentrations of the reacting
species.
We examine electrodeposition reactions that do not necessarily conform to such
an interpretation. The following discussion begins with lithium deposits in the liquid
electrolyte where a wider range of morphological observations has been made.
Successively more complex (and less understood) reactions describing electrochemical
growth of compounds such as lithium sulfide, Li2S, and lithium peroxide, Li2O2, are
discussed. The physicochemical response of these reactions changes, somewhat
unexpectedly, when liquid electrolytes are switched to solids and are debated as such.

Lithium Electrodes in Liquid Electrolytes

In contrast to the expected planar front, electrodepositing lithium is found to result in a


porous pattern that grows steadily (referred to as mossy deposits; Figure 1(d),(e)).
Additionally, it has been shown that the mossy lithium grows from the base, unlike the
dendritic lithium where successive electrodeposition events take place at the tips and
implies two distinct scenarios of root-growth and tip-growth respectively2. It is observed
that beyond a certain time, the nature of electrodeposition possibly switches from mossy
(Figure 1(d),(e)) to dendritic deposits (Figure 1(f),(g)) in the same experiment. The time
evolution of electrode potential also changes concurrently (Figure 1(b)). Mossy deposits
are characterized by steadily increasing potential, which jumps almost instantaneously to
a higher value at the onset of dendritic growth and subsequently shows a chaotic
signature. The slow increase during the mossy growth signifies a gradual decline of
cationic concentration near electrodepositing locations (schematic Figure 1(h)). Once the
cations are locally consumed, the electrochemical response is stochastic and refers to
uncontrolled deposition.
The sequential mossy-dendritic growth generalizes the conventional
interpretation of electrodeposition instability. The conventional interpretation identifies
the Sand’s time as the shift from planar front to dendritic, while herein it represents the
transition from mossy to dendritic growth 2. Given the presence of ample cations, mossy
geometry, though being irregular, does not qualify for a transport-limited instability. The
high reactivity of lithium (unlike copper) forms a surface layer when in contact with a
liquid electrolyte. It has been typically argued that this surface layer (commonly referred
to as Solid Electrolyte Interphase, SEI) breaks down locally, forming pinholes that lead
to the mossy growth2,13.

2845
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

Figure 1. Representative lithium electrodeposition measurement in liquid electrolyte optical cells 2: (a) experimental
geometry (b) evolution of electrode potential during a fixed current density deposition (c)-(g) visualization of growth at
characteristic times identified on voltage signature, and (h) schematic illustration of growth sequence. (reproduced with
permission from Royal Society of Chemistry2)

The electrodeposition sequence presented in Figure 12 has been visualized in


situ in a custom Li|Li symmetric cell configuration. In contrast, Lu et al. (Figure 27)
examine the electrodeposition in a Li|NCA (Nickel Cobalt Aluminum oxide – a typical
cathode) cell, representative of the battery. They operate multiple such cells at fixed rates
for 100 cycles (recording macroscopic descriptors throughout). Subsequently, the cells
are dissembled carefully for an ex situ visualization and materials characterization to
comprehend the morphological evolution of the lithium metal surface 7. In contradiction
with the mixed mossy-dendritic growth observed in optical cells 2,3, here the
electrodepositing surface grows into a porous structure comprised of electronically
isolated lithium (often known as dead lithium) 7. Materials characterization reveals that
SEI coverage is responsible for isolating lithium. Furthermore, analysis of lithium anode
surfaces that undergo different electrodeposition histories at different currents, suggests
that the development of porous structure is faster at when cycled at higher rates (SEM,
Scanning Electron Microscopy images; Figure 2(b)-(d)). A thicker structure results in a
greater resistance buildup.

2846
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

Figure 2. Repeated electrodeposition in Li|NCA cells reveal previously unseen deposition patterns, where lithium
electrodeposits as a porous structure containing electronically isolated lithium, rather than forming dendritic patterns 7: (a)
schematic of the electrodeposition sequence; (b)-(d) ex situ SEM for anodes electrodeposited-stripped-redeposited at
different rates for 100 cycles. Porous structure at the anode surface and leftover lithium are visible in (b)-(d). (reproduced
with permission from Wiley-VCH7)

Such an electrodeposition structure (Figure 27) is considerably different than


Figure 12. Despite repeated electrodeposition, no dendrites are observed here (in contrast
Figure 12 exhibits dendritic growth during the first deposition). Secondly, the surface
structure growth is inward from the surface of the pristine lithium. Thirdly, even if this
porous structure and mossy deposits contain pores, two are geometrically unalike. Mossy
deposits (Figure 1(c)-(g)) have a tentacled structure and grow into the electrolyte, while
the porous layer in Figure 2(b)-(d) is cavitied in nature and siphons electrolyte. The
former one manifests during electrodeposition, while the latter develops over dissolution.
Note that both lithium electrodeposition and SEI formation are simultaneously active.
Since porous structure grows over repeated deposition-stripping sequences, there appears
to be a characteristic degree of irreversibility associated with lithium electrodeposition 7.
When repeated deposition-stripping experiments are carried out in a Li|Li symmetric
cell3, a similar surface structure is observed. However, here, again the growth is outward
from the pristine lithium surface.

2847
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

Figure 3. Examining electrodeposition in typical cells with a few microns spacing between the two electrodes 16: (a)
dendritic growth is unlikely in such cells given smaller separations; (b) these lengths are comparable to nucleation and in
tun electrodeposition is a locally discrete event; (c) such microscopic irregularity traps electrolyte as a confined layer and
limits the critical flux. (reproduced with permission from American Chemical Society 16)

The aforementioned morphological differences hint at two possible


electrodeposition regimes. Lengthscale, specifically the interelectrode spacing, is an
apparent difference between the two experiments. The limiting current (equivalently the
Sand’s time) scales with interelectrode spacing (Figure 3(a)). Optical cells have much
larger spacing than the standard configurations (e.g., coin cell). Higher interelectrode
spacings reduce the limiting currents, as a consequence, dendritic growth is prevalent in
visual cells. On the other hand, for typical electrode spacings 7,16, smaller interelectrode
distances render dendrite formation quite unlikely as these small separations exclude
severe concentration gradients causing ionic depletion near the depositing electrode.
However, when growth patterns are visually examined in these cells 7,16, nonuniform
electrodeposition is observed (e.g., porous structure in Figure 2(b)-(d)). Since the
currents are well below the limiting current, one would expect planar growth, and the
microscopically nonuniform growth appears to be intriguing.
Such examples7,16 suggest an unidentified mechanism causing microscopic
unevenness. Examination of different lengthscales reveals that when the electrode
separations reduce to a couple of microns, they become comparable to the average
distance between neighboring nucleation sites. Given the comparable lengths, the
discrete nature of nucleation and growth becomes relevant (Figure 3(b)). Thenceforth,
electrodeposition is no longer a planar front, even if macroscopically it is a uniform
growth. The associated surface structure locally traps electrolyte, causing a short-range
transport limitation. Such electrolyte confinement worsens (Figure 3(c)) with
inhomogeneity as well as the growth of the surface structure 16. Consequently, the cation
concentration drops as electrodeposition progresses, resulting in a rapid voltage rise.

2848
Theoretical analysis of the confined electrolyte predicts signatures observed in controlled
experiments16, and justify the relevance of the new mechanism. Additionally, since the
geometry of such a surface structure is related to nucleation dynamics, and the confined
electrolyte limitation should scale with temperature. Further experimentation confirms
such trends16.
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

These studies2,7,13,16 investigate long term electrodeposition. Given the strongly


correlated evolutions, differences in the initial stages of electrodeposition are expected to
generate different growth patterns. Figure 4 reports representative short-time attributes
for lithium electrodeposition under varying conditions 8,17. Since experiments are carried
out in coin-cells, long-range transport limitation 2,3 is absent and observed geometrical
variations arise from local effects. The formation of a new phase on a dissimilar substrate
is an energy intensive process that involves the formation and disappearance of interfaces
(a characteristic of short-range interactions). To probe this interplay, lithium
electrodeposition is carried out on copper substrates 8,17.
When current densities are varied (Figure 4(a)-(j)) with the same electrolyte,
nucleation dynamics change in response to the deposition timescale. Slower deposition
facilitates nucleation growth and coalescence. Such a tendency diminishes at higher
currents, where a larger number of nuclei are formed. Figure 4(a)-(j) present ex situ SEM
images, and visually corroborate such trends 17. The size of nuclei and nucleation site
density are identified to scale with currents (equivalently overpotentials). Alternatively,
Figure 4(k)-(n) examines the effect of electrolyte environment on deposition morphology
at fixed currents (i.e., deposition rate) and capacities (i.e., the amount deposited). The
electrolytes are chosen to exemplify Li-ion (Figure 4(k)), Li-sulfur (Figure 4(l)-(m)) and
Li-oxygen (Figure 4(n)) chemistries. Notice that the resultant morphologies in Figure
4(k)-(n) are geometrically dissimilar8, unlike Figure 4(a)-(j) where electrodeposition
patterns are morphologically equivalent and vary in quantitative attributes. The growth
structures in Figure 4(a)-(j) are still distinguishable from dendritic tree-like growth (an
attribute of diffusive instability). An X-ray diffraction examination of the surfaces
Figure 4(k)-(n) reveals crystallographic differences. The geometrical details at such
smaller scales appear to be quite different from other deposition patterns presented
before.

2849
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

Figure 4. Analyzing lithium nucleation on a copper substrate as a function of currents(a)-(j)17. Studying lithium
deposition morphology with different electrolytes(k)-(n)8. A fixed amount of electrodeposition is carried out to ensure a
consistent comparison. (a)-(j) darker spots are lithium. (reproduced with permissions from American Chemical Society 17
and US National Academy of Sciences 8)

Figure 5 schematically compares the electrodeposition at different


interelectrode spacings. At smaller electrolyte thickness, a considerable amount is
trapped in the nonuniform structure created due to nucleation and growth of
electrodeposits. The transport in this confined layer is hindered as compared to the bulk
electrolyte and can lead to concentration depletion within the confined layer. When the
confined layer thickness is comparable, its resistance is higher than the bulk electrolyte
and in turn depositing locations are starved of cations even when the bulk layer
concentration drop is negligible (Figure 5(a)). The confined layer thickness evolves
during electrodeposition and is a couple of microns. On the other hand, when bulk
electrolyte thickness is larger (e.g., millimeters; Figure 5(b)), it dominates the electrolyte
resistance (confined layer is still present). The conventional interpretation of limiting
currents is related to transport resistance of the bulk electrolyte and remains valid when
the electrode thickness is increased.
Since the confined layer is related to irregular deposition characterized by
nucleation and growth events at smaller lengths, transport limitations in this layer
actively alter the associated geometrical evolution. Once the nuclei coalescence and form
continuous films (of possibly nonuniform thickness), the confined layer diminishes and
transport in bulk electrolyte becomes limiting (Figure 5(c)).

2850
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

Figure 5. Electrolyte transport limitation is length-scale dependent. (a) When the interelectrode spacing is comparable to
lengthscales of nucleation and growth events, electrolyte trapped in such nonuniform deposits dictates growth. (b)
Alternatively, when the two electrodes are far apart, concentration gradients are predominantly established in the bulk
electrolyte. (c) Since electrolyte confinement is due to nonuniform deposition during nucleation and discrete growth,
their evolution is affected by electrolyte transport. When the nuclei coalesce and form films, confined electrolyte layer
vanishes and the bulk transport limitations, if any, affect further growth.

Supplemental complications arise when structuring is present. For example,


lithium electrodeposition uniformity is found to improve when the incoming ionic flux is
ordered using a nano-channel construct18. Another somewhat mysterious observation is
homogenized lithium deposition in the presence of high atomic weight cation, e.g.,
cesium19. Without resolving the unpredictability of lithium metal anode, it is paired with
prototype cathodes to examine their limitations, for example, sulfur and air cathodes. A
recent study20 finds that polysulfides along with lithium nitrate present in sulfur
electrolytes form a protective surface layer on lithium surface that modulates the
irregular growth, implying a stabilizing effect of the surface layer.
In nutshell, lithium electrodeposition over microscopic diffusion lengths
displays geometrical nonuniformities that are distinct from the mass transport limited
dendritic deposition (diffusion length is equivalent to the interelectrode spacing). Such
dissimilarities are found to manifest at short-times8,17, long-times16 as well as over
repeated deposition-dissolution events7 (in contrast the dendritic irregularity develops
over time) and point to unexplained short-range interactions in small-scale
electrochemical growth.

Electrochemical Growth of Low Conductivity Compounds

Electrodeposition reactions require a supply of ions and electrons and take place at the
interface of solid and solution phases. Often the deposits are metals (lithium, copper,
etc.) and the availability of electrons is implicitly guaranteed. Hence, conventionally, any

2851
unexpected behavior is attributed to solution-phase limitations. Preceding discussion
points to complexities arising from the nature of the interface, especially in the limit of a
large active surface to electrolyte volume ratio. In terms of limiting processes, long-range
electrolyte transport resistance becomes less relevant at the smaller scales and short-
range resistances dominate electrodeposition response. If the electrodeposited phase is a
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

poor conductor of electricity, electrons cannot easily reach all the points at the deposit –
electrolyte interface (Figure 6) and reaction distribution and growth are skewed. When
growing a high conductivity phase, electrons are available at all the surface locations and
in turn, growth takes place along the two-phase interface. Alternatively, if the electronic
conduction is severely limited, electrons can only reach interface locations close to the
substrate. In other words, electrodeposition preferentially takes place close to the contact
of the substrate, deposit, and electrolyte. In this limit, growth is confined to the three-
phase contact and takes place laterally. The growth thickness would be limited, and
morphologically one would expect two-dimensional electrodeposition. Figure 6
schematically illustrates the competitive influence of electrodeposition reaction and
electronic conduction on resultant growth morphology.
Electrodeposition of lithium sulfide, Li2S, and lithium peroxide, Li2O2, are
representative low conductivity species relevant to futuristic battery chemistries (Li-
sulfur and Li-oxygen, respectively). Such an electrochemical growth differs from metal
deposition in various respects:
ƒ The electrodepositing phase is electronically insulating
ƒ As a result, electrodeposition takes place on a conductive substrate that is to
facilitate electrons for the reaction
ƒ Involves multiple electrolyte species (cationic, anionic or charge neutral)
These characteristics modify physicochemical interactions and the geometrical
manifestation of growth. Given the presence of multiple electrolyte phase species, the
limiting current may not be well defined. Since the substrate is required for electronic
transport, the nucleation of deposits on the substrate becomes relevant. The reaction
pathway probably involves multiple elementary steps and needs to be probed explicitly.

Figure 6. The electronic conductivity of the electrodeposited phase can alter the distribution of the electrochemical
reaction causing growth at the deposit – electrolyte interface. As electronic conductivity becomes limiting, electrons
cannot reach each of these surface locations and morphology changes accordingly.

Consider the electrochemical growth of Li2S in organic electrolytes. It is


formed upon electrochemical reduction of solution phase species, per the reaction

2852
Figure 7 reports potentiostatic measurements that examine the nucleation and growth
characteristics of Li2S electrodeposition11. Higher order polysulfides are first carefully
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

reduced to the same valence, S42-. Subsequently, the cell is held at a reducing potential
(compared to the equilibrium reaction voltage of 2.15 V) to trigger Li2S nucleation. As
an example, Figure 7(a),(c)-(e) presents electrodeposition at 130 mV overpotential.
Figure 7(a) plots corresponding current evolution containing the growth signature. The
Avrami law based analysis reveals that Li2S formation takes place through progressive
nucleation and 2D (i.e., planar) growth. Visual images Figure 7(c)-(e) confirm the
analytical interpretation. At an early instant (c), surface coverage is marginal (Figure
7(c)). As nuclei grow laterally, maximum surface coverage for unobstructed growth is
reached (Figure 7(d)) leading to the current peak (d). Beyond this instant,
electrodeposition current decays (e) as both nucleation and growth are hindered (Figure
7(e))11.
Since Li2S is electronically insulating, a three-phase boundary formed by Li2S,
substrate and electrolyte contact is argued to be the reaction zone. Over time, the length
of the three-phase boundary increases, partly due to the growth of old nuclei and partly
due to the formation of new ones, thus increasing the reaction current in Figure 7(a).
Surface diffusion of Li2S formed at the three-phase boundary is claimed to be the
mechanism for thickness growth (Figure 7(b)). Such an understanding of potentiostatic
(constant voltage) electrodeposition also translates to Li2S formation at a fixed rate. It is
found that slow electrodeposition results in smaller nucleation site density but greater
growth. On the other hand, fast discharge forms a higher nucleation density and
conformal film-type precipitates. When Li2S is grown on different substrates,
qualitatively similar behavior is observed 21.

Figure 7. Mechanism of Li2S electrodeposition examined under potentiostatic conditions 11: (a) current evolution at a 130
mV overpotential; (b) physicochemical scheme of Li 2S growth; (a-inset) SEM of pristine carbon substrate, (c)-(e) SEM
at 2.5, 4 and 6h during potentiostatic deposition. All scale bars are 1μm. (reproduced with permission from Wiley-
VCH11)

2853
Lithium peroxide, Li2O2, electrodeposition at constant current 10 exhibits certain
similarities to the aforementioned potentiostatic Li2S formation. At low currents, a few
particles nucleate and grow, while at high currents (i.e., greater nucleation tendency)
much larger number of particles are found with a broad particle size distribution. Early
time images of high current electrodeposition reveal smaller particles, i.e., as current
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

increases, a greater number of nucleation events take place leading to a higher number
density, while simultaneously experiencing growth to a smaller extent 10. This interplay
among electrodeposition current, nucleation site density, and extent of growth is fairly
equivalent to Li2S formation (possibly hinting universality for electrodepositing
insulating species). More interestingly, geometrical aspects of particle shape (Figure
8(b)) exhibit a monotonic dependence on both the electrodeposition currents and
capacities (Figure 8(a)).
A careful inspection of geometrical statistics (Figure 8(a)) and shape evolution
(Figure 8(b)) disclose that the initial particle growth is due to the stacking of Li2O2 disks
(crystallites). Upon further electrodeposition, lateral growth takes place, such that the
existing disks splay apart (Figure 8(c)) and additional plates nucleate in these crevices.
The resultant shape gradually transitions to toroidal. Despite the macroscopically
unfaceted shape, plates are found to be single crystallites through diffraction tests10.
Moreover, they consistently stack along the same crystal orientation, marking a layer-by-
layer growth. The elemental plate geometry corresponds to a compact surface energy
shape, i.e., the Wulff shape. Such observations underscore the dominance of interfacial
energies in driving electrodeposition geometry.

Figure 8. Mechanism of Li2O2 electrodeposition analyzed under galvanostatic conditions 10: (a) aspect ratio of Li2O2
particles at different currents; (b) SEM images for electrodeposition at (i) 10 mA/g C to 500 mAh/gC (ii) 100 mA/gC to 14
000 mAh/gC (iii) 10 mA/gC to 1100 mAh/gC; normalized to weight of carbon substrate, gC; (c) schematic of particle
growth sequence. (reproduced with permission from American Chemical Society10)

2854
Li2O2 also suffers from an electronically insulating nature 22, which is expected
to inhibit electrochemical growth when deposit thickness becomes comparable to
tunneling length of the electrons (further reaction cannot be sustained as electrons cannot
reach reaction sites). Following this chain of thought, the complex directional growth of
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

Li2O2 shapes as in Figure 8(b) appears somewhat mysterious. It has recently been
argued23 that such a nonplanar morphology is a consequence of multi-step
electrodeposition that progresses via solvated species. Put simply, there are two reaction
pathways23 (i) a surface mechanism that necessitates a simultaneous availability of ions
and electrons at the electrode-electrolyte interface, and (ii) a solution mechanism that
splits the overall electrodeposition into multiple steps such that electrochemical step and
physical growth are two separate reactions. The former mechanism produces insulating
films, whose thickness is limited by tunneling length for charge carriers, be it electrons
or hole polarons24, while the latter producing (macroscopically) three-dimensional
growth.

Electrodeposition with Solid Electrolytes

Aforesaid reports detail different instants of electrodeposition in liquid electrolytes.


Recent scientific interest in solid-state electrolytes is attributed to unique aspects of such
ionic conductors, specifically stiffness and selective ionic transport (ionic current is
composed of charges present in the crystal structure). Arguably, mechanical rigidity of
solid electrolyte could exert a regularizing influence over electrodeposition. Intuitively,
one would expect a critical stiffness of the electrolytes that generates enough mechanical
stresses to counter the inhomogeneous reaction distribution of a perturbed interface
(generated stresses scale positively with stiffness). Such a condition is theoretically
identified by Monroe and Newman25. However, experiments with multiple solid
electrolytes12,26,27, expected to inhibit irregular growth, report characteristic lithium
deposition patterns (e.g., Figure 9(a)-(c)). It is counterintuitive for soft lithium metal to
penetrate through much stiffer solids, let alone form ordered patterns (Figure 9(b),(c);
softer organic solid electrolytes experience irregular growth jointly due to mechanical
compliance and mass transport limitation 28–30, and are not discussed here).
Experiments (e.g.,12) with inorganic solid electrolytes sandwiched between
lithium (film) electrodes demonstrate the existence of two representative currents
ƒ a critical current density beyond which voltage growth deviates from linear
ƒ a maximum current density when short circuit takes place due to lithium filament
propagation across the electrolyte
The solid electrolytes are desired to behave as ohmic (ionic) resistors exhibiting a linear
voltage-current relation, however, beyond the critical current density, such linearity is
disfigured. When as operated LLZO is fractured to identify the morphology of lithium
deposits, SEM imaging reveals (Figure 9(b)) web-like lithium deposits along the grain
boundaries. The short-circuiting filaments propagate predominantly along the grain
boundaries. Other inorganic solid electrolytes also exhibit equivalent macroscopic
signatures and peculiar deposition patterns (e.g., Figure 9(c)26,27).

2855
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

Figure 9. Electrodeposition patterns in inorganic solid electrolytes: (a) schematic of intergranular filament propagation in
polycrystalline LLZO12; (b) SEM of the fractured surface reveals hexagonal lithium deposits; (c) transmission optical
microscopy of branched lithium growth in a polycrystalline LPS26; (d)-(f) schematic of Li2S electrodeposition with a
solid electrolyte; (g) spatial distribution of solid phases 31 as imaged via electron energy loss spectroscopy (EELS); (h)
material interactions in an equivalent solid-state intercalation electrode are devoid of moving interfaces and phase growth
dynamics. LLZO ≡ Li6.25Al0.25La3Zr2O12; LPS ≡ β-Li3PS4; AB ≡ Acetylene Black; SE ≡ Solid Electrolyte. (reproduced
with permissions from Elsevier12, Wiley-VCH26 and Royal Society of Chemistry31)

The mechanistic underpinnings of such electrodeposition patterns remain


elusive, in part due to the translucent nature of solid electrolytes. Even ex situ
nondestructive imaging is challenging. Two sets of arguments, preexisting flaws serving
as crack seeds26, and mechanical and ionic mismatch of grain – grain boundary
properties32, appear to be promising. However, neither provides a generalized
interpretation of electrodeposition patterning in solid electrolytes. The additional
dilemma stems from the fact that most of the inorganic solid electrolytes are single ion
conductors where cations are embedded in a negatively charged backbone. This atomic-
scale arrangement is quite different from liquid electrolytes where both anion and cation
are mobile. Such a structural difference, along with local charge neutrality, refutes the
classical mass transport limited instability. The stoichiometry of the atomic constituents
at the nanoscopic scale of repeating units (of the crystal structure) enforces the
composition of these solid electrolytes. In turn, unlike liquid electrolytes, charge carriers
and their local concentrations are predefined based on the crystal structure, and transport
of short-lived species is inhibited.
The use of solid electrolytes in Li-sulfur is primarily motivated by this
advantage31. It could be argued that the presence of only cationic charge carriers,
prohibits other ions to pass through. Hence, such electrolytes could be used to limit the
unwanted migration of anionic polysulfides. However, in addition to the characteristic
peculiarities of the insulating Li2S (pointed out earlier), a change from liquid to solid
electrolyte (i) alters the reaction pathway, bypassing the solution phase ionic species, to
electrodeposit Li2S at the expense of solid sulfur, S8, and (ii) volumetric difference
between S8 and Li2S causes stress inhomogeneities (on the other hand, liquid electrolyte
displaces to locally accommodate this volume change). The solid-state reaction pathway

2856
is schematically shown in Figure 9(d)-(f). Note that an intimate contact among
electronically conducting carbon (or other substrate material), ionically conducting solid
electrolyte and solid reactants is a prerequisite to facilitating the electrodeposition. Figure
9(g) records an exemplar phase distribution in a solid-state Li-sulfur electrode31. The
continuously evolving stress field due to the volumetric mismatch between S8 and Li2S is
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

to be relaxed to acclimatize the freshly grown phase. However, resultant geometrical


changes distort the three phase-contact necessary for reaction progression. Mechanistic
steps elucidating the geometrical evolution during electrodeposition of Li2S, or growth of
S8 when reoxidized, remain unexplained. For example, whether the geometrical
evolution is localized or comparatively long-ranged (to relax stress inhomogeneity) is
unclear. Notice that this complex interplay between reaction and mechanics becomes
much critical for insulator electrodeposition, given the need to ensure three-phase
contact. Metallic electrodeposition (Figure 9(a)-(c)) is comparatively simpler as only a
two-phase contact is required. On the other hand, intercalation reaction at the interface of
solid electrolyte and cathode material (Figure 9(h)) does not have to accommodate such a
high volume change (despite minimal volumetric changes, such contact does experience
severe mechanical issues33).

A discussion of unknowns

Electrodeposition reactions hold a great (theoretical) promise for futuristic energy


storage, however, characteristic uncontrollability plagues practical implementation.
Engineering specifications for an application are rather well-articulated: fast, reversible
and safe growth. A common source of ambiguity lies in its (relatively distant) analogy
with solidification, which only holds for the mass transport limited regime. An elemental
difference between the two is the reactive nature of the electrodepositing interface that
actively contributes to the evolution of solid-electrolyte interface (solidification or mass
transport limited electrodeposition is governed by a flux mismatch at the interface, heat
and diffusive fluxes, respectively). The short-range nature of these prominent
interactions remains marginally probed. Incomplete mechanistic understanding makes it
difficult to identify and characterize the material properties that govern the geometry and
time evolution of the electrodeposited phase.
The definition of unstable growth needs to be revisited. Macroscopic
interpretation qualifies nonuniform growth as a signature of electrodeposition instability.
As exemplified by preceding discussion, nonconformity is more likely a geometrical
attribute of electrodeposition at the small microscopic (but larger than atomic/molecular)
scales. As a matter of fact, while depositing a low conductivity phase, the active concern
is sustaining the three-dimensionality, e.g., 34.
We envisage the following questions to facilitate the much needed exposition.

Short-time dynamics

nucleation and confined growth represent the early time transients that are characteristic
of substrate and solution (solution ≡ electrolyte + additive) combination 8,17. How the
choice of substrate and solution that can rationally guide the short-time dynamics is
unclear.

Reversibility

macroscopically, reversibility of electrodeposition and dissolution operation is quantified


in terms of coulombic efficiency and capacity decay, however, it is a consequence of

2857
various elemental factors such as side reaction (e.g., SEI formation), irregular dissolution
(often results in dead lithium7) and possibly inherent semi-reversibility of the reaction
kinetics.
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

Ionic environment

the presence of non-plating cation19, anion20 and neutral species (e.g., water23) is found to
distinguishably alter deposition geometry. These species can alter the reaction pathway
(a chemical interaction) or alter the transport of species to reaction locations (a physical
interaction). In the presence of multiple reactions involving different species, the
chemical composition of the electrolyte continuously evolves and it can negatively alter
the transport effectiveness of the electrolyte 35 (note that the mass-transport limitations
relate to electrolyte transport).

Structuring

geometrical features such as an array of channels 18 or porous hosts are found to


considerably alter deposition uniformity with often unexpected results when extended to
seemingly equivalent geometries.

Insulating deposition

electrochemical formation of insulating species is either explained through tunneling


charges at nanoscale24 or surface diffusion11 of nanoscopic deposits. Tunneling alone
cannot account for nonplanar morphologies 10, while diffusion, microscopically a random
and entropy increasing influence, cannot cause directional aggregation.

Redox mediator

effectively splits electrodeposition into electron transfer at the interface and chemical
deposition, i.e., precipitation of the insulating phase 34,36. Correlation among its chemical
activity, physical transport, and deposition morphology is unexplained.

Nonequilibrium growth

very slow (near-equilibrium) phase growth results in equilibrium shapes (i.e., defined by
Wulff constructions10). A fast deposition may not allow enough time for a local
rearrangement towards the equilibrium shape. This intrinsic relation among deposition
rate, relaxation timescale, and irregularity, if any, needs to be fingerprinted.

Conjugate fields

mechanical stress and temperature alter deposition geometry, but the results are often
counterintuitive12. Mechanical effects are not only limited to solid electrolytes, the
incompressible nature of liquid electrolytes indirectly causes stresses when reaction
volumes differ (e.g., Li-sulfur) or deposition is irregular even in a fixed cell volume.

2858
Statistical attributes

the randomness of the small-scale interactions limits predictability. Relevant causes are
thermal fluctuation of molecular transport, discrete nature of reactions and statistics of
nucleation sites. Such intrinsic stochasticity can lead to varying geometrical evolutions
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

for the same macroscopic measurement constraints. The repeatability of experiments is


to be redefined. Since resolving spatial information is resource-intensive, stochastic
characteristics are a daunting burden.

Beyond lithium chemistries (zinc 37,38, calcium39,40, magnesium38) also attempt to


leverage electrodeposition. Herein active cations are multivalent ions, and equivalently
their reduction could be a multi-step electrochemical reaction. Additionally, the
multivalency promotes the formation of ion-pairs in the electrolyte41,42 and decreases the
availability of cations for electrochemical growth.

Conclusions

The functionality of electrodeposition as an energy storage mode implicitly relates to the


extent of control afforded at the (small) length scales relevant for batteries. We examine
reported spatiotemporal features for different electrodeposition reactions and identify the
lack of universality or mechanistic principles that could explain the peculiar occurrences.
The insufficient probing of the short-range interactions is identified to cause the
ambiguity.

Acknowledgments

The authors gratefully acknowledge support from the U. S. Department of Energy


(DOE), Vehicle Technologies Office. Argonne National Laboratory is operated for DOE
Office of Science by UChicago Argonne, LLC under contract number DE-AC02-
06CH11357.

References

(1) Bruce, P. G.; Freunberger, S. A.; Hardwick, L. J.; Tarascon, J.-M. Li-O2 and Li-S Batteries
with High Energy Storage. Nat. Mater. 2012, 11 (1), 19–29.
(2) Bai, P.; Li, J.; Brushett, F. R.; Bazant, M. Z. Transition of Lithium Growth Mechanisms in
Liquid Electrolytes. Energy Environ. Sci. 2016, 9 (10), 3221–3229.
https://doi.org/10.1039/C6EE01674J.
(3) Chen, K. H.; Wood, K. N.; Kazyak, E.; Lepage, W. S.; Davis, A. L.; Sanchez, A. J.; Dasgupta,
N. P. Dead Lithium: Mass Transport Effects on Voltage, Capacity, and Failure of Lithium
Metal Anodes. J. Mater. Chem. A 2017. https://doi.org/10.1039/c7ta00371d.
(4) Mullins, W. W.; Sekerka, R. F. Stability of a Planar Interface during Solidification of a Dilute
Binary Alloy. J. Appl. Phys. 1964. https://doi.org/10.1063/1.1713333.
(5) Sundström, L. G.; Bark, F. H. On Morphological Instability during Electrodeposition with a
Stagnant Binary Electrolyte. Electrochim. Acta 1995. https://doi.org/10.1016/0013-
4686(94)00379-F.
(6) Khoo, E.; Zhao, H.; Bazant, M. Z. Linear Stability Analysis of Transient Electrodeposition in
Charged Porous Media: Suppression of Dendritic Growth by Surface Conduction. J.
Electrochem. Soc. 2019. https://doi.org/10.1149/2.1521910jes.

2859
(7) Lu, D.; Shao, Y.; Lozano, T.; Bennett, W. D.; Graff, G. L.; Polzin, B.; Zhang, J.; Engelhard,
M. H.; Saenz, N. T.; Henderson, W. A. Failure Mechanism for FastǦ charged Lithium Metal
Batteries with Liquid Electrolytes. Adv. Energy Mater. 2015, 5 (3).
(8) Shi, F.; Pei, A.; Vailionis, A.; Xie, J.; Liu, B.; Zhao, J.; Gong, Y.; Cui, Y. Strong Texturing of
Lithium Metal in Batteries. Proc. Natl. Acad. Sci. U. S. A. 2017.
https://doi.org/10.1073/pnas.1708224114.
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

(9) Sun, F.; Osenberg, M.; Dong, K.; Zhou, D.; Hilger, A.; Jafta, C. J.; Risse, S.; Lu, Y.;
Markötter, H.; Manke, I. Correlating Morphological Evolution of Li Electrodes with
Degrading Electrochemical Performance of Li/LiCoO 2 and Li/S Battery Systems:
Investigated by Synchrotron X-Ray Phase Contrast Tomography. ACS Energy Lett. 2018.
https://doi.org/10.1021/acsenergylett.7b01254.
(10) Mitchell, R. R.; Gallant, B. M.; Shao-Horn, Y.; Thompson, C. V. Mechanisms of
Morphological Evolution of Li2O2 Particles during Electrochemical Growth. J. Phys. Chem.
Lett. 2013. https://doi.org/10.1021/jz4003586.
(11) Fan, F. Y.; Carter, W. C.; Chiang, Y. Mechanism and Kinetics of Li2S Precipitation in
Lithium–Sulfur Batteries. Adv. Mater. 2015, 27 (35), 5203–5209.
(12) Cheng, E. J.; Sharafi, A.; Sakamoto, J. Intergranular Li Metal Propagation through
Polycrystalline Li6.25Al0.25La3Zr2O12 Ceramic Electrolyte. Electrochim. Acta 2017.
https://doi.org/10.1016/j.electacta.2016.12.018.
(13) Bieker, G.; Winter, M.; Bieker, P. Electrochemical in Situ Investigations of SEI and Dendrite
Formation on the Lithium Metal Anode. Phys. Chem. Chem. Phys. 2015, 17 (14), 8670–8679.
(14) Léger, C.; Elezgaray, J.; Argoul, F. Dynamical Characterization of One-Dimensional
Stationary Growth Regimes in Diffusion-Limited Electrodeposition Processes. Phys. Rev. E -
Stat. Physics, Plasmas, Fluids, Relat. Interdiscip. Top. 1998.
https://doi.org/10.1103/PhysRevE.58.7700.
(15) Monroe, C.; Newman, J. Dendrite Growth in Lithium/Polymer Systems a Propagation Model
for Liquid Electrolytes under Galvanostatic Conditions. J. Electrochem. Soc. 2003, 150 (10),
A1377–A1384.
(16) Mistry, A.; Fear, C.; Carter, R.; Love, C. T.; Mukherjee, P. P. Electrolyte Confinement Alters
Lithium Electrodeposition. ACS Energy Lett. 2019, 4 (1).
https://doi.org/10.1021/acsenergylett.8b02003.
(17) Pei, A.; Zheng, G.; Shi, F.; Li, Y.; Cui, Y. Nanoscale Nucleation and Growth of
Electrodeposited Lithium Metal. Nano Lett. 2017.
https://doi.org/10.1021/acs.nanolett.6b04755.
(18) Liu, W.; Lin, D.; Pei, A.; Cui, Y. Stabilizing Lithium Metal Anodes by Uniform Li-Ion Flux
Distribution in Nanochannel Confinement. J. Am. Chem. Soc. 2016.
https://doi.org/10.1021/jacs.6b08730.
(19) Ding, F.; Xu, W.; Graff, G. L.; Zhang, J.; Sushko, M. L.; Chen, X.; Shao, Y.; Engelhard, M.
H.; Nie, Z.; Xiao, J. Dendrite-Free Lithium Deposition via Self-Healing Electrostatic Shield
Mechanism. J. Am. Chem. Soc. 2013, 135 (11), 4450–4456.
(20) Li, W.; Yao, H.; Yan, K.; Zheng, G.; Liang, Z.; Chiang, Y. M.; Cui, Y. The Synergetic Effect
of Lithium Polysulfide and Lithium Nitrate to Prevent Lithium Dendrite Growth. Nat.
Commun. 2015. https://doi.org/10.1038/ncomms8436.
(21) Fan, F. Y.; Chiang, Y.-M. Electrodeposition Kinetics in Li-S Batteries: Effects of Low
Electrolyte/Sulfur Ratios and Deposition Surface Composition. J. Electrochem. Soc. 2017, 164
(4), A917–A922.
(22) Viswanathan, V.; Thygesen, K. S.; Hummelshj, J. S.; Nrskov, J. K.; Girishkumar, G.;
McCloskey, B. D.; Luntz, A. C. Electrical Conductivity in Li 2O 2 and Its Role in
Determining Capacity Limitations in Non-Aqueous Li-O 2 Batteries. J. Chem. Phys. 2011.
https://doi.org/10.1063/1.3663385.
(23) Aetukuri, N. B.; McCloskey, B. D.; Garciá, J. M.; Krupp, L. E.; Viswanathan, V.; Luntz, A. C.
Solvating Additives Drive Solution-Mediated Electrochemistry and Enhance Toroid Growth
in Non-Aqueous Li-O2 Batteries. Nat. Chem. 2015. https://doi.org/10.1038/nchem.2132.
(24) Radin, M. D.; Monroe, C. W.; Siegel, D. J. Impact of Space-Charge Layers on Sudden Death
in Li/O<inf>2</Inf> Batteries. J. Phys. Chem. Lett. 2015.
https://doi.org/10.1021/acs.jpclett.5b01015.
(25) Monroe, C.; Newman, J. The Impact of Elastic Deformation on Deposition Kinetics at
Lithium/Polymer Interfaces. J. Electrochem. Soc. 2005, 152 (2), A396–A404.
(26) Porz, L.; Swamy, T.; Sheldon, B. W.; Rettenwander, D.; Frömling, T.; Thaman, H. L.;
Berendts, S.; Uecker, R.; Carter, W. C.; Chiang, Y. M. Mechanism of Lithium Metal
Penetration through Inorganic Solid Electrolytes. Adv. Energy Mater. 2017.
https://doi.org/10.1002/aenm.201701003.

2860
(27) Lewis, J. A.; Cortes, F. J. Q.; Boebinger, M. G.; Tippens, J.; Marchese, T. S.; Kondekar, N.;
Liu, X.; Chi, M.; McDowell, M. T. Interphase Morphology between a Solid-State Electrolyte
and Lithium Controls Cell Failure. ACS Energy Lett. 2019.
https://doi.org/10.1021/acsenergylett.9b00093.
(28) Barai, P.; Higa, K.; Srinivasan, V. Lithium Dendrite Growth Mechanisms in Polymer
Electrolytes and Prevention Strategies. Phys. Chem. Chem. Phys. 2017.
Downloaded from https://www.cambridge.org/core, on subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/adv.2019.443

https://doi.org/10.1039/c7cp03304d.
(29) Maslyn, J. A.; Loo, W. S.; McEntush, K. D.; Oh, H. J.; Harry, K. J.; Parkinson, D. Y.; Balsara,
N. P. Growth of Lithium Dendrites and Globules through a Solid Block Copolymer
Electrolyte as a Function of Current Density. J. Phys. Chem. C 2018.
https://doi.org/10.1021/acs.jpcc.8b06355.
(30) Gribble, D. A.; Frenck, L.; Shah, D. B.; Maslyn, J. A.; Loo, W. S.; Mongcopa, K. I. S.; Pesko,
D. M.; Balsara, N. P. Comparing Experimental Measurements of Limiting Current in Polymer
Electrolytes with Theoretical Predictions. J. Electrochem. Soc. 2019.
https://doi.org/10.1149/2.0391914jes.
(31) Nagao, M.; Hayashi, A.; Tatsumisago, M. High-Capacity Li 2S-Nanocarbon Composite
Electrode for All-Solid-State Rechargeable Lithium Batteries. J. Mater. Chem. 2012.
https://doi.org/10.1039/c2jm16802b.
(32) Barai, P.; Higa, K.; Ngo, A. T.; Curtiss, L. A.; Srinivasan, V. Mechanical Stress Induced
Current Focusing and Fracture in Grain Boundaries. J. Electrochem. Soc. 2019.
https://doi.org/10.1149/2.0321910jes.
(33) Koerver, R.; Aygün, I.; Leichtweiß, T.; Dietrich, C.; Zhang, W.; Binder, J. O.; Hartmann, P.;
Zeier, W. G.; Janek, J. Capacity Fade in Solid-State Batteries: Interphase Formation and
Chemomechanical Processes in Nickel-Rich Layered Oxide Cathodes and Lithium
Thiophosphate Solid Electrolytes. Chem. Mater. 2017.
https://doi.org/10.1021/acs.chemmater.7b00931.
(34) Gerber, L. C. H.; Frischmann, P. D.; Fan, F. Y.; Doris, S. E.; Qu, X.; Scheuermann, A. M.;
Persson, K.; Chiang, Y. M.; Helms, B. A. Three-Dimensional Growth of Li2S in Lithium-
Sulfur Batteries Promoted by a Redox Mediator. Nano Lett. 2016.
https://doi.org/10.1021/acs.nanolett.5b04189.
(35) Mistry, A. N.; Mukherjee, P. P. Electrolyte Transport Evolution Dynamics in Lithium-Sulfur
Batteries. J. Phys. Chem. C 2018, 122 (32). https://doi.org/10.1021/acs.jpcc.8b05442.
(36) Bergner, B. J.; Hofmann, C.; Schürmann, A.; Schröder, D.; Peppler, K.; Schreiner, P. R.;
Janek, J. Understanding the Fundamentals of Redox Mediators in Li-O2 Batteries: A Case
Study on Nitroxides. Phys. Chem. Chem. Phys. 2015. https://doi.org/10.1039/c5cp04505c.
(37) Parker, J. F.; Chervin, C. N.; Pala, I. R.; Machler, M.; Burz, M. F.; Long, J. W.; Rolison, D. R.
Rechargeable Nickel-3D Zinc Batteries: An Energy-Dense, Safer Alternative to Lithium-Ion.
Science (80-. ). 2017. https://doi.org/10.1126/science.aak9991.
(38) Ta, K.; See, K. A.; Gewirth, A. A. Elucidating Zn and Mg Electrodeposition Mechanisms in
Nonaqueous Electrolytes for Next-Generation Metal Batteries. J. Phys. Chem. C 2018.
https://doi.org/10.1021/acs.jpcc.8b00835.
(39) Wang, D.; Gao, X.; Chen, Y.; Jin, L.; Kuss, C.; Bruce, P. G. Plating and Stripping Calcium in
an Organic Electrolyte. Nat. Mater. 2017, 17, 16.
(40) Ta, K.; Zhang, R.; Shin, M.; Rooney, R. T.; Neumann, E. K.; Gewirth, A. A. Understanding
Ca Electrodeposition and Speciation Processes in Nonaqueous Electrolytes for Next-
Generation Ca-Ion Batteries. ACS Appl. Mater. Interfaces 2019.
https://doi.org/10.1021/acsami.9b04926.
(41) Rajput, N. N.; Qu, X.; Sa, N.; Burrell, A. K.; Persson, K. A. The Coupling between Stability
and Ion Pair Formation in Magnesium Electrolytes from First-Principles Quantum Mechanics
and Classical Molecular Dynamics. J. Am. Chem. Soc. 2015.
https://doi.org/10.1021/jacs.5b01004.
(42) Samuel, D.; Steinhauser, C.; Smith, J. G.; Kaufman, A.; Radin, M. D.; Naruse, J.; Hiramatsu,
H.; Siegel, D. J. Ion Pairing and Diffusion in Magnesium Electrolytes Based on Magnesium
Borohydride. ACS Appl. Mater. Interfaces 2017. https://doi.org/10.1021/acsami.7b15547.

2861

You might also like