You are on page 1of 18

Review

www.advenergymat.de

Anode-Free Full Cells: A Pathway to High-Energy Density


Lithium-Metal Batteries
Sanjay Nanda, Abhay Gupta, and Arumugam Manthiram*

layered-oxide cathodes to improve capaci-


The development of high-energy density batteries is critical to the decarboni- ties.[5–7] In contrast, there is still substan-
zation of the transportation and power generation sectors. For any given tial unfulfilled potential for increasing the
lithium-containing cathode system, the anode-free full cell configuration, energy density at the anode. The incum-
bent graphite anode, with a theoretical
which eliminates excess lithium and pairs the fully lithiated cathode with a
capacity of only 372 mAh g−1, has been
bare current collector, can deliver the maximum possible energy density. The used mostly unchanged since the intro-
absence of free lithium metal during cell assembly confers significant practical duction of commercial Li-ion batteries
advantages as well. It is also the ideal framework for developing a thorough by Sony Corporation in 1991.[8–10] Some
understanding of lithium deposition in conjunction with various cathode improvements in capacity have been deliv-
systems. However, the poor efficiencies of lithium plating and stripping ered by the incorporation of silicon into
graphite anodes.[11,12] However, lithium-
lead to rapid lithium inventory loss and poor cycle life. In the last few years, metal anodes show the greatest potential
multiple studies have demonstrated the application of advanced electrolytes, for enabling a quantum leap in the energy
modified current collectors, and optimized formation and cycling parameters density of lithium batteries. It is con-
to stabilize lithium deposition and improve cycle life (80% capacity retention) sidered the ideal anode material due to
to 100 cycles and beyond. This review provides an overview of the various having the lowest redox potential (-3.04 V
vs SHE) and the highest theoretical
strategies toward sustaining lithium inventory in anode-free full cells and sum-
capacity (3860 mAh g−1, 2061 mAh cm−3)
marizes the work undertaken in this nascent field. It is expected that further of all possible anode materials.[13–15] It is
improvement upon these strategies and a combinatorial approach can enable also indispensable for futuristic cathodes
cycle lives far in excess of what has been achieved so far. such as sulfur and oxygen/air. Despite its
advantages, the practical implementation
of lithium-metal anodes has been hin-
dered by the poor reversibility of lithium
1. Introduction plating and stripping.[16,17] The formation of high surface area
“mossy” deposits of lithium leads to its depletion by undesir-
High-energy-density rechargeable batteries have emerged as able side reactions.[18] This can be compensated for by using
one of the most important technologies of the 21st century, excess lithium, but that negates the energy density advantage
indispensable for a wide range of applications ranging from of lithium-metal anodes. Over the last few years, there have
miniaturized electronics to all-electric vehicles to grid storage. been considerable efforts by researchers to stabilize lithium
Many fundamental developments in solid-state chemistry deposition and improve the cyclability of lithium-metal
and physics led to the development of Li-ion batteries, which anodes.[19–21]
are now pre-eminent in most application spaces.[1] Commer- An intriguing and under-explored possibility for realizing lith-
cial Li-ion batteries have a typical energy density of less than ium-metal batteries is the anode-free full cell configuration.[22–25]
250 Wh kg−1 at the cell level.[2] Increasing the energy density of By pairing a fully lithiated cathode with a bare current collector
the batteries, to say 500 Wh kg−1, would enable one or a com- on the anode side, the anode to cathode capacity ratio can be
bination of increased device run-time (e.g., driving range) or kept exactly equal to 1. This enables the maximum possible
decreased device footprint.[3,4] Steady gains in energy density of energy density to be delivered from any given cathode system.
up to 40% have been realized by increasing the nickel content of Since there is no excess lithium to artificially inflate cyclability,
the electrochemical performance of anode-free full cells is lim-
ited almost entirely by the efficiency of lithium plating and
S. Nanda, A. Gupta, Prof. A. Manthiram stripping. This allows a more realistic assessment of the electro-
Materials Science and Engineering Program & Texas Materials Institute chemical performance of lithium-metal batteries. Various strat-
The University of Texas at Austin egies for stabilizing lithium deposition can also be effectively
Austin, TX 78712, USA evaluated using the anode-free configuration. In the last few
E-mail: manth@austin.utexas.edu
years, a number of groups have reported a variety of strategies
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/aenm.202000804.
for improving the cyclability of anode-free batteries, including
novel electrolyte formulations, artificial SEI layers, modified
DOI: 10.1002/aenm.202000804 current collector substrates, favorable formation procedures,

Adv. Energy Mater. 2020, 2000804 2000804 (1 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

and optimized cell testing protocols. This review provides an


overview of the different approaches reported in the literature Sanjay Nanda is currently a
and concludes with an outlook for this nascent and promising PhD candidate in Materials
field. Science and Engineering at
the University of Texas at
Austin. He received his B.Sc.
(2013) degree in Physics
2. Anode-Free Full Cells
from the Indian Institute
The anode-free full cell configuration promises a transforma- of Technology, Kanpur, and
tive enhancement in energy density to 500 Wh kg−1 and his M.Phil (2014) degree in
beyond. Li-ion batteries are typically assembled in the Cu/C6/ Nanotechnology from the
PP/LiXO/Al configuration, where Cu and Al are the current University of Cambridge.
collectors, C6 is the graphite anode, LiXO is the lithiated oxide His research is focused
cathode (LCO, NMC, NCA, LMO, LFP), and PP is the polymer on devising strategies for stabilizing lithium deposi-
separator. Anode-free full cells can be visualized as the same tion in lithium–sulfur batteries and understanding the
configuration as above, but with the C6 graphite anode com- degradation mechanisms that underlie the failure of
pletely eliminated to yield the Cu/PP/LiXO/Al configuration. lithium-metal anodes.
Other unconventional lithiated cathodes such as Li2S, LiI, and
LiMS2, where M is a transition metal, can also be employed. Abhay Gupta is a PhD
Assembled in the discharged state, Li+ ions from the cathode student in Materials Science
in the anode-free full cells are reduced and plated as lithium and Engineering at the
(Li0) metal on the Cu foil current collector during charge. Sub- University of Texas at Austin.
sequently, the plated lithium metal is stripped away during He received his B.S. from the
discharge. In its simplest form, the absence of a host to accom- Hildebrand Department of
modate the lithium deposited on the current collector make Petroleum and Geosystems
it a “hostless” anode. In one embodiment of a Li-ion battery, Engineering at the University
the removal of the graphite host, which occupies 27.8% of the of Texas at Austin in 2016. His
overall stack weight and 46.1% of the overall stack thickness, research primarily focuses
would boost gravimetric energy density by 38.5% and volu- on investigating solution-
metric energy density by 85.5% (at the stack level).[26] Figure 1 coordination behavior in
demonstrates the differences in cell configuration and the lithium–sulfur batteries, with an emphasis on low-tempera-
corresponding increases in both gravimetric and volumetric ture performance.
energy density.
It should be noted here that there are large volume changes in Arumugam Manthiram is
the cell stack associated with charge and discharge of an anode- the Cockrell Family Regents
free full cell. Even accounting for the ≈20% volume expansion Chair in Engineering and the
of the stack with the plating of lithium and assuming negligible Director of Texas Materials
volume contraction at the cathode, the volumetric energy den- Institute and Materials
sity is still boosted by 57.1% compared to a Li-ion battery with a Science and Engineering
graphite anode. Nevertheless, these volume changes will have Program at the University of
to be accommodated by appropriate cell or battery pack designs. Texas at Austin. His research
In regular full cells, an increase in cathode capacity due to interests are in the area
either increased loading or higher electrochemical utilization of batteries and fuel cells,
has to be accompanied by a corresponding increase in anode including new materials
capacity, typically by an increased loading of the anode active development, novel synthesis
material. This requirement is obviated in the anode-free full approaches, and structure–composition–performance
cell configuration. Furthermore, the negative to positive elec- relationships.
trode capacity (N/P) ratio, defined for anode-free full cells as the
ratio of lithium inventory to cathode inventory, is exactly equal
to 1. The need to use a slight excess of anode-active material
to prevent undesirable lithium plating is also obviated. Finally, Here, X can be any lithiated or nonlithiated cathode. While
the use of any anode material other than lithium metal would the use of lithium foils is required for nonlithiated cathodes
lower the voltage output of the cell due to the potential asso- such as sulfur, fully lithiated cathodes already carry the stoi-
ciated with lithiation of the anode. The anode-free configura- chiometric amount of lithium necessary for a complete charge/
tion ensures the highest possible voltage output for any cathode discharge cycle. In the example system shown in Figure 1 with
system. All of these factors combine to engender a dramatic a fully lithiated NCA cathode of capacity 6.1 mAh cm−2 and
increase in energy density with anode-free batteries. a N/P ratio of 3, the lithium-metal foil anode would have a
Lithium metal batteries are typically envisioned as employing capacity of 12.2 mAh cm−2, corresponding to a lithium-metal
thin lithium foil anodes, in the Cu/Li/PP/X/Al configuration. foil of thickness 59.2 µm and weight 3.16 mg cm−2. However,

Adv. Energy Mater. 2020, 2000804 2000804 (2 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

Figure 1. a) Schematic illustration of an example lithium-ion battery using an NCA cathode of thickness 100 µm and a loading 30.5 mg cm−2, with
the thickness of the various components drawn to scale. b) Schematic illustration of an anode-free full cell using an equivalent NCA cathode, in both
charged and discharged states. The difference in energy density can be clearly visualized. There is a ≈20% volume change associated with the plating
and stripping of lithium, which will have to be accommodated. c) Weight distribution of each stack component in a Li-ion battery, with the graphite
anode occupying 27.8% of the stack weight. d) Weight of each stack component in an equivalent anode-free full cell. The volume and weight distribu-
tions of the example lithium-ion battery were obtained from a 2018 work by Betz et al.[26]

manufacturing thin lithium foils (<100 µm thickness) over graphite anodes with minimal loss over thousands of cycles, the
large areas with acceptable control over chemical purity and same cannot be said for the plating and stripping of lithium.
microstructure has proven exceedingly challenging and expen- In liquid electrolytes, lithium metal undergoes two growth
sive.[27,28] Handling thin lithium foils is also difficult due to mechanisms during deposition: i) mossy and ii) dendritic.[18]
the lack of mechanical robustness and its high reactivity with The initial “mossy” growth is attributed to a nonuniform flux
moisture and air. Hence, the process of coating the surface of of Li+ ions on the current collector and leads to the forma-
lithium metal with passivating layers or stabilizing interphases tion of a large number of entangled thin filament-like lithium
is quite nontrivial. All of these problems are precluded with the microstructures. The highly reducing nature of lithium metal
use of the anode-free full cell configuration due to the absence causes the electrolyte to decompose on the lithium surface to
of any free lithium metal. The minimal and trivial changes in form a solid-electrolyte interphase (SEI). The large surface area
cell architecture when compared to existing Li-ion batteries of the porous mossy lithium deposits intensifies these para-
means that the production of anode-free full cells can be readily sitic side reactions. The insulating SEI layer is nominally pas-
integrated into existing battery manufacturing infrastructure sivating to further side reactions as it blocks electron transfer
with minimal disruption. Since anode-free full cells are assem- from metallic lithium to the electrolyte. However, since lithium
bled in the fully discharged condition, they are at a lower-energy metal is a “hostless” anode, the plating and stripping of lithium
state compared to the charged condition. Hence, they are inher- is accompanied by very large volume changes, which disinte-
ently safer to handle. Furthermore, since the cathode is fully grates the SEI layer and leads to continued decomposition of
lithiated, any further reaction in open-circuit conditions is the electrolyte on the lithium surface. Over many cycles, a thick
nonspontaneous (Gibbs free energy ΔG° > 0). Hence, the pos- layer of electrolyte decomposition products, which are typically
sibility of self-discharge, which is an internal energy-lowering poor Li+ ion conductors, is formed. This blocks electron and
mechanism, is precluded in as-assembled anode-free full cells. Li+-ion transfer to the unreacted metallic lithium and renders
This is a particular advantage over systems assembled in the it electrochemically inaccessible or “dead.”[29] The combina-
charged condition that show considerable self-discharge, such tion of parasitic side reactions and formation of “dead” lithium
as conventional lithium–sulfur batteries. leads to the irreversible loss of both the lithium inventory and
electrolyte supply. When the electrolyte salt concentration falls
to zero at the lithium surface and the applied current exceeds
3. Capacity Loss in Anode-Free Full Cells diffusion-based transport limitations, rapid dendritic growth
occurs.[30] This further intensifies parasitic side reactions with
The advantages of enhanced energy density and design sim- the electrolyte and can cause internal short circuits, which
plicity with anode-free full cells come with a major disadvan- increases the risk of catastrophic cell failure along with the
tage. While lithium can be inserted into and removed from associated safety concerns.

Adv. Energy Mater. 2020, 2000804 2000804 (3 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

The intrinsically low reversibility of lithium plating and strip- of Coulombic efficiency values for n cycles.[24] The amount of
ping leads to rapid loss of lithium inventory with cycling. This lithium that is recoverable as a fraction of the lithium plated
is especially true for lithium deposition on bare planar sub- during the first charge step is typically smaller compared to later
strates with unoptimized electrolytes. Since anode-free full cells cycles, due to which the initial discharge capacity may be sub-
(Cu || LiX, where LiX is a fully lithiated cathode) only have a lim- stantially smaller than the theoretical capacity of the anode-free
ited lithium inventory (N/P ratio = 1), the loss of lithium inven- full cell. The first cycle capacity loss can be accounted for by con-
tory leads to rapid capacity fade. This is particularly evident sidering the initial discharge capacity for the calculation above.
when comparing the performance of anode-free full cells to that Nevertheless, it needs to be clearly reported for accurate assess-
of half cells (Li || LiX), which employ a large excess of lithium. ment of electrochemical performance, as a significant first cycle
The cyclability of half cells is artificially inflated and does not capacity loss can diminish the realizable energy density, even if
truly reflect lithium inventory loss inside the cell, as any lithium lithium deposition is highly reversible during further cycling.
lost due to inefficiencies in cell operation is simply replenished Further assuming that the value of Coulombic efficiency is
from the lithium reservoir. In addition, the excess lithium constant for the n cycles at a value CE, the equation reduces
employed invalidates the energy density advantage of lithium- to Qn = Qin *CEn. Figure 2a shows the capacity retention
metal anodes. In contrast, lithium inventory loss in anode-free (Qn/Qin) as a function of cycle number for different values of
full cells renders an equivalent amount of the cathode material CE. It can be seen that cycle life, defined generously as the
electrochemically inaccessible, which leads to a corresponding number of cycles taken to reach 50% capacity retention, dis-
reduction in deliverable capacity and limits their cyclability. plays a strong dependence on Coulombic efficiency. The ratio
Improving the cycle life of anode-free full cells requires the of cycle lives for cells a and b with Coulombic efficiencies CEa
development of strategies for stabilizing lithium deposition and and CEb can be derived as
sustaining the limited lithium inventory with cycling. These
strategies are discussed in detail in Section 4. na log ( CEb ) (2)
=
Assuming that there is negligible loss of active material at nb log ( CEa )
the cathode, the capacity fade of anode-free full cells can be
attributed entirely to the loss of lithium inventory. It may be Using the formula, it can be shown that every jump in
further assumed that lithium inventory loss can be directly pre- Coulombic efficiency from 90% to 99%, and then to 99.9% and
dicted from the measured Coulombic efficiency, i.e., the ratio so on leads to a roughly 10-fold improvement in cycle life. At a
of discharge capacity to charge capacity as measured during Coulombic efficiency of 99.99%, the cycle life becomes compa-
electrochemical cycling. This relies on the presumption that rable to that of commercial cells with graphite anodes. All of
lithium inventory loss occurs during either the charge step or the formulas above apply even if there is loss of active mate-
discharge step, but not both. Under these assumptions, the dis- rial at the cathode, so long as it remains less than the lithium
charge capacity Qn of an anode-free full cell after n cycles can inventory loss throughout cycling.
be expressed as In practical systems, however, the measured values of Cou-
lombic efficiency can display significant variations with cycling.
n
Hence, different capacity fade regimes with significant varia-
Q n = Q in * ∏ CE
i=2
i = Q in * GCEn (1)
tions in the rates of lithium inventory loss may be observed (for
instance, if an electrolyte additive is depleted in the course of
where Qin is the initial discharge capacity, CEi is the Coulombic cycling). It is also possible that Coulombic efficiency could cease
efficiency for cycle number i, and GCE is the geometric mean to be a useful predictor of capacity fade. This is particularly

Figure 2. a) Capacity retention with cycle number of an anode-free full cell for different values of average Coulombic efficiency. 50% capacity retention
is generously used to determine the cycle life. b) Capacity retention with cycle number for hypothetical anode-free full cells that show a higher initial
cathode inventory loss (40%) compared to lithium inventory loss (20%). The subsequent cathode inventory loss rate is 0.1% per cycle. A capacity
plateau can be seen that is engendered due to the temporary excess of lithium inventory in the cell. The length of the capacity plateau depends on the
lithium inventory retention rates, as can be seen from the difference between 99% and 99.5%.

Adv. Energy Mater. 2020, 2000804 2000804 (4 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

applicable when comparable losses in lithium inventory are profile would directly overlap with the capacity plateau observed
registered during both the plating and stripping steps. Another in the anode-free full cell. Therefore, it may be useful to report
important exception to this rule occurs in systems with sol- electrochemical cycling data for equivalent half cells in con-
uble redox mediators that shuttle between the cathode and the junction with anode-free full cells in order to shed light on the
anode, which are discussed in detail in Section 5.[31,32] dynamics of inventory loss for both the anode and the cathode.
These problems can be circumvented by defining analogous The dependence of electrochemical performance of anode-
to Coulombic efficiency a new parameter, the lithium inventory free full cells on the Coulombic efficiencies of plating and strip-
retention, which for any given cycle number i can be defined ping lithium imparts them with poor cyclability, but also makes
as LIRi = Qi/Qi−1, where Qi is the discharge capacity for cycle them excellent diagnostic templates for effectively investigating
number i. This can be further used to derive another useful the dynamics of lithium deposition. This is especially true in
parameter, the lithium inventory retention rate (LIRR), which conjunction with certain cathode systems, such as NMC, LMO,
can be defined as a percentage and sulfur, where active material dissolution at the cathode and
crossover to the anode side has an observable impact on the
1
characteristics of lithium deposition.[34–37] The commonly uti-
 Qn  n
LIRR = 100 *  (3) lized Li || Cu half cells or Li || Li symmetric cells are inadequate
 Q in  for accurately replicating lithium plating and stripping in such
systems and can only provide incomplete information. Various
Here, the LIRR can be calculated by using a specific capacity strategies for stabilizing lithium deposition and improving the
retention (say 80%) and the corresponding number of cycles, or Coulombic efficiencies of lithium plating and stripping, such
conversely, by using a specific number of cycles (say 100 cycles) as electrolyte additives and artificial SEI layers, can be robustly
and the corresponding capacity retention. In either case, the and rapidly evaluated in the anode-free framework. Signifi-
LIRR, as a proxy for Coulombic efficiency and as a measure of cantly, the compatibility of all aspects of such strategies, espe-
the “average” loss in lithium inventory per cycle, can be a useful cially different electrolyte formulations, with various cathode
value for a comparison across different reported anode-free full systems can be determined. Detailed characterization of the
cells owing to its simple calculation. In this review, we have deposited lithium in anode-free full cells, both before and after
strived to calculate the LIRR for the best reported anode-free stripping, can generate useful insights into its morphology,
full cell in each work, and used n = 100 or 50 cycles, depending structure, and composition.[38] Considering the absence of
on the available information, and the corresponding capacity excess metallic lithium typically present in lithium-metal foil
retention. anodes, an accurate picture of the distribution of various SEI
For most systems, the overall inventory loss rates for the components and deposited lithium metal (both electrochemi-
cathode active material are generally less than that for lithium. cally active and “dead”) can be obtained. This can prove particu-
However, the first cycle inventory loss for the cathode can be larly useful for detailed quantitative studies on the mechanisms
quite significant and can even be greater than that for lithium. that lead to lithium inventory loss with cycling.[29]
In such cases, there is a partial excess of lithium inventory com- A major challenge with anode-free full cells is the large
pared to the cathode inventory after the first charge/discharge volume expansion of the cell stack that accompanies plating
cycle. Since the measured capacity in a full cell is a convolution and stripping of lithium. In the example system shown in
of the active inventory for both the cathode and the anode, a Figure 1b, 6.2 mAh cm−2 of plated lithium capacity corresponds
“capacity plateau” is observed until the lithium inventory excess to a 30 µm increase in stack thickness, or a volume expansion
is depleted and exponential capacity loss is resumed. Figure 2b of nearly 20%. However, this is still an underestimate, as the
shows the predicted capacity retention for two anode-free full actual thickness of porous high-surface area deposited lithium
cells with an LIRR equal to 99% and 99.5% respectively, given of a particular capacity would almost certainly exceed the the-
a first cycle lithium inventory loss of 20%, first cycle cathode oretical value.[39] The volume expansion of the cell stack with
inventory loss of 40%, and subsequent cathode inventory loss lithium plating may be partially offset by an accompanying con-
rate of 0.1% per cycle. In these example systems, the capacity traction of the fully lithiated cathode upon delithiation. This is
plateau is observed for 33 cycles at 99% LIRR and 73 cycles at especially true for cathodes such as Li2S that undergoes a 43%
99.5% LIRR, when the active lithium and cathode inventory volume contraction during charge. However, layered-oxide
reach parity. This behavior can also be observed when the lithi- NMC and NCA cathodes experience negligible contraction with
ated cathode is fully charged, but subsequently the delithiated lithium extraction. In such cases, novel strategies will have to
cathode is only partially discharged.[33] In these cases, the LIRR be devised to accommodate the large changes in volume of the
has to be extrapolated from the lithium-limited cycling regime, cell stack with cycling. Since the problem of volume expansion
i.e., beyond the capacity plateau. Therefore, it is critical to is made significantly worse by mossy or dendritic plating of
report complete cyclability information for anode-free full cells, lithium, it becomes even more critical to stabilize lithium depo-
until cell failure, to prevent an overestimation of the lithium sition and form compact, dense, and uniform lithium deposits.
inventory retention rates. The application of pressure, possibly in conjunction with
A comparison of capacity fade between a Li || LiX half-cell temperature controls, could also help alleviate this problem.
and an anode-free Cu || LiX full cell may be instructive here. Another strategy is to devise a 3D host to accommodate the
Assuming complete replenishment of lost lithium inventory deposited lithium and contain the extreme volume changes.
from the lithium reservoir in a half cell, its capacity is simply However, the extra weight and volume of the 3D current col-
a measure of the active cathode inventory. Hence, its cycling lector will have to be taken into consideration so as to not

Adv. Energy Mater. 2020, 2000804 2000804 (5 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

Figure 3. a) An outline of the three broad classes of strategies employed to improve anode-free full cell performance and b) scanning electron micros-
copy image of deposited lithium-metal with high surface area and low surface area morphologies.

invalidate the energy density advantage of anode-free full cells. (SEI), which ensures facile conduction of Li+ ions and pre-
They could also compromise energy density by way of the extra vents decomposition of the liquid electrolyte on the deposited
electrolyte needed for acceptable cell operation. Fundamentally lithium. The major tools in delivering these aims are i) modi-
new cell component designs and pack-level innovations could fication of the liquid electrolyte, ii) modification of the current
be necessary to surmount this problem. collector substrate upon which the lithium metal is deposited,
and iii) optimization of formation and cycling parameters to
prevent capacity fade. These strategies are outlined in Figure 3.
4. Strategies for Improving Performance
of Anode-Free Full Cells
4.1. Modification of the Electrolyte
The first reported work on anode-free full cells was by Neu-
decker et al.[40] They reported a thin-film battery employing The electrolyte is a critical component of any anode-free full
up to 3 µm thick LiCoO2 cathodes, a LiPON film as the elec- cell, as it plays an essential role in determining the morphology
trolyte, and a sputter deposited Cu current collector. Excellent of the deposited lithium as well as the composition of the
cyclability was obtained at current rates up to 5 mA cm−2, but SEI layer formed on the lithium surface. By precisely tuning
the capacity of the plated lithium was less than 0.1 mAh cm−2. the electrolyte to enable uniform and smooth lithium deposi-
This review primarily covers anode-free full cells employing tion, the growth of high surface area “mossy” lithium can be
traditional liquid electrolytes, except for a section on solid-state mitigated, inhibiting undesirable side reactions between the
anode-free batteries at the end. electrolyte and metallic lithium. In the same vein, the electro-
Given the considerable importance of accounting for and lyte can be tuned to decompose into a favorable SEI on the
sustaining lithium inventory in anode-free full cells, the major surface of the deposited lithium, with high ionic conductivity
aim of anode-free research over the past few years has been to ensure electrochemical accessibility to the enclosed lithium
focused on improving Coulombic efficiencies (CE) by miti- metal and effective passivation to mitigate further electrolyte
gating irreversible loss of lithium inventory. While cathode opti- decomposition. Both of these measures can go a long way
mization is a critical tool for achieving optimal electrochemical toward preserving lithium inventory over the course of cycling.
performance, anode-free full cells are primarily limited by the The seminal work on anode-free full cells with liquid elec-
poor efficiency of plating and stripping lithium at the anode trolytes, published by Qian et al.,[22] thoroughly investigated the
current collector. Hence, most studies have focused on two trade-offs between using a standard carbonate electrolyte, 1 m
critical aspects for stabilizing lithium deposition and sustaining LiPF6 in ethylene carbonate (EC)/dimethyl carbonate(DMC)
lithium inventory: promoting favorable lithium deposition mor- (1:2, v/v), versus a highly concentrated ether-based electrolyte, 4 m
phology and ensuring a favorable solid-electrolyte interphase lithium bis(fluorosulfonyl)imide (LiFSI) in dimethoxyethane

Adv. Energy Mater. 2020, 2000804 2000804 (6 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

(DME). These electrolytes were employed in an anode-free bat- capacity retention after 50 cycles (LIRR = 98.2%). The initial
tery with a copper current collector and LiFePO4 (LFP) cathode. lithium stripping capacity was ≈2 mAh cm−2, and the current
Lithium plating and stripping was effectively stabilized by rate was 0.2 mA cm−2. This work highlighted the use of anode-
the 4 m LiFSI in DME electrolyte, exhibiting both improved free full cells employing carbonate-based electrolytes and high-
deposition morphology and favorable SEI formation. This voltage cathodes and foreshadowed even greater focus on such
high-concentration electrolyte enabled 60% capacity reten- systems. Nilsson et al. reported stabilized lithium deposition
tion after 50 cycles (LIRR = 99%), maintaining an average CE with a highly concentrated electrolyte comprising 1.86 m LiTFSI
greater than 99%. The initial lithium stripping capacity was in EC solvent.[45] Anode-free Cu || LFP full cells with a rela-
1.7 mAh cm−2, and the current rate was 0.2 mA cm−2. The tively high initial lithium stripping capacity of 3.5 mAh cm−2
improvement in electrochemical performance was attributed and applied current of 1 mA cm−2 maintained 50% of their
to the greater degree of coordination between solvent and salt initial capacity at 30 cycles (LIRR = 97.7%). In comparison to
molecules, such that the majority of salt anions, cations, and the formation of lithium globules with ether-based electrolytes,
solvent molecules exist in contact-ion-pairs or solvate aggregate dense columnar growth was observed with the concentrated
states. Volumetrically, there is also a fourfold increase in the carbonate-based electrolyte containing LiTFSI.
concentration of Li+ ions available to reduce on the surface of Alvarado et al. introduced a bisalt ether electrolyte that
the current collector and form metallic lithium during charge. showed considerable performance improvements when
Further, the increased concentration of salt anions in the elec- employed with a high-nickel NMC622 cathode in the anode-free
tric double-layer adjacent to the negatively polarized current full cell configuration.[46] This strategy employed a concentrated
collector surface leads to a more uniform flux of Li+ ions, dual-salt formulation containing both LiTFSI and LiFSI in
leading to uniform lithium deposition. Finally, the reduction of ether-based solvents. This allowed for the use of high-capacity
FSI− anionic species leads to the formation of a favorable SEI NMC622 cathodes at high voltages of up to 4.4 V, expanding
for lithium deposition rich with inorganic components such upon the voltage window shown to date in anode-free full cells
as LiF.[17,41] This work outlined the fundamental design princi- and increasing the energy density achievable with such sys-
ples and considerations for scientific investigations into anode- tems. By carefully choosing a blend of 2.3 m LiTFSI and 4.6 m
free full cells, and it is these established tenets that have since LiFSI (molality was utilized instead of molarity), a favorable
guided work on electrolytes for these systems. SEI layer was formed on the deposited lithium. Using careful
Since this initial investigation, a multitude of diverse strat- computational and experimental studies, it was found that a
egies have been developed and implemented by varying the complex interplay of the salt anions and their selective decom-
composition and concentration of the electrolyte salts and sol- position on the lithium surface was responsible for forming a
vents in anode-free full cells. Beyene et al. reported a dual salt SEI layer with a unique chemistry that was well suited for sta-
formulation that developed upon the high concentration salt bilizing lithium deposition. Both half-cells and anode-free full
strategy, but employed an electrolyte containing 2 m LiFSI cells tested with the novel bisalt electrolyte displayed consider-
and 1 m lithium bis(trifluoromethanesulfonyl)imide (LiTFSI) able performance improvements, the anode-free full cell in par-
in DME/dioxolane (DOL) (1:1, v/v).[42] This work reported 50% ticular demonstrating capacity retention of 55% after 54 cycles
capacity retention after 50 cycles (LIRR = 98.6%) in anode-free (LIRR = 98.9%), compared to the baseline cells employing car-
Cu||LFP full cells. The initial lithium stripping capacity was bonate-based electrolyte that faded to 0% after 35 cycles. The ini-
1.6 mAh cm−2, and the current rate was 0.2 mA cm−2. While tial lithium stripping capacity was 1.44 mAh cm−2, and variable
it does not outperform the previous work employing 4 m LiFSI current rates of 0.144 mA cm−2 for charge and 0.48 mA cm−2
in DME electrolyte, this strategy does offer the advantage of for discharge were used. More so, the authors demonstrated
reducing the overall salt concentration and utilizing the rela- that the performance improvement was not simply from the
tively less expensive LiTFSI salt. Rodriguez et al. investigated increased lithium content in the highly concentrated elec-
a number of ether-based electrolyte formulations in anode-free trolyte, but due to the favorable chemistry of the SEI layer
Cu || LFP full cells and found that lowering the LiFSI salt con- that was engendered by an optimized LiFSI to LiTFSI ratio.
centration in DME from 4 to 1 m only led to a minor reduction Fan et al. reported stabilized lithium deposition with a highly
in capacity retention with cycling.[43] fluorinated electrolyte comprising 1 m LiPF6 in FEC/3,3,3-
Hagos et al. investigated the use of carbonate electrolytes, fluoroethylmethyl carbonate/1,1,2,2-tetrafluoroethyl-2’,2’,2’-
employing an electrolyte consisting of 2 m LiPF6 in EC/diethyl trifluoroethylether (FEC/FEMC/HFE, 2:6:2 by wt).[47] An impor-
carbonate (DEC) (1:1, v/v) diluted by 50% with fluoroethylene tant advantage of this electrolyte chemistry is its stability with
carbonate (FEC).[44] This is notable in that the use of carbonate high voltage cathodes, such as LiCoPO4 at up to 5 V. Anode-free
electrolytes enables the use of higher voltage cathodes such Cu || NMC811 full cells with an initial lithium stripping capacity
as high-nickel layered oxides, as carbonate-based electrolytes of 2.75 mAh cm−2 and cycled at 0.5 mA cm−2 maintained 35%
generally possess greater electrochemical stability windows of initial capacity after 30 cycles (LIRR = 96.6%). The improve-
than ether-based electrolytes. The large volume percent of ment over a nonfluorinated electrolyte was attributed to the
the fluorinated solvent FEC stabilizes lithium deposition due formation of a stabilizing LiF-rich SEI layer.
to the formation of a LiF-rich SEI layer, while reducing the In 2019, Weber et al. reported an anode-free Cu || NMC
overall salt amount and viscosity of the electrolyte compared to full cell consisting of an optimized dual-salt carbonate-
high concentration electrolytes. Anode-free full cells pairing a based electrolyte.[23] The electrolyte, consisting of 1 m lithium
LiNi1/3Mn1/3Co1/3O2 (NMC111) cathode with a copper foil cur- difluoro(oxalato)borate (LiDFOB) and 0.2 m LiBF4 in a blend of
rent collector and evaluated with this electrolyte displayed 40% FEC/DEC (1:2, v/v), enabled a significant improvement in the

Adv. Energy Mater. 2020, 2000804 2000804 (7 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

reversibility of lithium deposition, with 80% capacity reten- The role of electrolyte in determining both the morphology
tion at 90 cycles (and ≈90% at 50 cycles, corresponding to an of the deposited lithium as well as the composition of the
LIRR of 99.8%). The initial lithium stripping capacity was SEI layer underscores its importance in anode-free full cells.
2.4 mAh cm−2, and variable current rates of 0.48 mA cm−2 for While significant headway has been made in this area over the
charge and 1.2 mA cm−2 for discharge were employed. The per- last few years, there are clear opportunities for progress and
formance is even more remarkable when considering that it is improvement. A thorough understanding needs to be devel-
achieved in a practical large-area pouch cell format under lean oped of the relative importance of the different critical proper-
electrolyte conditions, with an electrolyte amount to capacity ties of the SEI layer that impact the efficiency of plating and
ratio of 2 g Ah−1. The cells were cycled at 40 °C under 75 kPa stripping lithium, including lithium-ion conductivity, elec-
of stack pressure. The effect of pressure on electrochemical tronic insulation, mechanical stability, chemical passivation,
performance is discussed later. The plated lithium in the opti- and interfacial energy. Promising SEI components that meet
mized electrolyte displayed highly favorable morphology, with the criteria above and stabilize lithium deposition should also
large, dense, columnar deposits and minimal surface area be identified. A considerable number of work in this section
exposed to the electrolyte. Through investigations with XPS, rely on the stabilizing properties of LiF in the SEI layer. Other
the authors showed that the SEI layer arising from decomposed SEI layer components might also be considered, including high
LiDFOB is rich in organic boronated and fluorinated compo- ionic-conductivity sulfides such as Li3PS4, halides such as LiI,
nents. In contrast, the SEI layer arising from the decomposi- polymeric or organic species, and in situ formed lithium–metal
tion of LiPF6 and LiBF4 is rich in inorganic components, such alloys. Capacity retention in anode-free full cells will need to
as LiF. By employing the dual salt strategy with an optimized last beyond the sub-100 cycle range demonstrated so far, par-
salt ratio, the resultant SEI is rich in both favorable organic and ticularly while accounting for practical engineering considera-
inorganic components. This is the probable cause of the excel- tions such as lean-electrolyte conditions, gas generation from
lent electrochemical performance obtained with this strategy. electrolyte decomposition, and nonuniform current densities in
Furthermore, it was found that the major driver of lithium large-area pouch cells.
inventory loss is the consumption of LiDFOB and LiBF4 salts
with cycling, suggesting that optimized dual-salt or even tri-
salt formulations may be able to extend cycle life further. The 4.2. Modification of the Current Collector
electrochemical performance shown in this work is one of the
best reported for anode-free full cells so far, and is even more Another route toward improving the cyclability of anode-free
impressive due to the use of a pouch cell format with lean-elec- full cells is through modifying the current collector substrate
trolyte conditions.[48] upon which the lithium metal is deposited. The current col-
Beyond modifying the composition and concentration of the lector plays a key role in determining the initial nucleation
electrolyte salts and solvents, a few studies have focused on sta- conditions and subsequent growth of the deposited lithium and
bilizing lithium deposition in anode-free full cells with the use is, therefore, an important lever in the development of an opti-
of electrolyte additives. Sahalie et al. reported using KNO3 as an mized anode-free full cell. Modification of the current collector
additive in a carbonate-based electrolyte, which was employed through pretreatments can help improve the morphology of
in an anode-free Cu || NMC full cell. The use of KNO3 enabled lithium deposition, while coatings can function as an artificial
more favorable nucleation and growth of lithium metal during SEI layer and prevent electrolyte decomposition on the depos-
charge, allowing for 40% capacity retention after 50 cycles ited lithium. While still a relatively nascent and developing sub-
(LIRR = 98.2%).[49] The initial lithium stripping capacity was field, the few investigations that have been conducted in this
1.7 mAh cm−2, and the current rate was 0.2 mA cm−2. The area have focused primarily on coatings on the current-collector
same group also reported a dual electrolyte additive strategy or modification/substitution of the underlying current-collector
consisting of KPF6 and tris (trimethylsilyl) phosphite, which substrate.
moderately enhanced capacity retention compared to a control Hwang and co-workers have reported a series of different
cell in Cu||NMC full cells.[50] Brown et al. reported the use of current-collector coatings films and artificial SEIs including pol-
triethylphosphate as an electrolyte cosolvent to enable the dis- yethylene oxide (PEO),[53] multilayer graphene (MLG),[54] garnet
solution of 0.2 m LiNO3 additive in 1 m LiDFOB in EC/DMC (Li7La2.75Ca0.25Zr1.75Nb0.25O12),[55] and graphene oxide.[56] The
electrolyte.[51] The high concentration of LiNO3 in the electro- use of PEO and multilayer graphene led to moderate improve-
lyte led to the formation of an SEI rich in nitrogenated species ments in capacity retention over the control case with LFP cath-
that allowed effective stabilization of lithium deposition. With odes and an unoptimized ether-based electrolyte. In particular,
an initial lithium stripping capacity of 1 mAh cm−2 and applied anode-free Cu@PEO || LFP full cells demonstrated 64% capacity
currents of 0.1 mA cm−2 for charge and 0.4 mA cm−2 for dis- retention at 50 cycles (LIRR = 99.1%), albeit with a low initial
charge, anode-free Cu || LFP full cells demonstrated roughly lithium stripping capacity of 0.71 mAh cm−2 and a current rate
35% capacity retention at 50 cycles (LIRR = 97.9%). The same of 0.14 mA cm−2. In the investigation into multilayer graphene
group also reported that adding 5% vinylene carbonate (VC) to (MLG), the authors employed a larger area current collector
1 m LiPF6 in EC/EMC (3:7, vol) electrolyte led to a significant relative to the cathode area to reduce the stress encountered at
improvement in the capacity retention of Cu || LFP full cells the interface of the deposited lithium metal. With the improve-
(showing a LIRR of 95.1%).[52] The improvement was attrib- ments from the MLG coating, anode-free Cu@MLG || LFP cells
uted to the formation of a polymeric SEI layer on the lithium showed an impressive 78% capacity retention at 50 cycles (LIRR
surface. = 99.5%) with an unoptimized ether-based electrolyte. The

Adv. Energy Mater. 2020, 2000804 2000804 (8 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

initial discharge capacity was 1.7 mAh cm−2 and the current rate (30% capacity retention at 50 cycles, corresponding to an LIRR
was 0.5 mA cm−2. The garnet and graphene oxide coatings were of 97.6%).[60] The initial plating capacity was ≈1 mAh cm−2,
employed to demonstrate moderate improvements in the elec- and the current rate was 0.5 mA cm−2. In the case of lithium,
trochemical performance of anode-free Cu || NMC cells with a Li2CuO2 additive irreversibly plates lithium metal onto the
an unoptimized carbonate-based electrolyte, showing LIRRs of current collector during the first cycle and subsequent lithium
98.2% and 98.3% respectively. These coatings behave as artifi- deposition occurs upon this thin in situ formed lithium–
cial SEIs and are thought to enable a uniform Li+ ion flux and metal substrate. A significant improvement in cyclability was
prevent the deposited lithium from reacting with the liquid observed with the pre-plating of lithium metal, and the cycle life
electrolyte. However, in the absence of an optimized electrolyte, increased with an increase in the mount of preplated lithium
these implementations display average Coulombic efficien- from the Li2CuO2 additive. While some of the improvement
cies ranging from 97% to 99%, which falls short of the values can be attributed to the additional reservoir of lithium available
needed for extending cyclability to the range necessary for prac- in the cell, it is mostly expected to originate with the minimal
tical consideration.[22] For instance, the authors demonstrated interfacial energy between the cycled lithium and the preplated
that even with the use of GO protective layers, the addition of lithium.[61] A final study evaluating current collector substrates
5% FEC had a dominant impact on the cyclability of anode-free for anode-free full cells was published in 2019 by Pande and
Cu || NMC full cells with a standard carbonate electrolytes.[56] Viswanathan.[62] By computationally screening candidates for
While the premise of artificial SEIs and robust protective or reduced activation barriers for nucleation of lithium metal
stabilizing layers for lithium deposition has been the focus of and surface diffusion of lithium atoms, the authors found
much research attention over the last decade,[20,21,57] their sole several lithium alloy substrates that could help form com-
application in anode-free full cells is yet to yield reversible pact and dense deposits of lithium. This would feasibly allow
lithium deposition over hundreds of cycles. for enhanced electrochemical performance for an extended
Rather than coating the current collector with a film at the number of cycles, even at high charge rates.
electrode–electrolyte interface, an alternative approach involves The current collector plays an essential role in determining
modifying the underlying conductive substrate upon which the morphology of the deposited lithium during cycling and
lithium metal nucleates. The overpotential for nucleating is thus a key lever for sustaining lithium inventory in anode-
metallic lithium is an important parameter as it plays a role free full cells. There have been important developments in
in determining the density and uniformity of lithium deposi- coatings, pretreatments, and modifications for current collec-
tion.[57,58] In a 2016 study, Cui and co-workers showed that the tors, but there is still a large room for improvement. Coating
nucleation overpotential for plating lithium metal is highly layers employed as artificial SEIs to stabilize lithium deposition
dependent on the deposition substrate.[59] A series of metallic must be robust, defect-free, and tolerant to the severe volume
substrates were evaluated for lithium deposition, and it was changes of the deposited lithium metal during cycling. Addi-
found that the substrates with the greatest crystalline mismatch tional opportunities lie in the replacement of copper with more
with lithium metal exhibited the highest nucleation overpoten- favorable conductive substrates for lithium nucleation and
tial. Metallic lithium possesses a body-centered cubic (BCC) diffusion. A crucial factor that needs to be considered in the
crystal structure and thus deposits favorably on substrates with context of anode-free full cells is the lithiation capacity of the
BCC structures. Lithium also deposits favorably on “lithio- “lithiophilic” substrate that is employed, which should be as
philic” materials, which partially solubilize or alloy with lithium low as possible to maximize the fraction of the limited lithium
metal. This includes metals such as gold, even though it pos- inventory that is available for plating and stripping. Prelithia-
sesses a face-centered cubic (FCC) crystal structure. This study tion of the current collector substrate could be considered, or
highlights that copper is among the least favorable deposition it could be engineered to minimize its lithiation capacity while
substrates for lithium metal, given that it possesses an FCC still maintaining the promised performance improvements. It
crystal structure and allows no solubility of lithium metal. should be ensured that the additional lithiophilic components
Hence, there is significant opportunity for improvement on the current collector substrate does not significantly under-
in the electrochemical performance of anode-free full cells mine the energy density gains with the anode-free configura-
through modification of the current collector substrate. The tion. Further work in this area, and possibly several strategies
vast majority of studies on anode-free full cells employ bare in conjunction, will allow the development of anode-free full
copper foils as the deposition substrate, but there have been a cells to ascend to a more practically deployable stage.
few selected studies evaluating other substrates. Zhang et al.
investigated thin layers of tin and lithium metal as deposition
substrates in anode-free full cells.[60,61] These layers are depos- 4.3. Optimization of Cell Formation and Cycling Parameters
ited onto copper foil as a pretreatment in the case of tin, or in
situ in the case of lithium. In the case of tin, lithium proceeds A final lever for sustaining lithium inventory in anode-free full
to alloy with the tin layer during initial deposition. This stabi- cells is the formation and cycling parameters employed during
lizes subsequent lithium deposition by lowering the interfacial operation. Such operational controls, including modified
energy between the substrate and lithium metal. Anode-free resting steps, optimized first cycle protocols, and the application
full cells employing NCA cathodes with 1.0 m (molality) LiPF6 of temperature and pressure can have drastic effects on cycle
in 1:4 FEC/EMC (by wt) as the electrolyte display improved life in lithium-ion and lithium metal batteries.[63–66] The initial
performance with the tin coating compared to those with bare formation steps are particularly crucial for forming a stable and
Cu foils, and the cycle life extended from 30 cycles to 80 cycles passivating SEI layer on the anode before prolonged cycling.

Adv. Energy Mater. 2020, 2000804 2000804 (9 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

As anode-free full cells lack a pre-existing anode, optimal cell


formation procedures may vary considerably from other con-
ventional systems. While optimization of the initial formation
procedure for anode-free full cells is still an emerging area of
research, it has nonetheless garnered considerable efforts from
the research community.
Beyene et al. introduced a resting and cycling protocol for
anode-free full cells, which when combined with a high-
concentration electrolyte, produced a considerably more robust
SEI on the surface of the in situ deposited lithium metal.[67]
They employed a twofold approach, where after cell assembly,
the cell was charged and lithium was deposited at a low rate
of 0.1 mA cm−2. After the first charge, the cell was rested for
24 h in order to allow the formation of a robust SEI layer rich
in fluorinated species, which originates from the LiFSI elec-
trolyte salt. Next, the cell was discharged, and the deposited
lithium stripped away at a higher rate of 0.5 mA cm−2. Finally,
the cell was galvanostatically cycled at up to 2 mA cm−2. The
authors claim that this increases the temperature of the depos-
ited lithium locally and allows uniform and smooth deposition
Figure 4. A summary of the investigations into anode-free lithium-ion
of lithium. When this approach was applied in anode-free Cu
cells and affiliated LIRR that have been conducted since 2016.
|| LFP full cells, stable cycling was achieved with an average
Coulombic efficiency of ≈99% and a capacity retention of 64%
after 50 cycles (LIRR = 99.1%). The initial plating capacity employed by Hwang’s group,[67] although the hot formation
was 1.7 mAh cm−2, and the current rate was 0.5–1 mA cm−2. procedure also leads to considerable gas generation in the cell.
This approach is interesting in that it combines an initial low- This work utilized a highly optimized LiDFOB/LiBF4 dual-salt
rate lithium deposition step, a rest protocol, and a high-rate electrolyte in FEC:DEC (1:2, v/v) solvent, with varying concen-
conditioning step. This beckons further investigation into ini- trations of LiDFOB and LiBF4. It is expected that the optimal
tial formation protocols for other electrolyte systems. formation procedures will vary with the electrolyte. Neverthe-
In early 2019, Louli et al. documented the effects of applying less, the hot formation procedure was further applied under a
mechanical stack pressure on the cycling stability of anode-free 1200 kPa of pressure with a concentrated salt electrolyte formu-
full cells.[68] They employed large-area anode-free Cu || NMC lation consisting of 1.8 m LiDFOB and 0.4 m LiBF4 in FEC:DEC
pouch full cells with a low electrolyte amount of 3 g Ah−1, which (1:2, v/v). The anode-free Cu || NMC pouch full cell employing
were cycled while applying a uniaxial stack pressure ranging this electrolyte retained 80% of its initial capacity at a remark-
from 75 to 2200 kPa. The cyclability of the cells generally able 195 cycles (95% retention at 100 cycles, or a LIRR = 99.9%),
improved with an increase in the applied pressure, although with an average Coulombic efficiency of 99.67%. The initial
the extent of this improvement varied with the physical prop- lithium stripping capacity was 2.5 mA cm−2, and the applied
erties of the electrolyte solvent. As with the previous resting current was 0.5 mA cm−2 for charge and 1.25 mA cm−2 for dis-
and cycling protocol outlined by Beyene et. al., favorable opti- charge. This represents one of the best achievements of elec-
mization of cycling parameters depends significantly on the trochemical performance not just in anode-free full cells, but
electrolyte employed.[67] Genovese et al. reported the effects of even in cells employing lithium-metal anodes. This milestone
a “hot formation” procedure, where the initial two cycles were in the development of anode-free batteries represents a combi-
carried out at an elevated temperatures of 40 °C.[69] They found nation of an optimized electrolyte with high salt concentration,
that the high-temperature formation cycles enable the deposited optimized formation procedure, and favorable cycling protocols.
lithium metal to form dense, columnar growths, when com- The development of anode-free full cells has brought about
pared with formation cycles carried out at 20 °C. This optimized a great deal of engineering efforts and innovation toward
high-temperature formation procedure was applied in anode- improving LIRR and sustaining lithium inventory over the
free Cu || NMC pouch full cells in conjunction with 75 kPa of course of cycling. The diversity of approaches, ranging from
stack pressure to realize drastically improved performance, with optimized electrolyte composition to modification of the current
over 55% capacity retention at 100 cycles compared to the same collector and formation and cycling protocols, have advanced
retention at only 18 cycles in cells with the formation procedure our understanding of anode-free systems and the critical con-
carried at room temperature (LIRR = 99.4%). The initial lithium siderations necessary for high Coulombic efficiency stripping
stripping capacity was 2.5 mA cm−2, and the applied current and plating of lithium metal. As Figure 4 shows, the research
was 0.5 mA cm−2 for charge and 1.25 mA cm−2 for discharge. progress to date in improving LIRR has been very promising,
Moreover, when the pressure was increased to 1200 kPa, anode- with some of the best-performing systems displaying retention
free full cells employing the hot formation procedure exhibited rates much greater than 99%. The retention rates plotted in this
85% capacity retention at 100 cycles (LIRR = 99.8%). It is pos- figure were calculated based on the number of cycles taken to
sible that this hot formation procedure stabilizes lithium depo- reach a given percentage of capacity retention, as described
sition in a manner similar to the high-rate conditioning steps in Section 3 of this review. Given that LIRR directly describes

Adv. Energy Mater. 2020, 2000804 2000804 (10 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

Figure 5. a) Electrochemical performance of a Li || Li2S half-cell and an anode-free Cu || Li2S full cell, with a 4 mg cm−2 cathode. c) Coulombic efficiencies
of plating and stripping lithium on Cu foil from a Li2S cathode and a lithium-metal counter-electrode. Reproduced with permission.[24] Copyright 2018,
Wiley-VCH.

retention rate of lithium inventory in anode-free full cells and Li2S cathode. These polysulfide species diffuse to the anode side,
determines overall cyclability, this metric gives the most holistic where they reduce on the lithium surface to form Li2S and Li2S2.
and comprehensive evaluation of the work on anode-free full The formation of these reduced sulfur species on the lithium
cells conducted thus far. surface, which exist in a dynamic equilibrium with the dissolved
polysulfide species, is believed to have a stabilizing effect on
lithium deposition, as demonstrated by multiple investigations
5. Anode-Free Sulfur Batteries on this topic.[36,37,70,71] In particular, a synergistic effect of poly-
sulfides and the LiNO3 electrolyte additive is thought to form a
Lithium–sulfur batteries have attracted a great deal of attention passivating SEI layer that is favorable for reversible lithium dep-
in the last decade because of the high theoretical gravimetric osition. Some of this effect can be attributed to the sulfide ani-
capacity of sulfur (1675 mAh g−1) and its low cost and abun- onic framework of the reduced sulfur species in the SEI layer,
dance compared to the transition metals employed in incum- which leads to intrinsically lower Li+ ion diffusion barriers, when
bent Li-ion batteries. As a nonlithiated cathode, the sulfur compared to an oxide anionic framework that dominates many
cathode has to be paired with a lithium-metal anode for opera- SEI components.[72–74] Figure 5b compares the Coulombic effi-
tion. Alternatively, fully lithiated Li2S can be used as cathodes ciencies of plating and stripping lithium metal on Cu foil from
and paired with lithium-free anodes. a Li2S counter electrode (Cu || Li2S) and a bare lithium metal
In 2018, our group reported an anode-free full cell pairing counter electrode (Cu || Li). While the Cu || Li control cell shows
a Li2S/CNT cathode with a bare Cu foil.[24] To the best of our CE values with a geometric mean of 62.01% over 50 cycles, the
knowledge, this is the first and only report of a Li–S battery in Cu || Li2S control cell shows CE values with a geometric mean
the anode-free full cell configuration. When compared with of 97.14% over 50 cycles. The difference can be attributed to the
a Li || Li2S half-cell containing a large excess of lithium, the stabilizing effect of polysulfides forming reduced sulfur species
anode-free Cu || Li2S full cell clearly demonstrated the impact on the lithium surface on the reversibility of lithium deposition.
of a limited lithium inventory on electrochemical performance Notwithstanding other issues, this effect bodes very well for the
(Figure 5a). Nevertheless, the Cu || Li2S full cell employing a commercial potential for Li–S batteries.
regular 2 m LiCF3SO3 + 0.1 m LiNO3 in DOL/DME (1:1 v/v) It is important to note here that the Coulombic efficiencies
electrolyte showed remarkable cycling stability, maintaining obtained for the anode-free Cu || Li2S full cell are not a useful
51.5% of its initial capacity over 100 cycles (LIRR = 99.35%). In predictor of the lithium inventory loss rate. CE values with a
contrast, the Li || Li2S half-cell with a large excess of lithium geometric mean of 97.14% over 50 cycles would correspond to
retained 70% of its initial capacity over 100 cycles. The initial 23.4% lithium inventory retention. However, the actual capacity
lithium stripping capacity was 2.33 mA cm−2 and the current retention obtained is 59.4%, corresponding to a lithium inven-
rate was 0.47 mA cm−2. Thus, restricting the initial lithium tory retention rate of almost 99% per cycle over 50 cycles. The
inventory in Li–S batteries has only a limited impact on elec- deviation between measured values of Coulombic efficiencies
trochemical performance. The performance of the unoptimized and the actual values of lithium inventory retention rates can be
anode-free Cu || Li2S full cell is comparable to some of the best attributed to the presence of polysulfide intermediates, which
performing anode-free full cells reported in the literature that behave as soluble redox mediators that shuttle and “leak” charge
employ NMC/LFP cathodes with highly optimized electrolytes. and active material between the two electrodes, effectively
This indicates an intrinsic stabilization of lithium deposition in behaving as an internal short. In systems with soluble redox
Li–S batteries. mediators that shuttle between the cathode and the anode, the
Lithium deposition in Li–S batteries is uniquely affected by lithium inventory loss rate is significantly lower than what the
the soluble polysulfide intermediates (Li2Sx, 2 < x ≤ 8), which measured value of Coulombic efficiency would suggest and has
are formed during the charge/discharge process of the sulfur/ to be estimated from capacity fade.[31,32,75–77] This phenomenon

Adv. Energy Mater. 2020, 2000804 2000804 (11 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

is not observed in anode-free full cells with cathodes such as cells based on Li2S cathodes and bare current collectors enable
LFP or LCO, which do not generate soluble redox mediators. complete circumvention of this problem. As-assembled cells
Future work on anode-free full cells based on any cathode that before any current has been applied are expected to show zero
displays shuttle phenomena such as Li2S or LiI should account self-discharge since the Li2S cathode is already fully discharged
for this discrepancy. and there is no free lithium metal in the system. This aspect
The presence of polysulfide species in the electrolyte of Li–S is a major practical advantage of anode-free Li2S batteries and
batteries also means that lithium deposition in such systems is merits further exploration.
very difficult to model extrinsically. The exact composition and
concentration of the in situ generated polysulfide species in
the electrolyte, which exist in a dynamic equilibrium with each 6. Anode-Free Sodium Batteries
other, the sulfur/Li2S cathode, and the reduced sulfur species
on the lithium surface, are unknown and depend on a number Sodium is an attractive replacement for lithium in various
of factors, including the amount of electrolyte, sulfur loading, rechargeable battery technologies due to its high gravimetric
cycle number, and state of charge.[78–84] Hence, lithium deposi- capacity (1165 mAh g−1), low redox potential (-2.71 V vs
tion in Li–S batteries cannot be accurately modeled using poly- SHE),[90,91] and significantly higher earth-abundance and much
sulfide additives of a fixed composition and concentration, for lower cost compared to lithium. Using sodium as the charge
instance 0.1 m Li2S6, in Cu || Li half cells or Li || Li symmetric carrier ion instead of lithium also confers other advantages,
cells. Furthermore, polysulfide species are highly reactive and including the use of cobalt-free sodium-ion cathodes, replacing
the compatibility of various strategies for lithium stabilization the copper anode current collector with cheaper and lighter alu-
(such as lithium alloy substrates) with the polysulfide-rich elec- minum, and better flame resistance compared to lithium bat-
trolyte is not guaranteed. These factors combine to make the teries.[92–94] One disadvantage of sodium batteries is the lack
anode-free Cu || Li2S full cell configuration an excellent tem- of a suitable anode host material with low redox potential and
plate for i) developing a thorough understanding of lithium good cycling stability.[95,96] Using metallic sodium as the anode
deposition in Li–S batteries and ii) effectively evaluating var- can be a good circumvention to this problem. However, sodium
ious approaches toward improving the reversibility of lithium metal shows very high reactivity to air and moisture, even
plating/stripping in Li–S batteries. more so than lithium. It would be particularly advantageous to
Recently, our group discovered that the addition of a small assemble full cells without any free sodium metal. The anode-
amount of tellurium to the Li2S cathode can have a dramatic free full cell configuration, pairing a fully sodiated cathode with
effect on the reversibility of lithium plating and stripping in a bare current collector for sodium plating and stripping, would
anode-free Ni || Li2S full cells (nickel was used instead of copper yield significant gains in energy density and resolve both of the
as the anode current collector).[85] Polysulfides (Li2Sn) gener- challenges described above.
ated during cell operation react with the elemental tellurium Sodium-metal anodes suffer from many of the same issues
to form polytellurosulfide species (Li2TexSy) that migrate to the as lithium-metal anodes described earlier. The deposition of
anode side and reduce on the lithium surface to form Li2TeS3 sodium metal leads to the growth of highly reactive porous
as a component of the SEI layer. Due to its higher ionic con- mossy deposits with large surface areas, which leads to poor
ductivity, Li2TeS3 is much more effective at stabilizing lithium efficiencies of sodium plating and stripping. Dendritic growth
deposition compared to Li2S, which is formed in a control of sodium metal can similarly occur when sodium ions get
cell without any additive. The initial discharge capacity of the depleted at the electrode surface and diffusion-based charge
anode-free Ni || Li2S full cells was 2.5 mAh cm−2, and they were transport limitations are reached. There has been much less
cycled at C/5 rate or 1 mA cm−2. The addition of tellurium to work on improving the reversibility of sodium-metal anodes
the cathode enabled the anode-free full cell to retain 50% of its compared to their lithium metal counterparts. Nevertheless,
initial capacity for 240 cycles, whereas the control cell cycled the solutions proposed to improve the Coulombic efficiencies
for only 34 cycles at 50% retention. Thus, the lithium inventory of sodium-metal batteries are similar in spirit to those pro-
retention rate was improved from 97.98% per cycle to 99.72% posed for lithium-metal batteries: optimized electrolyte for-
per cycle. It was verified with half cells that the addition of tel- mulations to form favorable interphasial components, coating
lurium has no impact on cathode electrochemical performance. layers to homogenize Na+ ion flux, and modified current col-
This work demonstrates that stabilizing lithium deposition in lector substrates with low nucleation overpotentials for sodium
anode-free sulfur batteries requires the development of novel deposition. However, there are significant differences in the
strategies that account for the unique chemistry of the system. chemical and physical properties of sodium and lithium such
A major challenge with Li–S batteries is the self-discharge that sodium deposition cannot be considered a direct analogue
phenomena inherent to the system.[86–89] When assembled cells of lithium deposition.[97] Sodium ions are larger than lithium
in the Li || S configuration are rested before any current has ions (116 pm vs 90 pm) but are less polarizing due to the lower
been applied, the S8 cathode gets partially reduced by Li+ ions charge density. This is expected to have an impact on the energy
in the electrolyte. This generates soluble polysulfide species, barrier for sodium ion migration through different SEI compo-
which diffuse to the lithium anode and reduce on its surface. nents. Sodium metal is also less reducing than lithium metal.
The generated lower-order polysulfide species shuttle back to The melting point of sodium (98 °C) is significantly lower than
the cathode and catalyze further reduction of the S8 cathode. lithium (181 °C), leading to the possibility of restricting mossy
This self-catalyzing process severely affects the shelf life of Li–S growths with a combination of pressure and temperature con-
batteries and limits their deliverable capacities. Anode-free full trols. Sodium metal exhibits significantly higher reactivity with

Adv. Energy Mater. 2020, 2000804 2000804 (12 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

the commonly used carbonate electrolyte solvents compared


to lithium.[98] At the same time, simple low salt concentra-
tion ether-based electrolytes, such as 1 m NaPF6 in diglyme
and 1 m NaTFSI in diglyme, have been identified that show
excellent Coulombic efficiencies for sodium plating and strip-
ping.[99] These electrolytes behave quite distinctly from their
lithium counterparts, and comparable Coulombic efficiencies
for lithium plating and stripping can typically only be achieved
with highly optimized high-concentration electrolytes.[100] The
retention of sodium ions during cycling can be accounted for
by the sodium inventory retention rate (NIRR), defined identi-
cally to LIRR.
In 2017, Cohn et al. demonstrated an anode-free sodium bat-
tery employing a sodiated iron pyrite cathode (Na1.5FeS2) paired
with an aluminum foil coated with a thin layer of carbon to
provide a low interfacial energy surface for sodium deposi-
tion.[25] The thin layer of carbon is shown to have a low sodia-
tion capacity of 0.05 mA cm−2, and enables the sodium metal
Figure 6. A visual summary of the investigations conducted thus far into
to plate as large, flat, and smooth hexagonal islands. With a 1 m anode-free sodium-ion cells with the associated NIRR.
NaPF6 in diglyme electrolyte, the authors demonstrated stable
capacities for 40 cycles. The initial sodium plating capacity was
1.7 mAh cm−2, although the applied currents were quite low at rate and a 1.26 mAh cm−2 sodium plating capacity. Importantly,
0.125 mA cm−2. Tang et al. have demonstrated that coating a the work employed no presodiation treatments on the current
thin 50 nm “sodiophilic” gold layer on the Cu foil current col- collectors and exhibited low first cycle losses. This work dem-
lector has a beneficial impact on the reversibility of sodium onstrates that the combination of a favorable nucleation layer
deposition. By forming Au2Na alloy upon initial sodiation, the with a favorable electrolyte such as 1 m NaPF6 in diglyme can
nucleation energy for subsequent sodium plating could be signif- enable excellent cyclability in anode-free systems. It is expected
icantly reduced.[101] The initial sodium plating capacity was about that further improvements in cyclability can be delivered with
0.5 mAh cm−2 and the current rate was 0.2 mA cm−2 for charge modified electrolytes (e.g., additives) and systematic application
and 2 mA cm−2 for discharge. In an anode-free full cell with a of more optimized formation procedures and cycling protocols
sodiated iron pyrite cathode paired with the Au/Cu current col- (e.g., pressure). A visual summary of all the work conducted
lector and using 1 m NaCF3SO3 in diglyme electrolyte, 55% of the into anode-free sodium batteries is presented in Figure 6.
initial capacity was maintained after 50 cycles (NIRR = 98.8%).
Rudola et al. demonstrated an anode-free Cu || Na2Fe2(CN)6
full cell with an initial sodium plating capacity of 1.14 mAh cm−2 7. Anode-Free Solid-State Batteries
and an applied current of 0.18 mA cm−2 during charge and
0.9 mA cm−2 during discharge.[102] The cell delivered 76% capacity One of the fundamental limitations of lithium-metal batteries
retention at 100 cycles with a 1 m NaBF4 in tetraglyme electrolyte employing liquid electrolytes, including the anode-free full
(NIRR = 99.7%). The cell was subject to initial formation cycling cells discussed throughout this paper, is the potential for den-
in a shallow voltage window of 3.25–2.0 V for 20 cycles to elimi- dritic lithium growth leading to internal shorting of the cell and
nate irreversible losses, before starting regular cycling between catastrophic cell failure. This has generated widespread interest
3.9 and 2.0 V. Mazzali et al. demonstrated an anode-free sodium in all-solid-state lithium-metal batteries as a means of circum-
battery based on sodiated iron hexacyanoferrate (Prussian Blue) venting the safety concerns associated with a flammable organic
as the cathode active material paired with a carbon-coated alu- liquid electrolyte.[105,106] By inhibiting the growth of lithium den-
minum foil.[103] Using 1 m NaPF6 in diglyme as the electrolyte, drites, solid-state electrolytes (SSEs) eliminate the risk of internal
they showed a first cycle loss of ≈40%, but subsequently retained shorting of the cell and consequent thermal runaway.[107,108] The
90% of their second cycle capacity at 150 cycles (NIRR = 99.6%), transference number of SSEs is close to 1, which also helps
albeit with a low sodium plating capacity of 0.17 mAh cm−2 and impede the formation of lithium dendrites.[109,110] Solid-state
an applied current of 0.22 mA cm−2. electrolytes can also enable higher energy density due to the
Cohn et al. have also demonstrated anode-free sodium bat- elimination of the polymer separator and the need for less elec-
teries pairing a sodiated Na3V2(PO4)3 cathode with an alu- trolyte loadings compared to liquid electrolytes.[26] An anode-free
minum current collector coated with various thin nucleation full cell employing SSEs would enable a dramatic enhancement
layers.[104] Among the different nongraphitic carbons and in achievable energy density. The decomposition of liquid elec-
sodium-alloying metals evaluated as nucleation layers, carbon trolytes leads to substantial gas evolution, which has a severely
black and bismuth showed the highest Coulombic efficiencies detrimental impact on electrochemical performance.[64,111] In
for sodium plating and stripping. With a 1 m NaPF6 in diglyme contrast, SSEs show limited gas evolution.[112] SSEs are espe-
electrolyte and carbon black as the nucleation layer, the authors cially promising for systems that undergo considerable active-
demonstrated a remarkable capacity retention of 82.5% after material crossover, such as NMC, spinel, and sulfur cathodes, as
100 cycles (NIRR = 99.8%) at a reasonable 0.5 mA cm−2 current they prevent the formation and passage of such species.

Adv. Energy Mater. 2020, 2000804 2000804 (13 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

Despite these advantages, the development of all-solid-state solid-state anode-free batteries in large-area pouch cell format.
lithium-metal batteries faces a number of challenges. Tra- With a cathode capacity of 6.8 mAh cm−2 and an applied cur-
ditional oxide-based solid-state electrolytes suffer from rela- rent density of 0.5C (=3.4 mA cm−2), 89% of the initial discharge
tively poor ionic conductivity at room temperature. Even opti- capacity was retained after 1000 cycles. This corresponds to a
mized SSE materials such as LIPON, perovskite-type LLTO, lithium inventory retention rate of 99.988%, which approaches
NASICON-type LAGP and LATP, and garnet-type LLZO display the Coulombic efficiency of commercial Li-ion batteries with
ionic conductivities that fall short of their liquid electrolyte graphite anodes. The pouch cell was pressurized to 490 MPa
counterparts.[113–115] This inhibits cell operation at acceptable during cell fabrication with a warm isostatic pressure system.
current densities with high cathode loadings. Hence, consider- During cell operation, the temperature was maintained at 60°C
able efforts have been devoted toward improving the ionic con- and a uniaxial stack pressure of 2– 4 MPa was applied. The
ductivity of SSEs. Over the last decade, sulfide-based solid elec- application of pressure and temperature controls is necessary
trolytes have been developed which show ionic conductivities for maintaining an intimate low-impedance contact between the
approaching liquid electrolytes at room temperature.[73,74,116,117] solid electrolyte and the deposited lithium metal and ensuring
Solid electrolytes with a sulfide anionic framework show facile ionic diffusion through the interfacial layers.
lower Li+-ion diffusion barriers compared to solid electrolytes The performance shown in this paper far exceeds any reported
with an oxide anionic framework due to the larger size and work on anode-free full cells and sets a new benchmark for what
higher polarizability of the sulfide anion. Another advantage is achievable with this configuration. It also reveals some of the
of sulfide-based SSEs is that the grain boundary resistance can design considerations that are necessary for achieving revers-
be reduced by a simple cold-pressing technique. In contrast, ible lithium deposition in solid-state anode-free batteries. The
oxide-based SSEs such as garnet-type LLZO require a high- solid electrolyte chosen should possess high ionic conductivity at
temperature sintering process for reduction of grain boundary room temperature and demonstrate excellent compatibility with
resistance.[117] Nevertheless, sulfide-based solid electrolytes metallic lithium. Engineering a suitable buffer layer between the
suffer from a number of disadvantages. Their high moisture solid electrolyte and the anode current collector is critical, as it
sensitivity leads to the formation of toxic H2S gas, and hence plays a vital role in ensuring dense and uniform lithium plating
material preparation and cell assembly have to be done in an while maintaining stable interfacial contact between the solid
inert atmosphere. Sulfide-based SSEs also display limited elec- electrolyte and deposited lithium during cell operation. Finally,
trochemical stability windows. They tend to reduce upon con- the application of pressure during cell assembly and operation
tact with metallic lithium to form a resistive decomposition is necessary for achieving optimal electrochemical performance.
layer at the lithium interface.[118] Argyrodite-type solid electro-
lytes, in which one of the sulfide anions is replaced by a halide
(Cl, Br, I), demonstrate significantly better compatibility with 8. Conclusions and Outlook
lithium metal.[119,120] A significant challenge with all-solid-state
lithium-metal batteries is achieving a conformal low-impedance The anode-free full cell configuration is an important mile-
interfacial contact between the solid electrolyte and lithium stone in the development of lithium-metal batteries. For any
metal. Furthermore, maintaining such an intimate contact over given lithium-containing cathode system, the anode-free full cell
extended cycling with the large volume changes that accom- delivers the highest possible energy density due to the complete
pany plating and stripping of lithium is even more difficult. elimination of all anode active material and utilization of the
Modifying the lithium–electrolyte interface with an artificial maximum possible voltage output from the cathode. Given the
SEI layer is a possible solution. These buffer layers could be high specific volume of lithium (1.87 cm3 g−1), the gains in volu-
composed of an organic polymer coating, a secondary inorganic metric energy density can be particularly significant. From a prac-
solid electrolyte, or even lithiophilic materials that alloy with tical perspective, the elimination of free lithium metal and facile
lithium, such as Si, Ge, and Au.[107] The use of polymer–ceramic cell assembly in the discharged state is a significant advantage.
composite solid electrolytes, which demonstrate significantly As we have discussed, anode-free cells can enable the diag-
better wettability with metallic lithium, may also be considered. nostic evaluation of capacity-loss mechanisms in their truest
Recently, Lee et al. demonstrated a solid-state battery in the sense, as lithium inventory in the cell can be directly monitored
anode-free full cell with a LiNi0.90Co0.05Mn0.05O2 (NMC) cathode over the course of cycling due to the lack of excess lithium. In
and an argyrodite-type Li6PS5Cl sulfide solid electrolyte.[121] this review, we have introduced the metric of lithium inven-
The solid electrolyte layer was uniformly applied onto the tory retention rate, or LIRR, to best quantify this loss across
cathode by employing pressure using a warm isostatic press. the broad range of work currently published in the literature.
The NMC cathode was coated with a thin layer of Li2O–ZrO2 Tables 1 and 2 summarize some of the important work reviewed
coating to maintain interfacial stability with the solid electro- in this paper on lithium-based and sodium-based anode-free
lyte. On the anode side, a 5 µm thick nanocomposite layer full cells, respectively, and compare the inventory retention
composed of lithiophilic Ag nanoparticles and carbon black rates based on the reported electrochemical performance.
was used as the buffer layer between the solid electrolyte and For meaningful progress, it is important to determine cer-
the stainless-steel current collector. The nanocomposite buffer tain cell design, testing, and performance parameters that are
layer enabled uniform, dense and dendrite-free lithium depo- uniformly applied to studies in this field. In particular, the theo-
sition on the current collector substrate and acted as a stable retical capacity of the cell, the first cycle lithium stripping (or
interface between the solid electrolyte and lithium metal. The discharge) capacity, the capacity retention at a specific number
authors demonstrated truly remarkable performance with their of cycles (say 100), or the number of cycles taken to reach a

Adv. Energy Mater. 2020, 2000804 2000804 (14 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

Table 1. A summary of the published work on lithium-based anode-free full cells discussed in this article.

Ref. Year Anode Cathode Electrolyte Initial areal capacity Current rate LIRR [%]
[mAh cm−2] [mA cm−2]
[22] 2016 Cu LFP 4 m LiFSI in DME 1.7 0.2 98.98
[52] 2017 Cu LFP 5% VC + 1 m LiPF6 in EC/EMC (3:7, v/v) 2 0.1 (charge), 95.07
0.5 (discharge)
[60] 2017 Sn–Cu NCA 1 m LiPF6 in FEC:EMC (1:4, wt/wt) 1 0.5 97.62
[24] 2018 Cu Li2S 2 m LiCF3SO3 + 0.1 m LiNO3 in DOL/DME (1:1, v/v) 2.33 0.47 99.35
[53] 2018 Cu@ PEO LFP 1 m LiTFSI + 2% LiNO3 in DOL:DME (1:1, v/v) 0.71 0.14 99.11
[47] 2018 Cu NMC 811 1 m LiPF6 in FEC:FEMC:HFE (2:6:2, by weight) 2.75 0.5 96.60
[54] 2019 Cu@ MLG LFP 1 m LiTFSI + 2% LiNO3 in DOL:DME (1:1, v/v) 1.56 0.2 99.50
[46] 2019 Cu NMC 622 2.3 m LiTFSI and 4.6 m LiFSI in DME 1.44 0.144 (charge), 98.90
0.48 (discharge)
[44] 2019 Cu NMC 111 1 m LiPF6 in EC:DEC:FEC (1:1:2, v/v) 2 0.2 98.18
[68] 2019 Cu NMC 532 1 m LiPF6 in FEC:TFEC (1:2, v/v) 3 0.6 (charge), 99.08
1.5 (discharge)
[42] 2019 Cu LFP 2 m LiFSI + 1 m LiTFSI in DOL/DME (1:1, v/v) 1.6 0.2 98.62
[50] 2019 Cu NMC 111 2 wt% KPF6 + 2 vol% TMSP + 1 m LiPF6 in 1.7 0.2 96.40
EC:DEC (1:1, v/v)
[23] 2019 Cu NMC 111 1 m LiDFOB + 0.2 m LiBF4 in FEC/DEC (1:2, v/v) 2.4 0.48 (charge), 99.75
1.2 (discharge)
[51] 2019 Cu LFP 0.2 m LiNO3 + 1 m LiDFOB in TEP:EC:DMC 1 0.1 (charge), 97.90
(8.4:8.4:83.2, v/v) 0.4 (discharge)
[49] 2019 Cu NMC 111 0.5 m KNO3 + 1 m LiPF6 in EC:DEC (1:1, v/v) 1.7 0.2 98.18
[67] 2019 Cu LFP 3 m LiFSI in DOL:DME (1:1, v/v) 1.7 0.1 and 0.5 (formation), 99.11
1 (cycling)
[55] 2019 Cu@ LLCZN/PVDF NMC 111 1 m LiPF6 in EC:DEC (1:1, v/v) 2.13 0.2 98.24
[69] 2019 Cu NMC 532 1.8 m LiDFOB + 0.4 m LiBF4 in FEC:DEC (1:2, v/v). 2.5 0.5 (charge), 99.89
1.25 (discharge)
[45] 2019 Cu LFP 1.86 m LiTFSI in EC 3.5 1 97.70
[56] 2020 Cu@ GO NMC 111 1 m LiPF6 + 5% FEC in EC:DEC (1:1, v/v) 2 0.2 98.37
[121] 2020 SS@ Ag-C NMC 900505 Solid Li6PS5Cl 6.8 0.68 and 1.36 (formation), 99.99
3.4 (cycling)
[85] 2020 Ni Li2S + 0.1 Te 1 m LiTFSI + 0.1 m LiNO3 in DOL/DME (1:1, v/v) 2.5 1 99.72

given capacity retention (say 50%), and the applied current ensure that the anode-free full cells are operating in a lithium-
densities should be clearly reported. The parameters described limited regime during cycling. Where possible, the cathodes
earlier should be used to calculate the value of the LIRR that should be cycled to 100% depth of discharge. Any lithiation
can be used for comparison across different works. Any initial capacities associated with the current collector substrate should
formation steps that are employed should be clearly described. also be clearly reported. It should be attempted to report the
Reasonable plating/stripping capacities (>2 mAh cm−2) and electrolyte amount employed in the cell, in units of mg mAh−1.
current densities (>1 mA cm−2) should also be employed. The Finally, reasonable control cells should be devised to demon-
performance of equivalent half cells may be reported in order to strate the performance improvements.

Table 2. A summary of the published work on sodium-based anode-free full cells discussed in this article.

Refs. Year Anode Cathode Electrolyte Initial areal capacity [mAh cm−2] Current rate [mA cm−2] NIRR [%]
[25] 2017 Al@C Na1.5FeS2 1 m NaPF6 in diglyme 1.7 0.125 N/A
[102] 2018 Cu Na2Fe2(CN)6 1 m NaBF4 in tetraglyme electrolyte 1.14 0.18 (charge), 0.9 (discharge) 99.73
[101] 2018 Cu@Au Na1.3FeS2 1 m NaCF3SO3 in diglyme 0.5 0.2 (charge), 2 (discharge) 98.81
[104] 2018 Al@C Na3V2(PO4)3 1 m NaPF6 in diglyme 1.26 0.5 99.81
[103] 2019 Al@C Na2Fe2(CN)6 1 m NaPF6 in diglyme 0.17 0.22 99.60

Adv. Energy Mater. 2020, 2000804 2000804 (15 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

Given the tremendous progress and insights showcased [1] A. Manthiram, Nat. Commun. 2020, 11, 1550.
throughout the scientific work published in this nascent [2] G. Zubi, R. Dufo-López, M. Carvalho, G. Pasaoglu, Renewable Sus-
area thus far, there is an encouraging path forward for tainable Energy Rev. 2018, 89, 292.
[3] M. M. Thackeray, C. Wolverton, E. D. Isaacs, Energy Environ. Sci.
future developments. Further optimization of the liquid
2012, 5, 7854.
electrolyte, anode-side current collector, and formation and [4] P. G. Bruce, L. J. Hardwick, K. M. Abraham, MRS Bull. 2011, 36, 506.
cycling parameters will continue leading the charge for [5] W. Li, E. M. Erickson, A. Manthiram, Nat. Energy 2020, 5, 26.
further progress in increasing LIRR to the levels needed [6] A. Manthiram, J. C. Knight, S. T. Myung, S. M. Oh, Y. K. Sun, Adv.
to ensure commercial deployment (a LIRR approaching Energy Mater. 2016, 6, 1501010.
99.99%). For the development of optimized electrolyte for- [7] J. Xu, F. Lin, M. M. Doeff, W. Tong, J. Mater. Chem. A 2017, 5, 874.
mulations, a clear understanding of the SEI layer formed [8] J. B. Goodenough, Nat. Electron. 2018, 1, 204.
on the lithium surface and the role played by individual SEI [9] A. Yoshino, Angew. Chem., Int. Ed. 2012, 51, 5798.
components is necessary. The formation of high ionic con- [10] M. S. Whittingham, Chem. Rev. 2004, 104, 4271.
ductivity interphasial species such as Li3PS4 and high sur- [11] S. Chae, S. H. Choi, N. Kim, J. Sung, J. Cho, Angew. Chem., Int. Ed.
2020, 59, 110.
face energy interphasial species such as LiF is particularly
[12] J. Wu, Y. Cao, H. Zhao, J. Mao, Z. Guo, Carbon Energy 2019, 1, 57.
effective at improving lithium cycling efficiency. Hence, [13] X. B. Cheng, R. Zhang, C. Z. Zhao, Q. Zhang, Chem. Rev. 2017, 117, 10403.
the use of fluorinated electrolytes and mixed polysulfide [14] W. Xu, J. Wang, F. Ding, X. Chen, E. Nasybulin, Y. Zhang,
additives may be considered. Localized high-concentration J.-G. Zhang, Energy Environ. Sci. 2014, 7, 513.
electrolytes, which combine the low viscosity and high wet- [15] D. Lin, Y. Liu, Y. Cui, Nat. Nanotechnol. 2017, 12, 194.
tability of regular electrolytes with the lithium stabilization [16] C. Fang, X. Wang, Y. S. Meng, Trends Chem. 2019, 1, 152.
of concentrated electrolytes, may be promising. Modifying [17] J. Qian, W. A. Henderson, W. Xu, P. Bhattacharya, M. Engelhard,
the anode-side current collector with lithiophilic materials O. Borodin, J. G. Zhang, Nat. Commun. 2015, 6, 6362.
is another useful strategy for improving the reversibility [18] P. Bai, J. Li, F. R. Brushett, M. Z. Bazant, Energy Environ. Sci. 2016,
of lithium deposition. Finally, the application of pressure 9, 3221.
[19] J. Xiang, L. Yang, L. Yuan, K. Yuan, Y. Zhang, Y. Huang, J. Lin,
is necessary to achieve high lithium inventory retention
F. Pan, Y. Huang, Joule 2019, 3, 2334.
rates and needs to be explored further. However, substan- [20] H. Wang, Y. Liu, Y. Li, Y. Cui, Electrochem. Energy Rev. 2019, 2, 509.
tial improvements will likely not occur in isolation, and the [21] H. Yang, C. Guo, A. Naveed, J. Lei, J. Yang, Y. Nuli, J. Wang, Energy
onus will be on the entire community to integrate devel- Storage Mater. 2018, 14, 199.
opments across a broad range of strategies to ensure the [22] J. Qian, B. D. Adams, J. Zheng, W. Xu, W. A. Henderson, J. Wang,
greatest retention of lithium inventory. Furthermore, these M. E. Bowden, S. Xu, J. Hu, J.-G. Zhang, Adv. Funct. Mater. 2016,
developments will need to be proven with practically rele- 26, 7094.
vant cell parameters, including large-area pouch cells and [23] R. Weber, M. Genovese, A. J. Louli, S. Hames, C. Martin, I. G. Hill,
lean electrolyte conditions to ensure the scalability of the J. R. Dahn, Nat. Energy 2019, 4, 683.
technology. The possibility of such a drastic increase in vol- [24] S. Nanda, A. Gupta, A. Manthiram, Adv. Energy Mater. 2018, 8,
1801556.
umetric and gravimetric energy densities could enable the
[25] A. P. Cohn, N. Muralidharan, R. Carter, K. Share, C. L. Pint, Nano
largest step-change in energy storage technology seen since Lett. 2017, 17, 1296.
the introduction of the lithium-ion battery, and we look for- [26] J. Betz, G. Bieker, P. Meister, T. Placke, M. Winter, R. Schmuch,
ward to seeing developments in this area moving forward. Adv. Energy Mater. 2019, 9, 1803170.
[27] O. Mashtalir, M. Nguyen, E. Bodoin, L. Swonger, S. P. O’Brien,
ACS Omega 2018, 3, 181.
[28] J. Becking, A. Gröbmeyer, M. Kolek, U. Rodehorst, S. Schulze,
Acknowledgements M. Winter, P. Bieker, M. C. Stan, Adv. Mater. Interfaces 2017, 4,
This work was supported by the Assistant Secretary for Energy 1700166.
Efficiency and Renewable Energy, Office of Vehicle Technologies of the [29] C. Fang, J. Li, M. Zhang, Y. Zhang, F. Yang, J. Z. Lee, M.-H. Lee,
U.S. Department of Energy through the Advanced Battery Materials J. Alvarado, M. A. Schroeder, Y. Yang, B. Lu, N. Williams, M. Ceja,
Research (BMR) Program (Battery500 Consortium) award number L. Yang, M. Cai, J. Gu, K. Xu, X. Wang, Y. S. Meng, Nature 2019,
DE-EE0007762. 572, 511.
[30] P. Barai, K. Higa, V. Srinivasan, J. Electrochem. Soc. 2018, 165,
A2654.
[31] J. Xu, R. D. Deshpande, J. Pan, Y.-T. Cheng, V. S. Battaglia, J. Elec-
Conflict of Interest trochem. Soc. 2015, 162, A2026.
The authors declare no conflict of interest. [32] S. Li, M. Jiang, Y. Xie, H. Xu, J. Jia, J. Li, Adv. Mater. 2018, 30,
1706375.
[33] M. Genovese, A. J. Louli, R. Weber, S. Hames, J. R. Dahn, J. Elec-
trochem. Soc. 2018, 165, A3321.
Keywords [34] W. Li, U. H. Kim, A. Dolocan, Y. K. Sun, A. Manthiram, ACS Nano
anode-free cells, full cells, lithium batteries, lithium deposition, lithium- 2017, 11, 5853.
metal anodes [35] H. Lee, H. S. Lim, X. Ren, L. Yu, M. H. Engelhard, K. S. Han,
J. Lee, H. T. Kim, J. Xiao, J. Liu, W. Xu, J. G. Zhang, ACS Energy
Received: March 2, 2020 Lett. 2018, 3, 2921.
Revised: April 17, 2020 [36] X.-B. Cheng, C. Yan, H.-J. Peng, J.-Q. Huang, S.-T. Yang, Q. Zhang,
Published online: Energy Storage Mater. 2018, 10, 199.

Adv. Energy Mater. 2020, 2000804 2000804 (16 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

[37] X.-B. Cheng, J.-Q. Huang, Q. Zhang, J. Electrochem. Soc. 2018, 165, [67] T. T. Beyene, B. A. Jote, Z. T. Wondimkun, B. W. Olbassa,
A6058. C.-J. Huang, B. Thirumalraj, C.-H. Wang, W.-N. Su, H. Dai,
[38] R. Weber, J.-H. Cheng, A. J. Louli, M. Coon, S. Hy, J. R. Dahn, B.-J. Hwang, ACS Appl. Mater. Interfaces 2019, 11, 31962.
J. Electrochem. Soc. 2019, 166, A3250. [68] A. J. Louli, M. Genovese, R. Weber, S. G. Hames, E. R. Logan,
[39] S. Chen, J. Zheng, D. Mei, K. S. Han, M. H. Engelhard, W. Zhao, J. R. Dahn, J. Electrochem. Soc. 2019, 166, A1291.
W. Xu, J. Liu, J. G. Zhang, Adv. Mater. 2018, 30, 1706102. [69] M. Genovese, A. J. Louli, R. Weber, C. Martin, T. Taskovic,
[40] B. J. Neudecker, N. J. Dudney, J. B. Bates, J. Electrochem. Soc. 2000, J. R. Dahn, J. Electrochem. Soc. 2019, 166, A3342.
147, 517. [70] C.-Z. Zhao, X.-B. Cheng, R. Zhang, H.-J. Peng, J.-Q. Huang, R. Ran,
[41] O. Borodin, J. Self, K. A. Persson, C. Wang, K. Xu, Joule 2020, 4, 69. Z.-H. Huang, F. Wei, Q. Zhang, Energy Storage Mater. 2016, 3, 77.
[42] T. T. Beyene, H. K. Bezabh, M. A. Weret, T. M. Hagos, C. J. Huang, [71] W. Li, H. Yao, K. Yan, G. Zheng, Z. Liang, Y.-M. Chiang, Y. Cui, Nat.
C. H. Wang, W. N. Su, H. Dai, B. J. Hwang, J. Electrochem. Soc. Commun. 2015, 6, 7436.
2019, 166, A1501. [72] Z. Xu, R. Chen, H. Zhu, J. Mater. Chem. A 2019, 7, 12645.
[43] R. Rodriguez, K. E. Loeffler, R. A. Edison, R. M. Stephens, [73] S. Chen, D. Xie, G. Liu, J. P. Mwizerwa, Q. Zhang, Y. Zhao, X. Xu,
A. Dolocan, A. Heller, C. B. Mullins, ACS Appl. Energy Mater. 2018, X. Yao, Energy Storage Mater. 2018, 14, 58.
1, 5830. [74] J. Lau, R. H. DeBlock, D. M. Butts, D. S. Ashby, C. S. Choi,
[44] T. T. Hagos, B. Thirumalraj, C.-J. Huang, L. H. Abrha, T. M. Hagos, B. S. Dunn, Adv. Energy Mater. 2018, 8, 1800933.
G. B. Berhe, H. K. Bezabh, J. Cherng, S.-F. Chiu, W.-N. Su, [75] H. D. Lim, B. Lee, Y. Zheng, J. Hong, J. Kim, H. Gwon, Y. Ko,
B.-J. Hwang, ACS Appl. Mater. Interfaces 2019, 11, 9955. M. Lee, K. Cho, K. Kang, Nat. Energy 2016, 1, 16066.
[45] V. Nilsson, A. Kotronia, M. Lacey, K. Edström, P. Johansson, ACS [76] S. Zhang, K. Zhao, T. Zhu, J. Li, Prog. Mater. Sci. 2017, 89, 479.
Appl. Energy Mater. 2020, 3, 200. [77] Q. Zhao, Y. Lu, Z. Zhu, Z. Tao, J. Chen, Nano Lett. 2015, 15, 5982.
[46] J. Alvarado, M. A. Schroeder, T. P. Pollard, X. Wang, J. Z. Lee, [78] L. E. Camacho-Forero, T. W. Smith, S. Bertolini, P. B. Balbuena,
M. Zhang, T. Wynn, M. Ding, O. Borodin, Y. S. Meng, K. Xu, J. Phys. Chem. C 2015, 119, 26828.
Energy Environ. Sci. 2019, 12, 780. [79] A. Gupta, A. Bhargav, A. Manthiram, Adv. Energy Mater. 2019, 9,
[47] X. Fan, L. Chen, O. Borodin, X. Ji, J. Chen, S. Hou, T. Deng, 1803096.
J. Zheng, C. Yang, S. C. Liou, K. Amine, K. Xu, C. Wang, Nat. [80] C. Dillard, A. Singh, V. Kalra, J. Phys. Chem. C 2018, 122, 18195.
Nanotechnol. 2018, 13, 715. [81] C. Park, A. Ronneburg, S. Risse, M. Ballauff, M. Kanduč,
[48] J. G. Zhang, Nat. Energy 2019, 4, 637. J. Dzubiella, J. Phys. Chem. C 2019, 123, 10167.
[49] N. A. Sahalie, A. A. Assegie, W. N. Su, Z. T. Wondimkun, B. A. Jote, [82] D. Zheng, X.-Q. Yang, D. Qu, ChemSusChem 2016, 9, 2348.
B. Thirumalraj, C. J. Huang, Y. W. Yang, B. J. Hwang, J. Power [83] S. S. Zhang, J. Power Sources 2016, 322, 99.
Sources 2019, 437, 226912. [84] C. Zu, A. Manthiram, J. Phys. Chem. Lett. 2014, 5, 2522.
[50] T. M. Hagos, G. B. Berhe, T. T. Hagos, H. K. Bezabh, L. H. Abrha, [85] S. Nanda, A. Bhargav, A. Manthiram, Joule 2020, 4, 1.
T. T. Beyene, C. J. Huang, Y. W. Yang, W. N. Su, H. Dai, B. J. Hwang, [86] S.-H. Chung, A. Manthiram, Adv. Mater. 2018, 30, 1705951.
Electrochim. Acta 2019, 316, 52. [87] H. S. Ryu, H. J. Ahn, K. W. Kim, J. H. Ahn, K. K. Cho, T. H. Nam,
[51] Z. L. Brown, S. Heiskanen, B. L. Lucht, J. Electrochem. Soc. 2019, Electrochim. Acta 2006, 52, 1563.
166, A2523. [88] S. H. Chung, A. Manthiram, ACS Energy Lett. 2017, 2, 1056.
[52] Z. L. Brown, S. Jurng, B. L. Lucht, J. Electrochem. Soc. 2017, 164, [89] L. Wang, J. Liu, S. Yuan, Y. Wang, Y. Xia, Energy Environ. Sci. 2016,
A2186. 9, 224.
[53] A. A. Assegie, J. H. Cheng, L. M. Kuo, W. N. Su, B. J. Hwang, [90] B. Sun, P. Xiong, U. Maitra, D. Langsdorf, K. Yan, C. Wang,
Nanoscale 2018, 10, 6125. J. Janek, D. Schröder, G. Wang, Adv. Mater. 2019, 1903891, https://
[54] A. A. Assegie, C. C. Chung, M. C. Tsai, W. N. Su, C. W. Chen, doi.org/10.1002/adma.201903891.
B. J. Hwang, Nanoscale 2019, 11, 2710. [91] X. Zheng, C. Bommier, W. Luo, L. Jiang, Y. Hao, Y. Huang, Energy
[55] L. H. Abrha, T. A. Zegeye, T. T. Hagos, H. Sutiono, T. M. Hagos, Storage Mater. 2019, 16, 6.
G. B. Berhe, C. J. Huang, S. K. Jiang, W. N. Su, Y. W. Yang, [92] C. Vaalma, D. Buchholz, M. Weil, S. Passerini, Nat. Rev. Mater.
B. J. Hwang, Electrochim. Acta 2019, 325, 134825. 2018, 3, 18013.
[56] Z. T. Wondimkun, T. T. Beyene, M. A. Weret, N. A. Sahalie, [93] Y. Lyu, Y. Liu, Z. E. Yu, N. Su, Y. Liu, W. Li, Q. Li, B. Guo, B. Liu,
C. J. Huang, B. Thirumalraj, B. A. Jote, D. Wang, W. N. Su, Sustainable Mater. Technol. 2019, 21, e00098.
C. H. Wang, G. Brunklaus, M. Winter, B. J. Hwang, J. Power Sources [94] J. Peters, A. Peña Cruz, M. Weil, Batteries 2019, 5, 10.
2020, 450, 227589. [95] W. Zhang, F. Zhang, F. Ming, H. N. Alshareef, EnergyChem 2019, 1,
[57] R. Xu, X.-B. Cheng, C. Yan, X.-Q. Zhang, Y. Xiao, C.-Z. Zhao, 100012.
J.-Q. Huang, Q. Zhang, Matter 2019, 1, 317. [96] D.-Y. Kim, D.-H. Kim, S.-H. Kim, E.-K. Lee, S.-K. Park, J.-W. Lee,
[58] X. Liang, Q. Pang, I. R. Kochetkov, M. S. Sempere, H. Huang, Y.-S. Yun, S.-Y. Choi, J. Kang, Nanomaterials 2019, 9, 793.
X. Sun, L. F. Nazar, Nat. Energy 2017, 2, 17119. [97] B. Lee, E. Paek, D. Mitlin, S. W. Lee, Chem. Rev. 2019, 119, 5416.
[59] K. Yan, Z. Lu, H. W. Lee, F. Xiong, P. C. Hsu, Y. Li, J. Zhao, S. Chu, [98] D. I. Iermakova, R. Dugas, M. R. Palacín, A. Ponrouch, J. Electro-
Y. Cui, Nat. Energy 2016, 1, 16010. chem. Soc. 2015, 162, A7060.
[60] S. S. Zhang, X. Fan, C. Wang, Electrochim. Acta 2017, 258, 1201. [99] Z. W. Seh, J. Sun, Y. Sun, Y. Cui, ACS Cent. Sci. 2015, 1, 449.
[61] S. S. Zhang, X. Fan, C. Wang, Electrochem. Commun. 2018, 89, 23. [100] Q. Zhang, Y. Lu, L. Miao, Q. Zhao, K. Xia, J. Liang, S. L. Chou,
[62] V. Pande, V. Viswanathan, ACS Energy Lett. 2019, 4, 2952. J. Chen, Angew. Chem., Int. Ed. 2018, 57, 14796.
[63] J. R. Belt, C. D. Ho, T. J. Miller, M. A. Habib, T. Q. Duong, J. Power [101] S. Tang, Z. Qiu, X. Y. Wang, Y. Gu, X. G. Zhang, W. W. Wang,
Sources 2005, 142, 354. J. W. Yan, M. Sen Zheng, Q. F. Dong, B. W. Mao, Nano Energy
[64] M. Broussely, P. Biensan, F. Bonhomme, P. Blanchard, S. Herreyre, 2018, 48, 101.
K. Nechev, R. J. Staniewicz, J. Power Sources 2005, 146, 90. [102] A. Rudola, S. R. Gajjela, P. Balaya, Electrochem. Commun. 2018, 86,
[65] J. Wang, W. Huang, A. Pei, Y. Li, F. Shi, X. Yu, Y. Cui, Nat. Energy 157.
2019, 4, 664. [103] F. Mazzali, M. W. Orzech, A. Adomkevicius, A. Pisanu, L. Malavasi,
[66] C. Niu, H. Lee, S. Chen, Q. Li, J. Du, W. Xu, J. G. Zhang, D. Deganello, S. Margadonna, ACS Appl. Energy Mater. 2019, 2,
M. S. Whittingham, J. Xiao, J. Liu, Nat. Energy 2019, 4, 551. 344.

Adv. Energy Mater. 2020, 2000804 2000804 (17 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advenergymat.de

[104] A. P. Cohn, T. Metke, J. Donohue, N. Muralidharan, K. Share, [114] F. Zheng, M. Kotobuki, S. Song, M. O. Lai, L. Lu, J. Power Sources
C. L. Pint, J. Mater. Chem. A 2018, 6, 23875. 2018, 389, 198.
[105] N. J. Dudney, W. C. West, J. Nanda, Mater. Energy 2015, 6, 836. [115] Y.-Z. Sun, J.-Q. Huang, C.-Z. Zhao, Q. Zhang, Sci. China: Chem.
[106] J. Schnell, T. Günther, T. Knoche, C. Vieider, L. Köhler, A. Just, 2017, 60, 1508.
M. Keller, S. Passerini, G. Reinhart, J. Power Sources 2018, [116] D. Liu, W. Zhu, Z. Feng, A. Guerfi, A. Vijh, K. Zaghib, Mater. Sci.
382, 160. Eng., B 2016, 213, 169.
[107] S. Xia, X. Wu, Z. Zhang, Y. Cui, W. Liu, Chem 2019, 5, 753. [117] Q. Zhang, D. Cao, Y. Ma, A. Natan, P. Aurora, H. Zhu, Adv. Mater.
[108] A. Manthiram, X. Yu, S. Wang, Nat. Rev. Mater. 2017, 2, 16103. 2019, 31, 1901131.
[109] M. Doyle, T. F. Fuller, J. Newman, Electrochim. Acta 1994, 39, 2073. [118] T. Swamy, X. Chen, Y. M. Chiang, Chem. Mater. 2019, 31, 707.
[110] K. Xu, Chem. Rev. 2014, 114, 11503. [119] F. Zhao, J. Liang, C. Yu, Q. Sun, X. Li, K. Adair, C. Wang, Y. Zhao,
[111] A. Jozwiuk, B. B. Berkes, T. Weiß, H. Sommer, J. Janek, S. Zhang, W. Li, S. Deng, R. Li, Y. Huang, H. Huang, L. Zhang,
T. Brezesinski, Energy Environ. Sci. 2016, 9, 2603. S. Zhao, S. Lu, X. Sun, Adv. Energy Mater. 2020, 10, 1903422.
[112] F. Strauss, J. H. Teo, A. Schiele, T. Bartsch, T. Hatsukade, [120] P. R. Rayavarapu, N. Sharma, V. K. Peterson, S. Adams, J. Solid
P. Hartmann, J. Janek, T. Brezesinski, ACS Appl. Mater. Interfaces State Electrochem. 2012, 16, 1807.
2020, https://doi.org/10.1021/acsami.0c02872. [121] Y. G. Lee, S. Fujiki, C. Jung, N. Suzuki, N. Yashiro, R. Omoda,
[113] W. Zhao, J. Yi, P. He, H. Zhou, Electrochem. Energy Rev. 2019, D. S. Ko, T. Shiratsuchi, T. Sugimoto, S. Ryu, J. H. Ku, T. Watanabe,
2, 574. Y. Park, Y. Aihara, D. Im, I. T. Han, Nat. Energy 2020, 5, 299.

Adv. Energy Mater. 2020, 2000804 2000804 (18 of 18) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like