Você está na página 1de 9

Electrochimica Acta 210 (2016) 206–214

Contents lists available at ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

Porous TiO2 urchins for high performance Li-ion battery electrode:


facile synthesis, characterization and structural evolution
Yi Caia , Hong-En Wanga,* , Shao-Zhuan Huanga , Muk Fung Yuenb , Heng-Hui Caia ,
Chao Wanga , Yong Yua , Yu Lia , Wen-Jun Zhangb,**, Bao-Lian Sua,c,***
a
State Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, 122 Luoshi Road, 430070, Wuhan,
Hubei, China
b
Department of Physics and Materials Science and Center Of Super-Diamond and Advanced Films (COSDAF), City University of Hong Kong, Hong Kong, China
c
Laboratory of Inorganic Materials Chemistry (CMI), University of Namur, 61 rue de Bruxelles, B-5500 Namur, Belgium

A R T I C L E I N F O A B S T R A C T

Article history:
Received 30 March 2016 Porous TiO2 urchins have been synthesized by a hydrothermal route using TiO2/oleylamine as precursors
Received in revised form 17 May 2016 with subsequent ion-exchange and calcination. The resultant material consists of porous spherical cores
Accepted 20 May 2016 and nanochains-constructed shells with straight channels. Electrochemical measurements indicate the
Available online 21 May 2016 TiO2 urchins deliver superior lithium storage capability in terms of high capacity (206.2 mA h g1 at 0.5 C),
superior rate performance (94.4 mA h g1 at 20 C) and stable cycling stability (94.3% capacity retention
Keywords: over 1000 cycles at 10 C versus the third cycle). Such performance enhancement is mainly due to the
TiO2 increased electrode/electrolyte contact interface, reduced Li+ diffusion pathways and improved mass
porous structure
transfer of electrolyte in the unique 3D interconnected hierarchical network. In addition, ex-situ XRD,
anode materials
SEM and TEM analyses further reveal high structure integrity of the porous TiO2 urchins during the
Li ion battery
electrochemistry electrochemical lithiation, leading to enhanced lithium storage stability. Moreover, we detected that
some anatase nanocrystals evolved into electrochemically inactive Li1TiO2 dots (10 nm in size) during
long-term electrochemical cycling. Our findings provide more insights for better understanding of the
structure evolution and capacity decay mechanism in porous TiO2 nanostructures.
ã 2016 Elsevier Ltd. All rights reserved.

1. Introduction small volume changes upon Li+ uptake/removal, and absence of


solid-electrolyte interphase layer for better operation safety [3–5].
Lithium ion batteries (LIBs) have played a crucial role in The electrochemical performance of TiO2 strongly relies on its
powering portable electronic devices in the past two decades [1]. polymorphs [6,7], crystallinity [8], grain size [9], morphology
They are now being further improved for emerging energy-storage [10,11], porosity and composition [12,13], which can determine the
applications in (hybrid, plug-in) electric vehicles, smart grids and Li+ diffusion and electron transport. In particular, much work has
intermittent renewable energies from solar and wind. Despite their been devoted to the controllable synthesis and electrochemical
commercial success, improvements in power, cost, durability and characterization of various TiO2 nanostructures with tunable
safety are still highly desired [2]. Among various electrode morphologies [14–16] and compositions [17,18], either in pure
materials, TiO2 has attracted considerable interest as alternative phases or mixed polymorphs [19–25], as well as nanocomposites
anode because of its high abundance, nontoxicity, high activity, hybridized with carbon nanotubes [26], and reduced graphene
oxides [6,27–29]. Among these, urchin-like TiO2 has gained much
attention due to its unique structure and electrochemical
performance [30–32]. For example, Park et al. synthesized
* Corresponding author. Tel.: +86 27 87855322; fax: +86 27 87879468.
urchin-like rutile TiO2 submicron spheres composed of single-
** Corresponding author. Tel.: +852 34427433; fax: +852 34420538. crystalline nanorods. After coating with amorphous carbon layer, it
*** Corresponding author at: Laboratory of Inorganic Materials Chemistry (CMI), could deliver a reversible capacity of 165.7 mA h g1 after 100
University of Namur, 61 rue de Bruxelles, B-5500 Namur, Belgium. cycles at 0.2 C [33]. Meanwhile, porous or hollow TiO2 nano-
Tel.: +32 81 724531; fax: +32 81 725414.
structures [34–43] can offer more active sites for Li+ storage due to
E-mail addresses: hongenwang@whut.edu.cn, hongen.wang@gmail.com
(H.-E. Wang), apwjzh@cityu.edu.hk (W.-J. Zhang), bao-lian.su@unamur.be their high surface area, good accommodation of volume change
(B.-L. Su). during (de)lithiation and shortened pathways for fast Li+ and

http://dx.doi.org/10.1016/j.electacta.2016.05.140
0013-4686/ã 2016 Elsevier Ltd. All rights reserved.
Y. Cai et al. / Electrochimica Acta 210 (2016) 206–214 207

electron transport, leading to enhanced electrochemical perfor- 2. Experimental


mance. Porous TiO2 nanostructures with well-defined morphol-
ogies and pore structures can be prepared by template-directing 2.1. Materials synthesis
approaches. For example, TiO2 nanocages with uniform shells and
tunable sizes have been fabricated by using Cu2O polyhedra as All the chemicals were analytically pure grade and used as
sacrifice templates [44]; however, conventional template-based received without further purification.
synthetic routes may require the removal of hard- or soft- Amorphous TiO2/oleylamine spheres were synthesized refer-
templates in the following steps, leading to increased synthetic ring to literature work [46] with little modification (oleylamine
costs and environmental concerns. was used as surfactant) and served as precursor for following
In addition, recent study discloses mesoporous TiO2 material experiments.
can undergo evident structure degradation and capacity loss upon Titanate urchins were prepared by hydrothermal reaction of
electrochemical cycling in opposition to what is expected for a long TiO2/oleylamine precursor in an alkaline solution. In a typical
time [45]. The structure change has been primarily ascribed to procedure, 0.2 g of precursor was dispersed in 40 mL of H2O and
irreversible lattice swelling/shrinkage caused by Li+ concentration 20 mL of ethanol containing 12 g of NaOH. Then, the suspension
gradient in porous framework during initial (de)lithiation. was transferred into an autoclave (with a PTFE liner, 100 mL
Therefore, design of rational pore structure and deeper under- capacity) and heated at 160  C for 12 h. The resultant white
standing of structure evolution of porous TiO2 electrode during precipitates were rinsed with ethanol and distilled water, and
cycling are very important and highly desired. dried at 80  C in air for 12 h. Protonated titanate (H-titanate)
In this work, we report the designed synthesis of porous TiO2 sample was obtained by soaking the titanate sample in 0.12 M of
urchins constructed from nanochains using TiO2/oleylamine HCl solution for 2 h to exchange the interlayer Na+ with H+ and then
composite as precursor via an in-situ self-sacrificing template thoroughly rinsed with deionized water and dried at 80  C for 12 h
method. The resultant TiO2 material exhibits superior electro- in air.
chemical performance as anode for LIBs. Post-mortem character- Porous TiO2 urchins were obtained after thermal treatment of
izations validate high structure stability of the TiO2 material during the H-titanate sample at 450  C for 2 h in air with a heating rate of
cycling. More importantly, we observe the evolution of trace 10- 2  C min1.
nm Li1TiO2 crystalline dots that are randomly distributed within
the TiO2 host after long-term cycling. Our findings can provide 2.2. Material characterizations
more insights into the deeper understanding of the high Li+ storage
capability of porous TiO2 nanostructures as well as their capacity Crystal structures of the samples were analyzed by powder X-
decay mechanism upon cycling. ray diffraction (XRD) patterns recorded on a Bruker Diffractometer
operated at 40 kV and 40 mA with Cu Ka radiation. Rietveld

Fig. 1. Schematic illustration of the fabrication process of the porous TiO2 urchins. (a) TiO2/oleylamine composites, (b) out-diffusion of oleylamine from precursor and partial
conversion of TiO2 nuclei into titanate fragments during hydrothermal process, (c) titanate nanosheets-assembled flowers, (d) titanate nanotubes-constructed urchins, (e)
porous TiO2 urchins, (f) magnified portion of a TiO2 nanochain constructed from connected nanocrystals.
208 Y. Cai et al. / Electrochimica Acta 210 (2016) 206–214

refinement of XRD pattern has been performed using Maud 0.01 Hz at an amplitude of 10 mV. All the electrochemical tests
program [47]. Morphologies and element composition data of the were performed at room temperature.
samples were obtained by a Hitachi S4800 scanning electron
microscope (SEM) equipped with an energy dispersive X-ray (EDX) 3. Results and discussion
analyzer. Transmission electron microscopy (TEM) and high
resolution transmission electron microscopy (HRTEM) images Fig. 1 schematically illustrates the fabrication process of the
were acquired on a JEOL JEM-2100F electron microscope operated porous TiO2 urchins using TiO2/oleylamine composite as precursor.
at an acceleration voltage of 200 kV. Fourier transform infrared The composite precursor is prepared by a modified sol-gel route
spectroscopy (FT-IR) was performed with a Bruker Vertex 80 V FT- [46] and consists of amorphous spheres with a mean size of
IR spectrometer. Nitrogen sorption experiments were performed 500 nm (Fig. S1). According to the previous report, the TiO2 nuclei
using a Micrometrics, TriStar II 3020 system operated at 77 K. Prior and oleylamine molecules hybridize together uniformly by
to the adsorption experiments, the samples were degassed at hydrogen-bonding interaction between TiO2 and NH2 groups
200  C for 12 h. (Fig. 1a and S2a) [46].
Next, the composite precursor can be converted into titanate
2.3. Electrochemical Measurements intermediates by hydrothermal reaction (Fig. 1b–d). Hydrothermal
treatment accelerates the out-diffusion and dissolution of oleyl-
Electrochemical properties of the as-synthesized porous TiO2 amine molecules in the ethanol, leading to the weakening of
urchins and reference Degussa P25 nanoparticles were measured hydrogen-bonding attraction between TiO2 and oleylamine (Fig. 1b
using CR2025 Li-half cells. The working electrode was prepared by and S2b). Meantime, the strong interaction between exposed TiO2
evenly spreading a slurry of 70 wt% TiO2, 20 wt% acetylene black colloids and hydroxyl ions (OH) disrupts some specific Ti-O-Ti
and 10 wt% polyvinylidenedifluoride binder onto a Cu current bonds and generates new Ti-O-H bonds [48], leading to the
collector. In both cases, a Li metal foil was used as the counter formation of titanate fragments (Fig. 1b). XRD patterns show the
electrode and reference electrode. 1 M LiPF6 dissolved in ethylene titanate crystallites can be formed after a short reaction time. As
carbonate/dimethyl carbonate (1:1 w/w) was used as the working shown in Fig. 2a-1, the titanate product obtained after 1 h has a
electrolyte, and Celgard 2400 polypropylene film served as the body-centered orthorhombic Na2Ti2O5H2O structure consisting of
separator. The coin cells were assembled in a glove-box filled with 2D edge-shared [TiO6] octahedra sheets [49]. The titanate product
ultrapure Ar gas with oxygen and water contents below 1 ppm. The obtained after 12 h retains the same phase structure with an
mass loading of the active material is about 11.5 mg. Galvano- enhanced crystallinity, characterized by the sharper diffraction
static discharge/charge experiments were performed on a peaks (Fig. 2a-2). In addition, the diffraction peak of (200) plane
multichannel battery testing system (LAND CT2001A) over a shifts slightly to a higher degree due to the contraction of interlayer
potential range of 31 V vs. Li/Li+. Cyclic Voltammetry (CV) spacing. SEM characterization reveals the titanate samples
measurements were recorded on a CHI 604E electrochemical undergo a series of morphology evolution during the hydrothermal
workstation at varying scan rates from 0.1 to 1 mV s1 within a process. As shown in Fig. 2b, the sample prepared after 0.5 h is
potential window of 31 V (vs. Li/Li+). Electrochemical impedance characterized by spherical particles consisting of thin laminar
spectra (EIS) were recorded on an electrochemical workstation sheets with a thickness of 3 nm and 30 nm lateral sizes. After
(Autolab PGSTAT302N) in the frequency range from 100 kHz to reaction for 1 h, the titanate sample is composed of nanotubes with

Fig. 2. Microstructure characterizations of the titanate samples obtained by hydrothermal reaction for different time. (a) XRD patterns of the titanate samples prepared for 1 h
(1), 12 h (2), and H-titanate sample via ion-exchange (3); (b–d) SEM images of titanate samples prepared for 0.5 h, 1 h, and 12 h, respectively; (e) TEM image of the titanate
sample prepared for 12 h; (f) SEM image of the H-titanate sample after ion-exchange.
Y. Cai et al. / Electrochimica Acta 210 (2016) 206–214 209

Fig. 3. (a) XRD pattern, (b) SEM image, (c) TEM and (d) HRTEM micrographs of the as-obtained TiO2 material. The insets on the top right and bottom right in (d) are the FFT
patterns of an anatase grain and TiO2-B grain labeled by the large and small dashed squares, respectively.

diameters of 12 nm and lengths of 100 nm (Fig. 2c). The shape The titanates can be readily converted into porous TiO2 urchins
evolution from laminar sheets to hollow tubes adheres to the “self- via ion-exchange and annealing. After ion-exchange, the resultant
curling” mechanism driven by reducing total system energy via protonated titanate (H-titanate) sample has an orthorhombic
surface energy minimization and elimination of crystal defects H2Ti2O5H2O phase (JCPDS Card No. 47-0124) with similar peak
[48]. After 12 h of reaction, the resultant porous titanate urchins positions to its parent titanate (Fig. 2a-3) [49]. The intensity change
are constructed by longer nanotubes with 8 nm diameters, of (200), (110) and (310) diffraction peaks is mainly due to the
2 nm wall thickness and lengths up to 2 mm (Fig. 2d, e). The displacement of Na+ by H+ [49]. SEM image (Fig. 2f) reveals the
similar tube diameters of the two samples prepared after 1 h and obtained H-titanate sample retains an urchin-like morphology.
12 h hint the titanate nanotubes preferentially grow larger by TEM and high-magnification TEM micrographs (Fig. S5) further
oriented-attachment of adjacent shorter ones between two tip confirm the as-obtained H2Ti2O5H2O product consists of urchin-
ends. Moreover, the EDX pattern (Fig. S3) further confirms the like porous spheres constructed by interlinked nanotubes with
existence of Na, Ti, O elements in the titanate sample. diameters of ca. 8 nm. The EDX spectrum (Fig. S6) further evidences
We note that the ethanol/water solvent and TiO2/oleylamine the Na+ ions have been virtually replaced by H+ ions. After
composite precursor play an important role in the formation of annealing, the resultant TiO2 sample (Fig. 3a) has a tetragonal
such unique titanate nanostructures. Without ethanol, tiny anatase structure with a space group of I41/amd (JCPDS Card No.
titanate nanosheets-assembled flowers are obtained in our 21-1272) [51]. The sharp diffraction peaks indicate the high
hydrothermal experiments (Fig. S4a). This result is possibly crystallinity of the TiO2 product. Rietveld refinement of the XRD
ascribed to the adsorption of oleylamine on the surfaces of the pattern was further performed and the obtained lattice parameter
laminar titanate sheets, which prevents their further curling into values are summarized in Table 1. The refined lattice parameters
nanotubes due to a steric hindrance effect [50]. In contrast, the are in good agreement with the published results [52]. The
ethanol in the solvent can dynamically dissolve the oleylamine refinement also suggests trace TiO2-B (0.49%) may coexist in the
molecules during hydrothermal process, facilitating the structure TiO2 host. SEM result (Fig. 3b) depicts the TiO2 product comprises
evolution of titanate fragments. In addition, random titanate 1D porous urchins with uniform particle sizes of 500 nm. These
nanostructures with irregular particle sizes are formed if urchins tend to densely pack together with a higher tap density in
commercial P25 nanoparticles are used as the starting precursor comparison to commercial P25 nanoparticles (Fig. S7). TEM image
(Fig. S4b). It is speculated that the oleylamine in the precursor can in Fig. 3c discloses the TiO2 urchins are constructed from a large
slow down the reaction rate between TiO2 and OH, featuring the
formation of uniform hierarchical nanostructures. Nonetheless, Table 1
our synthetic strategy herein provides a general approach for the Refined unit cell parameters of the as-obtained TiO2 material.
facile fabrication of an array of urchin-like titanate and TiO2 Phases Space group a (nm) b (nm) c (nm) Content (wt%) Rw (%)
nanostructures with tunable configurations and polymorphs by
anatase I41/amd 0.37807 0.37807 0.95013 99.51 8.98
further optimizing experimental parameters. TiO2-B R-3c: H 0.52406 0.52406 1.35806 0.49 8.98
210 Y. Cai et al. / Electrochimica Acta 210 (2016) 206–214

reversible capacity of 170 mA h g1 [3]. Porous TiO2 nano-


structures are expected to possess higher Li+ storage capacity
due to their large contact interface with electrolyte and increased
pseudocapacitive Li+ storage at grain surfaces.
Cyclic voltammograms (CVs) were firstly recorded to study the
reaction kinetics and reversibility of the TiO2 electrode. From
Fig. 5a, a pair of dominant redox peaks (A-peaks) can be observed
at ca. 1.7 V (cathodic sweep) and 2.1 V (anodic sweep) in the first
sweep for the TiO2 electrode. The small overpotential of 0.4 V
signifies the good reaction kinetics of the TiO2 electrode. In the
second cycle, the cathodic peak shifts to 1.67 V while the anodic
peak stays almost unaltered due to an activation process. The third
CV curve almost overlaps with the second one, suggesting the high
reversibility of the electrode in the following sweeps. In addition,
Fig. 4. Nitrogen sorption isotherms and pore size distribution plot (inset) of the two pairs of extra redox peaks (S1-, S2-peaks) are observed at
porous TiO2 urchins.
1.50/1.55 V (cathodic sweep) and 1.55/1.62 V (anodic sweep),
respectively, which can be attributed to the characteristic peaks of
number of interwoven nanochains, forming porous spherical cores trace TiO2-B phase [19]. The CV curves were further recorded at
and hairy shells. Each nanochain is comprised of connected 5 nm varying scan rates from 0.1 to 1 mV s1 (Fig. S8a), showing an
nanocrystals. The morphology transformation from titanate nano- increased potential separation of redox peaks due to the
tubes to TiO2 nanocrystals adheres to the local structure incremental polarization. In addition, the anodic peak currents
rearrangement of [TiO6] octahedra. The nanotubes with small of anatase (A-peaks) scale with the square root of scan rates due to
diameters can hardly stand the internal strains and become a diffusion-controlled Li+ insertion process (Fig. S8b) [51]. In
disconnected during structure reorganization [53]. HRTEM micro- contrast, the peak currents of TiO2-B (S1-, S2-peaks) linearly scale
graph shown in Fig. 3d further depicts a polyhedral anatase with the first power of scan rates, which is typical of pseudoca-
nanocrystal enclosed by (101) and (001) planes (labelled by dashed pacitive Li+ storage process (Fig. S8c) [37,55].
square in the middle region and FFT pattern in the inset on top Fig. 5b presents the first two discharge-charge curves of the
right). In addition, one TiO2-B (JCPDS card No. 74-1940) [37] grain porous TiO2 urchins at 0.5 C (1 C = 170 mA g1). Well-defined
is also observed by the HRTEM (labelled by dashed square on the voltage plateaus can be observed at 1.7 V and 1.9 V during
left bottom) and corresponding FFT pattern (inset on the right discharge and charge, which are in accordance with the CV
bottom). The coexistence of trace TiO2-B distributed in the anatase analyses. The Li+ insertion in anatase TiO2 can be divided into three
host is believed to improve the Li+ storage capability with different stages (regions I, II, and III) [56]. The first stage (region I)
enhanced electron/Li+ transfer capability at the interface of the shows a monotonous potential drop from open-circuit potential to
two-phase boundaries [19]. a plateau of 1.7 V, relating to a pseudocapacitive surface effect or
The unique porous network might endow the resultant TiO2 solid-solution insertion mechanism [57]. The second stage (region
material with a high surface area. To this end, nitrogen adsorption- II) is characterized by a horizontal plateau region, reflecting the
desorption experiments of the porous TiO2 urchins were mea- process of Li+ insertion into the vacant octahedral sites of TiO2
sured. As depicted in Fig. 4, the isotherms can be identified as type lattice. During this process, a two-phase reaction occurs with
IV with two hysteresis loops. The first loop located at relative equilibrium between tetragonal Li0.01TiO2 and orthorhombic
pressures of 0.650.85 is related to the capillary condensation in Li0.5TiO2 [3]. The third stage (region III) shows a slope potential
the mesopores while the second loop located at relative pressure of drop linking to the Li+ storage at the surface layer of TiO2 grains
0.91.0 indicates the presence of macropores among adjacent [56]. The very long slope region of region III suggests an increased
spheres. The sample yields a high specific surface area of surface Li+ storage of the porous TiO2 urchins due to their high
152 m2 g1 based on the Brunauer-Emmett-Teller (BET) model. surface area and small crystallite size. The initial discharge
Pore size distribution centers at ca. 7 nm based on the Barrett- capacity of the porous TiO2 urchins is 263.4 mA h g1 and the
Joyner-Halenda (BJH) model from the desorption branch, which subsequent charge capacity is 196.6 mA h g1, leading to a
apparently agrees with the TEM analysis (Fig. 3c). coulombic efficiency (CE) of 75%. The initial irreversible capacity
Hierarchically porous nanostructures with a high surface area loss is caused by insertion of Li+ into sites not extractable and
and pore structures at different length scales may have promising decomposition of trace moisture adsorbed on the electrode surface
applications in energy storage and conversion, as well as [52]. The CE value is increased further to 94.6% in the second cycle,
environment fields [54]. As a proof-of-concept, we next evaluate indicating the fast balance of the Li+ insertion and extraction in the
the electrochemical performance of the porous TiO2 urchins as following discharge-charge process.
anode in LIBs. Meantime, we study the structure and morphology The rate capability of the TiO2 material is shown in Fig. 5c. It
evolution of the TiO2 material upon electrochemical cycling. reveals the discharge and charge capacities of the TiO2 electrodes
The Li+ storage process in TiO2 lattice can be formulated by the decrease along with the increase of current rates. However, the
following equation porous TiO2 urchins still deliver highly reversible Li+ storage
TiO2 + xLi+ + xe ! LixTiO2 (0  x  1). capacities of 166, 146.1, 123.1 and 94.4 mA h g1 at 2, 5, 10 and 20 C,
respectively. After lowering the current rate to 0.5 C, the Li+ storage
Theoretically, the maximum Li+ storage capacity is 335 mA h capacity of the urchins can be recovered to 204 mA h g1 again. In
1
g when one Li+ is inserted in one TiO2 molecule accompanied by comparison, the commercial P25 nanoparticles display an inferior
a complete reduction of Ti4+ to Ti3+. However, the strong Li+ storage capacity of 91.4, 69.4 and 39.2 mA h g1 at 0.5, 1, and 2 C,
electrostatic repulsion between the inserted Li+ ions results in respectively. Further, the electrochemical performance of the
the random distribution of them over half of available interstitial porous TiO2 urchins is superior or comparable to an array of TiO2
octahedral sites. Therefore, from a practical point of view, the nanostructures and TiO2/carbon nanohybrids, such as porous
maximum reversible insertion level is limited to an intercalation microspheres composed of octahedra (157.3 mA h g1 at 1 C) [35],
coefficient of 0.5 Li+ per formula unit of TiO2, corresponding to a anatase nanosheets with exposed (001) facets (180 mA h g1 at
Y. Cai et al. / Electrochimica Acta 210 (2016) 206–214 211

Fig. 5. Electrochemical performances of the porous TiO2 urchins and commercial P25 nanoparticles. (a) cyclic voltammetry (CV) curves recorded at 0.1 mV s1, (b) the first
two discharge-charge curves of the TiO2 spheres at 0.5 C, (c) rate property, and (d) cycling performances, (e) electrochemical impedance spectroscopy data and equivalent
circuit model (inset), (f) long-term cycling performance of the porous TiO2 urchins at 10 C for 5500 cycles. (1 C = 170 mA g1).

0.5 C and 100 mA h g1 at 10 C) [10], peapod-like mesoporous TiO2- electrodes have similar Rs values (4 V). However, the electrode
C array (164 mA h g1 at 1 C and 124 mA h g1 at 10 C) [58], and made of porous TiO2 urchins has a much reduced Rct value of
mesoporous hollow spheres with conformal carbon coating 64 V in comparison to that of the P25 electrode (107 V). The
(137 mA h g1 at 5 C) [59]. Fig. 5d further displays the cycling smaller Rct value is a clear indication of much facile charge transfer
performance of the TiO2 electrodes at 0.5 C and 1 C, respectively. It occurring at the electrode-electrolyte interface as well as a reduced
is clear that the porous TiO2 urchins demonstrate high and stable total internal resistance in the porous electrode. Moreover, the
Li+ storage capacities of 206.2 mA h g1 and 172.3 mA h g1 over larger slope in the Warburg-like response of the electrode made of
100 cycles at 0.5 C and 1 C, respectively, which are much better porous urchins suggests a more rapid Li+ diffusion in the solid-
than that of commercial P25 nanoparticles and superior to the state. As a result, the EIS results unravel more effective electronic
mesoporous TiO2 microspheres composed of nanoparticles and ionic conduction in the porous TiO2 urchins for more efficient
(160.5 mA h g1 after 100 cycles at 1 C) [41]. Li+ insertion/extraction.
Electrochemical impedance spectroscopy (EIS) data are per- The long-term cycling performance of the porous TiO2 urchins
formed to give more information on the electrode reaction kinetics is further evaluated at 10 C after activation for two cycles at 0.5 C.
(Fig. 5e and Fig. S9). The EIS responses can be divided into high, As shown in Fig. 5f, the TiO2 material exhibits an initial capacity of
high-to-medium, and low frequency regions, corresponding to the 129.2 mA h g1 at 10 C and retains a reversible Li+ storage capacity
series resistance (Rs), charge transfer resistance (Rct) at electrode- of 121.8 mA h g1 over 1000 cycles with 94.3% capacity retention.
electrolyte interface, and Li+ diffusion resistance (W) in the solid Then, the porous TiO2 urchins gradually exhibit slight capacity loss
state, respectively [60]. Fig. S9 shows the Nyquist plots of the TiO2 after each cycle. The Li+ insertion capacity of the porous TiO2
urchins electrode before cycling and after 5 cycles at 0.5 C. The urchins is lowered to 89.1 mA h g1 after 5500 cycles with 69%
obvious decrease of Rct after initial cycling can be ascribed to the 3D capacity retention.
interconnected porous structure, which provides a better interface To probe the possible capacity decay and structure evolution
contact between the active material and electrolyte and thus mechanism of the material, ex-situ XRD, SEM, and TEM character-
reduces the charge transfer resistance. To further confirm that, the izations of the TiO2 electrodes after electrochemical testing for 50
Nyquist plots of the TiO2 urchins and commercial P25 were also cycles and 5500 cycles were performed, respectively. From XRD
studied and fitted using ZView software based on an equivalent pattern in Fig. 6a, all the visible diffraction peaks can be indexed to
circuit model shown in the inset of Fig. 5e. It depicts that the two anatase phase in addition to some strong peaks arising from the
212 Y. Cai et al. / Electrochimica Acta 210 (2016) 206–214

urchins after 50 cycles largely retain anatase structure, suggesting


the very high reversibility of Li+ insertion/extraction in/from the
anatase lattice at the early stage. In contrast, HRTEM micrograph
(Fig. 7f) reveals the TiO2 electrode after 5500 cycles consists of both
anatase nanocrystals and cubic Li1TiO2 dots (10 nm in size). The
fringe spacing of 0.35 nm and 0.21 nm can be indexed to the
(101) facet of anatase and (200) facet of cubic Li1TiO2 (JCPDS card
No. 016-0223) [61], respectively. The coexistence of TiO2 and
Li1TiO2 can be further verified by the corresponding FFT pattern
(inset in Fig. 7f) derived from the HRTEM micrograph. The TEM
result, together with the high Li+ storage capacity of the porous
TiO2 urchins (Fig. 5), suggests that cubic Li1TiO2 has been reversibly
formed and decomposed during the initial discharge-charge
cycles, leading to increased Li+ storage capability. Then, some of
Li1TiO2 phase formed during discharge gradually evolves into
Fig. 6. Ex-situ XRD patterns of the porous TiO2 urchins after galvanostatic cycling at
isolated nanodots and randomly distributed in the anatase host,
10 C for 50 cycles (a) and 5500 cycles (b), respectively.
which loses electrochemical activity in the subsequent cycling.
To further discern the origin of the trace Li1TiO2 dots, CV curves
underlying Cu current collector, suggesting the reversible Li+
of the porous TiO2 urchins after cycling at 10 C for 5500 times were
storage in the obtained TiO2 material. Compared to Fig. 6a, the TiO2
recorded and compared with that of a fresh TiO2 electrode before
electrode after 5500 cycles displays a slight reduction of diffraction
galvanostatic test. As depicted in Fig. 8, the two CV curves roughly
peak intensity and disappearance of (001) plane (Fig. 6b), which is
overlap with each other, suggesting the high redox reversibility of
an indication of the loss of anatase phase as well as the instability
the obtained TiO2 material. However, the intensity of redox peaks
of active (001) crystal facets of anatase upon electrochemical
for anatase slightly lowers while that for TiO2-B stays the same,
cycling.
implying some anatase has been irreversibly converted into
SEM observation in Fig. 7a, d depicts that the “urchin”
electrochemically inert Li1TiO2 and lost Li+ storage ability in the
morphology has been well retained and the urchins are in intimate
subsequent cycling. In contrast, the trace TiO2-B phase remains
contact without aggregation or cracking during the long-term
electrochemically active even after long-term cycling due to their
electrochemical tests. From the magnified SEM images in the insets
pseudocapacitive Li+ storage characteristics, more open channel
of Fig. 7a and d, it can be found that the nanochains-constructed
structure for Li+ diffusion, and low mass density than anatase,
shells have been largely retained. The high structure stability of the
rendering the formation and decomposition of Li1TiO2 phase more
porous urchins during electrochemical tests can be further
facile and reversible.
confirmed by TEM images (Fig. 7b, e). In particular, some
Based on the above electrochemical and structural data, the
nanocrystals-connected nanochains can still be observed on the
superior electrochemical properties of the porous TiO2 urchins can
outer layer of the urchins. However, HRTEM analyses clearly show
be mainly ascribed to their distinct structural characteristics: (1)
some structure evolution of the porous urchins after different
3D hierarchical porous network with large voids among the
electrochemical cycles. The lattice fringes in the HRTEM micro-
adjacent urchins and mesopores within each urchin is ideal for
graph and corresponding FFT pattern reveal the porous TiO2

Fig. 7. Ex-situ SEM (a, d), magnified SEM (insets of a and d), TEM (b, e) and HRTEM (c, f) micrographs and corresponding FFT patterns (insets of c and f) of the TiO2 urchins after
cycling at 10 C for 50 cycles (a–c) and 5500 cycles (d–f), respectively. The capital letters “T” and “L” in (c) and (f) denote TiO2 and Li1TiO2, respectively.
Y. Cai et al. / Electrochimica Acta 210 (2016) 206–214 213

Acknowledgements

This work is realized in the frame of a program for Changjiang


Scholars and Innovative Research Team (IRT_15R52) of Chinese
Ministry of Education. B. L. Su acknowledges the Chinese Central
Government for an “Expert of the State” position in the Program of
the “Thousand Talents”. H. E. Wang and Y. Li acknowledge the
Hubei Provincial Department of Education for the “Chutian
Scholar” program. H. E. Wang also thanks the financial support
from National Natural Science Foundation of China for Young
Scholars (No. 51302204) and China Scholarship Council (CSC) for
supporting as a Visiting Scholar at University of Washington,
Seattle, WA, US. W. J. Zhang thanks the financial support from the
Applied Research Grant of the City University of Hong Kong (ARG
Fig. 8. CV curves of a fresh electrode made of TiO2 urchins recorded in the second 9667111).
sweep (a) and after galvanostatic cycling at 10 C over 5500 cycles (b). (Scan rate:
0.2 mV s1).
Appendix A. Supplementary data

large electrode/electrolyte interface contact, better electrolyte Supplementary data associated with this article can be
permeation and mass transfer; (2) the connected mesochannels found, in the online version, at http://dx.doi.org/10.1016/j.
effectively shorten the Li+ and electron diffusion pathways; (3) the electacta.2016.05.140.
nanochains and straight pore channels efficiently buffer the minor
volume change during (de)lithiation; (4) the coexistence of trace References
TiO2-B nanophases marginally increases the Li+ storage capacity
and improve the charge transfer/transport at two-phase bound- [1] B. Dunn, H. Kamath, J.-M. Tarascon, Electrical energy storage for the grid: a
aries. battery of choices, Science 334 (2011) 928–935.
[2] J.B. Goodenough, Y. Kim, Challenges for rechargeable Li batteries, Chem. Mater.
However, the porous TiO2 urchins also exhibit slight capacity
22 (2009) 587–603.
loss after very long-term electrochemical cycling possibly due to [3] M. Wagemaker, A. Kentgens, F. Mulder, Equilibrium lithium transport between
the irreversible formation of trace isolated Li1TiO2 nanodots. It is nanocrystalline phases in intercalated TiO2 anatase, Nature 418 (2002) 397–
noted that part of the obtained TiO2 material can be fully lithiated 399.
[4] G.-N. Zhu, Y.-G. Wang, Y.-Y. Xia, Ti-based compounds as anode materials for Li-
into Li1TiO2 due to its unique porous and nanocrystalline ion batteries, Energy Environ. Sci. 5 (2012) 6652–6667.
characteristics of the porous urchins during the initial electro- [5] D. Deng, M.G. Kim, J.Y. Lee, J. Cho, Green energy storage materials:
chemical cycling. However, some Li1TiO2 phase formed during Nanostructured TiO2 and Sn-based anodes for lithium-ion batteries, Energy
Environ. Sci. 2 (2009) 818–837.
discharge gradually loses electrochemical reactivity and exists as [6] T. Zhou, Y. Zheng, H. Gao, S. Min, S. Li, H.K. Liu, Z. Guo, Surface engineering and
isolated nanodots randomly distributed in the TiO2 host in the later design strategy for surface-amorphized TiO2@graphene hybrids for high
recharge process during the long-term cycling. This fact is probably power Li-ion battery electrodes, Adv. Sci. 2 (2015) 1500027.
[7] Z. Hong, M. Wei, T. Lan, G. Cao, Self-assembled nanoporous rutile TiO2
due to either the sluggish Li+ diffusivity of Li1TiO2 [62] or mesocrystals with tunable morphologies for high rate lithium-ion batteries,
insufficient electronic wiring of the Li1TiO2 nanodots with Nano Energy 1 (2012) 466–471.
surrounding electrode matrix [63]. In addition, the adjacent TiO2 [8] H.-T. Fang, M. Liu, D.-W. Wang, T. Sun, D.-S. Guan, F. Li, J. Zhou, T.-K. Sham, H.-M.
Cheng, Comparison of the rate capability of nanostructured amorphous and
nanochains might gradually fuse together during (de)lithiation, anatase TiO2 for lithium insertion using anodic TiO2 nanotube arrays,
leading to the loss of some straight mesochannels and increase of Nanotechnology 20 (2009) 225701.
Li+ concentration gradient between the outer and inner of the [9] K. Shen, H. Chen, F. Klaver, F.M. Mulder, M. Wagemaker, Impact of particle size
on the non-equilibrium phase transition of lithium-inserted anatase TiO2,
porous urchins. However, no cracks or larger secondary particles
Chem. Mater. 26 (2014) 1608–1615.
are observed in our material even after very long-term electro- [10] J.S. Chen, Y.L. Tan, C.M. Li, Y.L. Cheah, D. Luan, S. Madhavi, F.Y.C. Boey, L.A.
chemical cycling [45], demonstrating the very high structure Archer, X.W. Lou, Constructing hierarchical spheres from large ultrathin
integrity of the porous TiO2 urchins. anatase TiO2 nanosheets with nearly 100% exposed (001) facets for fast
reversible lithium storage, J. Am. Chem. Soc. 132 (2010) 6124–6130.
[11] X. Hua, Z. Liu, P.G. Bruce, C.P. Grey, The Morphology of TiO2 (B) Nanoparticles, J.
4. Conclusions Am. Chem. Soc. 137 (2015) 13612–13623.
[12] Z. Lu, C.T. Yip, L. Wang, H. Huang, L. Zhou, Hydrogenated TiO2 nanotube arrays
as high-rate anodes for lithium-ion microbatteries, ChemPlusChem 77 (2012)
Porous TiO2 urchins with spherical cores and nanochains- 991–1000.
constructed shells have been fabricated using TiO2/oleylamine as [13] Y. Ma, G. Ji, B. Ding, J.Y. Lee, Facile solvothermal synthesis of anatase TiO2
precursor. The resultant TiO2 material exhibits superior electro- microspheres with adjustable mesoporosity for the reversible storage of
lithium ions, J. Mater. Chem. 22 (2012) 24380–24385.
chemical performance with a combination of high capacity, [14] X.Y. Yu, H.B. Wu, L. Yu, F.X. Ma, X.W.D. Lou, Rutile TiO2 submicroboxes with
superior rate capability and excellent cycling stability. This superior lithium storage properties, Angew. Chem. Int. Ed. 54 (2015) 4001–
improved electrochemical performance of the TiO2 material can 4004.
[15] Z. Wang, Z.C. Wang, S. Madhavi, X.W.D. Lou, One-step synthesis of SnO2 and
be attributed to their distinct pore structures and high structure TiO2 hollow nanostructures with various shapes and their enhanced lithium
stability. More importantly, ex-situ structure characterizations storage properties, Chem. Eur. J. 18 (2012) 7561–7567.
demonstrate the irreversible transformation of trace anatase [16] H. Hu, L. Yu, X. Gao, Z. Lin, X.W.D. Lou, Hierarchical tubular structures
constructed from ultrathin TiO2 (B) nanosheets for highly reversible lithium
nanocrystals into cubic Li1TiO2 nanodots, which explains the
storage, Energy Environ. Sci. 8 (2015) 1480–1483.
slight capacity decay of the porous TiO2 material after long-term [17] W. Jiao, N. Li, L. Wang, L. Wen, F. Li, G. Liu, H.-M. Cheng, High-rate lithium
cycling. Further work is in progress to further optimize the storage of anatase TiO2 crystals doped with both nitrogen and sulfur, Chem.
geometry and pore structure of the TiO2 material for more stable Commun. 49 (2013) 3461–3463.
[18] Z. Bi, M.P. Paranthaman, B. Guo, R.R. Unocic, H.M. Meyer III, C.A. Bridges, X.-G.
cycling stability. In addition, the porous TiO2 urchins and Sun, S. Dai, High performance Cr, N-codoped mesoporous TiO2 microspheres
(protonated) titanates may find potential applications in sensitized for lithium-ion batteries, J. Mater. Chem. A 2 (2014) 1818–1824.
solar cells, (photo)catalysis, and sorption. [19] Q. Wu, J. Xu, X. Yang, F. Lu, S. He, J. Yang, H.J. Fan, M. Wu, Ultrathin anatase TiO2
nanosheets embedded with TiO2-B nanodomains for lithium-ion storage:
214 Y. Cai et al. / Electrochimica Acta 210 (2016) 206–214

capacity enhancement by phase boundaries, Adv. Energy Mater. 5 (2015) [41] H.-Y. Wang, J. Chen, S. Hy, L. Yu, Z. Xu, B. Liu, High-surface-area mesoporous
1401756. TiO2 microspheres via one-step nanoparticle self-assembly for enhanced
[20] C. Chen, X. Hu, B. Zhang, L. Miao, Y. Huang, Architectural design and phase lithium-ion storage, Nanoscale 6 (2014) 14926–14931.
engineering of N/B-codoped TiO2(B)/anatase nanotube assemblies for high- [42] G. Zhang, H.B. Wu, T. Song, U. Paik, X.W.D. Lou, TiO2 hollow spheres composed
rate and long-life lithium storage, J. Mater. Chem. A 3 (2015) 22591–22598. of highly crystalline nanocrystals exhibit superior lithium storage properties,
[21] P. Cui, B. Xie, X. Li, M. Li, Y. Li, Y. Wang, Z. Liu, X. Liu, J. Huang, D. Song, Anatase/ Angew. Chem. Int. Ed. 126 (2014) 12798–12801.
TiO2-B hybrid microspheres constructed from ultrathin nanosheets: facile [43] Q. Tian, Y. Tian, Z. Zhang, C. Qiao, L. Yang, S.-i. Hirano, Facile template-free
synthesis and application for fast lithium ion storage, CrystEngComm 17 preparation of hierarchical TiO2 hollow microspheres assembled by
(2015) 7930–7937. nanocrystals and their superior cycling performance as anode materials for
[22] Q. Tian, Z. Zhang, L. Yang, S.-i. Hirano, Morphology-engineered and TiO2(B)- lithium-ion batteries, J. Mater. Chem. A 3 (2015) 10829–10836.
introduced anatase TiO2 as an advanced anode material for lithium-ion [44] Z. Wang, X.W.D. Lou, TiO2 nanocages: fast synthesis, interior functionalization
batteries, J. Mater. Chem. A 3 (2015) 14721–14730. and improved lithium storage properties, Adv. Mater. 24 (2012) 4124–4129.
[23] R. Wang, X. Xue, W. Lu, H. Liu, C. Lai, K. Xi, Y. Che, J. Liu, S. Guo, D. Yang, Tuning [45] B.T. Yonemoto, Q. Guo, G.S. Hutchings, W.C. Yoo, M.A. Snyder, F. Jiao, Structural
and understanding the phase interface of TiO2 nanoparticles for more efficient evolution in ordered mesoporous TiO2 anatase electrodes, Chem. Commun. 50
lithium ion storage, Nanoscale 7 (2015) 12833–12838. (2014) 8997–8999.
[24] Z. Yan, L. Liu, J. Tan, Q. Zhou, Z. Huang, D. Xia, H. Shu, X. Yang, X. Wang, One-pot [46] D. Chen, L. Cao, F. Huang, P. Imperia, Y.-B. Cheng, R.A. Caruso, Synthesis of
synthesis of bicrystalline titanium dioxide spheres with a core–shell structure monodisperse mesoporous titania beads with controllable diameter, high
as anode materials for lithium and sodium ion batteries, J. Power Sources 269 surface areas, and variable pore diameters (14–23 nm), J. Am. Chem. Soc. 132
(2014) 37–45. (2010) 4438–4444.
[25] S. Liu, Z. Wang, C. Yu, H.B. Wu, G. Wang, Q. Dong, J. Qiu, A. Eychmüller, A flexible [47] L. Lutterotti, D. Chateigner, S. Ferrari, J. Ricote, Texture, residual stress and
TiO2(B)-based battery electrode with superior power rate and ultralong cycle structural analysis of thin films using a combined X-ray analysis, Thin Solid
life, Adv. Mater. 25 (2013) 3462–3467. Films 450 (2004) 34–41.
[26] L. He, C. Wang, X. Yao, R. Ma, H. Wang, P. Chen, K. Zhang, Synthesis of carbon [48] J. Yang, Z. Jin, X. Wang, W. Li, J. Zhang, S. Zhang, X. Guo, Z. Zhang, Study on
nanotube/mesoporous TiO2 coaxial nanocables with enhanced lithium ion composition, structure and formation process of nanotube Na2Ti2O4(OH)2,
battery performance, Carbon 75 (2014) 345–352. Dalton Trans. (2003) 3898–3901.
[27] X. Xin, X. Zhou, J. Wu, X. Yao, Z. Liu, Scalable synthesis of TiO2/graphene [49] C.-C. Tsai, H. Teng, Structural features of nanotubes synthesized from NaOH
nanostructured composite with high-rate performance for lithium ion treatment on TiO2 with different post-treatments, Chem. Mater. 18 (2006)
batteries, ACS Nano 6 (2012) 11035–11043. 367–373.
[28] B. Qiu, M. Xing, J. Zhang, Mesoporous TiO2 nanocrystals grown in situ on [50] Z. Sun, T. Liao, Y. Dou, S.M. Hwang, M.-S. Park, L. Jiang, J.H. Kim, S.X. Dou,
graphene aerogels for high photocatalysis and lithium-ion batteries, J. Am. Generalized self-assembly of scalable two-dimensional transition metal oxide
Chem. Soc. 136 (2014) 5852–5855. nanosheets, Nat. Comm. 5 (2014) 3813.
[29] L. Shen, X. Zhang, H. Li, C. Yuan, G. Cao, Design and tailoring of a three- [51] H. Wang, Z. Lu, L. Xi, R. Ma, C. Wang, J.A. Zapien, I. Bello, Facile and rapid
dimensional TiO2-graphene-carbon nanotube nanocomposite for fast lithium synthesis of highly porous wirelike TiO2 as anodes for lithium-ion batteries,
storage, J. Phys. Chem. Lett. 2 (2011) 3096–3101. ACS Appl. Mater. Interfaces 4 (2012) 1608–1613.
[30] W. Hu, L. Li, W. Tong, G. Li, Supersaturated spontaneous nucleation to TiO2 [52] T.E. Weirich, M. Winterer, S. Seifried, H. Hahn, H. Fuess, Rietveld analysis of
microspheres: synthesis and giant dielectric performance, Chem. Commun. 46 electron powder diffraction data from nanocrystalline anatase, TiO2,
(2010) 3113–3115. Ultramicroscopy 81 (2000) 263–270.
[31] L. Xiang, X. Zhao, J. Yin, B. Fan, Well-organized 3D urchin-like hierarchical TiO2 [53] Y. Cai, S.-Z. Huang, F.-S. She, J. Liu, R.-L. Zhang, Z.-H. Huang, F.-Y. Wang, H.-E.
microspheres with high photocatalytic activity, J. Mater. Sci. 47 (2012) 1436– Wang, Facile synthesis of well-shaped spinel LiNi0.5Mn1.5O4 nanoparticles as
1445. cathode materials for lithium ion batteries, RSC Adv. 6 (2016) 2785–2792.
[32] E. Hosono, S. Fujihara, H. Imai, I. Honma, I. Masaki, H. Zhou, One-step synthesis [54] Y. Li, Z.Y. Fu, B.L. Su, Hierarchically structured porous materials for energy
of nano-micro chestnut TiO2 with rutile nanopins on the microanatase conversion and storage, Adv. Funct. Mater. 22 (2012) 4634–4667.
octahedron, ACS Nano 1 (2007) 273–278. [55] S. Liu, H. Jia, L. Han, J. Wang, P. Gao, D. Xu, J. Yang, S. Che, Nanosheet-
[33] K.-S. Park, K.-M. Min, Y.-H. Jin, S.-D. Seo, G.-H. Lee, H.-W. Shim, D.-W. Kim, constructed porous TiO2-B for advanced lithium ion batteries, Adv. Mater. 24
Enhancement of cyclability of urchin-like rutile TiO2 submicron spheres by (2012) 3201–3204.
nanopainting with carbon, J. Mater. Chem. 22 (2012) 15981–15986. [56] J.Y. Shin, D. Samuelis, J. Maier, Sustained lithium-storage performance of
[34] C. Han, D. Yang, Y. Yang, B. Jiang, Y. He, M. Wang, A.-Y. Song, Y.-B. He, B. Li, Z. Lin, hierarchical, nanoporous anatase TiO2 at high rates: emphasis on interfacial
Hollow titanium dioxide spheres as anode material for lithium ion battery storage phenomena, Adv. Funct. Mater. 21 (2011) 3464–3472.
with largely improved rate stability and cycle performance by suppressing the [57] G. Sudant, E. Baudrin, D. Larcher, J.-M. Tarascon, Electrochemical lithium
formation of solid electrolyte interface layer, J. Mater. Chem. A 3 (2015) 13340– reactivity with nanotextured anatase-type TiO2, J. Mater. Chem. 15 (2005)
13349. 1263–1269.
[35] Y. Liu, T. Lan, W. Zhang, X. Ding, M. Wei, Hierarchically porous anatase TiO2 [58] L. Peng, H. Zhang, Y. Bai, Y. Zhang, Y. Wang, A self-supported peapod-like
microspheres composed of tiny octahedra with enhanced electrochemical mesoporous TiO2-C array with excellent anode performance in lithium-ion
properties in lithium-ion batteries, J. Mater. Chem. A 2 (2014) 20133–20138. batteries, Nanoscale 7 (2015) 8758–8765.
[36] Y.G. Guo, Y.S. Hu, W. Sigle, J. Maier, Superior electrode performance of [59] H. Liu, W. Li, D. Shen, D. Zhao, G. Wang, Graphitic carbon conformal coating of
nanostructured mesoporous TiO2 (anatase) through efficient hierarchical mesoporous TiO2 hollow spheres for high-performance lithium ion battery
mixed conducting networks, Adv. Mater. 19 (2007) 2087–2091. anodes, J. Am. Chem. Soc. 137 (2015) 13161–13166.
[37] H. Liu, Z. Bi, X.G. Sun, R.R. Unocic, M.P. Paranthaman, S. Dai, G.M. Brown, [60] Y. Ma, C. Fang, B. Ding, G. Ji, J.Y. Lee, Fe-doped MnxOy with hierarchical porosity
Mesoporous TiO2-B microspheres with superior rate performance for lithium as a high-performance lithium-ion battery anode, Adv. Mater. 25 (2013) 4646–
ion batteries, Adv. Mater. 23 (2011) 3450–3454. 4652.
[38] S. Yoon, A. Manthiram, Hollow core-shell mesoporous TiO2 spheres for lithium [61] Q. Gao, M. Gu, A. Nie, F. Mashayek, C. Wang, G.M. Odegard, R. Shahbazian-
ion storage, J. Phys. Chem. C 115 (2011) 9410–9416. Yassar, Direct evidence of lithium-induced atomic ordering in amorphous TiO2
[39] H. Ren, R. Yu, J. Wang, Q. Jin, M. Yang, D. Mao, D. Kisailus, H. Zhao, D. Wang, nanotubes, Chem. Mater. 26 (2014) 1660–1669.
Multishelled TiO2 hollow microspheres as anodes with superior reversible [62] M. Wagemaker, W.J. Borghols, E.R. van Eck, A.P. Kentgens, G.J. Kearley, F.M.
capacity for lithium ion batteries, Nano Lett. 14 (2014) 6679–6684. Mulder, The influence of size on phase morphology and Li-ion mobility in
[40] Y. Tang, Y. Zhang, J. Deng, J. Wei, H.L. Tam, B.K. Chandran, Z. Dong, Z. Chen, X. nanosized lithiated anatase TiO2, Chem. Eur. J. 13 (2007) 2023–2028.
Chen, Nanotubes: mechanical force-driven growth of elongated bending TiO2- [63] S.K. Das, A.J. Bhattacharyya, Influence of mesoporosity and carbon electronic
based nanotubular materials for ultrafast rechargeable lithium ion batteries, wiring on electrochemical performance of anatase titania, J. Electrochem. Soc.
Adv. Mater. 26 (2014) 6046. 158 (2011) A705–A710.

Você também pode gostar