Designing Hierarchical Assembly of Carbon-Coated TiO2

Mar 4, 2019 - Despite the many benefits of hierarchical nanostructures of oxide-based electrode materials for lithium-ion batteries, it remains a chal...
3 downloads 0 Views 10MB Size
Subscriber access provided by WEBSTER UNIV

Energy, Environmental, and Catalysis Applications

Designing Hierarchical Assembly of Carbon-Coated TiO2 Nanocrystals and Unraveling the Role of TiO2/Carbon Interface in Lithium-Ion Storage in TiO2 Je Uk Ha, Jeongmin Lee, Muhammad A Abbas, Moo Dong Lee, Junghyun Lee, and Jin Ho Bang ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.8b21705 • Publication Date (Web): 04 Mar 2019 Downloaded from http://pubs.acs.org on March 4, 2019

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 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 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.

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 36 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 Applied Materials & Interfaces

Designing Hierarchical Assembly of Carbon-Coated TiO2 Nanocrystals and Unraveling the Role of TiO2/Carbon Interface in Lithium-Ion Storage in TiO2 Je Uk Ha,†,# Jeongmin Lee,†,# Muhammad A. Abbas,‡ Moo Dong Lee,† Junghyun Lee,† and Jin Ho Bang†, ‡,$,* Department of Bionano Technology, Nanosensor Research Institute, and Department of Chemical and Molecular Engineering, Hanyang University, 55 Hanyangdaehak-ro, Sangnok-gu, Ansan, Gyeonggi-do 15588, Republic of Korea AUTHOR INFORMATION Corresponding Author: Jin Ho Bang *

Email: [email protected]



Department of Bionano Technology, Hanyang University, 55 Hanyangdaehak-ro, Sangnok-gu,

Ansan, Gyeonggi-do 15588, Republic of Korea ‡

Nanosensor Research Institute, Hanyang University, 55 Hanyangdaehak-ro, Sangnok-gu, Ansan,

Gyeonggi-do 15588, Republic of Korea

ACS Paragon Plus Environment

1

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 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 2 of 36

Department of Chemical and Molecular Engineering, Hanyang University, 55 Hanyangdaehak-

ro, Sangnok-gu, Ansan, Gyeonggi-do 15588, Republic of Korea #

These authors contributed equally to this work.

KEYWORDS: titanium dioxide, hierarchical assembly, interface, carbon layer, oxygen vacancy

ABSTRACT. Despite the many benefits of hierarchical nanostructures of oxide-based electrode materials for lithium-ion batteries (LIBs), it remains a challenging task to fully exploit the advantages of such materials partly because of their intrinsically poor electrical conductivities. The resulting limited electron supply to primary particles inside secondary microparticles gives rise to significant variation in lithium-ion (Li+) storage capability within the nanostructured particles. To address this, facile annealing, where in situ generated carbon-coated primary particles were assembled into porous micro-agglomerates, is demonstrated to prepare nanostructured titanium dioxide (TiO2). A systematic study on the effect of the carbon coating reveals that it is exclusively governed by the characteristics of the TiO2/carbon interface rather than by the nature of the carbon coating. Depending on their number, oxygen vacancies (VO) created by carbothermal reduction on the TiO2 surface are detrimental to Li+ diffusion in the TiO2 lattice, and structural distortion at the interface profoundly influences the Li+ (de-)intercalation mechanism. This new insight serves as a stepping stone toward understanding an important, yet often overlooked effect of the oxide/carbon interface on Li+ storage kinetics, thereby demanding more investigation to establish a new design principle for carbon-coated oxide electrode materials.

ACS Paragon Plus Environment

2

Page 3 of 36 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 Applied Materials & Interfaces

1. Introduction Owing to its high chemical stability, structural robustness, abundance, and high working voltage (~1.75 V vs. Li+/Li), anatase TiO2 has extensively been investigated as an anode material for LIBs since the first report of the electrochemical insertion of Li+ into TiO2 by Ozhuku et al.1 However, these merits are largely compromised by its intrinsically poor electrical conductivity (~10 −9 Ω−1 m−1)2 and lithium diffusivity (~10−15 cm2 s−1),3 which are major hurdles for the widespread application of TiO2 to LIBs. Numerous studies to overcome this limitation have been carried out. Among them, strategies based on preparing nanostructured TiO2 stand out, and the nanostructuring is critical for the application of TiO2 to LIBs from two perspectives. First, it facilities Li+ transport by shortening diffusion length (τeq=L2/2DLi, τeq: equilibrium time for Li+ diffusion, L: particle size, and DLi: diffusion coefficient of Li+).4 Second, it alters the thermodynamics and kinetics of Li+ insertion into TiO2 (TiO2 + xLi+ + xe- → LixTiO2, x: mole fraction of Li+ in TiO2) much more favorably as Mulder et al. revealed earlier.5,6 Since the maximum Li+ insertion (typically x=0.5 for bulk TiO2) is size-dependent, decreasing the size of TiO2 particles to less than 7 nm can render them fully able to accommodate Li+.5 Although a wide variety of TiO2 nanostructures have proven their prowess, hierarchically structured TiO2––where nanoparticles are assembled into porous micro-agglomerates––has been considered as the ideal nanostructure. Hierarchically assembled TiO2 has often been prepared by burning sacrificial porogens off of organic/inorganic hybrid precursors while converting them into mesoporous TiO2.7-9 Another popular route is utilizing selfassembly by Ostwald ripening under solvothermal conditions.10-13 This nanoarchitecture can mitigate the disadvantages of TiO2 while not significantly compromising electrode packing density. However, the hierarchical structure promotes uneven electron mobility in each primary particle depending on the location in the agglomerate.14 Nanoparticles inside the micro-agglomerates are

ACS Paragon Plus Environment

3

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 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 36

likely to show higher resistance than those outside them because they are not directly in contact with conductive carbon additive; hence, electron transport within the micro-agglomerates primarily proceeds by hopping (Figure 1A).15 Given the importance of well-balanced transport between Li+ and electrons (e-) for effective Li+ migration into the TiO2 lattice (i.e., the concurrent hopping of Li+/e- pairs),16 this limited electron supply to the interior particles hampers Li+ intercalation and, hence, inevitably leads to high overpotentials for Li+ transport.4,17 To resolve this issue, we synthesized hierarchically assembled carbon-coated TiO2 nanocrystals (HAC TiO2), where all the primary particles were coated with a thin carbon layer. Our work demonstrates that the carbon network formed throughout the hierarchical agglomerate can significantly alleviate the spatially unbalanced electron supply, hence enhancing the storage and rate capability of TiO2. Another important contribution of this work is that it provides new insights into the influence of the TiO2/carbon interface on the Li+ storage of TiO2, which has long been overlooked. Our extensive literature survey led us to affirm the common belief in the carbon coating effect whereby a thin carbon layer overcoated on TiO2 particles enhances the electrical conductivity of TiO2, thus improving its LIB performance.18-25 However, we determined that this benefit intimately depended on the nature of the TiO2/carbon interface. In-depth surface analysis revealed that the thermal reduction of TiO2 by the in situ generated carbon removed some oxygen atoms, thereby introducing surface oxygen vacancies (VO) (Figure 1B). The degree of VO at the interface profoundly impacted the LIB performance of TiO2, thereby kindling our interest to understand the nature of the TiO2/carbon interface and its relationship with Li+ storage kinetics. Our systematic investigation on this undisclosed aspect of the interface is discussed in this report.

ACS Paragon Plus Environment

4

Page 5 of 36 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 Applied Materials & Interfaces

2. Experimental Section 2.1 Synthesis of HAC TiO2. TiO2 nanocrystals were synthesized by a nonhydrolytic sol–gel reaction as follows: 10.0 mL of titanium (IV) isopropoxide and 50 mL of benzyl alcohol were poured into a 100 mL C-276 liner. After stirring the solution at 500 rpm for 30 min, the liner was slid into a bench-type autoclave and tightly sealed. The autoclave was heated to 250 °C, held there for 24 h, and then naturally cooled. A white precipitate was collected by centrifugation, washed with 60 mL of tetrahydrofuran, and dried in a vacuum oven at 60 °C for 24 h. To produce the carbon-coated TiO2 agglomerate, the as-obtained TiO2 powder (1.0 g), denoted as TiO2-as made, was placed in a ceramic crucible and annealed in N2 for 1 h. Each sample was heated to 500, 600, or 700 °C at 10 °C min−1. The obtained HAC TiO2 will hereafter be referred to as TiO2-500, TiO2600, and TiO2-700, respectively. For comparison, the as-obtained TiO2 powder was heated under the same conditions except in O2. 2.2 Characterization. X-ray diffraction (XRD) patterns of the HAC TiO2 were obtained using an X-ray diffractometer (Bruker D2 PHASER) with Cu Kα irradiation and operating at 10 mA and 30 kV. Nitrogen adsorption/desorption isotherms were measured at 77 K, using a surface area analyzer (BELSORP-mini II, BEL Japan), and the specific surface area and pore size distributions were determined by the Brunauer–Emmett–Teller (BET) and Barrett–Joyner–Halenda (BJH) methods, respectively. Raman spectroscopy (Renishaw RM-1000) was performed to examine the carbon layer on the HAC TiO2. Surface analysis was performed by X-ray photoelectron spectroscopy (XPS, PHI Versa Probe system), and the oxygen vacancies of TiO2 surface were examined using an electron paramagnetic resonance (EPR) spectrometer (Bruker EMXplus9.5/2.7). Thermogravimetric analysis (TGA) were performed with a thermal analyzer (SDT-Q600, TA Instrument) ramped at 5 °C min−1 in air. Transmission electron microscopy (TEM, JEOL-

ACS Paragon Plus Environment

5

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 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 36

2100F) and scanning electron microscopy (SEM, Hitachi S-4800) were performed to acquire information about morphology and local crystal structure. 2.3 Electrode Fabrication and Evaluation. For electrochemical characterization, CR-2032-type coin cells were assembled in an argon-filled glove box and were evaluated using various electrochemical techniques. Each TiO2 sample was thoroughly mixed with carbon black (Super P) and polyvinylidene difluoride (PVDF) and was dissolved in N-methyl-2-pyrrolidone (NMP) in a weight ratio of 70:15:15, using a mixer (AR-100, THINKY Corp.). The prepared slurry was spread uniformly thick onto a current collector (Cu foil), using a micrometer adjustable film applicator, and the electrode film was dried at 120 °C for 12 h in a vacuum oven. The active material loading was set to ~1.0 mg cm−1. For the coin cell assembly, lithium metal foil, 1 M LiPF6 dissolved in ethylene carbonate/dimethyl carbonate (1:1 v:v), and Celgard 2320 membrane were employed for the counter electrode, electrolyte, and separator, respectively, and the working electrode was made of TiO2 film. Galvanostatic charge–discharge measurements and galvanostatic intermittent titration technique (GITT) analysis were carried out using battery cyclers (WBCS3000, WonATech Co., Ltd.). In situ Electrochemical impedance spectroscopy (EIS) was measured during charge/discharge cycles at various potentials using ZIVE BP2A electrochemical workstation WonATech Co., Ltd.).

3. Results and Discussion In our approach to preparing the HAC TiO2, crystalline anatase TiO2 nanoparticles were first produced by a nonhydrolytic sol–gel reaction,26 where Ti–O–Ti bonds were formed through the C–C coupling of titanium tetrabutoxide and benzyl alcohol, and the resulting nanoparticles were then annealed at 500, 600, or 700 °C under an inert atmosphere. Annealing the TiO2

ACS Paragon Plus Environment

6

Page 7 of 36 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 Applied Materials & Interfaces

nanocrystals formed porous agglomerates while converting surface species (i.e., residues of benzyl alcohol and other decomposed organic species) into thin carbon layers. This conversion of an organic stabilizer into carbon layer was quite successful in other reports as well.27,28 The annealing temperature was limited to up to 700 °C because the rutile phase (which is less efficient for Li + storage than the anatase phase) began to form at elevated temperatures (Figure S1). Figure 2 shows the XRD patterns of the TiO2-as made, TiO2-500, TiO2-600, and TiO2-700. All the XRD peaks matched the typical diffraction peaks of anatase TiO2 (JCPDS 21-1272), and none of the peaks were attributable to any impurities in the specimens. The TiO2 cell parameters and grain sizes were determined by Rietveld refinement and are summarized in Table 1. Although the grain size of the TiO2 annealed at 500 °C increased from 9.9 to 13 nm, further increase in the grain size was substantially suppressed even when the annealing temperature was raised to 700 °C. Unlike the TiO2 annealed in N2, however, that annealed in air showed a dramatic increase in grain size (14.9, 29.2, and 63.2 nm for TiO2-500, TiO2-600, and TiO2-700, respectively), highlighting another benefit of the in situ generated carbon layer over the TiO2 particles: preventing grain growth even during high-temperature annealing (Figures S2-3). Since there is always a compromise between crystallinity and surface area during high-temperature annealing,29,30 this simple, scalable process can be of great benefit for the preparation of high-performance electrode materials. The lattice parameters of TiO2-as made (a=3.7868 Å, c=9.4974 Å), on the other hand, slightly deviated from those of standard anatase TiO2 (a=3.783 Å, c=9.507 Å; JCPDS 21-1272), particularly in the [001] direction. However, the shorter lattice along the [001] direction of TiO2-as made lengthened upon annealing, and the lattice parameters of all the HAC TiO2 became closely matched to those of standard anatase TiO2.

ACS Paragon Plus Environment

7

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 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 36

The microtextural properties of the HAC TiO2 were characterized by N2 physisorption isotherms (Figure 3), and the obtained microtextural parameters are summarized in Table 1. The overall trend in the isotherms was notably different from that observed for the TiO2 samples annealed in air. Although the mesoporosity of the agglomerated TiO2 annealed in air abruptly decreased with temperature (Figure S4), the HAC TiO2 inherited the original porosity of TiO2-as made by and large. It is noteworthy that the thermal treatment offered sufficient robustness to the porous structure, which played a crucial role in stability during battery operation that will be discussed later. The hysteresis loop became shifted toward a high-pressure region after the annealing, implying that a mesoporous structure with larger pores had been formed at the expense of smaller mesopores during the annealing. The specific surface area of TiO2-as made was 70.92 m2 g−1, and those of TiO2-500, TiO2-600, and TiO2-700 were as high as 93.78, 88.69, and 83.69 m2 g−1, respectively. The pore size distribution (in the insets of Figure 3) showed that although annealing had slightly narrowed the multimodal mesopore distribution, the similar total pore volumes (~0.29 cm3 g−1) of the HAC TiO2 suggested that there was no significant difference in the porous structures of the three TiO2 agglomerates. This trend was strikingly different from that found in the TiO2 annealed in air; the surface area of TiO2 heated at 700 °C was merely 3.73 m2 g−1, and the mesoporosity completely disappeared when TiO2 was annealed at 700 °C. The nature of the carbon layers in the HAC TiO2 was examined by Raman spectroscopy. Figure 4A-C display the Raman spectra of TiO2-500, TiO2-600, and TiO2-700, where two characteristic Raman peaks designated as D and G bands appeared at ~1343 and 1589 cm−1, respectively. This spectral feature was different from that observed for TiO2-as made, where the distinctive peaks were not well resolved (Figure S5). These bands were carefully deconvoluted to gain more insight into the carbon nature, the results of which are presented in Figure 4. It is

ACS Paragon Plus Environment

8

Page 9 of 36 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 Applied Materials & Interfaces

generally accepted that D1 and G1 peaks appearing at ~1344 and ~1593 cm−1 are analogous to those of the D and G bands of crystalline graphite, and D2 and G2 peaks at ~1412 and ~1525 cm−1 are ascribed to those of the D and G bands of amorphous carbon.31,32 Note that additional peaks labeled D3 (~1275 cm−1) and S (~1131 cm−1) were included in the D band fitting as previously reported,33 which presumably stemmed from the very disordered nature of the carbon layer formed by the carbonization of organic residues at low temperatures.21 The D′ peak at ~1617 cm−1 has been known to be associated with disordered carbon edges.34-36 Shimodaira et al. proposed the G2/G1 ratio as a useful gauge to identify the relative disorderness in the sp2 network of activated carbons (a smaller ratio is associated with less disorderness),31 based on which we concluded that a more-disordered carbon layer had been formed on TiO2-700 than on TiO2-500 (Figure 4D and Table 1). The amount of carbon in each TiO2 sample was estimated using TGA, according to which the carbon contents were 4.5, 3.8, and 3.5 wt.% for TiO2-500, TiO2-600, and TiO2-700, respectively. The surface characteristics of the HAC TiO2 were investigated using XPS. Figure 5A exhibits the XPS Ti 2p spectra of TiO2-as made, TiO2-500, TiO2-600, and TiO2-700. The raw spectral data were deconvoluted into two doublet peaks (2p3/2 and 2p1/2), each of which was assigned to Ti4+ at 458.7 and 464.4 eV and to Ti3+ at 457.3 and 462.5 eV, respectively (See Table S1 for detailed fitting results). Interestingly, although there was no Ti3+ species in TiO2-as made, Ti3+ evolved upon annealing and, in particular, a remarkable amount of it was found in TiO2-500. With increasing annealing temperature, the intensity of the peaks attributed to Ti 3+ diminished. Since the formation of Ti3+ accompanies the creation of VO,37-39 this implied that carbonization had introduced defects (VO) to the TiO2 surface as follows: TiO2 (s) + C (s) → TiO2-δ (s) + CO2 (g)

(1)

ACS Paragon Plus Environment

9

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 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 36

In fact, the reducing power of carbon at elevated temperatures (i.e., carbothermal reduction) has long been utilized to extract metals from ores. The interfacial oxygen defects could survive with the aid of the carbon layer wrapping individual TiO2 nanocrystals.40-42 This could be part of the reason for the gradual decrease in the intensity of the peaks attributed to Ti 3+ with temperature because the carbon content decreased with increasing annealing temperature. In addition, the more disordered carbon layer formed in TiO2-700 may ease the diffusion of O2 through the carbon layer, leading to the decrease in VO. The XPS O 1s spectra of the four samples (Figure 5B, Table S2) showed the same pattern as the Ti 2p spectra. The peak at 530.32 eV, corresponding to VO, appeared upon annealing and faded with increasing annealing temperature. The appearance of a peak attributed to O2- at 527.7 eV along with the VO peak was in accordance with this observation because the adsorbed anionic oxygen could exist only when excess electrons provided by VO were available.43 The XPS spectrum of TiO2 annealed in air (Figure S7) did not show any VO signature peaks, confirming the role of the carbon layer as the reducing agent. The VO in the HAC TiO2 was affirmed by EPR spectroscopy. Figure 6 shows the EPR spectra of TiO2-as made, TiO2-500, TiO2-600, and TiO2-700, where although a characteristic peak at g=2.001 (corresponding to VO) was absent in the EPR spectrum of TiO2-as made, it appeared upon annealing, and its intensity diminished with increasing annealing temperature. The morphology and microstructural information about the TiO2 samples were investigated using electron microscopy. The TEM analysis of TiO2-as made (Figure S8A) showed that the TiO2 nanocrystals consisted of a mixture of rhombic and rectangular particles whose average sizes were in the range 15–25 nm. A high-resolution TEM (HRTEM) image combined with a fast Fourier transform (FFT) pattern, where an ordered dotted pattern was observed, revealed the single

ACS Paragon Plus Environment

10

Page 11 of 36 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 Applied Materials & Interfaces

crystalline nature of the TiO2 nanoparticles (Figure S8B). The TEM images along with the SEM images of the HAC TiO2 (Figure 7A-C and Figures S9-10) show that the TiO2 nanocrystals coated with the thin carbon layer were packed into dense micro-agglomerates. To acquire the detailed atomic structures of the HAC TiO2, the local crystal structures of the TiO2 surfaces were analyzed in depth. The HRTEM images in Figure 7D-F were taken along the [1̅1̅1] zone axis (as affirmed by analyzing the FFT patterns), hence allowing the inspection of the (101) plane of each HAC TiO2. The Ti–Ti interlayer distance (d) along the [101] direction (as depicted in the modeled crystal structure) was carefully measured using TEM intensity profiles (Figure 7G-I), where the spacings of five exterior atomic layers and five interior ones were examined for each sample. For TiO2-500 (Figure 7G), although the average interior interlayer distance was 3.36 Å, the average exterior one was 3.50 Å, implying structural distortion on the TiO2-500 surface.44-46 In contrast, the spatial difference in the interlayer distance became much less significant for TiO2-600 (Figure 7H), and there was nearly no variation in the interlayer distance for all the examined areas of TiO 2-700 (Figure 7I), which corroborated the XPS and EPR analysis results in that VO were preferentially created at the interface between the TiO2 surface and the carbon layer. The Li+ storage capability of the HAC TiO2 was assessed by galvanostatic charge/discharge measurements at various C rates in the range 1.0–3.0 V vs. Li+/Li. Figure 8A displays the voltage profiles of the four TiO2 electrodes measured at 1 C (168 mA g−1). The voltage profile of TiO2 has been well established by three characteristic regions (often referred to as Regions A, B, and C).47,48 In Region A (above 1.75 V), Li+ is initially stored in the TiO2 lattice by the formation of a lithiumpoor solid solution (i.e., no structural change occurs during Li+ insertion) denoted as LiδTiO2.49,50 Region B corresponds to the flat potential region at ~1.75 V, where biphasic domains of lithiumpoor tetragonal titanate (LiδTiO2) and lithium-rich orthorhombic titanate (Li0.55TiO2) coexist.51-53

ACS Paragon Plus Environment

11

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 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 36

The sloped region below 1.75 V is designated as Region C, where a diffusion-limited biphasic transition between the orthorhombic Li0.55TiO2 and fully lithiated TiO2 (Li1TiO2) occurs. Although there is a second pseudo-plateau at ~1.5 V, it is often indistinguishable at fast charge/discharge rates (Figure S11) because of the more sluggish diffusion kinetics for inserting Li+ into Li0.55TiO2 whose octahedral sites are more than half-filled.47,54-57 The specific capacities increased with increasing annealing temperature, and the capacity of TiO2-700 at 1 C reached as much as 225 mAhg−1. Figure 8B shows that the capacity of TiO2-700 can be sustained at as high as 188 and 162 mAhg−1 at 5 and 10 C, respectively, (and even 91 mAhg−1 at 50 C), which are comparable to the best sustained capacities reported in the literature (Table S3). On the contrary, with decreasing annealing temperature, the capacity declined more noticeably; for instance, the capacities of TiO2500 at 5 and 10 C were merely 85 and 56 mAhg−1, respectively. The LIB performance of TiO2 annealed in air (uncoated TiO2), on the other hand, was inferior to that of HAC TiO2 (Figure S12), highlighting the merits of the in situ generated thin carbon layer: preserving nanostructures that facilitated Li+ diffusion while providing better electrical conductivity. The electrochemical behavior of TiO2 is dictated by several factors including crystallite size, porous structure, surface area, and morphology. Although there were slight dissimilarities in these structural parameters, our analyses revealed that all three TiO2 agglomerates were nearly identical in their structural and textural characteristics. It is noteworthy that the lattice parameters of TiO2-as made were different from those of standard TiO2, implying structural disordering (i.e., lattice distortion) in TiO2-as made, but not preferentially on the surface. Strictly speaking, TiO2500 seemed to possess more favorable structural characteristics for LIB application than the other annealed TiO2 samples: smaller grain size, larger pore volume and surface area, and larger lattice spacing along the c axis (i.e., the principal Li+ diffusion channel),58,59 albeit slightly different

ACS Paragon Plus Environment

12

Page 13 of 36 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 Applied Materials & Interfaces

among the annealed samples. Despite these benefits of TiO2-500, its performance was inferior to that of TiO2-700, which led us to explore the influence of the TiO2/carbon interface on Li+ storage. Aside from the aforementioned TiO2 characteristics, the thin carbon layer could be an extrinsic factor that affected the LIB performance of TiO2. Since Li+ intercalates into graphitic layers at potentials below 1.0 V vs. Li+/Li,60 the major diffusion pathways for Li+ were likely to be an interconnected pore network created by various defects in the carbon layer. For this reason, the porous structure rather than the thickness and graphitic nature of the carbon layer, as noted previously, would be the most critical factor that determined the effect of the carbon layer.61,62 The TEM images of the carbon layers revealed that the layers were thin, rough, porous, and very similar in appearance, regardless of the annealing temperature (Figure S13). Although we speculated that the observed disparities in carbon content and the degree of graphitization may have impacted Li+ diffusion, the TEM analysis suggests that the impact would unlikely be so great to entirely dictate the LIB performance of TiO2 (as will be evidenced by electrochemical investigation). A noticeable difference in the discharge curves (Figure 8C) was the more elongated plateau region for TiO 2700 than for its other three counterparts. The capacity contributions from Region C were similar for all four electrodes (70–80 mAhg−1), and the slight differences in contribution (TiO2-500 > TiO2-600 > TiO2-700 > TiO2-as made) followed the trend in their specific surface areas. Given that the pseudocapacitive reaction is a dominant charge storage mechanism in Region C, this consistency sounded plausible because interfacial storage tended to be proportional to surface area. Hence, the similar capacity in Region C could be attributed to similar surface area and porous structure. It is noteworthy that if there were a significant difference in Li+/e- transport through the carbon layers, this would never be the case. Therefore, this observation partly supported our speculation on the effect of the carbon layer. In contrast, the capacity contributions from Regions

ACS Paragon Plus Environment

13

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 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 36

A and B to each electrode were distinct. For TiO2-700, it was as high as 150 mAhg−1, but it decreased with decreasing annealing temperature to ~90 mAhg−1 for TiO2-500 and TiO2-as made. In general, the relative contributions of Regions A and B to the overall capacity is intimately related to crystallite size.5,6,54 Given the similar crystallite sizes of all the TiO2 samples, however, the remarkably different capacity suggested that VO at the interface might play a critical role in dictating Li+ (de-)intercalation kinetics. The stability of HAC TiO2 was examined by a long-term cycling test (Figure S14), where all HAC TiO2 retained their original capacity over 500 cycles unlike TiO2-as made that showed gradual degradation. This revealed that the carbon layer could stabilize TiO2 upon repeated charge/discharge cycles. This was partly supported by the SEM images taken after the long-term cycling test (Figure S15). The majority of micrometer-sized particles were preserved in TiO2-500, TiO2-600, and TiO2-700. In the contrary, TiO2-as made were randomly agglomerated because of the lack of rigid interparticle connections and carbon coating as in TiO2-500, TiO2-600, and TiO2-700. The XRD analysis provided additional insight into the capacity degradation (Figure S16). While anatase TiO2 phase was preserved by and large in all the electrodes, there was a small peak at ~45° in the case of TiO2-as made and TiO2-500 that was determined to be bronze-type TiO2 phase (TiO2(B)). We speculated that the formation of this new crystal phase may be associated with the capacity fade. To gain insight into the influence of the TiO2/carbon interface, cyclic voltammetry was carried out, and the resulting cyclic voltammograms (CVs) are presented in Figure 8D-G. A salient feature in the CVs was a pair of new redox peaks (labeled as P2) appearing in the ranges 1.2–1.5 and 2.5–2.8 V in addition to the typical redox peaks corresponding to lithium insertion/extraction into the TiO2 lattice (labeled as P1). The new cathodic and anodic peaks appeared more prominent with increasing scan rate, and the peak currents associated with P2 became even higher than those

ACS Paragon Plus Environment

14

Page 15 of 36 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 Applied Materials & Interfaces

associated with P1 at the high scan rates for TiO2-as made. Another important trend in the CVs was the disappearance of P2 with increasing annealing temperature. According to our extensive literature survey, although P2 has rarely been reported, it seemed to appear when Li+ diffusion was severely restricted.30,54,63,64 The recent report by Passerini et al. disclosed that the advent of this alternative Li+ insertion/extraction pathway was intimately related to structural disorder, which impeded Li+ diffusion in TiO2.63 Given this report, the distorted lattice of TiO2-as made (i.e., the shrunken lattice along the [001] direction) was presumably responsible for the relatively high P2 current. However, the lattice distortion was significantly alleviated upon annealing, as evidenced in the XRD analysis, on the basis of which we attributed P2 in the CVs of the HAC TiO2 to the remarkably contrasting surface characteristics at the TiO2/carbon interface. In general, the Li+ diffusion coefficient (DLiδ) of nonstoichiometric TiO2 (i.e., TiO2-δ) is a function of the ambipolar conductivity of lithium (σLiδ), as expressed by the following relationship:4,65 σLiδ ∝ σLi+ × σe- / (σLi+ + σe-),

(2)

where σLi+ is the ionic conductivity and σe-, the electrical conductivity. The dependencies of σLi+ and σe on the degree of Vo (δ) run in opposite directions;4 hence, there exists an optimal δ, where Li+/e- transport is well balanced for efficient Li+ diffusion. Therefore, more surface Vo have often been reported as deleterious because irreversibly trapped Li+ on defects could significantly hinder Li+ diffusion,4,66,67 which apparently was the case for TiO2-500, whose surface accommodated a considerable number of Vo. The relatively low initial coulombic efficiency of TiO2-500 compared to those of TiO2-600 and TiO2-700 reflected greater irreversible Li+ trapping in TiO2-500 (Figure S14). With decreasing Vo, however, unfavorable Li+ trapping at the TiO2/carbon interface was mitigated, allowing TiO2 to better utilize the benefits of the carbon layer. The more hindered Li+ diffusion in TiO2-500 was also supported by the overshooting peak in the

ACS Paragon Plus Environment

15

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 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 36

potential profile: a peak appearing in the charging profile before the potential plateau begins.68,69 This peak is the signature of the TiO2 memory effect caused by sluggish Li+ diffusion;69 therefore, the overshooting peak in the charging profile for TiO2-700 was undeveloped compared to its counterparts, indicative of less hindered Li+ diffusion. The Li+ diffusion kinetics of each TiO2 electrode was further investigated using in situ EIS, where impedance spectra were measured during the charge/discharge process. The limitations of Li+ insertion/extraction originate from the hindrance of ionic (or electronic) charge transfer at the electrode/electrolyte interface and inside a Li+-hosting matrix (i.e., TiO2). These kinetic barriers were reflected in the Nyquist plots of the TiO2 electrodes measured during charge and discharge cycle at various potentials (Figure 9A,B), where distorted arcs associated with interfacial charge transfer processes appeared in the high frequency region, and the inclined linear responses related to Li+ diffusion inside TiO2 followed in the low frequency region. Several resistance components linked to the kinetics of Li+ insertion/extraction were extracted by fitting the spectra to an equivalent circuit (Figure S17) consisting of uncompensated ohmic resistance (Rs), mostly representing the electrolyte resistance, coupled with several resistance (R)/constant phase element (CPE) components. In the equivalent circuit, the resistance associated with charge transfer were represented by Rct, and the resistance of Li+ diffusion within the TiO2 lattice by RLi. The CPE— expressed by ZCPE = Q(jw)-α (0≤ α ≤1), where j is an imaginary unit, w is the angular frequency, and Q and α are frequency-independent CPE parameters—represented the nonideality associated with electrode roughness, which is common for battery electrodes.38 Figure 9C-F show the resistances derived from the EIS fitting results. Although the interfacial charge transfer resistance (Rct) were almost independent of the applied potentials, the bulk diffusion resistance (RLi) varied to some extent depending on the potentials in particular in TiO2-as made and TiO2-500, which was

ACS Paragon Plus Environment

16

Page 17 of 36 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 Applied Materials & Interfaces

consistent with the EIS analysis results of TiO2 reported previously.70 The Rct values of TiO2-700 measured during the discharging were notably smaller than those of TiO2-600, TiO2-500, and TiO2-as made (Figure 9C), implying that the kinetic limitation associated with Li+ diffusion through the interfacial Li+ intake from the electrolyte was mitigated more effectively in TiO2-700 than in its counterparts. The same Rct trend was observed in the charging (Figure 9E), revealing that the Li+ extraction from TiO2 host at the interface was more favorable in TiO2-700 as well. Since the nature of the carbon layer was primarily responsible for the Li+ diffusion and electron transport (i.e., electrical conductivity) that dictated these surface processes, we speculated that more a kinetically favorable carbon layer had been formed in TiO2-700. Given the greater RLi values and their variation among the electrodes, however, we speculated that the kinetic limitation of TiO2 mainly stemmed from Li+ diffusion within TiO2 particles. The overall trend in RLi was consistent with LIB performance (Figure 9D, F), and this trend in solid-state Li+ diffusion was also verified by the diffusion coefficients measured by the GITT analysis (Figure S18). As GarciaBelmonte et al. noted,70 the difference in RLi most likely resulted from the strain developed at the TiO2/carbon interface given that other structural characteristics were very similar. The interfacial strain was obviously associated with lattice distortion driven by Vo in our case. It is noteworthy that there is a remarkable difference in the potential-dependency of RLi in TiO2-as made and TiO2500 during the charging and discharging cycles. Unlike the RLi trend in the charging (Figure 9D), RLi rose greatly during the charging (Figure 9F), indicating the presence of more sluggish Li + diffusion during the Li+ extraction process. This suggested that the structural disordering had a more influence on Li+ extraction rather than Li+ insertion. To investigate the carbon layer effect on the solid-electrolyte interface (SEI), the in situ EIS spectroscopy was performed again during the first and second cycles and the resulting Nyquist plots were compared (Figure S19). Since the

ACS Paragon Plus Environment

17

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 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 36

resistance corresponding to the SEI layer (RSEI) was not resolved in the Nyquist plots as a separate arc, it was difficult to quantify the resistance. However, the difference in the diameter of the arcs measured in the first and second cycles can reflect the resistance resulting from the SEI formation. The RSEI values were in the order of 100–200 mΩ·g, and their variation among the TiO2 electrodes were very close, hinting that there was no significant influence on the SEI formation by the carbon layer. 4. Conclusion Carbon coating over metal oxides utilized as Li+ storage matrixes has been a common practice to mitigate their poor electrical conductivities. Despite the development of various carbon coating strategies, it has been a great challenge to carbon-coat individual nanoparticles, in particular when metal oxides form hierarchical nanostructures. To remedy this, TiO2 nanocrystals were prepared by a nonhydrolytic sol–gel reaction and then annealed. A thin carbon layer was formed in situ over primary particles while hierarchical agglomerates were produced. This approach enabled the formation of a uniform carbon network throughout the micro-agglomerates, thereby alleviating the spatial unevenness in electron supply often encountered in many hierarchically nanostructured oxides. Aside from devising this synthesis route, our study offered new insights into a critical, yet often overlooked effect of the TiO2/carbon interface on Li+ storage in TiO2. Carbothermal reduction introduced oxygen vacancies on the TiO2 surface, thereby profoundly affecting the Li+ (de-)intercalation mechanism and Li+ diffusion kinetics. Cyclic voltammetry unveiled that structural distortion caused by excess oxygen defects could significantly alter the Li+ (de-)insertion pathway, and in situ EIS analysis combined with galvanostatic charge/discharge measurements revealed sluggish Li+ diffusion due to Li+ trapping at the undesirably defected TiO2 surface. This work not only highlighted the importance of seeking

ACS Paragon Plus Environment

18

Page 19 of 36 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 Applied Materials & Interfaces

a comprehensive understanding of the oxide/carbon interface, but also called urgent attention to devising a delicate approach that could manipulate the characteristics of the interface to fully exploit the benefits of carbon coating.

ASSOCIATED CONTENT Supporting Information. Additional characterizations and electrochemical analysis results. This material is available free of charge via the Internet at http://pubs.acs.org. AUTHOR INFORMATION Corresponding Author: Jin Ho Bang *Email: [email protected] Notes The authors declare no competing financial interests. ACKNOWLEDGMENTS. This work was supported by the Samsung Research Funding Center of Samsung Electronics, under Project Number SRFC-MA1601-03. It was also supported by a grant from the Basic Science Research Program through the National Research Foundation (NRF) of Korea funded by the Ministry of Science and ICT (NRF-2016R1A1A1A05005038, NRF2018M3A7B8061494) and by the Ministry of Education (NRF-2018R1A6A1A03024231).

ACS Paragon Plus Environment

19

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 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 20 of 36

REFERENCES (1) Ohzuku, T.; Takehara, Z.; Yoshizawa, S. Nonaqueous Lithium/Titanium Dioxide Cell. Electrochim. Acta 1979, 24, 219-222. (2) Pomoni, K.; Sofianou, M. V.; Georgakopoulos, T.; Boukos, N.; Trapalis, C. Electrical Conductivity Studies of Anatase TiO2 with Dominant Highly Reactive {001} Facets. J. Alloys Compd. 2013, 548, 194-200. (3) Lindström, H.; Södergren, S.; Solbrand, A.; Rensmo, H.; Hjelm, J.; Hagfeldt, A.; Lindquist, S.-E. Li+ Ion Insertion in TiO2 (Anatase). 1. Chronoamperometry on CVD Films and Nanoporous Films. J. Phys. Chem. B 1997, 101, 7710-7716. (4) Shin, J.-Y.; Joo, J. H.; Samuelis, D.; Maier, J. Oxygen-Deficient TiO2−Δ Nanoparticles via Hydrogen Reduction for High Rate Capability Lithium Batteries. Chem. Mater. 2012, 24, 543551. (5) Wagemaker, M.; Borghols, W. J. H.; Mulder, F. M. Large Impact of Particle Size on Insertion Reactions. A Case for Anatase LixTiO2. J. Am. Chem. Soc. 2007, 129, 4323-4327. (6) Shen, K.; Chen, H.; Klaver, F.; Mulder, F. M.; Wagemaker, M. Impact of Particle Size on the Non-Equilibrium Phase Transition of Lithium-Inserted Anatase TiO2. Chem. Mater. 2014, 26, 1608-1615. (7) Lee, E. J.; Nam, I.; Yi, J.; Bang, J. H. Nanoporous Hexagonal TiO2 Superstructure as a Multifunctional Material for Energy Conversion and Storage. J. Mater. Chem. A 2015, 3, 35003510. (8) Xiu, Z.; Alfaruqi, M. H.; Gim, J.; Song, J.; Kim, S.; Thi, T. V.; Duong, P. T.; Baboo, J. P.; Mathew, V.; Kim, J. Hierarchical Porous Anatase TiO2 Derived from a Titanium Metal-Organic

ACS Paragon Plus Environment

20

Page 21 of 36 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 Applied Materials & Interfaces

Framework as a Superior Anode Material for Lithium Ion Batteries. Chem. Commun. 2015, 51, 12274-12277. (9) Park, S.; Kim, J. B.; Lee, S.; Bang, J. H.; Kim, Y. S. Synthesis of Titanium Oxide Nanocups by Electrospraying. Mater. Lett. 2012, 69, 34-36. (10) Jin, J.; Huang, S.-Z.; Li, Y.; Tian, H.; Wang, H.-E.; Yu, Y.; Chen, L.-H.; Hasan, T.; Su, B.L. Hierarchical Nanosheet-Constructed Yolk-Shell TiO2 Porous Microspheres for Lithium Batteries with High Capacity, Superior Rate and Long Cycle Capability. Nanoscale 2015, 7, 12979-12989. (11) Chen, J. S.; Tan, Y. L.; Li, C. M.; Cheah, Y. L.; Luan, D.; Madhavi, S.; Boey, F. Y. C.; Archer, L. A.; Lou, X. W. Constructing Hierarchical Spheres from Large Ultrathin Anatase TiO 2 Nanosheets with Nearly 100% Exposed (001) Facets for Fast Reversible Lithium Storage. J. Am. Chem. Soc. 2010, 132, 6124-6130. (12) Wu, R.; Shen, S.; Xia, G.; Zhu, F.; Lastoskie, C.; Zhang, J. Soft-Templated Self-Assembly of Mesoporous Anatase TiO2/Carbon Composite Nanospheres for High-Performance Lithium Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 19968-19978. (13) Qiu, S.; Xiao, L.; Ai, X.; Yang, H.; Cao, Y. Yolk–Shell TiO2@C Nanocomposite as HighPerformance Anode Material for Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2017, 9, 345353. (14) Lee, S.; Cho, Y.; Song, H.-K.; Lee, K. T.; Cho, J. Carbon-Coated Single-Crystal LiMn2O4 Nanoparticle Clusters as Cathode Material for High-Energy and High-Power Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2012, 51, 8748-8752. (15) Stevanovic, A.; Yates, J. T. Electron Hopping through TiO2 Powder: A Study by Photoluminescence Spectroscopy. J. Phys. Chem. C 2013, 117, 24189-24195.

ACS Paragon Plus Environment

21

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 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 22 of 36

(16) Sushko, M. L.; Rosso, K. M.; Liu, J. Mechanism of Li+/Electron Conductivity in Rutile and Anatase TiO2 Nanoparticles. J. Phys. Chem. C 2010, 114, 20277-20283. (17) Jamnik, J.; Maier, J. Nanocrystallinity Effects in Lithium Battery Materials Aspects of NanoIonics. Part IV. Phys. Chem. Chem. Phys. 2003, 5, 5215-5220. (18) Das, S. K.; Patel, M.; Bhattacharyya, A. J. Effect of Nanostructuring and Ex Situ Amorphous Carbon Coverage on the Lithium Storage and Insertion Kinetics in Anatase Titania. ACS Appl. Mater. Interfaces 2010, 2, 2091-2099. (19) Lee, G. H.; Lee, J. W.; Choi, J. I. L.; Kim, S. J.; Kim, Y.-H.; Kang, J. K. Ultrafast Discharge/Charge Rate and Robust Cycle Life for High-Performance Energy Storage Using Ultrafine Nanocrystals on the Binder-Free Porous Graphene Foam. Adv. Funct. Mater. 2016, 26, 5139-5148. (20) Liu, H.; Li, W.; Shen, D.; Zhao, D.; Wang, G. Graphitic Carbon Conformal Coating of Mesoporous TiO2 Hollow Spheres for High-Performance Lithium Ion Battery Anodes. J. Am. Chem. Soc. 2015, 137, 13161-13166. (21) Escamilla‐Pérez Angel, M.; Louvain, N.; Boury, B.; Brun, N.; Mutin, P. H. Ethers as Oxygen Donor and Carbon Source in Non‐Hydrolytic Sol–Gel: One‐Pot, Atom‐Economic Synthesis of Mesoporous TiO2–Carbon Nanocomposites. Chem. Eur. J. 2018, 24, 4982-4990. (22) Deng, X.; Wei, Z.; Cui, C.; Liu, Q.; Wang, C.; Ma, J. Oxygen-Deficient Anatase TiO2@C Nanospindles with Pseudocapacitive Contribution for Enhancing Lithium Storage. J. Mater. Chem. A 2018, 6, 4013-4022. (23) Chattopadhyay, S.; Maiti, S.; Das, I.; Mahanty, S.; De, G. Electrospun TiO2–rGO Composite Nanofibers with Ordered Mesopores by Molecular Level Assembly: A High Performance Anode Material for Lithium‐Ion Batteries. Adv. Mater. Interfaces 2016, 3, 1600761.

ACS Paragon Plus Environment

22

Page 23 of 36 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 Applied Materials & Interfaces

(24) Pol, V. G.; Kang, S.-H.; Calderon-Moreno, J. M.; Johnson, C. S.; Thackeray, M. M. Autogenic Reactions for Preparing Carbon-Encapsulated, Nanoparticulate TiO2 Electrodes for Lithium-Ion Batteries. J. Power Sources 2010, 195, 5039-5043. (25) Bresser, D.; Paillard, E.; Binetti, E.; Krueger, S.; Striccoli, M.; Winter, M.; Passerini, S. Percolating Networks of TiO2 Nanorods and Carbon for High Power Lithium Insertion Electrodes. J. Power Sources 2012, 206, 301-309. (26) Pinna, N.; Niederberger, M. Surfactant-Free Nonaqueous Synthesis of Metal Oxide Nanostructures. Angew. Chem., Int. Ed. 2008, 47, 5292-5304. (27) Tahir, M. N.; Oschmann, B.; Buchholz, D.; Dou, X.; Lieberwirth, I.; Panthöfer, M.; Tremel, W.; Zentel, R.; Passerini, S. Extraordinary Performance of Carbon-Coated Anatase TiO2 as Sodium-Ion Anode. Adv. Energy Mater. 2016, 6, 1501489. (28) Jiang, B.; Tang, Y.; Qu, Y.; Wang, J.-Q.; Xie, Y.; Tian, C.; Zhou, W.; Fu, H. Thin Carbon Layer Coated Ti3+-TiO2 Nanocrystallites for Visible-Light Driven Photocatalysis. Nanoscale 2015, 7, 5035-5045. (29) Bang Jin, H.; Helmich Richard, J.; Suslick Kenneth, S. Nanostructured ZnS:Ni2+ Photocatalysts Prepared by Ultrasonic Spray Pyrolysis. Adv. Mater. 2008, 20, 2599-2603. (30) Lee, L.; Kang, B.; Han, S.; Kim, H.-e.; Lee, M. D.; Bang, J. H. A Generalizable Top‐Down Nanostructuring Method of Bulk Oxides: Sequential Oxygen–Nitrogen Exchange Reaction. Small 2018, 14, 1801124. (31) Shimodaira, N.; Masui, A. Raman Spectroscopic Investigations of Activated Carbon Materials. J. Appl. Phys. 2002, 92, 902-909. (32) Ahn, S. H.; Lee, C. H.; Kim, M. S.; Kim, S. A.; Kang, B.; Kim, H.-e.; Lee, S. U.; Bang, J. H. Deciphering the Electrocatalytic Activity of Nitrogen-Doped Carbon Embedded with Cobalt

ACS Paragon Plus Environment

23

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 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 24 of 36

Nanoparticles and the Reaction Mechanism of Triiodide Reduction in Dye-Sensitized Solar Cells. J. Phys. Chem. C 2017, 121, 27332-27343. (33) Schito, A.; Romano, C.; Corrado, S.; Grigo, D.; Poe, B. Diagenetic Thermal Evolution of Organic Matter by Raman Spectroscopy. Org. Geochem. 2017, 106, 57-67. (34) Vidano, R.; Fischbach D, B. New Lines in the Raman Spectra of Carbons and Graphite. J. Am. Ceram. Soc. 1978, 61, 13-17. (35) Dresselhaus, M. S.; Jorio, A.; Souza Filho, A. G.; Saito, R. Defect Characterization in Graphene and Carbon Nanotubes Using Raman Spectroscopy. Philos. Trans. R. Soc., A 2010, 368, 5355. (36) Ferrari, A. C.; Basko, D. M. Raman Spectroscopy as a Versatile Tool for Studying the Properties of Graphene. Nat. Nanotechnol. 2013, 8, 235. (37) Han, J. H.; Bang, J. H. A Hollow Titanium Oxynitride Nanorod Array as an Electrode Substrate Prepared by the Hot Ammonia-Induced Kirkendall Effect. J. Mater. Chem. A 2014, 2, 10568-10576. (38) Lee, E. J.; Lee, L.; Abbas, M. A.; Bang, J. H. The Influence of Surface Area, Porous Structure, and Surface State on the Supercapacitor Performance of Titanium Oxynitride: Implications for a Nanostructuring Strategy. Phys. Chem. Chem. Phys. 2017, 19, 21140-21151. (39) Brumbarov, J.; Vivek, J. P.; Leonardi, S.; Valero-Vidal, C.; Portenkirchner, E.; KunzeLiebhauser, J. Oxygen Deficient, Carbon Coated Self-Organized TiO2 Nanotubes as Anode Material for Li-Ion Intercalation. J. Mater. Chem. A 2015, 3, 16469-16477. (40) Zhuang, J.; Tian, Q.; Liu, Q.; Liu, P.; Cui, X.; Li, Y.; Fan, M. New Insight into Binary TiO2@C Nanocomposites: The Crucial Effect of an Interfacial Microstructure. Phys. Chem. Chem. Phys. 2017, 19, 9519-9527.

ACS Paragon Plus Environment

24

Page 25 of 36 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 Applied Materials & Interfaces

(41) Favaro, M.; Leonardi, S.; Valero‐Vidal, C.; Nappini, S.; Hanzlik, M.; Agnoli, S.; Kunze‐ Liebhäuser, J.; Granozzi, G. In‐Situ Carbon Doping of TiO2 Nanotubes via Anodization in Graphene Oxide Quantum Dot Containing Electrolyte and Carburization to TiOxCy Nanotubes. Adv. Mater. Interfaces 2015, 2, 1400462. (42) Wu, Q.; Zheng, Q.; van de Krol, R. Creating Oxygen Vacancies as a Novel Strategy to Form Tetrahedrally Coordinated Ti4+ in Fe/TiO2 Nanoparticles. J. Phys. Chem. C 2012, 116, 7219-7226. (43) Setvín, M.; Aschauer, U.; Scheiber, P.; Li, Y.-F.; Hou, W.; Schmid, M.; Selloni, A.; Diebold, U. Reaction of O2 with Subsurface Oxygen Vacancies on TiO2 Anatase (101). Science 2013, 341, 988-991. (44) Lee, D. K.; Choi, J. I.; Lee, G. H.; Kim, Y.-H.; Kang, J. K. Energy States of a Core-Shell Metal Oxide Photocatalyst Enabling Visible Light Absorption and Utilization in Solar-to-Fuel Conversion of Carbon Dioxide. Adv. Energy Mater. 2016, 6, 1600583. (45) Lee, M.-J.; Lho, E.; Bai, P.; Chae, S.; Li, J.; Cho, J. Low-Temperature Carbon Coating of Nanosized Li1.015Al0.06Mn1.925O4 and High-Density Electrode for High-Power Li-Ion Batteries. Nano Lett. 2017, 17, 3744-3751. (46) Rabatic, B. M.; Dimitrijevic, N. M.; Cook, R. E.; Saponjic, Z. V.; Rajh, T. Spatially Confined Corner Defects Induce Chemical Functionality of TiO2 Nanorods. Adv. Mater. 2006, 18, 10331037. (47) Lafont, U.; Carta, D.; Mountjoy, G.; Chadwick, A. V.; Kelder, E. M. In Situ Structural Changes Upon Electrochemical Lithium Insertion in Nanosized Anatase TiO2. J. Phys. Chem. C 2009, 114, 1372-1378.

ACS Paragon Plus Environment

25

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 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 26 of 36

(48) Shin, J.-Y.; Samuelis, D.; Maier, J. Sustained Lithium-Storage Performance of Hierarchical, Nanoporous Anatase TiO2 at High Rates: Emphasis on Interfacial Storage Phenomena. Adv. Funct. Mater. 2011, 21, 3464-3472. (49) Sudant, G.; Baudrin, E.; Larcher, D.; Tarascon, J.-M. Electrochemical Lithium Reactivity with Nanotextured Anatase-Type TiO2. J. Mater. Chem. 2005, 15, 1263-1269. (50) Hardwick, L. J.; Holzapfel, M.; Novák, P.; Dupont, L.; Baudrin, E. Electrochemical Lithium Insertion into Anatase-Type TiO2: An in Situ Raman Microscopy Investigation. Electrochim. Acta 2007, 52, 5357-5367. (51) Cava, R. J.; Murphy, D. W.; Zahurak, S.; Santoro, A.; Roth, R. S. The Crystal Structures of the Lithium-Inserted Metal Oxides Li0.5TiO2 Anatase, LiTi2O4 Spinel, and Li2Ti2O4. J. Solid State Chem. 1984, 53, 64-75. (52) Bonino, F.; Busani, L.; Lazzari, M.; Manstretta, M.; Rivolta, B.; Scrosati, B. Anatase as a Cathode Material in Lithium—Organic Electrolyte Rechargeable Batteries. J. Power Sources 1981, 6, 261-270. (53) Wagemaker, M.; van de Krol, R.; Kentgens, A. P. M.; van Well, A. A.; Mulder, F. M. Two Phase Morphology Limits Lithium Diffusion in TiO2 (Anatase):  A 7Li MAS NMR Study. J. Am. Chem. Soc. 2001, 123, 11454-11461. (54) Mathew, V.; Gim, J.; Alfaruqi, M. H.; Kim, S.; Song, J.; Baboo, J. P.; Kim, S.; Park, S.; Kim, D.; Kim, J. A Porous TiO2 Electrode Prepared by an Energy Efficient Pyro-Synthesis for Advanced Lithium-Ion Batteries. J. Electrochem. Soc. 2015, 162, A1220-A1226. (55) Zachau-Christiansen, B.; West, K.; Jacobsen, T.; Atlung, S. Lithium Insertion in Different TiO2 Modifications. Solid State Ionics 1988, 28-30, 1176-1182.

ACS Paragon Plus Environment

26

Page 27 of 36 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 Applied Materials & Interfaces

(56) Macklin, W. J.; Neat, R. J. Performance of Titanium Dioxide-Based Cathodes in a Lithium Polymer Electrolyte Cell. Solid State Ionics 1992, 53-56, 694-700. (57) Belak, A. A.; Wang, Y.; Van der Ven, A. Kinetics of Anatase Electrodes: The Role of Ordering, Anisotropy, and Shape Memory Effects. Chem. Mater. 2012, 24, 2894-2898. (58) Hengerer, R.; Kavan, L.; Krtil, P.; Grätzel, M. Orientation Dependence of Charge‐Transfer Processes on TiO2 (Anatase) Single Crystals. J. Electrochem. Soc. 2000, 147, 1467-1472. (59) Froschl, T.; Hormann, U.; Kubiak, P.; Kucerova, G.; Pfanzelt, M.; Weiss, C. K.; Behm, R. J.; Husing, N.; Kaiser, U.; Landfester, K.; Wohlfahrt-Mehrens, M. High Surface Area Crystalline Titanium Dioxide: Potential and Limits in Electrochemical Energy Storage and Catalysis. Chem. Soc. Rev. 2012, 41, 5313-5360. (60) Zhu, Z.; Cheng, F.; Chen, J. Investigation of Effects of Carbon Coating on the Electrochemical Performance of Li4Ti5O12/C Nanocomposites. J. Mater. Chem. A 2013, 1, 9484-9490. (61) Guo, X.; Xiang, H. F.; Zhou, T. P.; Ju, X. K.; Wu, Y. C. Morphologies and Structures of Carbon Coated on Li4Ti5O12 and Their Effects on Lithium Storage Performance. Electrochim. Acta 2014, 130, 470-476. (62) Zhu, G.-N.; Wang, C.-X.; Xia, Y.-Y. A Comprehensive Study of Effects of Carbon Coating on Li4Ti5O12 Anode Material for Lithium-Ion Batteries. J. Electrochem. Soc. 2011, 158, A102A109. (63) Bresser, D.; Oschmann, B.; Tahir, M. N.; Tremel, W.; Zentel, R.; Passerini, S. Stabilizing Nanostructured Lithium Insertion Materials via Organic Hybridization: A Step Forward Towards High-Power Batteries. J. Power Sources 2014, 248, 852-860. (64) Saravanan, K.; Ananthanarayanan, K.; Balaya, P. Mesoporous TiO2 with High Packing Density for Superior Lithium Storage. Energy Environ. Sci. 2010, 3, 939-948.

ACS Paragon Plus Environment

27

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 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 28 of 36

(65) Shin, J.-Y.; Samuelis, D.; Maier, J. Defect Chemistry of Lithium Storage in TiO2 as a Function of Oxygen Stoichiometry. Solid State Ionics 2012, 225, 590-593. (66) Zheng, J.; Liu, Y.; Ji, G.; Zhang, P.; Cao, X.; Wang, B.; Zhang, C.; Zhou, X.; Zhu, Y.; Shi, D. Hydrogenated Oxygen-Deficient Blue Anatase as Anode for High-Performance Lithium Batteries. ACS Appl. Mater. Interfaces 2015, 7, 23431-23438. (67) Zheng, J.; Liu, L.; Ji, G.; Yang, Q.; Zheng, L.; Zhang, J. Hydrogenated Anatase TiO2 as Lithium-Ion Battery Anode: Size–Reactivity Correlation. ACS Appl. Mater. Interfaces 2016, 8, 20074-20081. (68) Madej, E.; La Mantia, F.; Schuhmann, W.; Ventosa, E. Impact of the Specific Surface Area on the Memory Effect in Li-Ion Batteries: The Case of Anatase TiO2. Adv. Energy Mater. 2014, 4, 1400829. (69) Ventosa, E.; Loffler, T.; La Mantia, F.; Schuhmann, W. Understanding Memory Effects in Li-Ion Batteries: Evidence of a Kinetic Origin in TiO2 Upon Hydrogen Annealing. Chem. Commun. 2016, 52, 11524-11526. (70) Acevedo-Peña, P.; Haro, M.; Rincón, M. E.; Bisquert, J.; Garcia-Belmonte, G. Facile Kinetics of Li-Ion Intake Causes Superior Rate Capability in Multiwalled Carbon Nanotube@TiO2 Nanocomposite Battery Anodes. J. Power Sources 2014, 268, 397-403.

ACS Paragon Plus Environment

28

Page 29 of 36 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 Applied Materials & Interfaces

(A)

Li+

Li+ Li+ Li+

e

e

Li+

Li+

Hierarchical Assembly of TiO2 Nanocrystals

Hierarchical Assembly of Carbon-Coated TiO2 Nanocrystals

(B) Carbon

Carbon Layer Titanium

TiO2/Carbon Interface

Oxygen Oxygen Vacancy

TiO2

TiO2

Figure 1. (A) Schematic of Li+/e- transport in the hierarchical assembly of TiO2 nanocrystals and the hierarchical assembly of carbon-coated TiO2 nanocrystals (blue arrows represent electron supply from carbon additive and red arrows, electron hopping through TiO2 particles). (B) Schematic of TiO2/carbon interface formed in the carbon-coated TiO2.

ACS Paragon Plus Environment

29

ACS Applied Materials & Interfaces

TiO2-as made Raw data Calculated data Diffrence Bragg positions

TiO2-500

Intensity (a.u.)

Raw data Calculated data Diffrence Bragg positions

TiO2-600 Raw data Calculated data Diffrence Bragg positions

TiO2-700 Raw data Calculated data Diffrence Bragg positions

140

120

100

80

60

40

20

2-Theta (Degree)

Figure 2. XRD patterns and Rietveld refined patterns of TiO2-as made, TiO2-500, TiO2-600, and TiO2-700.

-1

200

100

0.5

0.2 0.1 0

1

10

100

dp (nm)

50

Adsorption Desorption

0 0.2

0.4

0.6

0.8

100

0.4 0.3 0.2 0.1 0

1

Adsorption Desorption

0 0

0.2

-1

0.4

0.6

0.8

1.0

Relative Pressure (P/P0) 0.6

200

0.5

100

0.4 0.3 0.2 0.1 0

1

10

100

dp (nm)

50

Adsorption Desorption

0 0

0.2

0.4

0.6

0.8

1.0

150

dVp/dlogdp

150

dVp/dlogdp

3

0.5

100

dp (nm)

(D)

0.6

10

50

1.0

Relative Pressure (P/P0) 200

150

dVp/dlogdp

0.3

Volume Adsorbed (cm g )

dVp/dlogdp

150

0.4

(C) Volume Adsorbed (cm g )

0.6

200

3

0.5

0

-1

Volume Adsorbed (cm g )

(B) 0.6

3

-1

Volume Adsorbed (cm g )

(A)

3

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 30 of 36

100

0.4 0.3 0.2 0.1 0

1

10

100

dp (nm)

50

Adsorption Desorption

0 0

Relative Pressure (P/P0)

0.2

0.4

0.6

0.8

1.0

Relative Pressure (P/P0)

Figure 3. N2 physisorption isotherms of (A) TiO2-as made, (B) TiO2-500, (C) TiO2-600, and (D) TiO2-700. Insets show the pore size distributions.

ACS Paragon Plus Environment

30

Page 31 of 36

800

Raw Data D' G1 G2 D1 D2 D3 S Fitting Line Background

1000

1200

TiO2- 500

1400

Raman Shift Shift Raman

800

1600

1800

1200

800

Raman Shift Shift Raman

1600

1200

TiO2- 600

1400

Raman Shift Raman TiO2- 700

1400

1000

(cm (cm-1))

Raw Data D' G1 G2 D1 D2 D3 S Fitting Line Background

1000

2000

Raw Data D' G1 G2 D1 D2 D3 S Fitting Line Background

-1

(D) Relative Amount Amount (% (%)) Relative

(C)

(B) Intensity (a.u.) Intensity (a.u.)

Intensity (a.u.) Intensity (a.u.)

(A)

Intensity (a.u.) Intensity (a.u.)

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 Applied Materials & Interfaces

1800

2000

1600

1800

2000

-1

(cm (cm-1)) D' G1 G2 D1 D2 D3 S

30 25 20 15 10 5 0

500

-1

(cm (cm-1))

600

700

Temperature ((oC) ) Temperature ℃

Figure 4. Raman spectra of (A) TiO2-500, (B) TiO2-600, and (C) TiO2-700. (D) Relative amount of each deconvoluted peak (G1: E2g mode associated with the in-plane stretching vibration from graphite-like basal planes, G2: E2g mode associated with the in-plane stretching vibration from the basal planes of disordered carbon, D1: A1g mode associated with the in-plane breathing vibration from the basal planes of disordered carbon, D2 and D3: A1g mode associated with the in-plane breathing vibration from the basal planes of disordered carbon, D′: signal from the disordered edges of carbon, S: signal associated with various sources of disorderness.

ACS Paragon Plus Environment

31

ACS Applied Materials & Interfaces

Ti 2p TiO2-pristine

Ti4+ 2p1/2 464.40 eV

Ti4+ 2p3/2 458.70 eV

2p1/2 464.46 eV

2p1/2 462.56 eV

Ti3+ 2p3/2 457.36 eV

5.7eV

Ti4+ 2p1/2 464.45 eV

Ti4+ 2p3/2 458.75 eV

5.2eV Ti3+ 2p1/2 462.55 eV

Ti3+ 2p3/2 457.35 eV

5.7eV

TiO2-700 2p1/2 464.41 eV

-OH 531.32 eV

464

462

460

Ti-O 529.52 eV

Ov 530.32 eV

O2527.77 eV

TiO2-600 -OH 531.31 eV

5.2eV

-OH 531.33 eV

Ti3+ 2p1/2 462.51 eV

466

Ti-O 529.63 eV

Ti-O 529.51 eV

Ov 530.32 eV

O2527.71 eV

TiO2-700

Ti4+ 2p3/2 458.72 eV

Ti4+

TiO2-pristine -OH 531.53 eV

Ti4+ 2p3/2 458.76 eV

Ti3+

TiO2-600

O 1s

TiO2-500

5.2eV

Ti4+

468

(B)

5.7eV

TiO2-500

Intensity (a.u.)

5.7eV

Intensity (a.u.)

(A)

Ti-O 529.56 eV

Ov 530.36 eV

O2527.90 eV

Ti3+ 2p3/2 457.31 eV

458

456

454

452

536

534

Binding Energy (eV)

532

530

528

526

524

Binding Energy (eV)

Figure 5. XPS spectra of TiO2-as made, TiO2-500, TiO2-600, and TiO2-700: (A) Ti 2p spectra and (B) O 1s spectra.

TiO2-700

Intensity (a.u.)

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 32 of 36

TiO2-600

TiO2-500 TiO2- as made

1.90

1.95

2.00

2.05

2.10

g value

Figure 6. EPR spectra of TiO2-as made, TiO2-500, TiO2-600, and TiO2-700 measured at 110 K.

ACS Paragon Plus Environment

32

Page 33 of 36 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 Applied Materials & Interfaces

Figure 7. TEM images of (A) TiO2-500, (B) TiO2-600, and (C) TiO2-700 (Insets are SEM images). HRTEM images of the (101) lattice plane of (D) TiO2-500, (E) TiO2-600, and (F) TiO2-700 taken along the [1̅1̅1] zone axis, FFT patterns, and anatase TiO2 crystal structure viewed along the [1̅1̅1] direction showing the Ti–Ti distance (d). Lattice distance along the yellow arrow ([101] direction) in the HRTEM images: (G) TiO2-500, (H) TiO2-600, and (I) TiO2-700.

ACS Paragon Plus Environment

33

ACS Applied Materials & Interfaces

(C)

(A)

3.0

TiO2-700 TiO2-600 TiO2-500 TiO2-as made

2.5

2.0

1.5

+

1.0 0

40

80

120

160

200

240

-1

Specific Capacity (mAhg )

(B)

TiO2-700 TiO2-600 TiO2-500 TiO2-as made

-1

Specific Capacity (mAhg )

300 0.5 C

250

1C 2C

200

0.5 C

3C 5C 10 C

150

30 C

100

50 C

50 0

0

10

20

30

40

50

Region C

Region C

0.8 0

Region C

P1

P2

P1

-2.4

Region C

0

40

80

(E)

4 3 2 1 0

2.0

2.5

3.0 1.5

(G) 10 -1

-1.5 -3.0 -4.5 2.5

1.5

2.0

2.5

3.0

+

0

2.0

P2

P2 1.0

8

1.5

TiO2-500

P1

Voltage (V vs. Li /Li) TiO2-600

0.1 mV 0.5 mV 1 mV 2 mV 3 mV 5 mV 10 mV

1.0

240

P1

-3

3.0

Current Density (Ag )

-1

4.5

200

-2

+

6.0

160

-1

Voltage (V vs. Li /Li)

(F)

120

0.1 mV 0.5 mV 1 mV 2 mV 3 mV 5 mV 10 mV

-4 1.5

A+B: 150.40 mAhg -1 C: 74.77 mAhg-1

Region B

P2 1.0

TiO2-700

Region A

-1

-0.8 -1.6

A+B: 110.32 mAhg -1 C: 76.34 mAhg -1

Region B

-1

1.6

TiO2-600

Region A

2.5 2.0 1.5 1.0

Specific Capacity (mAhg ) TiO2-as made

0.1 mV 0.5 mV 1 mV 2 mV 3 mV 5 mV 10 mV

A+B: 90.54 mAhg -1 C: 78.13 mAhg-1

Region B

Current Density (Ag )

-1

Current Density (Ag )

2.4

TiO2-500

Region A

2.5 2.0 1.5 1.0 3.0

Cycle Number

(D)

A+B: 89.46 mAhg -1 C: 70.86 mAhg-1

Region B

3.0 2.5 2.0 1.5 1.0

20 C

TiO2-as made

Region A

2.5 2.0 1.5 1.0 3.0

Voltage (V vs. Li /Li)

+

Voltage (V vs. Li /Li)

3.0

Current Density (Ag )

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 34 of 36

3.0

6 4 2

TiO2-700

0.1 mV 0.5 mV 1 mV 2 mV 3 mV 5 mV 10 mV

0 -2 -4 -6 -8

1.0

+

Voltage (V vs. Li /Li)

1.5

2.0

2.5

3.0

+

Voltage (V vs. Li /Li)

Figure 8. (A) Galvanostatic charge/discharge voltage profiles of TiO2-as made, TiO2-500, TiO2600, and TiO2-700 measured at 1 C and (B) corresponding rate capabilities measured at various C

ACS Paragon Plus Environment

34

Page 35 of 36

rates. (C) Discharge curves of TiO2-as made, TiO2-500, TiO2-600, and TiO2-700 measured at 1 C with relative capacities measured at different voltage regions: Region A (E > 1.75 V), Region B (E ≈ 1.75 V), Region C (E < 1.75 V). Cyclic voltammograms of (D) TiO2-as made, (E) TiO2-500, (F) TiO2-600, and (G) TiO2-700 measured at various scan rates.

(A) Nyquist plots during discharge (2.55 V to 1.53 V)

0.4

TiO2-as made TiO2-500

0.2

0.6 0.4

TiO2-as made TiO2-500

0.2

TiO2-600

0

TiO2-as made

0.5

TiO2-700

0.6 0.4

TiO2-as made

1.78 V

TiO2-600

TiO2-600

0.4

TiO2-700

0.3 0.2

0

0

0

0.2 0.4 0.6 0.8 1.0

Z (g)

0

0

0.2 0.4 0.6 0.8 1.0

Z (g)

TiO2-700

0.3 0.2 0.1

0.1

TiO2-700

0.2 0.4 0.6 0.8 1.0

1.53 V

TiO2-500

0.4

TiO2-500

TiO2-600

0.2

TiO2-600

TiO2-700

2.03 V

TiO2-500

0.8

-Z (g)

0.6

0

TiO2-as made

0.8

-Z (g)

-Z (g)

0.8

0.5

0.6

1.0 2.35 V

-Z (g)

1.0 2.55 V

-Z (g)

1.0

0

0.15 0.30 0.45 0.60

Z (g)

0

0

0.1 0.2 0.3 0.4 0.5

Z (g)

Z (g)

(B) Nyquist plots during charge (1.48 V to 2.55 V)

TiO2-500

TiO2-500

TiO2-500

0.30 0.15

TiO2-700

0.30 0.15

0

0.15 0.30 0.45 0.60

0

0

10

(D)

10

1

10

0

10

TiO2-as made

2.4

0

-1

TiO2-as made

1.6

Potential (V)

1.2

10

-2

2.8

TiO2-as made TiO2-500

0

10

2.4

2.0

10

TiO2-700

0.3 0.6 0.9 1.2 1.5

TiO2-as made TiO2-500

0

TiO2-600 TiO2-700

0

(F)

10

1

10

0

10

TiO2-as made

0.3

0.6

0.9

-1

TiO2-as made TiO2-500

TiO2-600

TiO2-600

TiO2-600

TiO2-700

TiO2-700

1.6

Potential (V)

1.2

1.2

1.6

2.0

Potential (V)

2.4

1.2

Z (g)

Z (g)

0

-1

0.4 0.2

TiO2-600

0

0.6

TiO2-500

TiO2-500

2.0

0.4

0.15 0.30 0.45 0.60

(E)

TiO2-600

2.8

0.6

Z (g)

TiO2-500 TiO2-700

0.8

0.2

Z (g)

0

-1

0

0.15 0.30 0.45 0.60

RLi (  g)

10

TiO2-700

0.15

Z (g)

(C)

TiO2-600

0.30

Rct (  g)

0

0.45

TiO2-600

-Z (g)

TiO2-700

2.55 V

0.8

-Z (g)

0.45

TiO2-600

1.0

2.10 V

1.95 V TiO2-as made

-Z (g)

-Z (g)

0.45

1.0

0.60 1.71 V TiO2-as made

-Z (g)

0.60 1.48 V TiO2-as made

RLi (  g)

0.60

Rct (  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 Applied Materials & Interfaces

10

TiO2-700

-2

1.2

1.6

2.0

2.4

Potential (V)

Figure 9. In situ EIS measurement results. (A) Nyquist plots of TiO2-as made, TiO2-500, TiO2600, and TiO2-700 measured during discharge at various potentials (2.55, 2.35, 2.03, 1.78, 1.53 V) (B) Nyquist plots of TiO2-as made, TiO2-500, TiO2-600, and TiO2-700 measured during charge at various potentials (1.48, 1.71, 1.95, 2.10, 2.55 V). Various resistances obtained from the fitting at different voltages: (C) Rct and (D) RLi (during the discharge) and (E) Rct and (F) RLi (during the charge). All potentials in the plots are with respect to the standard potential of Li+/Li).

ACS Paragon Plus Environment

35

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 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 36 of 36

Table 1. Lattice Parameters, Grain Size, Surface Area, Total Pore Volume, and ID/IG Ratio of TiO2-as made, TiO2-500, TiO2-600, and TiO2-700. Lattice Parameter (Å)

Grain Size

Total Pore Volume

a

c

(nm)a)

Surface Area (m2g-1)b)

TiO2-as made

3.7868

9.4974

9.9

70.93

0.198

-

TiO2-500

3.7873

9.5077

13.0

93.78

0.295

1.97

TiO2-600

3.7867

9.5067

13.6

88.69

0.290

1.35

TiO2-700

3.7865

9.5066

14.5

83.69

0.287

1.06

IG2/IG1 c)

3 -1

(cm g )

a)

Grain size was determined using the Scherrer equation; b) Surface area was determined using the BET method; c) Ratio of G2 band-to-G1 band peak intensities in the Raman spectra.

Table of Contents Graphic Li+

Carbon

TiO2/Carbon Interface

Li+ e

TiO2

Titanium Oxygen Oxygen Vacancy

Li+

ACS Paragon Plus Environment

36