Anton MA 2812

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

Article
Direct Imprinting of Scalable, High-Performance Woodpile
Electrodes for Three-Dimensional Lithium-Ion Nanobatteries
Wenhao Li, Yiliang Zhou, Irene R. Howell, Yue Gai, Aditi R.
Naik, Shengkai Li, Kenneth R. Carter, and James J. Watkins
ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.7b14649 • Publication Date (Web): 25 Jan 2018
Downloaded from http://pubs.acs.org on January 26, 2018

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society.
1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 22 ACS Applied Materials & Interfaces

1
2
3
4
5
Direct Imprinting of Scalable, High-Performance
6
7
8
Woodpile Electrodes for Three-Dimensional
9
10
11
Lithium-Ion Nanobatteries
12
13
14
15
16
17 Wenhao Li, Yiliang Zhou, Irene R. Howell, Yue Gai, Aditi R. Naik, Shengkai Li, Kenneth R.
18
19 Carter, James J. Watkins
20
21
Department of Polymer Science and Engineering, University of Massachusetts Amherst,
22
23
24 Amherst, Massachusetts 01003, USA.
25
26 
27 W. Li and Y. Zhou contributed equally to this work.
28
29
30
31
KEYWORDS: nanoimprint lithography, direct patterning, 3D structures, nanomaterials, lithium-
32
33 ion batteries
34
35
36
37
38 ABSTRACT: The trend of device downscaling drives a corresponding need for power
39
40 source miniaturization. Though numerous microfabrication methods lead to successful
41
creation of submillimeter-scale electrodes, scalable approaches that provide cost-effective
42
43 nanoscale resolution for energy storage devices such as on-chip batteries remain elusive.
44
45 Here, we report nanoimprint lithography (NIL) as a direct patterning technique to
46
47 fabricate high-performance TiO2 nanoelectrode arrays for lithium-ion batteries (LIBs)
48 over relatively large areas. The critical electrode dimension is below 200 nm, which enables
49
50 the structure to possess favorable rate capability even under discharging current densities
51
52 as high as 5000 mAg-1. In addition, by sequential imprinting, electrodes with three-
53
54
dimensional (3D) woodpile architecture were readily made in a “stack-up” manner. The
55 height of architecture can be easily controlled by the number of stacked layers while
56
57
58 1
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 2 of 22

1
2
3 maintaining nearly constant surface-to-volume ratios. The result is a proportional increase
4
5 of areal capacity with the number of layers. The structure-processing combination leads to
6
7 efficient use of the material and the resultant specific capacity (250.9 mAhg-1) is amongst
8
9
the highest reported. This work provides a simple yet effective strategy to fabricate
10 nanobatteries and can be potentially extended to other electroactive materials.
11
12
13
14
15
16 1. Introduction
17
18
19
The advent of small devices, from microelectromechanical systems (MEMSs)1 to more recently,
20
21 nanoelectromechanical systems (NEMSs),2 requires power sources with commensurate
22
23 dimensions. The use of on-chip batteries, e.g. micro- and even nanobatteries, effectively avoids
24
25
interconnection problems and unnecessary signal noise encountered by devices that are
26
27
28 externally powered.3 The downscaling of battery size will significantly enhance the overall
29
30 portability and drive MEMS/NEMS toward fully autonomous devices.
31
32
33 A primary challenge for constructing small batteries is to fine-tune the electrode 3D architectures
34
35 in order to enable both high capacity and high power using a limited footprint area.4 To date,
36
37
38 several techniques have been applied to create such electrodes including colloidal templating,5-7
39
40 vapor deposition on 3D substrates8-11 and direct ink writing.12,13 Though these methods result in
41
42 delicate electrode fabrication, some key issues remain to be solved. First, the dimensions of most
43
44
45
of the reported electrodes are on the order of tens of microns. Further downscaling to sub-
46
47 micrometer scale will enhance the reaction-diffusion kinetics however, the cost of fabrication
48
49 using traditional subtractive clean room processing techniques soars as dimensions are driven to
50
51
the smaller dimensions. Moreover, from a commercial point of view, any chosen fabrication
52
53
54 technique needs to be facile, scalable and cost-effective. Current methods either require multi-
55
56
57
58 2
59
60 ACS Paragon Plus Environment
Page 3 of 22 ACS Applied Materials & Interfaces

1
2
3 step processing, extensive post-treatments or are restricted to relatively low throughput
4
5
6 manufacturing as limited by the deposition chamber size or the writing speed.
7
8
9 Nanoimprint lithography (NIL) is widely considered to be a relatively low-cost, high-resolution
10
11 and high-throughput surface patterning technique.14-16 Materials that can be patterned range from
12
13 self-assembled monolayers (SAMs),17 polymers and colloidal materials,18,19 to inorganic sol-gel
14
15
16
materials.20 More recently, a number of reports have explored NIL to directly pattern inorganic
17
18 nanoparticles/tubes into various architectures.21-23 For example, well dispersed gold and indium
19
20 tin oxide (ITO) nanoparticles can be directly molded into designed patterns via solvent-assisted
21
22 NIL.24,25 Nevertheless, due to the low modulus of elastomeric stamps,26 insufficient mass
23
24
25 transfer,27 and the shrinkage upon solvent drying, the aspect ratio of the imprinted structures are
26
27 generally small especially for non-polymer imprinting.28,29 This makes NIL more of a 2D
28
29 patterning technique and limits its use in 3D electrode fabrication. Recently our group has
30
31
32 designed a general imprinting approach to 3D metal oxide nanostructures, including the
33
34 woodpile structure.30 We utilized inks comprised predominantly of crystalline nanoparticles and
35
36 sequential imprint-planarization cycles, followed by removal of sacrificial planarization layers
37
38
through heating. This “stack-up” protocol enables the attainment of 3D architectures with high,
39
40
41 and importantly, user-defined aspect ratios.
42
43
44 Here, we presented a rapid and scalable approach for the fabrication of high-performance TiO2
45
46 woodpile nanoelectrodes for lithium-ion battery (LIB) using solvent-assisted NIL. The electrodes
47
48 demonstrate superior rate capability and exhibited a specific capacity of 250.9 mAhg-1, which is
49
50
51 among the highest reported. As the number of stacked layers increases, the electrode’s areal
52
53 capacity exhibits a proportional enhancement, demonstrating great potential as an on-chip power
54
55 source. To the best of our knowledge, this is the first report on 3D battery electrode fabricated by
56
57
58 3
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 4 of 22

1
2
3 direct NIL of electroactive nanoparticles. The compatibility of NIL with the current
4
5
6 micro/nanofabrication process make it potentially a platform technology toward scaled
7
8 production and integration of nanobatteries into MEMS/NEMS.
9
10
11 2. Experimental Section
12
13
14 2.1. Fabrication of PDMS Mold: Patterned PDMS stamps were fabricated by casting the
15
16 prepolymer mixture against silicon master molds with a line grating pattern (linewidth ~ 425 nm;
17
18
pitch~ 950 nm; height~ 480 nm). The prepolymer mixture (10:1 weight ratio of Sylgard 184
19
20
21 silicone elastomer base and curing agent) was mixed and degassed in vacuum oven for 20 min at
22
23 room temperature. Then the mixture was poured onto the master mold and placed at 70 ℃ for 5 h
24
25
in an oven. After curing, PDMS stamps were obtained via peeling off from master mold.
26
27
28 2.2. Preparation of TiO2 Nanoparticle Ink: To obtain TiO2 nanoparticle ink, a commercial
29
30
31 titanium oxide (anatase, 20 wt%) nanoparticle dispersion in 1, 2 propanediol (US Research
32
33 Nanomaterials, Inc.) was mixed with 1, 2 propanediol and methanol in 1:1:5 weight ratio. This
34
35 mixture dispersion was further vortex-mixed and sonicated for a few minutes to obtain a stable,
36
37
38
well-dispersed ink. Particle size and distribution were checked by TEM (JEOL 2000FX) and
39
40 Malvern Nano Zetasizer.
41
42
43 2.3. Fabrication of Woodpile Structure and the Planar Control Samples: Silicon dioxide
44
45 wafer substrates were sonicated in ethanol and acetone for 5 min. Then, these substrates were
46
47
washed with deionized water, and dried under nitrogen gas. A Ti/Au (5 nm/50 nm) bilayer was
48
49
50 deposited on the silicon dioxide wafer with a designed pattern, controlled by a customized
51
52 shadow mask. The TiO2 nanoparticle ink was spin-coated onto the silicon dioxide substrate
53
54 (3000 rpm, 90 s) in a glove box with a 5 % relative humidity environment. Then, PDMS stamp
55
56
57
58 4
59
60 ACS Paragon Plus Environment
Page 5 of 22 ACS Applied Materials & Interfaces

1
2
3 was placed on the ink and dried on the hot plate at 55 °C for 5 min. After drying, PDMS stamp
4
5
6 was peeled off and the patterned film was obtained. A UV-crosslinkable thiol-ene acrylate
7
8 prepolymer (NOA60, Norland Products Inc.) was used for planarization. After imprinting the
9
10 first layer, the patterned structure was first pretreated with UV-ozone for 15 min. Then two
11
12
13
layers of NOA60 (10 wt% in propylene glycol monomethyl ether acetate) were deposited by
14
15 consecutively spin coating at 1000 rpm for 1 min and curing. The imprinting-planarization
16
17 process was simply repeated for the desired number of layers. Finally, TiO2 with designed layers
18
19
was calcined at 750 °C for 5 min. Structure dimensions were checked under SEM (Magellan
20
21
22 400). All control samples were fabricated by multi-time spin coating and thermal annealing. The
23
24 film was thermally annealed at lower temperature (400 ℃) for 5 min after each spin coating
25
26 except for the last time that the sample was calcined at 750 ℃ for 5 min like the woodpiles. Film
27
28
29 thickness and porosity were checked by ellipsometry.
30
31
32 2.4. Electrochemical Tests: All electrochemical measurements were conducted in the half-cell
33
34 configuration, assembled in an argon glove box. LiClO4 (1M, in EC/DMC=1/1v) was used as the
35
36 liquid electrolyte. A piece of lithium foil served as both the counter and reference electrode.
37
38
39 Charge and discharge profiles of the electrodes were measured by galvanostatic tests (Maccor
40
41 4304) under different C-rates, within a voltage window of 1.0-3.0 V. Extended cycling tests were
42
43 performed with voltage window of 0.4-3.0 V. Control samples of un-patterned TiO2 films with
44
45
varied thickness were tested under same conditions and used for comparison.
46
47
48
49
50
51
3. Results and Discussions
52
53
54
55
56
57
58 5
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 6 of 22

1
2
3 The TiO2 ink used in this study was first engineered to provide imprintability. Nanoparticle size,
4
5
6 dispersity, solid concentration, and ink viscosity are among the key factors. Here, the imprint ink
7
8 was made from a 20 wt% commercial TiO2 dispersion in 1, 2-propanediol. Transmission electron
9
10 microscopy (TEM) shows that the particle size is below 10 nm (Figure 1a). This is comparable
11
12
13 with the dynamic light scattering (DLS) measurement where the number- and volume-averaged
14
15 particle sizes are 13.6 and 15.7 nm (Figure 1b). The slight difference results mainly from the
16
17 inclusion of solvation effect and ligands in DLS. The ink was then diluted in a 1,2-
18
19
20
propanediol/methanol mixture solvent to have a 3 wt% solid concentration. The incorporation of
21
22 methanol is critical in three ways: 1). Methanol lowers the ink viscosity and improves wettability
23
24 to facilitate film formation; 2). The low boiling-point nature (b.p. 64.7 ℃ ) leads to fast
25
26
27
evaporation during spin coating and quickly concentrates the ink while the high boiling-point
28
29 1,2-propanadiol (b.p. 188.2℃) provides reasonable fluidity for molding; 3). The evaporation
30
31 induced composition change of ink does not lead to severe particle aggregation as a result of
32
33
good miscibility. As shown in Figure 1c, the ink exhibits very low viscosity of 1.8 mPa ∙ s which
34
35
36 is close to the viscosity of the mixture solvent (1.37 mPa ∙ s at 25 ℃ ). This relatively low
37
38 concentration and viscosity is critical to mold filling and residual layer-free imprinting. The ink
39
40
was stable with negligible precipitation observed over several months (zeta potential measured to
41
42
43 be +13.4 mV). In fact, with 5-min sonication, ink stored for extended periods of time can be used
44
45 as successfully as fresh ink in the imprinting process. The electron diffraction (ED) pattern of the
46
47 TiO2 nanoparticles (Figure 1d) demonstrates a series of distinct diffraction rings, indicative of
48
49
50 the polycrystalline feature and is indexed to anatase phase. The calculated d-spacing of the (101)
51
52 planes is 0.36 nm as confirmed by high resolution TEM imaging (Figure 1e).
53
54
55
56
57
58 6
59
60 ACS Paragon Plus Environment
Page 7 of 22 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25 Figure 1. (a) TEM image of TiO2 nanoparticles showing particle size below 10 nm. (b) DLS
26
27 measurement of nanoparticle size in 1,2-propanediol/methanol mixture solvent at room temperature. (c)
28 Shear viscosity of the TiO2 ink measured at room temperature under shear rate between 100-1000 s-1.
29
30 (inset: optical image of the well-dispersed TiO2 ink). (d) Electron diffraction pattern of TiO2 nanoparticles.
31
(e). High resolution TEM image showing TiO2 crystal lattice.
32
33
34
35
36
37 The fabrication process of the multilayered TiO2 electrode is depicted in Scheme 1. The effective
38
39 footprint area was defined to be 1.51.5 cm2. To do this, a bilayer of Ti/Au (5 nm/50 nm)
40
41
current collector was e-beam evaporated onto the silicon dioxide wafer through a shadow mask
42
43
44 (Figure S1). The current collector is connected to the contact pad by a thin trace of gold, the
45
46 dimensions of which are negligible relative to the current collection area. The imprinting
47
48 followed a typical spin coating-molding-demolding process. The patterned TiO2 structure
49
50
51 solidifies as excessive solvent permeates through poly (dimethyl siloxane) (PDMS) stamp upon
52
53 heating. The structure is robust even before calcination due to the strong cohesive forces between
54
55 TiO2 particles. This allowed quick demolding in less than 5 min. In order to make multilayered
56
57
58 7
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 8 of 22

1
2
3 structures, we introduced the planarization step. A commercial thiol-ene based acrylate
4
5
6 photoresist (NOA60, Norland, Inc) is selected as the planarization material. During spin coating,
7
8 it fills the trenches between TiO2 line patterns, forming a solid, planar surface upon UV exposure,
9
10 which then acts as the new substrate. By repeating steps of imprinting and planarization,
11
12
13
additional layers can be readily stacked onto previous ones in an orthogonal orientation.
14
15 Conceivably, as the structure volume increases with the number of stacked layers, a nearly
16
17 constant surface-to-volume ratio is maintained. The woodpile architecture was obtained after a
18
19
short time calcination in air at 750 ℃ for 5 min, during which the organic planarization layer was
20
21
22 completely removed, as verified by thermal gravimetric analysis (TGA) (Figure S2). The neck
23
24 formation and ripening of particles enhance the mechanical strength of the architecture and
25
26 create nanoscale voids within TiO2 lines, which favors ion transfer.31 The porosity is estimated to
27
28
29 be 31.3% based on the difference of refractive index from bulk TiO2 (Figure S3, Supplementary
30
31 Information).
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53 Scheme 1. Schematic illustration of the TiO2 woodpile fabrication process. (i) Single layer imprinting.
54 TiO2 ink spin coated on gold charge collector followed by PDMS stamp molding-drying-demolding
55
56
57
58 8
59
60 ACS Paragon Plus Environment
Page 9 of 22 ACS Applied Materials & Interfaces

1
2
3 process. (ii) Planarization with cross-linkable thiol-ene based acrylate resin. (iii) Multilayer structure
4
5 fabrication by repeating (i) and (ii). (iv) Woodpile electrode obtained after calcination.
6
7
8
9
10 Scanning electron microscopy (SEM) and atomic force microscopy (AFM) of the attained single
11
12
layer imprint show that well-defined line pattern of approximately 200 nm in width and height
13
14
15 (Figure 2a, 2b) is obtained, generating an aspect ratio of 1. Barely any residual layer was
16
17 observed. The deviation from stamp dimension results from the shrinkage caused by solvent
18
19 evaporation and calcination. The dimensions remained consistent in the multilayered woodpile
20
21
22 structure (Figure 2c to 2g). Consequently, stacking up to 2, 3, 4 and 6 layers resulted in effective
23
24 aspect ratios of 2, 3, 4 and 6 respectively. Notably, the line shapes were well maintained even
25
26 though they spanned wide gaps over the underlying layers. A larger area view of the defect-free
27
28
29
imprint (Figure S4) demonstrates the robustness of the imprinting. X-ray diffraction (XRD)
30
31 indicates that after calcination, the sub-10 nm TiO2 particles merged into bigger crystallites
32
33 confirmed by the sharper (101) peak (Figure 2h). The crystallite size is calculated to increase
34
35
from 3.8 nm to 23.6 nm. In addition, we noticed that small amount of TiO2 converted from
36
37
38 anatase to rutile phase. Note that both phases are electroactive, and possess similar theoretical
39
40 specific capacities (~335 mAhg-1) when nanoscale materials are used,32,33 no extra effort on
41
42 crystal phase refinement was offered. To demonstrate the readiness of scaled fabrication, an
43
44
45 inch-scale imprint was made (Figure 2i). The optical color results from the light interaction with
46
47 the periodic patterns on the substrate.
48
49
50
51
52
53
54
55
56
57
58 9
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 10 of 22

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35 Figure 2. Characterizations of the calcined imprinted structures. (a, b) SEM top view and AFM height
36
37 profile of the single layer imprint. (c, d) SEM top views of 2-layer and 3-layer woodpile structure. (e)
38
39 Zoom-out view of the 3-layer woodpile structure. (f) SEM top view of 4-layer woodpile structure (inset:
40
41
cross sectional view). (g) SEM cross section of the 6-layer woodpile structure. (h) XRD measurement of
42
43
44 TiO2 before and after calcination. (i) Photograph of a larger area, 1.5’’ 2.5’’ imprint.
45
46
47
48
49
To evaluate the electrochemical performance of the imprinted electrodes, a Li-TiO2 half-cell was
50
51
52 assembled by immersing the TiO2 woodpile into liquid electrolyte and using lithium foil as
53
54 counter electrode (Figure 3a). Figure 3b shows the cyclic voltammogram (CV) of a 6-layer TiO2
55
56
57
58 10
59
60 ACS Paragon Plus Environment
Page 11 of 22 ACS Applied Materials & Interfaces

1
2
3 woodpile electrode at different scan rates. At 0.1 mVs-1, characteristic cathodic peak at 1.74V
4
5
6 and anodic peak at 2.04V were observed, corresponding to Li+ insertion and extraction (xLi
+
7
8 TiO ↔ Li TiO ). By correlating peak current (i) to scan rates (v) (Figure 3c), we find that
9
10 i ∝ v . , indicating diffusion controlled kinetics. 34
As scan rate increased to 10 mVs-1, cathodic
11
12
13 and anodic peak shifted slightly to1.65 V and 2.1 V. This scan rate-dependence is indicative of
14
15 the deviation from Nernstian system in the fast scan region. Note that the woodpile architecture
16
17 provides sufficient pathways for ion transport. We believe the kinetic limitation arises from
18
19
20 either electron transport across electrode or from the relatively slow ion diffusion within TiO2
21
22 particles.
23
24
25 To demonstrate sequential imprinting as an efficient way to enhance areal capacity, a series of
26
27 un-patterned TiO2 film electrodes with comparable loading and porosity were fabricated as
28
29
control samples. For example, a TiO2 film made by 6 spin coating-calcination cycles is used to
30
31
32 compare with the 6-layer woodpile. As shown in Figure 3d, areal capacity of woodpile electrodes
33
34 increases linearly with the stacking of layers. The imprinted electrodes exhibit areal capacities of
35
36 3.6, 7.1, 11.4, 15.5 and 21.3 μAhcm-2 for 1-, 2-, 3-, 4- and 6-layer electrode respectively under
37
38
39 discharge current density of 500 mAg-1. This confirms that the woodpile stacking is effective
40
41 towards multiplied capacity. Conceivably, when normalized to the mass loading of TiO2,
42
43 woodpile electrodes with one to six layers possess identical specific capacity of approximately
44
45
46 250.9 mAhg-1, corresponding to 0.74 Li uptake per formula unit. This value is larger than the
47
48 bulk anatase TiO2 lattice, which at most accommodates 0.5 Li per formula unit, but it is not
49
50 unexpected. Prior literature has revealed that decreasing particle size to the nanoscale induces
51
52
53
size-dependent expansion Li uptake capacity and that the maximum Li uptake was reported to
54
55 approach 1.0 per formula unit for particle sizes approaching 11 nm.32 Differential capacity curves
56
57
58 11
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 12 of 22

1
2
3 of charge and discharge (Figure 3e) of woodpile electrode shows sharp and single peaks at 1.74
4
5
6 V and 2.10 V. This corresponds well with the CV measurement, indicating the redox reaction
7
8 occurs predominantly at these two potentials. In contrast, for planar TiO2 electrodes, increasing
9
10 film thickness failed to achieve proportional enhancement of areal capacity (Figure 3f). A 6-layer
11
12
13
film generates less than quadruple amount of capacity of a single layer. This indicates a
14
15 significant waste of TiO2 due to inefficient material packing. The difference can be explained in
16
17 terms of the reaction- diffusion kinetics,3, 35
the impact of which may not be evident when
18
19
structure dimensions are small, as shown by the overlap of discharging profiles of patterned and
20
21
22 un-patterned single layers. Nevertheless, when thickness increases, the slow diffusion rate of Li-
23
24 ion across the electrode becomes significant. As a result, ion insertion into the inner regions of
25
26 the TiO2 film is impeded and only the top layer is effective. On the contrary, woodpile
27
28
29 architecture circumvents this limitation by maintaining a constant surface-to-volume ratio and
30
31 providing an infiltrated ionic pathway into the structure. This finding demonstrates the potential
32
33 of using TiO2 woodpile as a high-performance on-chip nanoelectrode.
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 12
59
60 ACS Paragon Plus Environment
Page 13 of 22 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
Figure 3. Electrochemical performance of TiO2 woodpiles and control samples. (a) Electrochemical test
27
28 set-up and illustrative view of electrolyte permeation into woodpile structure. (b) CVs of a 6-layer TiO2
29
30 woodpile electrode under scan rates of 0.1, 0.5, 1.0, 2.0, 5.0 and 10 mVs-1. (c) Plot of scan rate
31 dependence of anodic peak current. (d) Galvanostatic discharge profiles of imprinted architectures with 1,
32
33 2, 3, 4 and 6 layers under current density 500 mAg-1. (e) Differential charge/discharge capacity curves of
34
TiO2 woodpile electrode. (f) Comparison of capacity enhancement in TiO2 woodpiles and in planar films.
35
36
37
38
39
40 The imprinted woodpile architecture becomes even more advantageous when performance under
41
42 higher discharge rates is investigated. We define discharge current density of 1 C as 335 mAg-1.
43
44 As shown in Figure 4a, at 3 C (~ 1000 mAg-1), the planar TiO2 electrode lost half of its capacity
45
46
47 while the woodpile electrode demonstrated a capacity retention of 75%. From 1000 mAg-1 to
48
49 higher discharge rates, the woodpile electrode consistently exhibited doubled to tripled capacity
50
51 retention relative to the planar electrode. Even at rate as high as 15 C (5000 mAg-1), the
52
53
54
woodpile electrode is able to deliver 25% of its initial capacity whereas the flat sample suffers
55
56 almost complete capacity degradation. Notice that the absolute values of their initial capacities at
57
58 13
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 14 of 22

1
2
3 500 mAg-1 are offset by a factor of 1.7 (Figure 3f), the woodpile exhibits 3-5 folds improvement
4
5
6 of the actual areal capacity. Figure 4b compares the specific capacity and the discharge current
7
8 density with TiO2 electrodes reported from recent literature.32,33, 36-43 The selected works mainly
9
10 focus on nanostructured TiO2 (e.g. nanotube, nanocage, hollow shell, etc) or composites of TiO2
11
12
13
and conductive additives including carbon nanotubes (CNTs) and graphene. It is worth noting
14
15 that although we used commercially available particles and added no carbonaceous additives,
16
17 these relatively simple and scalable woodpile structures achieved comparable-to-better
18
19
electrochemical performance relative to the best-in-class reports. This comparison highlights the
20
21
22 impact of nanoscale patterning on electrode’s performance.
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47 Figure 4. Rate capability and electrochemical performance comparison. (a) Capacity retention of 6-layer
48 woodpile electrode and control film sample under discharge rates from 1.5 C (500 mAg-1) to 15 C (5000
49
50 mAg-1). (b) Comparison of specific capacity and cycling current density of the 6-layer woodpile electrode
51 with recently reported TiO2 electrodes. Voltage windows of all tests are between 1.0-3.0 V versus Li+/Li.
52
53
54
55
56
57
58 14
59
60 ACS Paragon Plus Environment
Page 15 of 22 ACS Applied Materials & Interfaces

1
2
3 Remarkably, the woodpile electrode demonstrated stable performance even during harsh testing
4
5
6 conditions, such as the 15 C discharging. To further investigate the stability of the imprinted
7
8 electrode, we extended the lower cut-off voltage to 0.4 V. As shown in Figure 5a, the woodpile
9
10 electrodes still followed linear relationship as stacking of layers increased and demonstrated
11
12
13
expanded capacity values. Areal capacity of 6-layer electrode increased from 21.3 μAhcm-2 to
14
15 35.1 μAhcm-2, resulting in specific capacity as high as 413.4 mAhg-1, which is even larger than
16
17 the theoretical value of 335 mAhg-1. The exact reason is not conclusive yet, but we propose that
18
19
20
surface charge storage mechanisms may be involved. The randomly oriented small TiO2
21
22 crystallites provide nanovoids to adsorb additional Li+, which is similar to the “house of cards”
23
24 model used to explain larger Li+ storage capacity in amorphous carbon (relative to graphite).44-46
25
26
In fact, Ferris et al recently reported that an anomalous extension of voltage window was also
27
28
29 observed in nanosupercapacitors.47 The atypical electrochemical properties in nanodevices need
30
31 more fundamental investigation. In this extended voltage window, TiO2 is likely to suffer over-
32
33 discharge and may go through irreversible structural changes. Surprisingly, the imprinted TiO2
34
35
36 woodpile electrode could be stably cycled under current densities from 500-5000 mAg-1 (Figure
37
38 5b). This may be attributed to the thermal treatment and the nanoscale line pattern for facile
39
40 strain release.
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 15
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 16 of 22

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22 Figure 5. Electrochemical tests of woodpile electrodes in extended voltage window from 0.4 V-3.0 V
23
24 versus Li+/Li. (a) Expanded areal capacities of 1-layer to 6-layer structures. (b) Capacity retention of 6-
25
layer woodpile electrode under discharge rates from 500-5000 mAg-1 in the extended voltage window.
26
27
28
29
30
31 It can be clearly seen in Figure 6a that the woodpile remained intact after the test without evident
32
33 cracks or delamination. It is noted however, that compared to the untested woodpiles, the surface
34
35 structure of the tested samples became coarser and line width changed from 200 nm to 380 nm
36
37
38 (Figure 6b). This results from the volume change of TiO2 and forming of the solid electrolyte
39
40 interphase (SEI). It has been reported that SEI tends to form on TiO2 surface at lower potential.48
41
42 To determine the surface elemental composition, electrodes after testing were rinsed in
43
44
45 propylene carbonate (PC) to clean the residual liquid electrolyte and dried in an oven. Energy-
46
47 dispersive X-ray spectroscopy (EDX) shows that besides Ti and O, Cl and C also existed (Figure
48
49 6c). A more detailed X-ray photoelectron spectroscopy (XPS) elemental scan reveals that the
50
51
atomic ratio of O: Li: Cl: C is 52.4%: 27.3%:17.0%: 3.1% (Figure 6d to 6g). As the penetration
52
53
54 depth (less than 20 nm) of XPS is much smaller than EDX, barely any signal from Ti was
55
56 collected, which is indicative of complete coverage of TiO2 under the surface SEI layer. In fact,
57
58 16
59
60 ACS Paragon Plus Environment
Page 17 of 22 ACS Applied Materials & Interfaces

1
2
3 capacity fade can result from continuous cracking and reformation of the SEI. The stable
4
5
6 performance of the woodpile electrodes confirms that the SEI was intact during the tests.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29 Figure 6. SEM imaging and elemental analysis of the woodpile electrodes after electrochemical tests. (a)
30
31 Cross-sectional view of a 3-layer woodpile after cycling. (b) Zoom-in view of a 3-layer woodpile showing
32
33 the structure’s dimension and surface morphology changes. (c) EDX measurement of the woodpile. (d-g)
34
35 XPS elemental scan of Li 1s, Cl 2p, O 1s and C 1s.
36
37
38
39
40
41 4. Conclusions
42
43
44 In summary, we fabricated the 3D TiO2 woodpile electrodes comprised designated numbers of
45
46
stacked layers for lithium-ion nanobatteries using solvent-assisted NIL. This technique allows
47
48
49 scalable fabrication of delicate electrode architectures with nanoscale resolution and yields
50
51 specific capacities as high as 250.9 mAhg-1 and stable performance even over an extended
52
53 voltage window. In addition, the sequential imprinting enabled the areal capacity to be readily
54
55
56 multiplied while maintaining superior rate capability. In this work, we demonstrated up to 6
57
58 17
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 18 of 22

1
2
3 layers of stackings. Conceivably, by using the method described, even thicker electrodes can be
4
5
6 fabricated. The upper limit is determined by the mechanical strength and conductivity of the
7
8 electroactive material. As a solution-based, additive manufacturing method, it is possible to
9
10 avoid such limitations by formulating composite inks to improve rheological, mechanical and
11
12
13
electrical properties. The method here is general and may inspire investigations into patterning a
14
15 broad array of electroactive materials.
16
17
18
19
20
21 Associated Content
22
23
24 Supporting Information. Shadow mask design for depositing current collector, thermal
25
26 gravimetric analysis (TGA) of NOA60 and TiO2 nanoparticles in air, ellipsometry of TiO2 films
27
28 after thermal treatment, SEM image showing the larger area, defect-free imprinted 4-layer
29
30
31 woodpile structure.
32
33
34 This material is available free of charge via the internet at http://pubs.acs.org.
35
36
37 Author Information
38
39
40
Corresponding Author
41
42 
e-mail: watkins@polysci.umass.edu
43
44
45
Author Contributions
46
47
48 
W. Li and Y. Zhou contributed equally to this work.
49
50
51 Notes
52
53
54 The authors declare no competing financial interest.
55
56
57
58 18
59
60 ACS Paragon Plus Environment
Page 19 of 22 ACS Applied Materials & Interfaces

1
2
3
4
5
6 Acknowledgement
7
8
9 W. L., Y. Z., I. R. H., Y. G., A. R. N., S. L., K. R. C., and J. J. W. were supported by the NSF
10
11
12
Center of Hierarchical Manufacturing at the University of Massachusetts, Amherst (CMMI-
13
14 1025020). The Materials Research Science and Engineering Center (MRSEC) at University of
15
16 Massachusetts Amherst supported facilities used in this study. W. L. acknowledges J. Harris, K.
17
18
Martin, A. Chang, L. Bartlett and A. Ribbe for their help with experimental procedures.
19
20
21
22
23
24
References
25
26
27 (1) Spearing, S. M., Materials issues in microelectromechanical systems (MEMS). Acta Mater. 2000, 48
28 (1), 179-196.
29 (2) Craighead, H. G., Nanoelectromechanical systems. Science 2000, 290 (5496), 1532-1535.
30
(3) Long, J. W.; Dunn, B.; Rolison, D. R.; White, H. S., Three-dimensional battery architectures. Chem.
31
Rev. 2004, 104 (10), 4463-4492.
32
33
(4) Rolison, D. R.; Long, J. W.; Lytle, J. C.; Fischer, A. E.; Rhodes, C. P.; McEvoy, T. M.; Bourga, M.
34 E.; Lubers, A. M., Multifunctional 3D nanoarchitectures for energy storage and conversion. Chem. Soc.
35 Rev. 2009, 38 (1), 226-252.
36 (5) Zhang, H. G.; Yu, X. D.; Braun, P. V., Three-dimensional bicontinuous ultrafast-charge and -
37 discharge bulk battery electrodes. Nat. Nanotechnol. 2011, 6 (5), 277-281.
38 (6) Ergang, N. S.; Lytle, J. C.; Lee, K. T.; Oh, S. M.; Smyrl, W. H.; Stein, A., Photonic crystal structures
39 as a basis for a three-dimensionally interpenetrating electrochemical-cell system. Adv. Mater. 2006, 18
40 (13), 1750-1753.
41 (7) McNulty, D.; Carroll, E.; O'Dwyer, C., Rutile TiO2 Inverse Opal Anodes for Li-Ion Batteries with
42 Long Cycle Life, High-Rate Capability, and High Structural Stability. Adv. Energy Mater. 2017, 7 ,
43 1602291.
44 (8) Wang, X. R.; Yushin, G., Chemical vapor deposition and atomic layer deposition for advanced
45 lithium ion batteries and supercapacitors. Energy Environ. Sci. 2015, 8 (7), 1889-1904.
46 (9) Notten, P. H. L.; Roozeboom, F.; Niessen, R. A. H.; Baggetto, L., 3-D integrated all-solid-state
47 rechargeable batteries. Adv. Mater. 2007, 19 (24), 4564-4567.
48 (10) Talin, A. A.; Ruzmetov, D.; Kolmakov, A.; McKelvey, K.; Ware, N.; El Gabaly, F.; Dunn, B.;
49
White, H. S., Fabrication, Testing, and Simulation of All-Solid-State Three-Dimensional Li-Ion Batteries.
50
ACS Appl. Mater. Interfaces 2016, 8 (47), 32385-32391.
51
52
(11) Cheah, S. K.; Perre, E.; Rooth, M.; Fondell, M.; Harsta, A.; Nyholm, L.; Boman, M.; Gustafsson, T.;
53 Lu, J.; Simon, P.; Edstrom, K., Self-Supported Three-Dimensional Nanoelectrodes for Microbattery
54 Applications. Nano Lett. 2009, 9 (9), 3230-3233.
55 (12) Sun, K.; Wei, T. S.; Ahn, B. Y.; Seo, J. Y.; Dillon, S. J.; Lewis, J. A., 3D Printing of Interdigitated
56 Li-Ion Microbattery Architectures. Adv. Mater. 2013, 25 (33), 4539-4543.
57
58 19
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 20 of 22

1
2
3 (13) Dokko, K.; Sugaya, J.; Nakano, H.; Yasukawa, T.; Matsue, T.; Kanamura, K., Sol-gel fabrication of
4 lithium-ion microarray battery. Electrochem. Commun. 2007, 9 (5), 857-862.
5 (14) Xia, Y. N.; Whitesides, G. M., Soft lithography. Angew. Chem.-Int. Edit. 1998, 37 (5), 550-575.
6
(15) Tormen, M.; Sovernigo, E.; Pozzato, A.; Pianigiani, M.; Tormen, M., Sub-100 mu s nanoimprint
7
lithography at wafer scale. Microelectron. Eng. 2015, 141, 21-26.
8
(16) Ahn, S. H.; Guo, L. J., Large-Area Roll-to-Roll and Roll-to-Plate Nanoimprint Lithography: A Step
9
10 toward High-Throughput Application of Continuous Nanoimprinting. ACS Nano 2009, 3 (8), 2304-2310.
11 (17) Love, J. C.; Estroff, L. A.; Kriebel, J. K.; Nuzzo, R. G.; Whitesides, G. M., Self-assembled
12 monolayers of thiolates on metals as a form of nanotechnology. Chem. Rev. 2005, 105 (4), 1103-1169.
13 (18) Leitgeb, M.; Nees, D.; Ruttloff, S.; Palfinger, U.; Gotz, J.; Liska, R.; Belegratis, M. R.; Stadlober,
14 B., Multilength Scale Patterning of Functional Layers by Roll-to-Roll Ultraviolet-Light Assisted
15 Nanoimprint Lithography. ACS Nano 2016, 10 (5), 4926-4941.
16 (19) Park, J. Y.; Hendricks, N. R.; Carter, K. R., Solvent-Assisted Soft Nanoimprint Lithography for
17 Structured Bilayer Heterojunction Organic Solar Cells. Langmuir 2011, 27 (17), 11251-11258.
18 (20) Li, M. T.; Tan, H.; Chen, L.; Wang, J.; Chou, S. Y., Large area direct nanoimprinting of SiO2-TiO2
19 gel gratings for optical applications. J. Vac. Sci. Technol. B 2003, 21 (2), 660-663.
20 (21) Ahn, J. H.; Kim, H. S.; Lee, K. J.; Jeon, S.; Kang, S. J.; Sun, Y. G.; Nuzzo, R. G.; Rogers, J. A.,
21 Heterogeneous three-dimensional electronics by use of printed semiconductor nanomaterials. Science
22 2006, 314 (5806), 1754-1757.
23 (22) Ko, S. H.; Park, I.; Pan, H.; Grigoropoulos, C. P.; Pisano, A. P.; Luscombe, C. K.; Frechet, J. M. J.,
24 Direct nanoimprinting of metal nanoparticles for nanoscale electronics fabrication. Nano Lett. 2007, 7 (7),
25 1869-1877.
26
(23) Paik, T.; Yun, H.; Fleury, B.; Hong, S. H.; Jo, P. S.; Wu, Y. T.; Oh, S. J.; Cargnello, M.; Yang, H.;
27
Murray, C. B.; Kagan, C. R., Hierarchical Materials Design by Pattern Transfer Printing of Self-
28
29
Assembled Binary Nanocrystal Superlattices. Nano Lett. 2017, 17 (3), 1387-1394.
30 (24) Lee, S. H.; Ha, N. Y., Nanostructured indium-tin-oxide films fabricated by all-solution processing
31 for functional transparent electrodes. Opt. Express 2011, 19 (22), 21803-21808.
32 (25) Yu, X.; Pham, J. T.; Subramani, C.; Creran, B.; Yeh, Y. C.; Du, K.; Patra, D.; Miranda, O. R.;
33 Crosby, A. J.; Rotello, V. M., Direct Patterning of Engineered Ionic Gold Nanoparticles via Nanoimprint
34 Lithography. Adv. Mater. 2012, 24 (47), 6330-6334.
35 (26) Huang, Y. G. Y.; Zhou, W. X.; Hsia, K. J.; Menard, E.; Park, J. U.; Rogers, J. A.; Alleyne, A. G.,
36 Stamp collapse in soft lithography. Langmuir 2005, 21 (17), 8058-8068.
37 (27) Kim, E.; Xia, Y. N.; Whitesides, G. M., Micromolding in capillaries: Applications in materials
38 science. J. Am. Chem. Soc. 1996, 118 (24), 5722-5731.
39 (28) Suh, K. Y.; Kim, Y. S.; Lee, H. H., Capillary force lithography. Adv. Mater. 2001, 13 (18), 1386-
40 1389.
41 (29) Zhang, Y.; Lo, C. W.; Taylor, J. A.; Yang, S., Replica molding of high-aspect-ratio polymeric
42 nanopillar arrays with high fidelity. Langmuir 2006, 22 (20), 8595-8601.
43 (30) Kothari, R.; Beaulieu, M. R.; Hendricks, N. R.; Li, S. K.; Watkins, J. J., Direct Patterning of Robust
44 One-Dimensional, Two-Dimensional, and Three-Dimensional Crystalline Metal Oxide Nanostructures
45
Using Imprint Lithography and Nanoparticle Dispersion Inks. Chem. Mater. 2017, 29 (9), 3908-3918.
46
(31) Vu, A.; Qian, Y. Q.; Stein, A., Porous Electrode Materials for Lithium-Ion Batteries - How to
47
Prepare Them and What Makes Them Special. Adv. Energy Mater. 2012, 2 (9), 1056-1085.
48
49 (32) Gentili, V.; Brutti, S.; Hardwick, L. J.; Armstrong, A. R.; Panero, S.; Bruce, P. G., Lithium
50 Insertion into Anatase Nanotubes. Chem. Mater. 2012, 24 (22), 4468-4476.
51 (33) Rai, A. K.; Anh, L. T.; Gim, J.; Mathew, V.; Kang, J.; Paul, B. J.; Song, J.; Kim, J., Simple
52 synthesis and particle size effects of TiO2 nanoparticle anodes for rechargeable lithium ion batteries.
53 Electrochimi. Acta 2013, 90, 112-118.
54 (34) Lindstrom, H.; Sodergren, S.; Solbrand, A.; Rensmo, H.; Hjelm, J.; Hagfeldt, A.; Lindquist, S. E.,
55 Li+ ion insertion in TiO2 (anatase) .2. Voltammetry on nanoporous films. J. Phys. Chem. B 1997, 101
56 (39), 7717-7722.
57
58 20
59
60 ACS Paragon Plus Environment
Page 21 of 22 ACS Applied Materials & Interfaces

1
2
3 (35) Roberts, M.; Johns, P.; Owen, J.; Brandell, D.; Edstrom, K.; El Enany, G.; Guery, C.; Golodnitsky,
4 D.; Lacey, M.; Lecoeur, C.; Mazor, H.; Peled, E.; Perre, E.; Shaijumon, M. M.; Simon, P.; Taberna, P. L.,
5 3D lithium ion batteries-from fundamentals to fabrication. J. Mater. Chem. 2011, 21 (27), 9876-9890.
6
(36) Yu, X. Y.; Wu, H. B.; Yu, L.; Ma, F. X.; Lou, X. W., Rutile TiO2 Submicroboxes with Superior
7
Lithium Storage Properties. Angew. Chem.-Int. Edit. 2015, 54 (13), 4001-4004.
8
(37) Liu, J. H.; Chen, J. S.; Wei, X. F.; Lou, X. W.; Liu, X. W., Sandwich-Like, Stacked Ultrathin
9
10 Titanate Nanosheets for Ultrafast Lithium Storage. Adv. Mater. 2011, 23 (8), 998-1002.
11 (38) Zhang, G. Q.; Wu, H. B.; Song, T.; Paik, U.; Lou, X. W., TiO2 Hollow Spheres Composed of
12 Highly Crystalline Nanocrystals Exhibit Superior Lithium Storage Properties. Angew. Chem.-Int. Edit.
13 2014, 53 (46), 12590-12593.
14 (39) Liu, H. S.; Bi, Z. H.; Sun, X. G.; Unocic, R. R.; Paranthaman, M. P.; Dai, S.; Brown, G. M.,
15 Mesoporous TiO2-B Microspheres with Superior Rate Performance for Lithium Ion Batteries. Adv. Mater.
16 2011, 23 (30), 3450-3454.
17 (40) Xu, W.; Wang, Z. Y.; Guo, Z. D.; Liu, Y.; Zhou, N. Z.; Niu, B.; Shi, Z. J.; Zhang, H., Nanoporous
18 anatase TiO2/single-wall carbon nanohorns composite as superior anode for lithium ion batteries. J.
19 Power Sources 2013, 232, 193-198.
20 (41) Park, S. J.; Kim, Y. J.; Lee, H., Synthesis of carbon-coated TiO2 nanotubes for high-power lithium-
21 ion batteries. J. Power Sources 2011, 196 (11), 5133-5137.
22 (42) Chen, J. S.; Lou, X. W., Unusual rutile TiO2 nanosheets with exposed (001) facets. Chem. Sci.
23 2011, 2 (11), 2219-2223.
24 (43) Wang, J.; Zhou, Y. K.; Xiong, B.; Zhao, Y. Y.; Huang, X. J.; Shao, Z. P., Fast lithium-ion insertion
25 of TiO2 nanotube and graphene composites. Electrochim. Acta 2013, 88, 847-857.
26
(44) Zheng, T.; Xing, W.; Dahn, J. R., Carbons prepared from coals for anodes of lithium-ion cells.
27
Carbon 1996, 34 (12), 1501-1507.
28
29
(45) Kaskhedikar, N. A.; Maier, J., Lithium Storage ion Carbon Nanostructures. Adv. Mater. 2009, 21
30 (25-26), 2664-2680.
31 (46) Pan, D. Y.; Wang, S.; Zhao, B.; Wu, M. H.; Zhang, H. J.; Wang, Y.; Jiao, Z., Li Storage Properties
32 of Disordered Graphene Nanosheets. Chem. Mater. 2009, 21 (14), 3136-3142.
33 (47) Ferris, A.; Reig, B.; Eddarir, A.; Pierson, J.; Garbarino, S.; Guay, D.; Pech, D., Atypical Properties
34 of FIB-Patterned RuOx Nanosupercapacitors. ACS Energy Lett. 2017, 2, 1734-1739.
35 (48) Petkovich, N. D.; Wilson, B. E.; Rudisill, S. G.; Stein, A., Titania-Carbon Nanocomposite Anodes
36 for Lithium Ion Batteries-Effects of Confined Growth and Phase Synergism. ACS Appl. Mater. Interfaces
37 2014, 6 (20), 18215-18227.
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 21
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 22 of 22

1
2
3 Table of Content
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 22
59
60 ACS Paragon Plus Environment

You might also like