You are on page 1of 11

Author’s Accepted Manuscript

High Capacity 3D Structured Tin-Based


Electroplated Li-ion Battery Anodes

Pengcheng Sun, Jerome Davis, Luoxia Cao,


Zhelong Jiang, John B. Cook, Hailong Ning, Jinyun
Liu, Sanghyeon Kim, Feifei Fan, Ralph G. Nuzzo,
Paul V. Braun
www.elsevier.com/locate/ensm

PII: S2405-8297(18)30674-3
DOI: https://doi.org/10.1016/j.ensm.2018.11.017
Reference: ENSM564
To appear in: Energy Storage Materials
Received date: 4 June 2018
Revised date: 14 November 2018
Accepted date: 16 November 2018
Cite this article as: Pengcheng Sun, Jerome Davis, Luoxia Cao, Zhelong Jiang,
John B. Cook, Hailong Ning, Jinyun Liu, Sanghyeon Kim, Feifei Fan, Ralph G.
Nuzzo and Paul V. Braun, High Capacity 3D Structured Tin-Based Electroplated
Li-ion Battery Anodes, Energy Storage Materials,
https://doi.org/10.1016/j.ensm.2018.11.017
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
High Capacity 3D Structured Tin-Based Electroplated Li-ion Battery Anodes

Pengcheng Suna1*, Jerome Davis IIIc1, Luoxia Caod, Zhelong Jianga, John B. Cookc, Hailong
Ningc, Jinyun Liue, Sanghyeon Kima, Feifei Fand, Ralph G. Nuzzob, Paul V. Brauna,b*

a
Department of Materials Science and Engineering, Materials Research Laboratory, Beckman Institute for Advanced Science and
Technology, University of Illinois at Urbana-Champaign, Urbana IL 61801, USA
b
Department of Chemistry, University of Illinois at Urbana-Champaign, Urbana IL 61801, USA
c
Xerion Advanced Battery Corporation, 3100 Research Boulevard St. 320, Kettering OH 45420, USA
d
Department of Mechanical Engineering, University of Nevada, Reno NV 89557, USA
e
Key Laboratory of Functional Molecular solids of the Ministry of Education, College of Chemistry and Material Science, Anhui
Normal University, Wuhu, Anhui 241000, P.R. China

psn@illinois.edu
pbraun@illinois.edu

*
Corresponding authors:

ABSTRACT:
3D structured porous electrodes have been considered as a possible solution for accommodating
the volume change of alloying lithium ion battery anode materials during cycling. However, lab-
scale porous electrodes tend to be thin, and the loading of the activity materials is also small, the
combination of which results in electrodes with impractically low areal and volumetric capacities.
Here, we develop a high areal and volumetric capacity 3D-structured Sn/C anode by using a two
steps electroplating process. An electrode with a 20%v/v Sn loading exhibits a high
volumetric/areal capacity of ∼879 mAh/cm3/6.59 mAh/cm2 after 100 cycles at 0.5C and a good
rate performance of about 750mAh/cm3 and 5.5 mAh/cm2 (delithiation) at 10C in a half-cell
configuration. The 3D Sn/C anode also shows good compatibility with a commercial LCO
cathode in a full cell configuration.

Graphical Abstract
Here we show how a simple two step electroplating process enables synthesis of 75 µm thick 3D
structured Sn/carbon anodes with high Sn loadings. The electrode-based volumetric and areal
capacity of a 20% v/v Sn loaded scaffolded anode reached 1352 mAh cm-3 and 10.14 mAh cm-2.
Enabled by the electrode mesostructure, the structured anode provides an outstanding power
performance, delivering up to 750 mAh cm-3 and 5.6 mAh cm-2 at 10C, which to the best of our
knowledge, is the highest rate capability reported for such a high-energy anode.

1
These authors contributed equally to this work.

1
Keywords: volumetric capacity, structured anode, lithium ion battery; electroplating

1. Introduction
Secondary (rechargeable) lithium-ion batteries are widely used for applications ranging from
mobile electronics to electric vehicles and grid-scale energy storage. While Li-ion batteries have
improved significantly over the past decade, significant advances in energy and power density
are still desired.[1-5] While reports of advances in power and energy are common, too often the
advances are reported based on specific gravimetric capacities of the active material, and the
weight and volume of inactive components (e.g., current collectors and electrolyte) are ignored.[4]
Given the importance of the total cell mass and volume for most applications, specific capacity
values can be misleading.[4] Another important driver of the cell-level energy and power density
is the areal capacity of the electrode, as high areal capacity electrodes require fewer separators
and current collectors.[6]
To increase the cell-level energy density, one approach is to replace graphite, which has a
rather low volumetric capacity (theoretical 818 mAh cm−3, and practical ∼550 mAh cm−3)[7,8]
with tin, which has a theoretical volumetric capacity of 1991 mAh cm−3 on a Li4.4Sn lithiated
volume basis.[9] An additional potential advantage of Sn is that the lithiation potential is about
0.5 V higher than graphite, reducing the risk of Li-plating during charging.[10] Sn however cracks
and pulverizes during cycling because of its large volume change (260%) during the charge-
discharge process, leading to loss of electrical connectivity within the electrode and damage to
the solid electrode interface (SEI), and thus a poor cycling performance.[11-12]
We and others have considered three-dimensionally structured porous electrodes as a path for
accommodating the volume change of alloying anode materials such as tin and silicon during
cycling.[13-17] However, lab-scale porous electrodes tend to be thin, due to limitations in the
structured electrode fabrication process, and the active layer on the electrodes also tends to be
thin relative to the size of the pores within the electrodes. These factors generally result in
electrodes with low areal and volumetric capacities. [4] We have already shown how
electroplating can be used to deposit cathode (LiCoO2, LiMn2O4 and MnO2[14, 18]) and anode
(NiSn[15]) materials, however in some cases, for example, in our previous work on scaffolded
NiSn alloy-based anodes,[15] because the scaffold was rather thin (order 10 µm), the areal energy
density was low. Here we show how a simple two step electroplating process enables synthesis
of high power density 75 µm thick 3D structured Sn/carbon(Sn/C) anodes with high Sn loadings.
The electrode-based volumetric and areal capacity of a 20% v/v Sn loaded scaffolded anode was
up to 1352 mAh cm-3 and 10.14 mAh cm-2 which is around 4 and 2.5 times, respectively, the
capacity of current commercial high energy graphite electrodes. [19] Due to the porous electrode
2
structure, the structured Sn/C anode shows high power performance, delivering 750 mAh cm-3
and 5.6 mAh cm-2 at 10C, which to our knowledge, is the highest rate capability reported for
such a high-energy anode. [6] [19-21] A 20% v/v Sn Sn/C anode, when matched with a commercial
LCO electrode exhibited 74.2% capacity retention over 50 cycles in a full cell configuration.
While challenges remain to improve the cycling performance we find a carbon coating improves
the cycling performance, and improvements by the community in electrolytes and electrode
coating processes may enable further improvements in cycle life.

2. Result and discussion


Figure 1a-c shows the two-step fabrication process of the structured Sn/C anode. A Sn thin
film is first conformally electroplated on a Ni scaffold. This is followed by a conformal carbon
electrodeposition step via electrolysis of ethanol at high voltage.[22] SEM images the structured
Ni scaffold current collector are shown in figure 1d and e. The porous Ni scaffold is about 75 µm
thick and about 5% Ni by volume. The diameter of the Ni tubes making up the porous structure
is about 800 nm and their wall thickness is about 25 nm (Fig. 1e). The Sn layer is about 200 nm
thick, giving a total Sn loading of about 20% v/v (Fig. 1f). Energy dispersive x-ray spectroscopy
(EDS) mapping of the Sn distribution in the Ni scaffold is shown in Figure S1. SEM images of
the structured Sn/C anodes with different Sn loadings are shown in Figure S2. As shown in
figure 1g, the rough surface of the porous structure appears smoother and denser after carbon
deposition, probably due to the presence of the carbon layer.

a b c
Sn Carbon
plating plating

Ni scaffold Sn/Ni scaffold C/Sn/Ni scaffold

Fig 1. (a-c) Schematic illustration of the structured Sn/C anode fabrication process. (d,e) Low and high
magnification SEM images of the Ni scaffold. (f) SEM image of the Sn coated Ni scaffold. (g) SEM image of the
carbon and Sn coated Ni scaffold. Scale bars in the insets are 500 nm.

X-ray diffraction on (XRD) of the structured Sn/C anode is shown in Figure 2a. Using
Rietveld refinement, the experimental XRD data (black curve) could be fit using a combination
of β-Sn (ICDD PDF#04-004-6229) and Ni (ICDD PDF#04-006-4262) (red curve). The pattern
refines to tetragonal Sn (I41/amd space group) with a=5.8295(22)Å, c=3.1839(13)Å (blue curve),
and cubic Ni (Fm-3m space group) with a=3.5256(15)Å (green curve). The difference (grey
curve) between the experimental data and the fit is small and mainly due to texturing in the film.
Transmission electron microscopy (TEM) is used to characterize the microstructure of the 3D
Sn/C anode. As shown in Figure 2b, a 25 nm thick carbon film is conformally coated on the Sn.

3
Selected area electron diffraction (SAED), obtained from a 50 nm diameter region near the edge
of the sample shown in figure 2b, indicates the Sn is crystalline (inset Fig. 2b). Raman
spectroscopy shows two broad peaks around 1350 and 1600 cm-1(Fig. 2c). These two peaks are
related to the disordered graphite band at 1355 cm-1(D line) and the single-crystal graphite band
at 1580 cm-1(G line). The presence of both peaks indicates the carbon is a mixture of sp2 and sp3
carbon and generally amorphous. [5] [22] The weight fraction carbon in the Sn/C layer estimated by
EDS and TGA are approximately 9.5%, 8.2% and 5.4% for the 10%, 15% and 20% Sn loaded
electrodes, respectively (Fig. S3 and Fig.S4).

a Experimental data
Refinement Fit
Intensity (a.u)

(200)

(101)

(220)
(211)

(400)
(112)

(411)
(420)
(321)

(312)

(501)
(301)

c
Carbon film
Sn
Intensity (a.u)
(111)

(200)

(220)

(311)

Ni
Difference

Mo-Kα 2θ (°) Raman shift(cm-1)


Fig 2. Characterization of the structured Sn/C anode. (a) XRD and its Rietveld refinement. (b) TEM image, scale bar
is 10 nm-1 in SAED. (c) Raman spectrum.

Figure 3a presents representative cyclic voltammograms (CVs) of the initial two cycles at a
sweep rate of 0.1 mV/s between 0.01 and 2.0 V for structured Sn/C anodes with different Sn
loadings (using lithium as the cathode). In the first cathodic scan, there is no peak at about 1.5 V
confirming Sn is well passivated by the amorphous carbon. If the Sn was in direct contact with
electrolyte, a peak due to electrolyte decomposition at about 1.5 V would be observed. [23,24] The
peaks at about 0.8 V are ascribed to SEI formation on carbon surface.[24,25] The distinct peaks
between 0.3 and 0.7 V in the cathodic scan are attributed to the alloying of lithium into tin
forming LixSn, while in the anodic scan the peaks between 0.5 and 0.8 V are assigned to the
dealloying process (Sn + xLi+ + xe− ↔LixSn). [26] No obvious SEI formation peak is observed in
the second cathodic scans for the structured Sn/C anodes, indicating the initial SEI is stable on
the carbon coated electrode.

4
Fig 3. Electrochemical performance of the structured Sn/C anodes in half cell configuration. (a) CV curves of
structured Sn/C anodes with different Sn loadings over the potential range of 0−2 V vs Li/Li + at a scan rate of 0.1
mV s−1. (b) 1st and 2nd cycle galvanostatic charge−discharge curves of Sn/C anodes with different Sn loadings at
0.1C from 0.01 to 2 V. (c) Areal, volumetric capacity and coulombic efficiency of structured Sn/C anodes with
different Sn loadings over 100 cycles at 0.5C (pre-cycled at 0.1C and 0.2C for 2cycles respectively). (d) Areal and
volumetric capacities of structured Sn/C anodes with different Sn loadings at varying C-rates. All volumetric
capacities are calculated using the full Sn/C electrode volume (including the void space).

Galvanostatic discharge−charge curves of the first two cycles at 0.05C (1C corresponds to
complete charge or discharge of the Sn theoretical capacity in 1 h) for structured Sn/C anodes
with different Sn loadings are shown in Figure 3b. The plateaus at around 0.7 to 0.3 V during the
discharge (lithiation) are consistent with the CV profiles and previous reports on Sn-based
anodes and can be ascribed to alloying between tin and lithium to form Li xSn. [15,16] During
charge (delithiation), the plateaus at ∼0.4 to 0.8 V are assigned to dealloying of Li xSn. The 10,
15 and 20% v/v Sn electrodes (full electrode basis) exhibit first cycle charge volumetric
capacities of 682, 1037 and 1352 mAh cm−3, respectively. For all loadings, the specific
gravimetric capacity of the Sn in the structured Sn/C anode is about 1170 and 940 mAh g-1 for
the 1st discharge and charge cycles, indicating the electroplated Sn can deliver its near theoretical
capacity when within the structured nickel scaffold. [27]
Prior to cyclability testing (100 cycles at 0.5C), the Sn/C anodes are pre-cycled twice at 0.05C
and twice at 0.1C. Pre-cycling proved beneficial for improving the electrode stability. [28,29] The
first cycle charge areal capacities (0.05C) of electrodes with 10, 15, 20% v/v Sn are 5.1, 7.8 and
10.1 mAh cm-2, with corresponding first cycle columbic efficiencies (CE) of 82.5%, 82.6% and
80.5%, respectively. The first cycle CEs of the structured Sn/C anodes are much higher than
many other Sn anodes in the literature[12] which we suspect is a direct result of both the

5
structured Ni support and the conformally coated carbon layer, which combine to reduce stress
accumulation and SEI formation. The CEs all increase to greater than 90% in the second cycle.
For 10 and 15% v/v Sn loadings, the CE increases to about 99% after about 10 cycles. The CE of
the 20% v/v Sn electrode settles at about 98% after 10 cycles; the lower CE may be caused by
the larger volume change in the active materials during cycling, which causes additional SEI
formation. After 100 cycles at 0.5C, the capacity retention of 10, 15 and 20% v/v Sn anodes are
81.9%, 77.6%, and 68.8%, respectively. The areal capacities after 100 cycles are 3.92, 5.58 and
6.59 mAh cm-2. While the 20% v/v Sn Sn/C anode shows the largest capacity decay during
cycling, it still delivers an areal capacity much higher than previously reported graphite, Sn, and
Si-based anodes with similar rate capability. [19] [21] [30] As suggested by one of the reviewers, one
possible reason for the increase in capacity decay with Sn loading could be the instability of the
lithium metal counter electrode. At higher Sn loadings, there is greater deplating and plating at
the Li anode with each cycle, which may lead to more pulverization and SEI formation. Given
the high Sn loading of the Sn/C anodes, degradation of the lithium electrode thus might also
contribute to the observed capacity decay during the cycling.[31] We plan to investigate this
further in a future report.
Figure 3d gives the rate performance of structured Sn/C anodes with different Sn loading.
Both discharge and charge capacities decrease with increasing C-rates for all electrodes. At 10C,
the charge volumetric and areal capacities are about 430, 595, 740 mAh cm−3 and 3.2, 4.4, 5.5
mAh cm-2 for 10, 15, and 20%v/v Sn loadings, respectively. When the C rate returned to 0.1C,
the charge volumetric capacities recovered to 610, 874 and 1110 mAh cm-3 and the areal
capacities recovered to about 4.5, 6.4 and 8.4 mAh cm-2. As a comparison, the rate capability of
high-power MCMB graphite anodes with similar areal capacities at 0.1 C were also conducted,
and as shown in Figure S5, these graphite anodes with areal capacities of 4.5, 7.0 and 9.5 mAh
cm-2 exhibited significantly lower charge volumetric capacities (256.3, 267.0 and 311.8 mAh cm-
3
) when cycled at 5, 1 and 0.5C, respectively. Note, if higher C rate is used for the graphite
anodes, a large overpotential is observed and the electrode capacities becomes quite small. To
match the areal capacities of the 10, 15, and 20% v/v Sn Sn/C anodes, the graphite anodes were
required to be about 100, 180, and 240 µm thick; this thickness is one reason for their poor rate
capability. Although the solid-state diffusion coefficient of Li is several orders of magnitude
larger in graphite than Sn, polarization caused by the thick graphite anode (which includes binder
and conductive additive) apparently limits the rate capability. [2] [9] [19] The good rate performance
of structured Sn/C anode is probably because the Sn is in near contact with both the conductive
Ni scaffold and the electrolyte, and also because the mesoporous 3D structure provides a
continuous and relatively low tortuosity network for liquid phase diffusion. [14]
A full cell consisting of the pre-cycled Sn/C anode and a commercial LiCoO2 (LCO) cathode
was constructed. The first two charge-discharge curves for different Sn loading are shown in
Figure 4a. Charge corresponds to the lithiation of Sn/C anode, discharge corresponds to
delithiation. The LCO cathode possesses an operating voltage of 3.7−4.2 V (Fig. S6), and the Sn-
based anode exhibits lithiation and delithiation plateaus between 0.7−0.3 V and 0.4−0.8 V,
respectively. The stages from ∼3.5 V to ∼3.7 V during charge are ascribed to the Li−Sn alloying
and the delithiation of LCO, while the stages from ∼3 to 3.5 V during discharge are attributed to
dealloying. The first cycle discharge volumetric capacities, 648, 961, 1259 mAh cm-3 for the 10,
15, and 20% v/v Sn Sn/C anodes are slightly smaller than in the half-cell, perhaps due to the
limited lithium available from LCO. [1]

6
Fig 4. Electrochemical performance of the structured Sn/C anodes in full cell configuration. (a) The 1st and 2nd
cycling galvanostatic charge−discharge curves of full cells consisting of LCO cathodes and structured Sn/C anodes
with different Sn loadings at 0.5 C. (b) Areal, volumetric capacity and coulombic efficiency of the full cells over 50
cycles at 0.5C. All volumetric capacities are calculated using the full Sn/C electrode volume (including void space).

To further examine the compatibility of the Sn/C anode with a commercial LCO anode, full
cell cycling is conducted. As shown in Figure 4b, the first cycle discharge areal capacities (0.05C)
of the electrodes with 10, 15 and 20% v/v Sn are 4.85, 7.20 and 9.44 mAh cm-2, respectively. The
corresponding first cycle CEs are 90.45%, 90.23% and 90.36%. The CEs all increase to 95% in
the second cycle and after about 8 cycles, the CE stabilizes at 99% for lower Sn loading
electrodes (10 and 15% v/v), and 98% for the 20% v/v Sn electrode. After 50 cycles, the
discharge capacity retention is 82.9%, 78.4% and 74.2% for the 10, 15 and 20% v/v Sn Sn/C
anodes.
To understand the effects of Sn loading on cycle performance, structured Sn/C anodes with
different Sn loadings are investigated using electrochemical impedance spectroscopy (EIS) after
100 cycles at 0.5C. One cell made of a 10% v/v Sn structured anode formed without the carbon
coating step is also tested to evaluate the effects of carbon coating. The results are displayed in
Figure 5a. The Nyquist plots for the Sn/C anodes consist of at least two semicircles in the high
and medium frequency regions and a straight line in the low frequency region. The Nyquist plot
for the bare structured Sn anode shows a large semicircle with a depression in the high frequency
region, is a sign of multiple overlapping semicircles. [15] We use the same equivalent circuit to fit
all the curves (Fig. S7). The semicircle at high frequency provides the contact resistance (Re) and
SEI resistance (Rs), the semicircle across the medium-frequency region provides the charge-
transfer impedance (Rct) of the electrode/electrolyte interface, and the low-frequency linear tail
corresponds to the Warburg impedance (Zw) associated with diffusion of lithium ions in the bulk
electrode.[32] The Re, Rs and Rct of structured Sn anodes are shown in Table S1. Rs for Sn/C
anodes with 10, 15 and 20% v/v Sn are 13.23, 20.18 and 38.33 Ω, respectively, indicating higher
Sn loading may result in a thicker SEI after cycling, which is consistent with the cycling
performance of the electrodes. The Rs of the bare Sn anode with 10% loading is 134.10 Ω, which
is about 10 times larger than the Sn/C anode, indicating significantly more SEI formation when
the carbon layer is not present. The smaller Rs of 3D Sn/C anodes might be benefit from the
carbon coating. [23] [33]
The morphology of the structured Sn/C and Sn anodes after 100 cycles was investigated by
SEM. A typical cycled 10%Sn(v/v) loaded 3D Sn/C anode is shown in Figure 5b, and although
there are white spots (perhaps SEI[34]) on the Sn/C surface, no delamination is observed, and the

7
tubular Ni scaffold structure is maintained despite the significant volume changes of the Sn
during cycling. In contrast, there are many fractures on the bare Sn coated 3D Ni scaffold and
what appears to be a thick SEI layer present everywhere (Fig. 5c).[34] SEM images of cycled 3D
Sn/C anodes with different Sn loadings (Fig. S8) show the carbon coating helps maintain the
integrity of the 3D Sn/C anode, but also that SEI formation and the density of fractures increase
with higher Sn loadings (which agrees with the larger capacity decay for the more highly loaded
3D Sn/C anodes). Combined with the electrochemical data, SEM images indicate that the
carbon-coated Sn is relatively stable to cycling compared to the bare Sn and the Ni scaffold
remains robust to the volume change of the Sn. The control cycling data for 10% Sn(v/v) loaded
3D Sn anodes with and without carbon coating is shown in Figure. S9. The effect of the carbon
coating on mechanical deformation, which is associated with the continuous formation of SEI
with cycling, was investigated through finite element simulations (see details and Fig. S10 in the
Supporting Information). Lithiation of the bare Sn anode induces large surface hoop tension
which can cause fracture and pulverization of the Sn, leading to continuous formation of SEI on
new surfaces. In comparison, simulations indicate the carbon coating reduces the outward
swelling. Presumably at some degree of swelling, the carbon coating cracks, but even should that
happen, the carbon will reduce the exposure of tin to the electrolyte and thus reduce the rate of
SEI formation.

a
a

Fig 5. Characterization of the structured Sn/C anodes after 100 cycles at 0.5C; a, EIS from 100 kHz to 0.01 Hz; b-c,
Typical SEM images of structured Sn/C anodes with (b) and without (c) carbon coating.

Conclusion
We use a two-step electrodeposition process to fabricate a high volumetric and areal capacity
structured Sn/C anodes starting with a thick structured 3D Ni scaffold. A 20% v/v Sn anode
provides volumetric and areal capacities of ∼879 mAh cm-3 and 6.59 mAh cm-2 after 100 cycles
at 0.5C, and about 750 mAh cm-3 and 5.5 mAh cm-2 even at 10C. The Sn/C anodes show good
compatibility with commercial LCO cathodes. EIS and finite element simulations are used to
investigate the effects of Sn loading and carbon coating on cycling and power performance of the
Sn/C anodes. Higher Sn loadings do cause a larger capacity decay with cycling, but the carbon

8
coating delays the capacity decay, perhaps because it reduces continuous SEI formation on the
Sn with cycling. Additional research is still necessary to reduce capacity decay, and we hope this
work stimulates additional work in this area.

Acknowledgements
Research at the University of Illinois supported by the U.S. Department of Energy, Office of
Basic Energy Sciences, Division of Materials Sciences and Engineering under Award # DE-
FG02-07ER46471, through the Materials Research Laboratory at the University of Illinois at
Urbana-Champaign. We thank Qiye Zheng for collecting Raman spectroscopy from the 3D Sn/C
sample.

Appendix A. Supporting information


Supplementary data associated with this article can be found in the online version

References
[1] M.-S. Balogun, W. Qiu, Y. Luo, H. Meng, W. Mai, A. Onasanya, T. K. Olaniyi, Y. Tong,
Nano Research 2016, 9, 2823.
[2] H. Buqa, D. Goers, M. Holzapfel, M. E. Spahr, P. Novák, Journal of the Electrochemical
Society 2005, 152, A474.
[3] H. Chen, T. N. Cong, W. Yang, C. Tan, Y. Li, Y. Ding, Progress in Natural Science 2009,
19, 291.
[4] J. W. Choi, D. Aurbach, Nature Reviews Materials 2016, 1, 16013.
[5] P. K. Chu, L. Li, Materials Chemistry and Physics 2006, 96, 253.
[6] A. M. Gaikwad, B. V. Khau, G. Davies, B. Hertzberg, D. A. Steingart, A. C. Arias,
Advanced Energy Materials 2015, 5, 1401389.
[7] N. Nitta, F. Wu, J. T. Lee, G. Yushin, Materials Today 2015, 18, 252.
[8] I. H. Son, J. Hwan Park, S. Kwon, S. Park, M. H. Rümmeli, A. Bachmatiuk, H. J. Song, J.
Ku, J. W. Choi, J.-m. Choi, S.-G. Doo, H. Chang, Nature Communications 2015, 6, 7393.
[9] N. Nitta, G. Yushin, Particle & Particle Systems Characterization 2014, 31, 317.
[10] H. Tian, F. Xin, X. Wang, W. He, W. Han, Journal of Materiomics 2015, 1, 153.
[11] L. Xu, C. Kim, A. K. Shukla, A. Dong, T. M. Mattox, D. J. Milliron, J. Cabana, Nano
Letters 2013, 13, 1800.
[12] W.-M. Zhang, J.-S. Hu, Y.-G. Guo, S.-F. Zheng, L.-S. Zhong, W.-G. Song, L.-J. Wan,
Advanced Materials 2008, 20, 1160.
[13] J. H. Pikul, H. Gang Zhang, J. Cho, P. V. Braun, W. P. King, Nature Communications
2013, 4, 1732.
[14] H. Zhang, X. Yu, P. V. Braun, Nature Nanotechnology 2011, 6, 277.
[15] H. Zhang, T. Shi, D. J. Wetzel, R. G. Nuzzo, P. V. Braun, Advanced Materials 2016, 28,
742.
[16] J. B. Cook, E. Detsi, Y. Liu, Y.-L. Liang, H. Kim, X. Petrissans, B. S. Dunn, S. H.
Tolbert, ACS Applied Materials & Interfaces 2016, 9, 293.
[17] J. B. Cook, T. C. Lin, E. Detsi, J. N. Weker, S. H. Tolbert, Nano Letters 2017, 17, 870.

9
[18] H. Zhang, H. Ning, J. Busbee, Z. Shen, C. Kiggins, Y. Hua, J. Eaves, J. Davis, T. Shi, Y.-
T. Shao, J.-M. Zuo, X. Hong, Y. Chan, S. Wang, P. Wang, P. Sun, S. Xu, J. Liu, P. V. Braun,
Science Advances 2017, 3, e1602427
[19] K. G. Gallagher, S. E. Trask, C. Bauer, T. Woehrle, S. F. Lux, M. Tschech, P. Lamp, B. J.
Polzin, S. Ha, B. Long, Q. Wu, W. Lu, D. W. Dees, A. N. Jansen, Journal of The
Electrochemical Society 2015, 163, A138.
[20] R. Yi, J. Zai, F. Dai, M. L. Gordin, D. Wang, Nano Energy 2014, 6, 211.
[21] M. Tian, W. Wang, Y. Wei, R. Yang, Journal of Power Sources 2012, 211, 46.
[22] Z. Sun, Y. Sun, X. Wang, Chemical Physics Letters 2000, 318, 471.
[23] I. Meschini, F. Nobili, M. Mancini, R. Marassi, R. Tossici, A. Savoini, M. L. Focarete, F.
Croce, Journal of Power Sources 2013, 226, 241.
[24] Y. S. Jung, K. T. Lee, J. H. Ryu, D. Im, S. M. Oh, Journal of The Electrochemical
Society 2005, 152, A1452.
[25] H. Zhang, H. Song, X. Chen, J. Zhou, The Journal of Physical Chemistry C 2012, 116,
22774.
[26] M. N. Obrovac, V. L. Chevrier, Chemical Reviews 2014, 114, 11444.
[27] D. H. Youn, A. Heller, C. B. Mullins, Chemistry of Materials 2016, 28, 1343.
[28] N. Nitta, G. Yushin, Particle & Particle Systems Characterization 2014, 31, 317.
[29] Y. M. Lee, J. Y. Lee, H.-T. Shim, J. K. Lee, J.-K. Park, Journal of The Electrochemical
Society 2007, 154, A515.
[30] Q. Xiao, Y. Fan, X. Wang, R. A. Susantyoko, Q. Zhang, Energy Environ. Sci. 2014, 7,
655.
[31] Z. Li, J. Huang, B. Yann Liaw, V. Metzler, J. Zhang, Journal of Power Sources 2014, 254,
168.
[32] S. S. Zhang, K. Xu, T. R. Jow, Electrochimica Acta 2006, 51, 1636.
[33] Y. Xu, Q. Liu, Y. Zhu, Y. Liu, A. Langrock, M. R. Zachariah, C. Wang, Nano Letters
2013, 13, 470.
[34] L. Zhong, C. Beaudette, J. Guo, K. Bozhilov, L. Mangolini, Scientific Reports 2016, 6,
30952.

10

You might also like