Comparative Study of Ni-Rich Layered Cathodes for Rechargeable

Jun 20, 2016 - Compositionally graded Li[Ni0.84Co0.06Mn0.09Al0.01]O2 (TSFCG-Al) with two-step concentration gradients was prepared as a high-capacity ...
0 downloads 7 Views 3MB Size
Subscriber access provided by The Chinese University of Hong Kong

Letter

Comparative Study of Ni-rich Layered Cathodes for Rechargeable Lithium Batteries: Li[Ni Co Al ]O and Li[Ni Co Mn Al ]O with Two-Step Full Concentration Gradients 0.85

0.84

0.06

0.09

0.01

0.11

0.04

2

2

Byung-Beom Lim, Seung-Taek Myung, Chong S. Yoon, and Yang-Kook Sun ACS Energy Lett., Just Accepted Manuscript • DOI: 10.1021/acsenergylett.6b00150 • Publication Date (Web): 20 Jun 2016 Downloaded from http://pubs.acs.org on June 20, 2016

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 Energy Letters 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 23

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

Comparative Study of Ni-rich Layered Cathodes for Rechargeable Lithium Batteries: Li[Ni0.85Co0.11Al0.04]O2 and Li[Ni0.84Co0.06Mn0.09Al0.01]O2 with Two-Step Full Concentration Gradients Byung-Beom Lim†⊥, Seung-Taek Myung‡⊥, Chong S. Yoon*,∥, Yang-Kook Sun*,† †

Department of Energy Engineering and ∥Department of Materials Science and Engineering,

Hanyang University, Seoul, 04763, South Korea ‡

Department of Nanotechnology and Advanced Materials Engineering, Sejong University, Seoul

05006, South Korea

1 Environment ACS Paragon Plus

ACS Energy Letters

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract Compositionally-graded Li[Ni0.84Co0.06Mn0.09Al0.01]O2 (TSFCG-Al) with two-step concentration gradients was prepared as a high-capacity cathode for Li batteries. The concentration gradients were introduced within a single particle to maximize the Ni fraction; this increased the discharge capacity, while ensuring cyclic stability with a Mn-enriched surface layer. The concentration gradients also produced a unique morphology, in which rod-shaped primary particles were radially aligned in a spoke-like pattern. The fundamental electrochemical performance of TSFCG-Al is compared against commercial Li[Ni0.85Co0.11Al0.04]O2 (NCA). The TSFCG-Al cathode exhibits a higher discharge capacity and better cycling stability than the NCA cathode, even when they are charged to 4.5 V. Structural analysis of the cycled TSFCG-Al and NCA cathodes shows that the TSFCG-Al keeps its original structural integrity, while the NCA particles undergo serious particle degradation due to the accumulation of strain in the grain boundaries upon cycling. TOC Graphic

2 Environment ACS Paragon Plus

Page 2 of 23

Page 3 of 23

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

Recent environmental concerns have led to an increased amount of interest in clean renewable energy technologies. Lithium-ion batteries (LIBs) have become crucial for harnessing new energy sources because they enable portability with high-energy density. Currently, LIBs are the dominant energy power sources for various applications ranging from portable electronic devices to electric vehicle applications.1-4 In particular, for plug-in hybrid and electric vehicles (P-HEVs and EVs), high capacity (i.e., high energy density) is the key requirement for the wide deployment of LIBs; hence, much effort has been spent to increase the capacity of cathodes. For instance, Ni-rich layered cathodes are known to deliver a large discharge capacity, reasonable rate capability, and low material cost due to their limited use of relatively expensive Co. Specifically, Ni-rich Li[Ni1-x-yCoxMny]O2 (NCM) and Li[Ni1-x-yCoxAly]O2 (NCA), where Ni ≥ 80 %, are typically able to deliver capacities as high as 200 mAh g-1 at a cutoff voltage of 4.3 V.5-7 However, these cathodes suffer from gradual capacity fading upon cycling due to their structural instability, which is ascribed to the successive phase transitions from hexagonal 1 to hexagonal 2 phases when large amounts of Li+ are extracted from the host structure.8,9 Finding of Abraham et al.10 from a LiNi0.8Co0.2O2 faded 43% of power was formation of NiO-like or LixNi1-xO-like (rock-salt structure) byproduct (Figure b below) on the surface of LiNi0.8Co0.2O2 particle. NiO is known to have poor conductivities of lithium ions and electrons, and the resistance layer formed on the primary particles. This surface change would entail increase in charge transfer resistance during cycling. This was confirmed by neutron and synchrotron XRD studies. Mori et al.

11

correlated the power fade and cationic disordering of NCA electrode. According to their XANES analysis, the valence of Ni ion did not vary with degradation, and the structural change from the layer structure to disordered cubic structure was located near the surface of the NCA materials. Sasaki et al.

12

and Muto et al.13 found that inactive NiO-like phase was detected near the grain

3 Environment ACS Paragon Plus

ACS Energy Letters

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 23

surface and boundaries in NCA by a spatial-distribution map. Consecutive microscopic work by Zheng et al.

14

demonstrated that the region with a disordered rock-salt structure, presumably

NiO-like or LixNi1-xO-like phase, is about 5 nm in width, while the transition region with a partially ordered structure is about 20 nm in width at the first cycle. The important is that a substantial volume of material with the transitional structure formed near the grain boundary. They further confirmed the same result using conducting agent and binder free NCA electrode.15 Conventional NCA and NCM materials have a spherical morphology with average particle size of 10 µm; in particular NCA sometimes shows large primary particles ~ 1µm .6,16,17 As a result, particle cracking or breakup of particles progresses gradually into the particle interior during cycling when 100 % depth of discharge (DOD) was used.14,17 The byproducts produced on the surface of the cathode particle, as well as particle rupture, lead to electrochemical inactivity; thus, the observed capacity fading is unavoidable in both Ni-rich NCM and NCA cathodes. Another concern is oxidation of electrolyte at high voltage. Selection of proper electrolyte by addition of additives significantly enhances electrochemical performances such as less swelling,

metal

dissolution,

and

gas

generation

at

fully

charged

state

for

graphite/Li[Ni0.8Co0.1Mn0.1]O2 full cell.18 Thin and less formation of LiF in the surface film of Li[Ni1/3Co1/3Mn1/3]O2 achieved by addition of trimethylboroxine (TMB) could have excellent cycling performance when cycled in the voltage range of 3 – 4.5 V,19 Hence, it is likely that a significant breakthrough in the high-voltage electrolyte20,21 and ionic liquid22 to widen the operating voltage window of the NCM cathodes. To resolve the aforementioned drawbacks, we recently proposed a new concept for Nirich NCM cathodes by confining stable Mn4+ in the outer surface region. Additionally, electroactive Ni2+ and Ni3+ species in the core region were responsible for activating the cathodes and

4 Environment ACS Paragon Plus

Page 5 of 23

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

delivering a high capacity, reaching approximately 200 mAh g-1 or higher, in the voltage range of 2.7 – 4.3 V.23 We also reported Ni-rich Li[Ni0.8Co0.06Mn0.14]O2 with a two-slope full concentration gradient (TSFCG), which was composed of radially-aligned, nanorod-shaped primary particles. This structure delivers a discharge capacity of 210 mAh g-1 at a rate of 0.1 C with capacity retention of 94 % after 100 cycles.24 Since EVs require a capacity that is considerably higher than 210 mAh g-1, the possibility of increasing the rechargeable capacity by increasing the Ni content in the TSFCG cathode was explored.

5 Environment ACS Paragon Plus

ACS Energy Letters

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 23

Figure 1. SEM images of (a) Li[Ni0.85Co0.11Al0.04]O2 NCA and (b) Li[Ni0.84Co0.06Mn0.09Al0.01]O2 TSFCG-Al; Rietveld refinement results of the XRD patterns of (c) Li[Ni0.85Co0.11Al0.04]O2 NCA and (d) Li[Ni0.84Co0.06Mn0.09Al0.01]O2 TSFCG-Al; (e) EPMA line scan results from the particle center

towards

the

surface

and

a

summary

of

the

Li[Ni0.84Co0.06Mn0.09Al0.01]O2 TSFCG-Al.

6 Environment ACS Paragon Plus

data;

and

(f)

scheme

of

Page 7 of 23

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

In addition to increasing the Ni content, Al was incorporated into the TSFCG cathode to form Li[Ni0.84Co0.06Mn0.09Al0.01]O2 (hereafter referred to as TSFCG-Al) via co-precipitation. This was done because the presence of Al3+ in the transition metal layer tends to lower the cationic disorder.25,26 It is also assumed that the Gibbs energy for the formation of Al2O3 at 25 oC (-1582.3 kJ mol-1 27) is sufficiently low to stabilize the crystal structure (from a thermodynamic viewpoint), which helps stabilize the crystal structure. Scanning

electron

microscopy

(SEM)

images

of

the

as-synthesized

Li[Ni0.85Co0.11Al0.04]O2 (hereafter referred to as NCA) and TSFCG-Al in Figures 1a and b show that both cathodes possess a spherical morphology, in which much smaller primary particles are tightly packed into a spherical shape. Although the differences in the size and morphology of the secondary particles were negligible, the primary particle size appeared to be smaller for the TSFCG-Al. It is notable that voids on the surface are less prominent in the TSFCG-Al particles as compared to the NCA particles. Both NCA and TSFCG-Al were crystallized into an O3-type structure with R-3m space groups (Figures 1c and d). Despite the spatial variation of the composition, the XRD data for TSFCG-Al were refined based on the average composition in order to compare the cation mixing in the Li layer for both compounds (Figures 1c, d, and Table 1). The refinement results indicated that the difference in the occupation of Ni2+ in the Li layer was negligible; however, the lattice parameters were greater for TSFCG-Al as compared to NCA. The average oxidation states of Ni, Co, and Mn in Ni-rich NCM compounds are usually 3+ for each transition metal element in Li[Ni0.8Co0.1Mn0.1]O2.28 Also, Ni2+ (0.69 Å) and Mn4+ (0.53 Å) could be partially produced in the surface region, of which the chemical composition was expressed as Li[Ni0.800Co0.052Mn0.148Al0.010]O2. However, the larger lattice parameters is

7 Environment ACS Paragon Plus

ACS Energy Letters

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 23

likely due to the prevalence of Mn3+ (0.645 Å) and Ni2+ (0.69 Å) though a small amount of Mn4+ could be formed near TSFCG-Al surface, which have a substantially larger ionic radii in TSFCG-Al relative to the ionic radius of Al3+ (0.535 Å) in NCA that does not have Mn in the compound. Measure tap densities were approximately 2.47 g cm-3 for NCA and 2.51 g cm-3 for TSFCG-Al. More physical properties of both NCA and TSFCG-Al are shown in Figures SI 1 and SI 2.

Table 1. Rietveld refinement results of XRD data for as-synthesized Li[Ni0.85Co0.11Al0.04]O2 NCA and Li[Ni0.84Co0.06Mn0.09Al0.01]O2 TSFCG-Al. Ni2+ in Li layer / %

a/Å

c/Å

Rwp / %

NCA

2.8

2.8688(1)

14.1709(2)

13.3

TSFCG-Al

2.6

2.8743(1)

14.2124(2)

13.1

TSFCG-Al was analyzed with EPMA to determine the local compositional change from the particle center to the surface. The concentrations of transition metals are plotted as a function of the distance from the particle center to the surface (Figure 1e). We carefully checked the concentration of transition metal elements using EPMA and the data are highly reproducible. The most important fact is selection of the heat-treatment temperature that does not show interdiffusion of transition metals. That is why low temperature is preferred to minimize such interdiffusion. For the reason, we used lithium hydroxide as the lithium source to impregnate lithium at low temperature, here 750 oC, instead of lithium carbonate, which requires high temperature to yield single phase above 900 oC, to avoid the interdiffusion of transition metal

8 Environment ACS Paragon Plus

Page 9 of 23

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

elements. As designed, the concentration of Ni decreased gradually from 90.0 at. % (at the center) to 88.1 at. % (up to 3.3 µm away from the center), which corresponds to the first gradient slope. Alternatively, the variations of the Co and Mn compositions were negligible. A constant level of Al was also detected throughout the particle. The concentration variation then became steeper for Ni and Mn; this corresponds to the second gradient slope. Hence, TSFCG-Al is regarded

as

a

continuous

solid

solution

of

Li[Ni0.90Co0.05Mn0.04Al0.01]O2



Li[Ni0.79Co0.05Mn0.15Al0.01]O2, as graphically described in Figure 1f, in which the concentrations of Ni and Mn are varied in two steps. The electrochemical properties of the NCA and the TSFCG-Al cathodes were compared using 2032 coin-type half-cells in the voltage ranges of 2.7 to 4.3 V and 2.7 to 4.5 V at 30 oC (Figure 2). The first charge and discharge data indicate that the TSFCG-Al cathode delivered a higher discharge capacity with improved Columbic efficiency: 221 mAh g-1 with 96.4 % efficiency up to 4.3 V and 241 mAh g-1 with 95.0 % efficiency up to 4.5 V (Figure 2a bottom). The NCA cathode delivered a discharge capacity of 214 mAh g-1 with 93.4 % efficiency up to 4.3 V and 224 mAh g-1 with 92.3 % efficiency up to 4.5 V as shown in the top of Figure 2a. Large polarization between charge and discharge is observed for both cells when charged to 4.5 V compared to the cells charged to 4.3 V. The behavior would be ascribed to increase in impedance due to electrolyte decomposition when the cells were charged to 4.5 V at the first cycle.

9 Environment ACS Paragon Plus

ACS Energy Letters

a

4.0

b

4.5

4.0

NCA

3.5

3.5

1st 30th 50th 100th

NCA 4.3 V cut-off NCA 4.5 V cut-off

3.0

4.3 V cut-off

4.5

4.0

4.5 V cut-off TSFCG-Al

3.5

1st 30th 50th 100th

TSFCG-Al 4.3 V cut-off TSFCG-Al 4.5 V cut-off

3.0 0

50

100

150

200

-1

Capacity / mAh g

250

0.9 0.6 0.3 0.0 -0.3

Voltage / V

4.3 V cut-off 4.5 V cut-off

3.0

4.5

4.0

3.5

Voltage / V

Voltage / V

4.5

Voltage / V

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 23

3.0

0.9 0.6 0.3 0.0 -0.3 -0.6 -1 -1

dQ/dV / mAh g V

Figure 2. (a) Initial charge-discharge curves at a current rate of 20 mA g-1 (0.1 C rate) and (b) differential capacity vs. voltage curves at a current rate of 100 mA g-1 (0.5 C rate) for Li[Ni0.85Co0.11Al0.04]O2 NCA and Li[Ni0.84Co0.06Mn0.09Al0.01]O2 TSFCG-Al in the voltage range of 2.7-4.3 V and 2.7-4.5 V.

The electrochemical reactions of both the NCA and TSFCG-Al cathodes are related to the successive phase transitions from hexagonal 1 to hexagonal 2 phases in a similar way 8,9, and no apparent variation was observed in the derivative curves of the TSFCG-Al upon cycling (when

10 Environment ACS Paragon Plus

Page 11 of 23

cycled to 4.3 V and 4.5 V), as shown in the bottom of Figure 2b. Alternatively, the intensities of the peaks became polarized and shifted apart for the NCA cathode upon cycling; this was presumably due to the degradation of the cathode, which was found to be particularly serious

-1

240

a

210 180 150

NCA 4.3 V cut-off NCA 4.5 V cut-off

120 0

20

40

60

80

100

Discharge capacity / mAh g

-1

when cycled at 4.5 V (the top of Figure 2b).

Discharge capacity / mAh g

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

240

b

210 180 150

TSFCG-Al 4.3 V cut-off TSFCG-Al 4.5 V cut-off

120 0

Number of cycle

20

40

60

80

100

Number of cycle

Figure 3. Discharge capacity vs. cycle number of (a) Li[Ni0.85Co0.11Al0.04]O2 NCA and (b) Li[Ni0.84Co0.06Mn0.09Al0.01]O2 TSFCG-Al in the voltage range of 2.7-4.3 V and 2.7-4.5 V at a current of 100 mA g-1 (0.5 C rate).

In consideration of the chemical compositions for both NCA (Li[Ni0.85Co0.11Al0.04]O2) and TSFCG-Al (Li[Ni0.84Co0.06Mn0.09Al0.01]O2), more than 96 % of Ni, Co, and Mn moieties are available for participation in the redox reaction. Accordingly, both cathodes should deliver similar discharge capacities. In particular, for the NCA cathode, the incorporated Al3+ element may significantly assist the stabilization of the crystal structure, which should impart good

11 Environment ACS Paragon Plus

ACS Energy Letters

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 23

cycling performance. However, the cycling performance of the NCA cathode measured at a rate of 0.5 C (100 mA g-1) was very disappointing and showed a capacity retention of only 73.4 % (146 mAh g-1) at 4.3 V and 65.3 % (128 mAh g-1) at 4.5 V after 100 cycles (Figure 3a). Our prior work indicated that Ni-rich Li[Ni0.8Co0.06Mn0.14]O2 TSFCG material could deliver 210 mAh g-1 at a rate of 0.1 C with capacity retention of 94 % after 100 cycles.24 Also, Li[Ni0.85Co0.05Mn0.10]O2 TSFCG maintained 92% of its initial capacity for 100 cycles at the same operation window. 29 For the present TSFCG-Al cathode, the cycling performance was markedly improved (Figure 3b), compared to the NCA (Figure 3a) and our earlier TSFCG materials that contain Ni not less than 80 % in the compounds. Specifically, the cathode was able to deliver a high capacity, even after 100 cycles; this value was 95.1 % (197 mAh g-1) at 4.3 V and 88.9 % (202 mAh g-1) at 4.5 V (Figure 3b). Even the coulombic efficiency at the first cycle reached approximately 96.4 % to 4.3 V cutoff voltage, of which such a high efficiency has ever achieved so far as we know. These data substantialize cycling stability of TSFCG, especially when substituted with Al. The role of the Al-substitution in improvement in capacity retention of the NCM

cathodes

has

been

previously

documented

for

Li[Ni1/3Co1/3Mn1/3]O2. 30

12 Environment ACS Paragon Plus

Li[Ni0.8Co0.1Mn0.1]O226

and

Page 13 of 23

ACS Energy Letters

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

13 Environment ACS Paragon Plus

ACS Energy Letters

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 23

Figure 4. (a) SEM image and (b) TEM image of the Li[Ni0.85Co0.11Al0.04]O2 NCA electrode after 100 cycles with an upper cutoff voltage of 4.5 V. (c) SEM image, (d, e) TEM images, and (f) high-resolution TEM in the 100 zone of the Li[Ni0.84Co0.06Mn0.09Al0.01]O2 TSFCG-Al electrode after 100 cycles with an upper cutoff voltage of 4.5 V; the inset in Figure f shows the Fourierfiltered image of the marked region.

The difference in the cycling performance was prominent when the upper cutoff voltage was raised to 4.5 V. Hence, the cycled electrodes were analyzed by SEM, transmission electron microscopy (TEM), and Rietveld refinement in order to understand the reasons for the capacity fade in both cathodes. As shown in Figures 1a and b, both cathodes in the fresh state showed a spherical morphology composed of primary particles. The NCA cathode cycled to 4.3 V (100 cycles) underwent particle cracking and this can loosen the intimate contact between primary particles was lost (Figure S1b); this is similar to other NCA cathodes that have been discussed in the literature.14,17,31 The degradation was much worse when NCA was cycled up to 4.5 V, demonstrating progress of particle rupture of the spherical morphology (Figures 4a and b and Figure SI 3). The broken parts shown in these figures are not likely to participate in the electrochemical reaction, resulting in an increase in the cell resistance (Figure SI 4, and Table S1), which explains the poor capacity retention of the NCA cathode. The Rietveld refinement results also indicate that, although the crystal structure seemed to be maintained, the resulting cation mixing (i.e., occupation of Ni2+ in the Li layer) increased to 6.8 % (Figure S 5 and Table S2), as compared to just 2.8 % in the fresh state. Alternatively, TSFCG-Al exhibited a smaller cation mixing change from 2.6 % to 4.1 % (Figure 1c and Table 1). In comparison of the c-axis parameters before and after cycling test, the c-axis was elongated from 14.1709 Å to 14.2459 Å,

14 Environment ACS Paragon Plus

Page 15 of 23

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

approximately ∆c = 0.075 Å for the cycled NCA to 4.5 V, whereas the ∆c was calculated to be 0.0122 Å for the cycled TSFCG-Al to 4.5 V. Even in comparison with Al-free TSFCG which showed c-axis expansion of 0.0158 Å when cycled to 4.3 V

29

, the c-axis for the TSFCG-Al

varied less approximately 0.0061 Å at the same operation condition. The markedly reduced caxis variation after the cycling indicates that crystal structure supported by the presence of strong Al-O bond delays the crack propagation at the grain boundary, hence enabling substantial improvement in the cyclability of the TSFCG-Al as shown in Figure 3. Surprisingly, for TSFCG-Al, the original spherical morphology was maintained, even after cycling at 4.3 V and 4.5 V, as can be seen in SEM images of the TSFCG-Al cathode (Figure S1b and 4c). A cross-sectional TEM image of a single TSFCG-Al particle cycled at 4.5 V (Figure 4d) contained only a small number of cracks in the particle interior. The presence of the long rod-shaped primary particles demonstrates that the original radial spoke-like morphology of the TSFCG-Al cathode (Figure 4e) was maintained, which is a unique feature of the compositionally-graded cathodes.32,33 A high-resolution image taken from the 100 zone showed no noticeable damage or localized structural transition; this was accomplished by minimizing micro-crack formation within particles upon cycling, as shown in the Fourier-filtered image in the inset (Figure 4f). The SEM and TEM results clearly demonstrate that the TSFCG-Al cathode experiences minimal structural damage and that the structural integrity and overall particle morphology are preserved upon extensive cycling at 4.5 V. The capacity loss observed at 4.5 V for the TSFCG-Al cathode can be partially ascribed to the breakdown of the electrolyte at the high voltage 34,35; hence, it is surmised that if one used an improved high-voltage electrolyte that was specifically designed for Ni-rich NCM cathodes, the capacity retention for the TSFCGAl cathode at 4.5 V would be greatly enhanced.

15 Environment ACS Paragon Plus

ACS Energy Letters

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 23

It has been reported that capacity fading is related to surface degradation of the cathode particle surface.10-17,31 To verify this, electrochemical impedance spectroscopy (EIS) was used for both cathodes, as shown in Figure S4. The smaller increase in the impedance of TSFCG-Al is understandable; this impedance rise was lower than that of NCA because TSFCG-Al has a lower specific surface area (0.336 m2 g-1) than NCA (0.491 m2 g-1). Once a cathode particle is exposed to the electrolyte, infiltration of the electrolyte proceeds via the capillary effect along the grain boundaries of the primary particles. The reduced contact area with the electrolyte leads to the smaller degree of degradation of the surface of primary particles. Furthermore, balancing the Ni and Mn contents towards the surface also helps improve the performance by reducing the amount of Ni from 90 % to 80 % and by increasing the amount of Mn from 5 % to 16 % (particle surface composition: Li[Ni0.79Co0.05Mn0.15Al0.01]O2). These synergistic effects prevented destructive particle degradation upon cycling, despite the fact that both cathodes contained a similar total average amount of Ni. The thermal stabilities of electrochemically-delithiated, wet TSFCG-Al and NCA were evaluated by differential scanning calorimetry (DSC), as shown in Figure S4. For the TSFCG-Al cathode, the exothermic decomposition was initiated at 200 oC and the amount of generated heat was maximized at 220 oC. Here, the reaction temperature and heat generation were somehow delayed, as compared with those of NCA. The thermal behavior difference appears to be related to the unique rod-shaped primary particle morphology of TSFCG-Al, which has a smaller surface area; less contact with the electrolyte can generate less heat, even in a highly-delithiated state. In summary, we developed a compositionally-graded Li[Ni0.84Co0.06Mn0.09Al0.01]O2 with a two-step concentration gradient of Ni and Mn throughout the particle. This cathode can deliver

16 Environment ACS Paragon Plus

Page 17 of 23

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

high reversible capacities of 221 (4.3 V cutoff) and 241 mAh g-1 (4.5 V cutoff) with outstanding capacity retentions of 95.1 % and 88.9 %, respectively, after 100 cycles. SEM and TEM analyses revealed that no significant structural changes were observed between the pristine and cycled TSFCG-Al cathodes due to the unique rod-shaped primary particle morphology. We believe that the present two-step full concentration gradient cathode represents a method to safely harness the high capacity of Ni-rich cathodes.

ASSOCIATED CONTENT Experimental details, additional SEM images, XRD patterns, electrochemical impedance spectroscopy, and differential scanning calorimetry data for the synthesized NCA Li[Ni0.85Co0.11Al0.04]O2 and TSFCG-Al Li[Ni0.84Co0.06Mn0.09Al0.01]O2. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author Yang-Kook Sun *E-mail: [email protected] Chong S. Yoon *E-mail: [email protected] Present Address †

Department of Energy Engineering, Hanyang University, Seoul, 04763, South Korea

17 Environment ACS Paragon Plus

ACS Energy Letters

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 23

Author Contributions ⊥Byung-Beom

Lim and Seung-Taek Myung contributed equally to this work.

Notes The authors have no competing financial interests to declare. ACKNOWLEDGMENT This work was mainly supported by the Global Frontier R&D Program (2013M3A6B1078875) of the Center for Hybrid Interface Materials (HIM) funded by the Ministry of Science, ICT & Future Planning and a National Research Foundation of Korea (NRF) grant funded by the Korean government (MEST) (No. 2014R1A2A1A13050479).

REFERENCES (1)

Goodenough, J. B.; Kim, Y. Challenges for Rechargeable Li Batteries†. Chem. Mater.

2010, 22, 587-603. (2)

Sun, Y. K.; Myung, S. T.; Park, B. C.; Prakash, J.; Belharouak, I.; Amine, K., High-

energy cathode material for long-life and safe lithium batteries. Nat. Mater. 2009, 8, 320-324. (3)

Liu, C.; Li, F.; Ma, L. P.; Cheng, H. M. Adv. Energy mater. 2010, 22, E28-62.

(4)

Ritchie, A. G. Recent developments and likely advances in lithium rechargeable batteries.

J. Power Sources 2004, 136, 285-289. (5)

Hayashi, T.; Okada, J.; Toda, E.; Kuzuo, R.; Oshimura, N.; Kuwata, N.; Kawamura, J.

Degradation Mechanism of LiNi0.82Co0.15Al0.03O2 Positive Electrodes of a Lithium-Ion Battery by a Long-Term Cycling Test. J. Electrochem. Soc. 2014, 161, A1007-A1011.

18 Environment ACS Paragon Plus

Page 19 of 23

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

(6)

Noh, H.-J.; Youn, S.; Yoon, C. S.; Sun, Y.-K. Comparison of the structural and

electrochemical properties of layered Li[NixCoyMnz]O2 (x = 1/3, 0.5, 0.6, 0.7, 0.8 and 0.85) cathode material for lithium-ion batteries. Journal of Power Sources 2013, 233, 121-130. (7)

Liao, J.-Y.; Manthiram, A. Surface-modified concentration-gradient Ni-rich layered

oxide cathodes for high-energy lithium-ion batteries. J. Power Sources 2015, 282, 429-436. (8)

Yoon, W.-S.; Chung, K. Y.; McBreen, J.; X.-Q. Yang, A comparative study on structural

changes of Li[Co1/3Ni1/3Mn1/3]O2 and Li[Ni0.8Co0.15Al0.05]O2 during first charge using in situ XRD, Electrochem. Commun. 2006, 8, 1257-1262. (9) Sun, H.-H.; Choi, W.; Lee, J. K.; Oh, I.-H.; Jung, H.-G. Control of electrochemical properties of nickel-rich layered cathode materials for lithium ion batteries by variation of the manganese to cobalt ratio, J. Power Sources 2015, 275, 877-883. (10)

Abraham, D. P.; Twesten, R. D.; Balasubramanian, M.; Kropf, J.; Fischer, D.; McBreen,

J.; Petrov, I.; Amine, K. Microscopy and Spectroscopy of Lithium Nickel Oxide-Based Particles Used in High Power Lithium-Ion Cells. J. Electrochem. Soc. 2003, 150, A1450-A1456. (11)

Mori, D.; Kobayashi, H.; Shikano, M.; Nitani, H.; Kageyama, H.; Koike, S.; Sakaebe, H.;

Tatsumi, K. Bulk and surface structure investigation for the positive electrodes of degraded lithium-ion cell after storage test using X-ray absorption near-edge structure measurement. J. Power Sources 2009, 189 (1), 676-680. (12)

Sasaki, T.; Nonaka, T.; Oka, H.; Okuda, C.; Itou, Y.; Kondo, Y.; Takeuchi, Y.; Ukyo, Y.;

Tatsumi, K.; Muto, S., Capacity-Fading Mechanisms of LiNiO2 - Based Lithium-Ion Batteries. Journal of The Electrochemical Society 2009, 156 (4), A289.

19 Environment ACS Paragon Plus

ACS Energy Letters

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(13)

Page 20 of 23

Muto, S.; Sasano, Y.; Tatsumi, K.; Sasaki, T.; Horibuchi, K.; Takeuchi, Y.; Ukyo, Y.,

Capacity-Fading Mechanisms of LiNiO2-Based Lithium-Ion Batteries. J. Electrochem. Soc. 2009, 156 (5), A371. (14)

Zheng, S.; Huang, R.; Makimura, Y.; Ukyo, Y.; Fisher, C. A. J.; Hirayama, T.; Ikuhara,

Y. Microstructural Changes in LiNi0.8Co0.15Al0.05O2 Positive Electrode Material during the First Cycle. J. Electrochem. Soc. 2011, 158 (4), A357. (15)

Makimura, Y.; Zheng, S.; Ikuhara, Y.; Ukyo, Y., Microstructural Observation of

LiNi0.8Co0.15Al0.05O2 after Charge and Discharge by Scanning Transmission Electron Microscopy. J.Electrochem. Soc. 2012, 159 (7), A1070-A1073. (16)

Watanabe, S.; Kinoshita, M.; Hosokawa, T.; Morigaki, K.; Nakura, K. Capacity fading of

LiAlyNi1−x−yCoxO2 cathode for lithium-ion batteries during accelerated calendar and cycle life tests (effect of depth of discharge in charge–discharge cycling on the suppression of the microcrack generation of LiAlyNi1−x−yCoxO2 particle). J. Power Sources 2014, 260, 50-56. (17)

Watanabe, S.; Kinoshita, M.; Hosokawa, T.; Morigaki, K.; Nakura, K. Capacity fade of

LiAlyNi1−x−yCoxO2 cathode for lithium-ion batteries during accelerated calendar and cycle life tests (surface analysis of LiAlyNi1−x−yCoxO2 cathode after cycle tests in restricted depth of discharge ranges). J. Power Sources 2014, 258, 210-217. (18)

Qiu, W.; Xia, J.; Chen, L.; Dahn, J. R. A study of methyl phenyl carbonate and diphenyl

carbonate as electrolyte additives for high voltage LiNiMnCoO/graphite pouch cells. J. Power Sources 2016, 318, 228-234. (19)

Yu, Q.; Chen, Z.; Xing, L.; Chen, D.; Rong, H.; Liu, Q.; Li, W. Enhanced high voltage

performances of layered lithium nickel cobalt manganese oxide cathode by using trimethylboroxine as electrolyte additive. Electrochim. Acta 2015, 176, 919-925.

20 Environment ACS Paragon Plus

Page 21 of 23

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

(20)

Lee, E.-H.; Park, J.-H.; Cho, J.-H.; Cho, S.-J.; Kim, D. W.; Dan, H.; Kang, Y.; Lee, S.-Y.

Direct ultraviolet-assisted conformal coating of nanometer-thick poly(tris(2-(acryloyloxy)ethyl) phosphate) gel polymer electrolytes on high-voltage LiNi1/3Co1/3Mn1/3O2 cathodes. J. Power Sources 2013, 244, 389-394. (21)

Kang, K. S.; Choi, S.; Song, J.; Woo, S.-G.; Jo, Y. N.; Choi, J.; Yim, T.; Yu, J.-S.; Kim,

Y.-J. Effect of additives on electrochemical performance of lithium nickel cobalt manganese oxide at high temperature. J. Power Sources 2014, 253, 48-54. (22)

Hofmann, A.; Schulz, M.; Indris, S.; Heinzmann, R.; Hanemann, T. Mixtures of Ionic

Liquid and Sulfolane as Electrolytes for Li-Ion Batteries. Electrochim. Acta 2014, 147, 704-711. (23)

Sun, Y. K.; Chen, Z.; Noh, H. J.; Lee, D. J.; Jung, H. G.; Ren, Y.; Wang, S.; Yoon, C. S.;

Myung, S. T.; Amine, K. Nanostructured high-energy cathode materials for advanced lithium batteries. Nat. Mater. 2012, 11, 942-7. (24)

Park, K.-J.; Lim, B.-B.; Choi, M.-H.; Jung, H.-G.; Sun, Y.-K.; Haro, M.; Vicente, N.;

Bisquert, J.; Garcia-Belmonte, G., A high-capacity Li[Ni0.8Co0.06Mn0.14]O2 positive electrode with a dual concentration gradient for next-generation lithium-ion batteries. J. Mater. Chem. A 2015, 3, 22183-22190. (25)

Guilmard, M.; Rougier, A.; Grune, M.; Croguennec, L.; Delmas, C. Effects of aluminum

on the structural and electrochemical properties of LiNiO2. J. Power Sources 2003, 115, 305-314. (26)

Woo, S. W.; Myung, S. T.; Bang, H.; Kim, D. W.; Sun, Y. K. Improvement of

electrochemical and thermal properties of Li[Ni0.8Co0.1Mn0.1]O2 positive electrode materials by multiple metal (Al, Mg) substitution. Electrochim. Acta 2009, 54, 3851-3856. (27)

Dean, J. A. Lange’s Handbook of Chemistry, 11th ed. McGraw-Hill: New York, New

York, USA 1979.

21 Environment ACS Paragon Plus

ACS Energy Letters

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(28)

Page 22 of 23

Lee, K. S.; Myung, S. T.; Amine, K.; Yashiro, H.; Sun, Y. K. Structural and

Electrochemical Properties of Layered Li[Ni1−2xCoxMnx]O2 (x=0.1–0.3) Positive Electrode Materials for Li-Ion Batteries. J. Electrochem. Soc. 2007, 154 (10), A971. (29)

Lee, J. H.; Yoon, C. S.; Hwang, J.-Y.; Kim, S.-J.; Maglia, F.; Lamp, P.; Myung, S.-T.;

Sun, Y.-K., High-energy-density lithium-ion battery using a carbon-nanotube–Si composite anode and a compositionally graded Li[Ni0.85Co0.05Mn0.10]O2 cathode. Energy Environ. Sci. 2016, 9, 2152-2158. (30) Zhou, F.; Zhao, X.; Goodbrake, C.; Jinag, J.; Dahn, J. R. Solid-State Synthesis as a Method for the Substitution of Al for Co in LiNi1 ⁄ 3Mn1 ⁄ 3Co ( 1 ⁄ 3 − z ) Al z O2, J. Electrochem. Soc. 2009, 156, A796-A801. (31)

Miller, D. J.; Proff, C.; Wen, J. G.; Abraham, D. P.; Bareño, J. Observation of

Microstructural Evolution in Li Battery Cathode Oxide Particles by In Situ Electron Microscopy. Adv. Energy Mater. 2013, 3, 1098-1103. (32)

Noh, H.-J.; Chen, Z.; Yoon, C. S.; Lu, J.; Amine, K.; Sun, Y.-K. Cathode Material with

Nanorod Structure—An Application for Advanced High-Energy and Safe Lithium Batteries. Chem. Mater. 2013, 25, 2109-2115; (33)

Lim, B.-B.; Yoon, S.-J.; Park, K.-J.; Yoon, C. S.; Kim, S.-J.; Lee, J. J.; Sun, Y.-K.

Advanced Concentration Gradient Cathode Material with Two-Slope for High-Energy and Safe Lithium Batteries. Adv. Funct. Mater. 2015, 25, 4673-4680. (34)

Demeaux, J.; Lemordant, D.; Caillon-Caravanier, M.; Galiano, H.; Claude-Montigny, B.

New insights into a high potential spinel and alkylcarbonate-based electrolytes. Electrochim. Acta 2013, 89, 163-172.

22 Environment ACS Paragon Plus

Page 23 of 23

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 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

(35)

Zuo, X.; Fan, C.; Liu, J.; Xiao, X.; Wu, J.; Nan, J. Lithium Tetrafluoroborate as an

Electrolyte Additive to Improve the High Voltage Performance of Lithium-Ion Battery. J. Electrochem. Soc. 2013, 160, A1199-A1204.

23 Environment ACS Paragon Plus