You are on page 1of 13

Articles

https://doi.org/10.1038/s41560-021-00797-7

Regulating electrodeposition morphology in


high-capacity aluminium and zinc battery anodes
using interfacial metal–substrate bonding
Jingxu Zheng 1, David C. Bock2, Tian Tang 1, Qing Zhao 3, Jiefu Yin 3, Killian R. Tallman 2
,
Garrett Wheeler 2, Xiaotun Liu3, Yue Deng 1, Shuo Jin3, Amy C. Marschilok 2,4,5,
Esther S. Takeuchi2,4,5, Kenneth J. Takeuchi4,5 and Lynden A. Archer 1,3 ✉

Although Li-based batteries have established a dominant role in the current energy-storage landscape, post-Li chemistries (for
example, Al or Zn) are emerging as promising candidates for next-generation rechargeable batteries. Electrochemical cells
using Al or Zn metal as the negative electrode are of interest for their potential low cost, intrinsic safety and sustainability.
Presently, such cells are considered impractical because the reversibility of the metal anode is poor and the amount of charge
stored is miniscule. Here we report that electrodes designed to promote strong oxygen-mediated chemical bonding between Al
deposits and the substrate enable a fine control of deposition morphology and provide exceptional reversibility (99.6–99.8%).
The reversibility is sustained over unusually long cycling times (>3,600 hours) and at areal capacities up to two orders of
magnitude higher than previously reported values. We show that these traits result from the elimination of fragile electron
transport pathways, and the non-planar deposition of Al via specific metal–substrate chemical bonding.

C
reating conformal metal coatings through electrochemical but ohmic heat generated by shorting poses added safety concerns
methods is important in multiple fields, including semicon- associated with thermal runaway14.
ductor manufacturing1, energy storage2, metal plating3 and so In theory, metal anodes that take advantage of micro- and/or
on. Although the governing chemistry is in principle simple—sol- nanopatterned conductive substrates that guarantee full access to
vated metal cations in a typically liquid electrolyte capture electrons electron and ion transport pathways throughout the plating/strip-
from an electrically conductive substrate, and are thereby reduced ping processes can surmount all of these challenges5,15. In conven-
into solid metallic form, for example, Mn+ (sol.) + ne− → M(s) tional battery configurations, ion transport occurs via interfacial
—the control over metal electrodeposition morphology remains contact between solid metal deposits and liquid electrolyte, which
challenging2,4. The search for fundamental solutions has recently means that control of the electrolyte volume and interface chem-
re-emerged as an area of scientific and technological interest because istry are sufficient to ensure a full ionic access. In contrast, elec-
of the role that the stable electrodeposition of metals in closed electro- tron transport relies solely on physical, solid–solid contact between
chemical cells plays in the stability and safety of rechargeable batteries the metal deposits and a conductive substrate. The hypothesis that
that utilize metal anodes, for example, Al, Li, Zn, Na and so on, to guides the current study is that the fragility of the electron trans-
achieve a higher energy and the low-cost storage of electrical energy. port pathways is the fundamental barrier to fully reversible metal
The working principle of such electrochemical cells is as follows: electrodeposition processes (Supplementary Fig. 1)6,10. Earlier
the reversible plating and stripping of metal at the anode enables work showed that, although electrodeposition of metals, such as
the reversible storage and release of electrical energy. In prac- Cu in patterned trenches for integrated circuits1,16 and Li and Na
tice, however, repeated formation and dissolution of the metallic in three-dimensional (3D) metal or carbon foams5,6, constrain the
phase is confounded by multiple problems that result in a porous, size of the electrodeposits, the transport length L and therefore the
non-planar deposition morphology5,6. A number of mechanisms electron transport timescale τ are broadly distributed, as the metal
have been proposed to explain these observations. The list includes is free to deposit in coarse, non-planar morphologies5.
intrinsic crystal anisotropy7, uneven solid–electrolyte interphase In this study, we investigated the electrodeposition of metals on
that drives heterogeneous growth4, concentration of electric field patterned substrates designed to strongly coordinate with the metal
lines at ‘hot spots’ for growth8 and growth limited by mass trans- deposit via surface chemical bonding (Fig. 1). Hypothetically,
port9. Porous metal deposits created in one plating cycle also have strong metal–substrate bonding increases the driving force for
a strong propensity to only partially dissolve in the next, to form nucleation to yield a uniform distribution of small metal nuclei.
the so-called ‘dead’ and/or ‘orphaned’ metals6,10–12. Over repeated The uniformly distributed nuclei enable compact, uniform deposi-
cycling, the accumulated metal deposits bridge the interelectrode tion, which produces deposits with a controlled thickness, L, dur-
space to short-circuit the cell10,13. All of these processes are obviously ing the growth phase. The straightforward effect of such control
fatal for the stable long-term operation of an electrochemical cell, is that the electron transport timescale τ ≈ L2/D, which sets the

Department of Materials Science and Engineering, Cornell University, Ithaca, NY, USA. 2Energy and Photon Sciences Directorate, Brookhaven National
1

Laboratory, Upton, NY, USA. 3Robert Frederick Smith School of Chemical and Biomolecular Engineering, Cornell University, NY, USA. 4Department of
Chemistry, State University of New York at Stony Brook, Stony Brook, NY, USA. 5Department of Materials Science and Chemical Engineering,
State University of New York at Stony Brook, Stony Brook, NY, USA. ✉e-mail: laa25@cornell.edu

398 NatUre Energy | VOL 6 | April 2021 | 398–406 | www.nature.com/natureenergy


Nature Energy Articles
a b c

Nano-Al L Micro-Al
Carbon
Nano-Al
Carbon Carbon

Al–O–C covalent bond


Al atom/cluster

Fig. 1 | Metal–substrate bonding-induced regulation of electrodeposition. Strong surface bonding between Al and a conductive fibrillar carbon substrate
facilitates electronic transport and influences the morphological evolution of the Al anode. a, At the onset of electrodeposition, Al forms the interfacial
Al–O–C chemical bonds with the surface of carbon fibres. b, Guided by the chemical bonding interaction, Al grows laterally to form a uniform, compact
deposition layer that comprises nanoscale grains on the carbon fibres. c, After the available carbon surface is fully covered, the subsequent Al deposits
form microsized particles embedded among the carbon fibres. L denotes the electron transport length.

effective exchange current density at the electrode, can be read-


ily manipulated. Here 1/D is the electron mobility15. We therefore Table 1 | Metallic Al foil anode plating/stripping throughput per
hypothesized that the presence of a strong metal–substrate bonding cycle for Al-ion electrochemical cells
provides a mechanism to narrow the distribution to produce com-
pact deposits17, which allows us to control τ and thereby maintain Reference Year Areal capacity Lifetime Configuration
of Al deposition (h)
a high reversibility. We experimentally tested this hypothesis using
(mAh cm–2)
Al metal as a model system. On substrates for which interfacial
chemical bonding was not detected, Al persistently formed coarse, 13 2015 0.10 240 Pouch
heterogeneous electrodeposits, regardless of the geometry of the 41 2015 0.14–0.42 500 Coin
substrates. By contrast, the Al deposition morphology on carbon 42 2016 0.15 200 Pouch
fibres treated by ionic liquid and AlCl3 was highly uniform across
multiple length scales. The uniform Al deposition was enabled by 43 2016 0.08 120 Pouch
Al–O–C bonds formed on the substrate surface, as evidenced by 44 2017 0.1–0.3 500 Coin
X-ray photoelectron spectroscopy. Such a precise regulation on 45 2017 0.12 600 Pouch
morphology was shown to markedly improve the electrochemical
23a 2017 1.25 600 Coin
performance of the Al electrodes.
46 2018 0.008–0.08 200 Pouch
Electrodeposition of Al metal in conventional battery anodes 47 2018 0.015–0.06 40 Coin
The rationale for choosing Al is straightforward—intrinsically safe, 48a 2018 0.4 1000 Coin
low cost and energy dense Al batteries are considered prospective
49 2019 0.0675 150 Pouch
alternatives to Li-ion batteries (Supplementary Table 1)17 in appli-
cations such as backup storage for electric power generation from 50 2019 0.03–0.08 600 Coin
intermittent, renewable sources, where cost and scalability are criti- 51 2020 0.18 30 Pouch
cal. Additionally, the reversible plating/stripping of Al was demon- 24a 2020 0.1–0.5 170 Coin
strated quite recently in 201118, which motivated a strong revival of
This work a
2021 8 3600 Coin
research interest in Al metal anodes for rechargeable batteries. As
a final point, we note that innovations in the design of reversible This worka 2021 0.4–0.5 100 Anode-free pouch
cathode materials compatible with Al metal anodes have produced As representative references, we selected typical studies that disclose the area-specific
batteries of exceptional promise13, albeit at impractically low areal parameters. aModified Al metal anodes were used in these studies.
specific capacities (0.01–0.18 mAh cm–2), that is, nearly two orders
of magnitude lower than that of state-of-the-art Li-ion batteries
(1–3 mAh cm–2) (Table 1). We assessed the performance of metallic Al anodes at elevated
The electrode areal specific capacity is an important—but areal capacities (for example, 0.8–8 mAh cm–2) by plating/strip-
oftentimes overlooked—battery parameter. Fundamentally, using ping Al metal on a conventional, planar stainless-steel substrate.
an impractically low areal capacity means that the intrinsic anode Surprisingly, the electrochemical cells fail quickly (<20 hours) by
failure modes discussed above might be obscured by an interelec- short-circuiting at these high capacities (Supplementary Fig. 2). To
trode spacing that is too large. From an applications perspective, the evaluate the plating/stripping reversibility of Al at practical areal
disproportionately large contribution from inactive battery parts to capacities—without disruptions due to battery shorting—we fol-
the overall weight and volume leads to a lower overall cell energy lowed a practice first reported in Lin et al. and intentionally placed
density and an elevated cost per unit of energy stored19–21. The slow multiple layers of separator in the interelectrode space13. The plat-
progress in addressing this issue raises concern about the actual ing/stripping reversibility (Coulombic efficiency (CE)) is thereby
potential of Al-based batteries as next-generation energy storage measured as approximately 85%, which means that 15% of the
systems19. Building Al metal anodes that sustain prolonged cycling Al deposits are irreversibly lost in each cycle (Fig. 2a and voltage
at practical areal capacities is therefore a critical step towards com- profile in Supplementary Fig. 3). As depicted in Supplementary
mercially relevant Al batteries22,23. Fig. 4, a practical Al anode would require a CE of 99%, preferably

NatUre Energy | VOL 6 | April 2021 | 398–406 | www.nature.com/natureenergy 399


Articles Nature Energy

a b c
Time (h)
0 10 20 30 40 50 60
Al plating/stripping CE (%) 110

100 Short

90

80 d

70

60 100 µm
0 2 4 6 8 10 12 14 16
Cycle no.

e f g
e– e– e– e–

100 µm
Al GF SS Al GF SS

Fig. 2 | The propensity of Al anodes to exhibit heterogeneous growth on conventional substrates. a, CE measured in Al plating/stripping reactions at a
capacity of 3.2 mAh cm–2. Two layers of a GF separator were needed to prevent shorting. b, SEM analysis of Al electrodeposits formed on a planar steel foil
electrode at 4 mA cm–2 and for a capacity of 0.8 mAh cm–2. c,d, Elemental maps of Al (c) and Fe (d) from an EDS analysis of the electrode surface shown
in b. e, Schematic diagram showing the failure mechanism of Al electrodes. A higher areal deposition capacity and the accumulation of dead metal in the
GF membrane over multiple plating/striping cycles result in the proliferation of non-planar metal deposits in the interelectrode space. f, SEM image of Al
electrodeposits formed on a non-planar Ni foam electrode at a rate of 4 mA cm–2 and capacity of 1.0 mAh cm–2. g,h, Elemental maps of Al (g) and Ni (h)
from an EDS analysis of the electrode surface shown in f. SS, inert stainless-steel electrode.

above 99.5%. We note also that even with multiple separators, the Ni foam substrate (Fig. 2f–h) has, at most, a negligible impact on
3.2 mAh cm–2 cell fails after only ~60 hours. Together, (1) the rapid the deposition morphology. Consequently, using Ni foam as the
battery shorting and (2) the low plating/stripping reversibility at anode substrate exhibited a rather limited improvement on the elec-
elevated areal capacities impose fundamental, pressing challenges trochemical performance (Extended Data Fig. 1). The first clue to
for practical Al batteries. understanding the aggressive non-planar deposition of Al is that
To diagnose the underlying problems, we performed scanning glass fibre (GF) (SiO2) membranes, rather than organic polymers,
electron microscopy (SEM) to characterize the electrodeposit are used as the separator. This choice stems from the fact that typi-
morphology. The SEM results in Fig. 2b show that the Al deposits cal polyolefin separators are chemically unstable in the imidazolium
exhibit a heterogeneous, non-planar morphology. Consistent with chloride + AlCl3 electrolyte melt. Early literature reports show that
prior reports24, the Al deposits were as large as 20–50 μm, sparsely there is a possibility for a strong chemical affinity between metallic
distributed on the substrate and isolated from each other. This led Al and SiO2 via an interfacial reaction: 4Al + 3SiO2 → 2Al2O3 + 3Si
to large and broadly distributed L and τ values, which produced (refs 26,27), which implies that, in addition to the other factors that
sluggish, incomplete metal stripping, as illustrated in Fig. 2b–d. The may drive non-planar metal electrodeposition, a chemical driving
low plating/stripping efficiency and the battery shorting in the Al force favours the growth of Al electrodeposits into the separator.
anodes are therefore concluded to be consequences of the local- This growth mode was confirmed using SEM and energy dispersive
ized, aggressive growth of the Al electrodeposits normal to the sub- spectroscopy (EDS) analysis on either a stainless steel or a Ni foam
strate plane. We here follow the literature21 and loosely refer to the substrate (Supplementary Figs. 5 and 6), which explains the speed
non-planar Al electrodeposit growth as dendritic, but show below with which Al is able to short-circuit cells even when multiple layers
that it is unusually aggressive and possesses previously unreported of separator membranes are used or when the deposition is per-
characteristics that appear to arise from the strong affinity of Al for formed using a Ni foam substrate.
the separator (Fig. 2e).
The speed with which our Al electrodes fail by short-circuiting— Uniform Al deposition enabled by O-mediated metal–
even at conventional electrode areal capacities—exceeds typical val- substrate bonding
ues for Li anodes21, which implies that the materials physics that This discovery poses a perhaps obvious question: could the pro-
control the non-planar electrodeposition of Al are quite different. pensity of Al to grow into the separator be countered by creating
It also explains the rather low areal capacity limits explored in pre- metal–substrate bonding that is strong enough to chemically usher
vious studies. We note further that whereas deposition of Li in a Al into a patterned substrate? This question can be theoretically
3D porous material, such as a Ni foam, can substantially extend reasoned by an analysis of the nucleation process in electrodeposi-
the time-to-failure due to short-circuiting25, deposition of Al in a tion, which plays a pivotal role in determining the size, uniformity

400 NatUre Energy | VOL 6 | April 2021 | 398–406 | www.nature.com/natureenergy


Nature Energy Articles
a b

50

40

No. of particles
30

20

10

0
500 nm 50 100 150 200 250 300 350
Particle diameter (nm)

c d

500 nm 500 nm

e f

Al

5 µm 5 µm
C

g h

C(002)
Intensity (a.u.)

Al(200)
C(100) Al(111)
90 Al(200)
Al(111) C(100)

60

25 30 35 40 45 50
C(002) 30
2θ (°)

Fig. 3 | Microstructure of Al metal deposits formed on a substrate with strong metal–substrate bonding. a,c–f, SEM analysis of the Al deposit
morphology at varying areal capacities: 0.2 (a), 1 (c), 3 (d), 3 (e) and 4 (f) mAh cm–2. Inset (e): SEM and corresponding EDS mapping of a selected region
in e. b, Particle size distribution at 0.2 mAh cm–2 (average = 139 nm). g, 2D XRD pattern of Al deposits on carbon fibres. h, Corresponding integrated XRD
line scan for the material in g. a.u., arbitrary units.

and distribution of electrodeposits8. On the application of a volt- can be dissociated into two steps: (1) electrochemical reduction of
age, the initial process is the desolvation of the coordinated species the metal, Mn+ + ne− → M(free), and (2) binding with the sub-
in the electrolyte as they approach the negatively charged electrode strate, M(free) + ∗ → M∗, which directly depends on the interaction
surface28,29. In this desolvation process, the metal-containing spe- between the metal and the substrate; the nucleation overpotential is
cies are converted into a cationic form by losing the coordinating determined by the free energy change ΔG of the latter30. Classical
ions and/or molecules. The subsequent step is the electrochemical nucleation theory then implies that strong metal–substrate bonding
adsorption of the species: Mn+ + ne− + ∗ → M∗, where * refers results in a smaller critical nucleate size, rc = −2γ/ΔG, where γ is the
to the adsorption site and M* to the adsorbed metal. This reaction energy penalty for creating a unit area of the metal surface and ΔG

NatUre Energy | VOL 6 | April 2021 | 398–406 | www.nature.com/natureenergy 401


Articles Nature Energy

a b c
Data Data 24,000 Data
12,000 Peak 1 Peak 1 Peak 1
Peak 2 3,000 Background Background
Intensity (c.p.s.)

Intensity (c.p.s.)

Intensity (c.p.s.)
Peak 3 Fit 18,000 Fit
Peak 4
8,000 Background
Fit 2,000
12,000

4,000 1,000
6,000
d e f
Data Data 24,000 Data
12,000 Peak 1 Peak 1 Peak 1
Peak 2 3,000 Peak 2 Peak 2
Intensity (c.p.s.)

Intensity (c.p.s.)

Intensity (c.p.s.)
Peak 3 Peak 3 Background
Background 18,000
Peak 4 Fit
Peak 5 Fit
8,000 2,000
Background
Fit 12,000

4,000 1,000
6,000

292 290 288 286 284 282 80 78 76 74 72 538 536 534 532 530 528
Binding energy (eV) Binding energy (eV) Binding energy (eV)

Fig. 4 | XPS characterization of the metal–substrate interaction. a–f, C 1s (a,d), Al 2p (b,e) and O 1s (c,f) spectra of Al samples in carbon fibre electrodes
in the as-deposited state (a–c) and the sputtered state (d–f). The peaks that correspond to the Al–O–C bond are shaded by grey and indicated in the plots
by vertical black lines. On sputtering, surface materials are removed from the samples by the incident Ar+ ion beam. Peak assignments are tabulated in
Supplementary Table 2. c.p.s., cycles per second.

is the Gibbs free energy associated with the formation of the con- altering the microstructure of the substrate. On these bases, in a
densed metal phase during the earliest stages of nucleation. non-planar architecture made of interwoven carbon fibres, the pos-
The implementation of this concept in a battery anode is not sible chemical metal–substrate bonding between C and Al mediated
straightforward, because the nucleation occurs on an electroni- by oxygen is then hypothetically able to promote uniform Al growth
cally conducting substrate, which means it does not strictly become on individual carbon fibres (Fig. 1), as opposed to the macroscopic
homogeneous. As discussed above, the formed metal nuclei may outer surface of the architecture, as shown in Supplementary Figs.
also make strong bonds with the separator, substrate or both, which 5 and 6.
will affect their shape. Additionally, to achieve a high electrode Figure 3 reports the main results of the present study. The Al
reversibility, the geometrical and chemical integrity of the substrate metal electrodeposition morphology on interwoven carbon fibres
must remain intact over repeated cycles of metal plating/stripping. was investigated using SEM. In stark contrast to the Al deposited
On this basis, a screening framework for promising substrate mate- on planar stainless steel and on non-planar Ni foam, for which
rial is proposed: the substrate should be electrically conductive, the deposit sizes are quite variable with average values of L on the
mechanically robust and capable of forming interfacial metal–sub- order tens of micrometres, the Al electrodeposits on the carbon
strate bonding. It is important that the chemical reaction between fibre surface with L typically in the range 100–200 nm in the nucle-
the substrate and the metal is limited within the interface, rather ation stage (Fig. 3a,b); this that for a fixed D, τ could be reduced
than penetrates into the bulk. A bulk reaction, owing to the crys- by a factor of 10,000 or more. As the deposition capacity increases,
tallographic and volume changes associated, inevitably gives rise to the nanoscale Al crystals grow laterally (Fig. 3c–e and EDS maps
morphological instabilities of the substrate over cycling. in Supplementary Fig. 9), that is expand along the carbon sur-
In screening for potential substrate materials, free-standing face to generate a conformal Al coating layer of thickness on the
interwoven carbon fibres (SEM image in Supplementary Fig. 7) are order of 102 nm (for example, 160 nm, see Supplementary Fig.
singled out because previous studies show that these materials form 10) that appear to strictly follow the surface curvature (Fig. 3d).
multiple types of interfacial covalent bonds, which include Al–C, Supplementary Fig. 10 also shows the morphology of the Al deposit
Al–O–C and so on, with Al metal during electroplating31,32. The cal- layer that faces the carbon fibre—nanoscale Al grains are seen to
culated binding energy of Al–O–C is as high as a few electronvolts31, merge into a compact layer. The Al deposition morphology is highly
which means that the driving force for Al to grow on the substrate uniform across a macroscopic area, as shown in Supplementary Fig.
surface is rather large. This characteristic was previously utilized 11, without any observable evidence of aggressive and/or dendritic
to electroplate a uniform Al precoating layer on carbon, which growth. Microstructural characterization of the morphology after
improved the Al/C wettability in subsequent metallurgical manu- thermal annealing shows that surface diffusion is not as effective a
facturing of Al/C composite materials31. This carbon-based mate- mechanism for smoothing the Al electrodeposits (Supplementary
rial is also of specific interest here because the bonding occurs via Fig. 12); rather, the initial uniform nucleation promoted by metal–
interfacial interaction at oxygen-enriched defective sites on the sur- substrate bonding plays a dominant role. These results confirm that
face rather than via bulk phase transition. Of note is that the Raman the strong metal–substrate bonding is able to (1) guide a metal elec-
spectrum of the interwoven carbon fibres shows both the D band trodeposition precisely using the surface topography of a strongly
and G band (Supplementary Fig. 8). This means that the carbon interacting non-planar architecture and (2) promote the uniform
contains a considerable amount of defects, which reportedly pro- nanoscale nucleation and lateral growth of metal deposits to create
motes the bonding with Al (ref. 31). X-ray photon electron spectros- a compact film.
copy (XPS) characterization (Extended Data Fig. 2) shows that the We increased the deposition capacity beyond 3 mAh cm–2 and
exposure of carbon fibres to the imidazolium chloride + AlCl3 elec- observed that the growth of Al enters a second stage (Fig. 1): growth
trolyte remarkably increases the level of oxygen enrichment without of microscale Al crystals on the compact nanoscale Al layer (Fig. 3f).

402 NatUre Energy | VOL 6 | April 2021 | 398–406 | www.nature.com/natureenergy


Nature Energy Articles
a 0 2,000 4,000 6,000 8,000 Cycle no.
d

Al CE(%)
105 0.6
100 no.1
95 CE = 99.8% no.50 (500 h)
90
0.4 no.125 (1,250 h)
0 1,000 2,000 3,000 Time (h)
no.200 (2,000 h)
b

Potential (V)
0 200 400 600 800 Cycle no. no.360 (3,600 h)
105 0.2
Al CE(%)

100
95 CE = 99.6%
90 0.0
0 1,000 2,000 3,000 Time (h)

c 0 50 100 150 200 250 300 350 Cycle no.


105 –0.2
Al CE(%)

100
95 CE = 99.6% 0 2 4 6 8
90
0 1,000 2,000 3,000 Time (h) Capacity (mAh/cm2)

Fig. 5 | Electrochemical cycling behaviour of structured Al electrodes in galvanostatic plating/stripping experiments. a–c, Al plating/stripping efficiency
measured at areal capacities of 0.8 (a), 3.2 (b) and 8 (c) mAh cm–2. d, Voltage profiles measured during Al plating/stripping at a very high areal capacity
of 8 mAh cm–2. The current densities used for the respective measurements are 4 (a) and 1.6 (b,c) mA cm–2.

X-ray diffraction (XRD) (Fig. 3g,h) verified that only Al crystals on the sample surface was etched away. In this case, obvious new
were present on the carbon substrate fibres. The microscale Al peaks emerged (C 1s, 283.5 eV; Al 2p, 74.7–74.9 eV; O 1s, 531.9 eV),
deposits gradually filled the space among the carbon fibres (mor- which are ascribed to Al–O–C bonding. XPS peak assignments are
phology at 8 mAh cm–2 shown in Supplementary Fig. 13). The tabulated in Supplementary Table 2. The observation is consistent
microscale Al deposits were intimately connected to the compact with that of previous work31. The metal–substrate bonding is also
nanoscale Al layer and to each other. This observation confirms evidently mediated by the surface oxidation species of the carbon.
that the metal–substrate bonding has a decisive influence on the In contrast, XPS measurements of Al electrodeposits on control
Al deposition morphology. After the substrate surface was fully substrates suggest that the Al–O–C bonding and other metal–sub-
occupied by the compact nanoscale Al layer, the newly deposited Al strate bonding do not exist on stainless steel and Ni foam (Extended
crystals resumed the intrinsic growth mode at the micrometre scale. Data Fig. 4).
As a more direct control experiment, we artificially precoated the
carbon fibres with an interphase that blocked the possible Al–O–C Electrochemical performance of regulated metal anodes
bonding, and evaluated the Al electrodeposition morphology on To evaluate the reversibility of the Al deposited at these high areal
this ‘deactivated’ carbon fibre matrix (Supplementary Fig. 14). The capacities, we measured the plating/stripping CE of the Al elec-
SEM characterization shows that the Al deposition was no longer trodes in electrochemical cells. The CE reveals the ratio between the
uniform over the surface of the carbon fibres. This negative control amount of metal that can be stripped and the amount that is origi-
experiment validates our hypothesis that oxygen-enriched surface nally plated. The results are reported in Fig. 5. In stark contrast to
chemistry plays an indispensable role to realize the control over the low CE and rapid battery failure observed on a planar stainless
deposition morphology. steel substrate, anodes produced by Al electrodeposition in a pat-
Moreover, we devised physical simulation experiments to more terned substrate composed of carbon fibres manifested a high level
quantitatively understand the role played by a large driving force of reversibility (99.4–99.8%, stripped morphology in Supplementary
at the nucleation stage and its impact on the ultimate deposition Fig. 16) over a broad range of capacities (0.8–8 mAh cm–2). The
morphology (Extended Data Fig. 3, Supplementary Note 1 and results also showed that the cells maintain stable operations for
Supplementary Fig. 15). It was shown that with a nucleation driven more than 2,000 hours (Fig. 5a–c). The voltage profile in Fig. 5d
by a large overpotential of 2 V, uniform Al growth was achieved confirms that the plating/stripping reaction was stable, as evidenced
at the subsequent stage of deposition under conventional con- by the narrowing overpotential over the 3,600 hours of cycling, and
stant current conditions. This set of results confirms that a large does not show any trend towards deterioration. It is also notewor-
driving force imposed in the initial nucleation process serves as a thy that the CE at 8 mAh cm–2 was not compromised by the pres-
determinant factor in achieving uniform Al deposition. However, ence of microscale Al deposits, owing to what we suspect is a robust,
this physical approach is fundamentally associated with chemical interconnecting electronic pathway in the conductive non-planar
and hydrodynamic instabilities, which makes it unsuitable for bat- architecture that guarantees a good electrochemical access to the
tery applications (detailed discussion after Supplementary Fig. 15). metal and prevents the formation of dead or orphaned Al. When
Further molecular-level simulation of this process could be obtained paired with a cathode, this improvement could be clearly reflected
by joint density function theory methods33 that capture the detailed in proof-of-concept full cells at an elevated capacity of ~1 mAh cm–2,
kinetic steps near a solid–liquid interface, for example, desolvation, which is 1–2 orders of magnitudes higher than the values reported
bonding and interphase formation33–35. in the literature (Supplementary Figs. 17 and 18). Consistent with
To more thoroughly interrogate the nature of the metal–sub- the plating/stripping measurement, batteries using Al foil as the
strate bonding, we performed X-ray photoelectron spectroscopy. anode fail rapidly in <100 cycles. By contrast, ‘anode-free’ full bat-
Figure 4a–c shows the XPS results of the as-deposited sample teries with a carbon-fibre anode without any pre-stored Al showed
(3 mAh cm–2), which are consistent with the conclusion that no a stable cycling and excellent capacity retention over thousands of
possible metal–substrate bonding is detected on the surface. This cycles. These results of the full battery tests further confirm the high
observation is expected because the substrate architecture is fully reversibility of Al plating/stripping on carbon fibres when subject
covered by the compact deposition layer, which means that the to cycling.
metal–substrate interface is not exposed. Figure 4d–f shows the XPS As an extreme test of the approach, we performed plating/strip-
results of the sample after Ar+ sputtering, during which the material ping experiments at an unprecedentedly high current density, that

NatUre Energy | VOL 6 | April 2021 | 398–406 | www.nature.com/natureenergy 403


Articles Nature Energy

a b

5 µm 5 µm

c d
105
Shorting-induced

Al CE(%)
0.5 100
endless charging
95
40 mA; 0.4 mAh cm–2
0 90
Potential (V)

0 20,000 40,000 60,000

–0.5 Cycle no.


0.4 mAh cm–2 plating
on inert substrate 105
Al CE(%)
–1.0 100

2nd cycle 95
40 mA; 1 mAh cm–2
–1.5 90
0.0 0.2 0.4 0.6 0.8 0 200 400 600 800 1,000 1,200
Areal capacity (mAh cm–2) Cycle no.

Fig. 6 | Ultrahigh current density plating/stripping of Al. a,b, SEM analysis of Al plated at a current density of 40 mA cm–2 on stainless steel (a) and on
carbon fibres (b). c, Voltage profile showing that the Al||stainless steel cells fail through short-circuiting by the second cycle. d, CE versus cycle index for
the Al plating/stripping process on carbon fibres with metal–substrate bonding.

is, 40 mA cm–2. SEM characterization revealed that diffusion-limited prior literature reports. It is apparent that even without taking into
ramified electrodeposit structures with classical fractal and/or den- account the intrinsic advantages of Al metal, which include safety,
dritic morphologies formed on the planar electrode (Fig. 6a). In cost, ease of manufacturing and high earth-crust abundance, Al bat-
contrast, nanoscale, compact Al electrodeposition was maintained teries based on the proposed anode design show enormous promise.
on the carbon-based fibres (Fig. 6b). As expected, the cells using To explore the broader utility of our concept, we studied the elec-
the planar stainless-steel foil failed quickly, after <5 minutes, due trodeposition of Zn, another metallic battery anode material that
to the short-circuiting process (Fig. 6c). This can be compared with has aroused enormous research interest recently36–38, in a patterned
the electrochemical results (Fig. 6d), which showed high levels of substrate composed of interwoven carbon fibres (Supplementary
reversibility and long lifetimes (for example, CE = 99.96% over Fig. 19). The galvanostatic plating/stripping results demonstrate
60,000 cycles at a current density of 40 mA cm–2 and an areal capac- that, unlike Al, Zn has a low plating/stripping efficiency and slow
ity of 0.4 mAh cm–2) with the regulation produced by the metal– kinetics on carbon fibres, which indicates that the substrate archi-
substrate bonding. tecture alone is insufficient to achieve highly reversible deposition.
The stable plating/stripping behaviours of Al under the regula- We artificially introduced a strong metal–substrate interaction by
tion of metal–substrate bonding can be understood in a quantitative coating the substrate with graphene, which we recently reported
manner. As Al forms a compact layer on the surface of the carbon strongly coordinates with Zn by an epitaxial mechanism owing to
fibres, the electron transport length scale is maintained as small. As a minimized interfacial energy36. The plating/stripping reversibility
a quantitative estimation, the characteristic relaxation time τ of the and lifetime are improved remarkably by this coating. SEM results
nanoscale Al deposits formed in the patterned substrate is four orders showed that the presence of the strong metal–substrate interac-
of magnitudes smaller than that of the dendritic Al (L ≈ 160 versus tion via the graphene interphase can effectively promote the uni-
20 μm) formed in the planar case. This means that the nanoscale form deposition of nanoscale, plate-like Zn metal with a minimized
Al can be stripped around four orders of magnitude faster. In addi- electron transport length scale, in comparison with the microscale
tion to considerations based on transport, the non-planar and/or Zn deposits observed on bare carbon fibres (Supplementary Fig.
dendritic metal deposit is associated with a stronger morphologi- 19). Moreover, the effectiveness of the metal–substrate bonding
cal instability—as stripping occurs at the bottom of the deposits, can be demonstrated on planar substrates, which is a more con-
the whole structure detaches from the electrode6. Our observations, ventional configuration in commercial batteries. We evaluated the
therefore, suggest that an Al anode in which the metal is deposited electrochemical plating/stripping of Al on stainless steel coated
in a patterned substrate capable of forming strong bonds with Al are by carbon nanotubes with and without carboxylic side groups.
of potential immediate interest for applications in Al batteries that As shown in Supplementary Fig. 20, the planar substrates coated
operate under various conditions. In Table 1, we compare the elec- with carboxylic-functionalized carbon nanotubes manifested stable
trochemical performance achieved for our Al anodes with those of cycling and a high CE, in comparison with those of bare stainless

404 NatUre Energy | VOL 6 | April 2021 | 398–406 | www.nature.com/natureenergy


Nature Energy Articles
steel and even non-planar Ni foam. On this basis, we conclude that by one layer of GF/B (Whatman), unless otherwise specified. Stainless steel and
the concept to regulate electrodeposition morphology can be readily carbon cloth were washed by ultrasonication in deionized water and acetone.
In each coil cell, ~120 μl of electrolyte was added by pipette to make sure the
extended to other electrodeposition systems of different chemistries electrodes and the separator were wetted.
or geometries by rationally designing an artificial metal–substrate In the Al metal plating/stripping CE measurements, the electrochemical
interphase that provides strong interactions. cells had two electrodes: an Al source electrode and a substrate electrode (for
example, Al||stainless steel, Al||carbon fibres and so on). In the measurement,
a certain amount of Al was plated on the substrate electrode of interest, and
Conclusion a reverse potential was applied to strip the deposited Al metal back to the Al
In summary, we report that specific interfacial metal–substrate source electrode, CE = plating stripping capacity
capacity on the substrate × 100%. In the Al||stainless
bonding can be used to achieve a fine control of metal electrodeposit
morphology and a uniform, compact and exceptionally reversible steel cells, a commercial Al foil was used as the Al source. In the Al||carbon fibre
cells, Al electrodeposited on carbon fibre was used as the Al source because the
metal deposition. Using metallic Al as an example, we first showed battery short was caused by uncontrollable Al growth during both the plating on
how such bonds can be used to overcome the metal’s natural affin- the substrate electrode and on the Al source electrode. As a result, both of the
ity for the separator and to prevent aggressive, non-planar electro- electrodes needed to be carbon fibres to measure the true stabilizing effect of using
deposition. Extension of the concept to create patterned Al anodes carbon fibres as the deposition substrate.
Full cells were assembled to further evaluate the Al plating/stripping behaviour
revealed that it is possible to achieve highly reversible metal anodes
in the anode. In ‘anode-free’ full cells constructed using carbon fibres, no Al metal
with combinations of areal specific capacity and cycle life that are was prestored in the anode. This means that Al metal solely comes from battery
one or more orders of magnitude higher than those in previous charging, 4Al2Cl7– + 3e– → Al(s) + 7AlCl4–. To offset the initial capacity loss possibly
literature reports. Our findings confirm that the reversibility of owing to side reactions, as can be seen in Figs. 5 and 6, carbon fibre electrodes were
metal anodes requires a continuous access to ionic and electronic subjected to a formation process: Al was plated and striped on carbon fibres at 40
mA cm–2, 0.4 mAh cm–2 for 50 cycles. After the formation process, the carbon fibre
transport pathways in the electrode and is strongly correlated with electrode was charged to 0.7 V versus Al3+/Al to fully remove any electrochemically
control of the electron transport length scale. Successful extension active Al metal, and was harvested for the subsequent assembly of full cells. In
to Zn metal anodes indicates that the concept may be generalized control full cells, commercial 0.25-mm-thick Al foil was used as the anode. This
to achieve a high reversibility in other metallic anodes in batteries. means that a large amount of extra Al metal was prestored in the anode, as can be
quantified by an N:P ratio of 100:1 (70 mAh cm–2 anode, 0.7 mAh cm–2 cathode).
The graphite-based cathodes that offered a high areal capacity were designed
Methods according to principles and methods discussed in a prior study20. Non-planar,
Materials. [EMIm]Cl (1-ethyl-3-methylimidazolium chloride, 98%), AlCl3
interwoven carbon fibre electrodes were used as the matrix to ensure a fast
(99.99%), Whatman glass fibre filters (GF/B), 0.25 mm Zn foil (99.9%),
transport and the physical integrity of the thick, high-areal-capacity cathodes.
carboxymethyl cellulose and ZnSO4·7H2O were purchased from Sigma Aldrich.
Carboxymethyl cellulose was added into a water dispersion of few-layer graphene
Al foil (0.25 mm thick, 99.99%) and tantalum foil (0.127 mm thick) were bought
to form a viscoelastic fluid. A strain was applied to periodically agitate the fluid to
from Alfa Aesar. Ni foam was manufactured by MTI. Stainless steel 304 was
drive the active materials into the porous medium, that is, the interwoven carbon
used as a conventional, planar substrate for electrodeposition. Deionized water
fibres. The graphene-infused carbon fibre electrodes (diameter: 1/4 inch; 0.64 cm)
was obtained from a Milli-Q water purification system. Graphene dispersion in
were dried at 60 °C overnight in an oven. Tantalum foils were used as the spacer to
N-methyl-2-pyrrolidone (4 wt%), few-layer graphene dispersion in water (4 wt%),
prevent possible corrosion in the cathode.
carboxylic-functionalized carbon nanotubes and graphitized carbon nanotubes
were purchased from ACS Material. Free-standing interwoven carbon fibres (Plain
carbon cloth 1071 HCB) were purchased from Fuel Cell Store. Data availability
The datasets generated during the current study are available in the article and its
Preparation of electrolytes. Al anode electrolyte. In an Ar-filled glovebox Supplementary Information.
(<0.1 ppm H2O, <1 ppm O2), AlCl3 powder was added slowly into [EMIm]
Cl in a glass vial with 500 r.p.m. stirring to form a yellowish liquid electrolyte. Received: 20 August 2020; Accepted: 15 February 2021;
The molar ratio was ionic liquid:AlCl3 = 1:1.3, which resulted in a Lewis acidic Published online: 5 April 2021
electrolyte in which reversible Al plating/stripping was possible. Al foil was placed
in the as-prepared ionic liquid–AlCl3 electrolyte to remove impurities such as
hydrochloric acid in the electrolyte39. After resting for 7 days, the electrolyte turned References
colourless and was used as an electrolyte for Al cells. 1. Josell, D., Wheeler, D., Huber, W. H. & Moffat, T. P. Superconformal
electrodeposition in submicron features. Phys. Rev. Lett. 87, 016102 (2001).
Zn anode electrolyte. ZnSO4·7H2O was slowly dissolved into the deionized water in 2. Zachman, M. J., Tu, Z., Choudhury, S., Archer, L. A. & Kourkoutis, L. F.
a glass vial to prepare the 2 M ZnSO4 electrolyte for Zn cells. The transparent, clear Cryo-STEM mapping of solid–liquid interfaces and dendrites in
electrolyte was rested for 1 day before use. lithium-metal batteries. Nature 560, 345 (2018).
3. Baker, H. The metal of the future. Nature 42, 537 (1890).
Characterization of materials. Field-emission scanning electron microscopy 4. Tikekar, M. D., Choudhury, S., Tu, Z. & Archer, L. A. Design principles for
(FESEM) was carried out on Zeiss Gemini 500 SEM equipped with a Bruker EDS electrolytes and interfaces for stable lithium-metal batteries. Nat. Energy 1,
detector to study the electroplate/strip morphology. An accelerating voltage of 5 kV 16114 (2016).
was used for imaging, and 10 kV was used for the EDS measurement. 2D-XRD 5. Zheng, J. et al. Physical orphaning versus chemical instability: Is dendritic
was performed on a Bruker D8 General Area Detector Diffraction System with a electrodeposition of Li fatal?. ACS Energy Lett. 4, 1349–1355 (2019).
Cu Kα X-ray source. The incident beam angle and the detector angle were both 6. Deng, Y. et al. On the reversibility and fragility of sodium metal electrodes.
set at 18°. Raman spectroscopy was performed on a Renishaw InVia Confocal Adv. Energy Mater. 9, 1901651 (2019).
Raman microscope using a 785 nm laser. XPS experiments were carried out in a 7. Shi, F. et al. Strong texturing of lithium metal in batteries. Proc. Natl Acad.
ultrahigh vacuum chamber equipped with a SPECS Phoibos 100 MCD analyser. Sci. USA 114, 12138–12143 (2017).
A non-monochromatized Al Kα (hν = 1,486.6 eV) X-ray source was used for 8. Biswal, P., Stalin, S., Kludze, A., Choudhury, S. & Archer, L. A. On the
the analysis. The accelerating voltage was 15 kV and the current was 20 mA. nucleation and early-stage growth of Li electrodeposits. Nano Lett. 19,
The chamber had a base pressure of 2 × 10−9 torr. Electrodes were pressed onto a 8191–8200 (2019).
conductive copper tape and mounted on the sample holder. The charge correction 9. Bai, P., Li, J., Brushett, F. R. & Bazant, M. Z. Transition of lithium growth
for the data was done to by adjusting the C 1s binding energy to 284.8 eV for C–C mechanisms in liquid electrolytes. Energy Environ. Sci. 9, 3221–3229 (2016).
and C–H bonds40. The XPS analysis regions measured were 65–90 eV for Al 2p, 10. Fang, C. et al. Quantifying inactive lithium in lithium metal batteries. Nature
520–545 eV for O 1s and 265–300 eV for C 1s. A pass energy of 20 eV, step size of 572, 511–515 (2019).
0.05 eV and scan number of 10 were applied to measure each individual region. 11. Zheng, J. et al. Regulating electrodeposition morphology of lithium: towards
Ar sputtering was performed at room temperature with a pressure of 2 × 10−5 torr commercially relevant secondary Li metal batteries. Chem. Soc. Rev. 49,
using an energy of 1.5 keV. 2701–2750 (2020).
12. Zheng, J. & Archer, L. A. Controlling electrochemical growth of metallic zinc
Electrochemical measurements. The galvanostatic plating/stripping performances electrodes: toward affordable rechargeable energy storage systems. Sci. Adv. 7,
of metals in coin-type electrochemical cells (CR2032) were tested on Neware eabe0219 (2021).
battery test systems at room temperature. The diameter of the electrodes in the 13. Lin, M.-C. et al. An ultrafast rechargeable aluminium-ion battery. Nature 520,
plating/stripping measurements was 0.5 inch (1.27 cm). Electrodes are separated 324–328 (2015).

NatUre Energy | VOL 6 | April 2021 | 398–406 | www.nature.com/natureenergy 405


Articles Nature Energy
14. Higashi, S., Lee, S. W., Lee, J. S., Takechi, K. & Cui, Y. Avoiding short circuits 42. Zhang, X., Tang, Y., Zhang, F. & Lee, C. S. A novel aluminum–graphite
from zinc metal dendrites in anode by backside-plating configuration. Nat. dual-ion battery. Adv. Energy Mater. 6, 1502588 (2016).
Commun. 7, 11801 (2016). 43. Wang, S. et al. A novel aluminum-ion battery: Al/AlCl3-[EMIm]Cl/Ni3S2@
15. Zhu, C., Usiskin, R. E., Yu, Y. & Maier, J. The nanoscale circuitry of battery graphene. Adv. Energy Mater. 6, 1600137 (2016).
electrodes. Science 358, eaao2808 (2017). 44. Chen, H. et al. A defect-free principle for advanced graphene cathode of
16. Reid, J. Copper electrodeposition: principles and recent progress. Jpn J. Appl. aluminum-ion battery. Adv. Mater. 29, 1605958 (2017).
Phys. 40, 2650 (2001). 45. Chen, H. et al. Ultrafast all-climate aluminum–graphene battery with
17. Elia, G. A. et al. An overview and future perspectives of aluminum batteries. quarter-million cycle life. Sci. Adv. 3, eaao7233 (2017).
Adv. Mater. 28, 7564–7579 (2016). 46. Liu, Z. et al. Carbon nanoscrolls for aluminum battery. ACS Nano 12,
18. Jayaprakash, N., Das, S. K. & Archer, L. A. The rechargeable aluminum-ion 8456–8466 (2018).
battery. Chem. Commun. 47, 12610–12612 (2011). 47. Zhang, E. et al. A novel aluminum dual-ion battery. Energy Storage Mater. 11,
19. Faegh, E., Ng, B., Hayman, D. & Mustain, W. E. Practical assessment of the 91–99 (2018).
performance of aluminium battery technologies. Nat Energy 6, 21–29 (2021). 48. Wu, F., Zhu, N., Bai, Y., Gao, Y. & Wu, C. An interface-reconstruction effect
20. Zheng, J. et al. Nonplanar electrode architectures for ultrahigh areal capacity for rechargeable aluminum battery in ionic liquid electrolyte to enhance
batteries. ACS Energy Lett. 4, 271–275 (2018). cycling performances. Green Energy Environ. 3, 71–77 (2018).
21. Albertus, P., Babinec, S., Litzelman, S. & Newman, A. Status and challenges in 49. Zhang, E. et al. Unzipped carbon nanotubes for aluminum battery.
enabling the lithium metal electrode for high-energy and low-cost Energy Storage Mater. 23, 72–78 (2019).
rechargeable batteries. Nat. Energy 3, 16–21 (2018). 50. Xu, H. et al. Low-cost AlCl3/Et3NHCl electrolyte for high-performance
22. Levy, N. R. & Ein-Eli, Y. Aluminum-ion battery technology: a rising star or a aluminum-ion battery. Energy Storage Mater. 17, 38–45 (2019).
devastating fall? J. Solid State Electrochem. 24, 2067–2071 (2020). 51. Tan, B. et al. Synthesis of RGO-supported layered MoS2 with enhanced
23. Chen, H. et al. Oxide film efficiently suppresses dendrite growth in electrochemical performance for aluminum ion batteries. J. Alloy Compd 841,
aluminum-ion battery. ACS Appl. Mater. Interfaces 9, 22628–22634 (2017). 155732 (2020).
24. Zhao, Q., Zheng, J., Deng, Y. & Archer, L. Regulating the growth of
aluminum electrodeposits: towards anode-free Al batteries. J. Mater. Chem. A Acknowledgements
8, 23231–23238 (2020). We thank F. Zhu, L. Luo, R. Luo, M. Pfeifer and X. Ren for valuable discussions. This
25. Chi, S. S., Liu, Y., Song, W. L., Fan, L. Z. & Zhang, Q. Prestoring lithium into work was supported as part of the Center for Mesoscale Transport Properties, an Energy
stable 3D nickel foam host as dendrite-free lithium metal anode. Adv. Funct. Frontier Research Center supported by the US Department of Energy, Office of Science,
Mater. 27, 1700348 (2017). Basic Energy Sciences, under award no. DE-SC0012673. This work made use of the
26. Roberts, S. & Dobson, P. J. Evidence for reaction at the Al–SiO2 interface. Cornell Center for Materials Research Shared Facilities which are supported through
J. Phys. D 14, L17 (1981). the NSF MRSEC program (DMR-1719875). The pouch cell assembly line used for the
27. Bauer, R. S., Bachrach, R. Z. & Brillson, L. J. Au and Al interface reactions anode-free electrodes with a 9 cm2 area was supported under award 75039 from the New
with SiO2. Appl. Phys. Lett. 37, 1006–1008 (1980). York State Energy Research and Development Authority (NYSERDA) and award 76890
28. Bard A. J. & Faulkner L. R. Electrochemical Methods: Fundamentals and from the New York State Department of Economic Development (DED), which were
Applications (John Wiley and Sons, 2002). provided as matching funds to the Center for Mesoscale Transport Properties under
29. Yang, H. et al. The rechargeable aluminum battery: opportunities and award no. DE-SC0012673. E.S.T. acknowledges support as the William and Jane Knapp
challenges. Angew. Chem. Int. Ed. 58, 11978–11996 (2019). Chair of Energy and the Environment.
30. Pande, V. & Viswanathan, V. Computational screening of current collectors
for enabling anode-free lithium metal Batteries. ACS Energy Lett. 4,
2952–2959 (2019). Author contributions
31. So, K. P. et al. Improving the wettability of aluminum on carbon nanotubes. L.A.A. directed the research. J.Z. and L.A.A. conceived and designed this
Acta Mater. 59, 3313–3320 (2011). work. J.Z. and L.A.A. wrote the paper with inputs from other authors. J.Z. and T.T.
32. So, K. P., Biswas, C., Lim, S. C., An, K. H. & Lee, Y. H. Electroplating performed the electrodeposition, electrochemical measurements and structure
formation of Al–C covalent bonds on multiwalled carbon nanotubes. characterizations. J.Z., Q.Z., G.W. and D.C.B. assembled the pouch cells. D.C.B. and
Synth. Met. 161, 208–212 (2011). K.R.T. collected and interpreted the XPS data with input from A.C.M.; Q.Z., J.Y., X.L.,
33. Choudhury, S. et al. Designing solid–liquid interphases for sodium batteries. Y.D. and J.S. analysed the data. All the authors contributed to the review and editing
Nat. Commun. 8, 898 (2017). of the manuscript.
34. Jadhav, A. L., Xu, J. H. & Messinger, R. J. Quantitative molecular-Level
understanding of electrochemical aluminum–ion intercalation into a Competing interests
crystalline battery electrode. ACS Energy Lett. 5, 2842–2848 (2020). The authors declare no competing interests.
35. Zhao, Q. et al. Solid electrolyte interphases for high-energy aqueous
aluminum electrochemical cells. Sci. Adv. 4, eaau8131 (2018).
36. Zheng, J. et al. Reversible epitaxial electrodeposition of metals in battery Additional information
anodes. Science 366, 645 (2019). Extended data is available for this paper at https://doi.org/10.1038/s41560-021-00797-7.
37. Wang, Z. et al. A metal–organic framework host for highly reversible Supplementary information The online version contains supplementary material
dendrite-free zinc metal anodes. Joule 3, 1289–1300 (2019). available at https://doi.org/10.1038/s41560-021-00797-7.
38. Zheng, J. et al. Spontaneous and field-induced crystallographic reorientation
of metal electrodeposits at battery anodes. Sci. Adv. 6, eabb1122 (2020). Correspondence and requests for materials should be addressed to L.A.A.
39. Vaughan, J. & Dreisinger, D. Electrodeposition of aluminum from aluminum Peer review information Nature Energy thanks the anonymous reviewers for their
chloride–trihexyl(tetradecyl)phosphonium chloride. J. Electrochem. Soc. 155, contribution to the peer review of this work.
D68–D72 (2008). Reprints and permissions information is available at www.nature.com/reprints.
40. Barr, T. L. & Seal, S. Nature of the use of adventitious carbon as a binding
energy standard. J. Vac. Sci. Technol. A 13, 1239–1246 (1995). Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
41. Paranthaman, P. M. A high performance hybrid battery based on aluminum published maps and institutional affiliations.
anode and LiFePO4 cathode. Chem. Commun. 52, 1713–1716 (2016). © The Author(s), under exclusive licence to Springer Nature Limited 2021

406 NatUre Energy | VOL 6 | April 2021 | 398–406 | www.nature.com/natureenergy


Nature Energy Articles

Extended Data Fig. 1 | Electrochemical plating/stripping behavior of Al metal on nonplanar nickel foam substrate. a, Coulombic efficiency obtained at
0.8 mAh, 4 mA/cm2. Voltage profiles of Al plating/stripping: (b) 0.8 mAh, 4 mA/cm2;(c) 3.2 mAh, 1.6 mAh/cm2 and (d) 8.0 mAh, 1.6 mA/cm2. The results
mean that the improvement made by using a nonplanar, inert architecture is very limited, particularly at practical capacities, that is 3.2 and 8 mAh/cm2.

NatUre Energy | www.nature.com/natureenergy


Articles Nature Energy

a b C=O C-O
c
Data Data Data
45,000
Intensity (cps)

Peak 1 Peak 1 Peak 1


Peak 2 Peak 2 Peak 2
Peak 3 Peak 3 Peak 3
30,000 Background
Fit
Background
Fit
Background
Fit

15,000

0
290 288 286 284 290 288 286 284 290 288 286 284
Binding Energy (eV) Binding Energy (eV) Binding Energy (eV)
Extended Data Fig. 2 | XPS spectrum of interwoven carbon fibers. a, Pristine, (b) after exposure to IL+AlCl3 electrolyte, (c) after exposure to dimethyl
carbonate as a negative control. Peak assignments: 284.8 eV (C-C, C-H), 286 eV (C-O), 287.5–288 eV (C=O, O-C-O), 289 eV (O=C-O). The intensities at
286 and 287.5 ~ 288 eV suggest that the exposure to IL+AlCl3 notably increases the level of oxygen enrichment on carbon fibers.

NatUre Energy | www.nature.com/natureenergy


Nature Energy Articles

Extended Data Fig. 3 | SEM images and EDS mapping of Al deposition morphology obtained using a sequential, two-step protocol. In Step I, a small
capacity (0.05 mAh/cm2) of Al is deposited at fixed overpotential, η values (for example, η = 0.05, 0.3, 1.0, 2.0, 3.0 V); In Step II, a greater areal capacity
(that is 0.45 mAh/cm2) of Al is galvanostatically deposited at a current density (that is 4 mA/cm2). SEM images of the Al deposition morphology obtained
using this protocol, with the η value equal to (a) 0.05 V, (b) 0.3 V, (c) 1.0 V, (d) 2.0 V, and (e) 3.0 V. f–o, the corresponding EDS mapping results. See
Supplementary Note 1 for a detailed discussion.

NatUre Energy | www.nature.com/natureenergy


Articles Nature Energy

a Al on carbon fiber d Al on steel g Al on Ni foam


40,000 40,000
Data Data Data
12,000 Al-O-C
Intensity (cps)

Intensity (cps)

Intensity (cps)
Peak 1 Peak 1 Peak 1
Peak 2 Peak 2 Peak 2
Peak 3 30,000 Peak 3 30,000 Peak 3
Peak 4 Peak 4 Peak 4

8,000 Background Background Background


Fit
20,000 Fit
20,000 Fit

4,000 10,000 10,000


C 1s Data Data Data
12,000
Intensity (cps)

Intensity (cps)

Intensity (cps)
Peak 1 Peak 1 Peak 1
Peak 2
Peak 3
16,000 Peak 2
Peak 3
16,000 Peak 2
Peak 3
Peak 4 Peak 4 Peak 4

8,000 Peak 5 Background Background


Background
Fit 12,000 Fit
12,000 Fit

4,000
8,000 8,000
292 290 288 286 284 282 292 290 288 286 284 282 292 290 288 286 284 282

b Binding Energy (eV)


e Binding Energy (eV)
h Binding Energy (eV)
24,000 Data 60,000 Data Data
Al-O-C
Intensity (cps)

Intensity (cps)

Intensity (cps)
Peak 1 Peak 1 Peak 1
Background Background 36,000 Background
18,000 Fit Fit Fit

40,000
12,000 24,000
20,000
6,000 12,000
O 1s 24,000 Data 60,000 Data Data
Intensity (cps)

Intensity (cps)

Intensity (cps)
Peak 1 Peak 1 Peak 1
Peak 2 Peak 2 36,000 Background
18,000 Background
Fit
Background
Fit
Fit

40,000
12,000 24,000
20,000
6,000 12,000
538 536 534 532 530 528 538 536 534 532 530 528 538 536 534 532 530 528

c Binding Energy (eV) f Binding Energy (eV) i Binding Energy (eV)


Data 12,000 Data Data
Al-O-C
Intensity (cps)

Peak 1
Intensity (cps)

Intensity (cps)

Peak 1 Peak 1
3,000 Background Background 9,000 Background
Fit Fit Fit
8,000
2,000 6,000
4,000
1,000 3,000
Al 2p
Data 12,000 Data Data
Intensity (cps)

Peak 1
Intensity (cps)

Intensity (cps)

Peak 1 Peak 1
3,000 Peak 2 Peak 2 Peak 2
Peak 3 Background 9,000 Background
Background Fit Fit

2,000
Fit 8,000
6,000
1,000
4,000 3,000
80 78 76 74 72 80 78 76 74 72 80 78 76 74 72
Binding Energy (eV) Binding Energy (eV) Binding Energy (eV)

Extended Data Fig. 4. | XPS spectrum of Al electrodeposits on carbon fibers, stainless steel and nickel foam. XPS of Al deposited on (a)–(c) carbon
fibers, (d)–(f) stainless steel, and (g)–(i) Ni foam. (a)(d)(g) C 1 s spectra; (b)(e)(h) O 1 s spectra; (c)(f)(i) Al 2p spectra. Upper panels and lower panels
show spectra before and after Ar+ sputtering, respectively. After sputtering, the Al-O-C bonding was observed on samples where Al was deposited on
carbon fibers. On other samples, no characteristic metal-substrate covalent bonding is observable.

NatUre Energy | www.nature.com/natureenergy

You might also like