Вы находитесь на странице: 1из 10

www.afm-journal.

de
www.MaterialsViews.com

MoS2-Quantum-Dot-Interspersed Li4Ti5O12 Nanosheets

FULL PAPER
with Enhanced Performance for Li- and Na-Ion Batteries
Guobao Xu, Liwen Yang,* Xiaolin Wei, Jianwen Ding, Jianxin Zhong, and Paul K. Chu*

materials such as Si,[5] transition metal


Rational nanoscale surface engineering of electroactive nanoarchitecture oxides,[6] transition metal dichalcoge-
is highly desirable, since it can both secure high surface-controlled energy nides,[2] and NASICON-type compounds,[7]
storage and sustain the structural integrity for long-time and high-rate production of suitable electrode materials
cycling. Herein, ultrasmall MoS2 quantum dots (QDs) are exploited as sur- with high-rate capability and ultrastability
face sensitizers to boost the electrochemical properties of Li4Ti5O12 (LTO). is still quite challenging. In addition, there
have been few investigations on the dual
The LTO/MoS2 composite is prepared by anchoring 2D LTO nanosheets
storage properties of lithium and sodium
with ultrasmall MoS2 QDs using a simple and effective assembly technique. based electrode materials.[8]
Impressively, such 0D/2D heterostructure composites possess enhanced Spinel Li4Ti5O12 (LTO) is a competi-
surface-controlled Li/Na storage behavior. This unprecedented Li/Na storage tive anode material for LIBs benefiting
process provides a LTO/MoS2 composite with outstanding Li/Na storage from stable charge/discharge platform
properties, such as high capacity and high-rate capability as well as long-term at 1.5 V, “zero-strain” in the lattice on
charging/discharging, and environmental
cycling stability. As anodes in Li-ion batteries, the materials have a stable benignity.[9,10] Recently, spinel LTO as an
specific capacity of 170 mAhg−1 after 20 cycles and are able to retain 94.1% anode material for NIBs has also been
of this capacity after 1000 cycles, i.e., 160 mAhg−1, at a high rate of reported.[11,12] Particularly, its relatively
10 C. Due to these impressice performance, the presented 0D/2D heterostruc- higher storage voltage versus Na/Na+ for
ture has great potential in high-performance LIBs and sodium-ion batteries. NIBs results in better cycle-life together
with improved safety compared with hard
carbon anode.[8] Various forms of LTO
nano/microstructures with improved LIB
1. Introduction performance have been prepared, for instance, nanoparticles,
nanowires, nanosheets, and hollow microspheres.[10,13] Further-
With ever-growing demand in automotive and stationary more, to overcome the hurdles of poor electrical conductivity,
energy storage applications, devices that can be recharged many strategies have been contrived to ameliorate the rate
swiftly, have long lifetime and high capacitance, and are inex- capability of LTO by means of elemental doping[14] and surface
pensive, are highly desirable.[1] Li-ion batteries (LIBs) have been modification with highly conductive substances.[15–24] Although
a favorable power source device in electronics and electrical LTO-based nanomaterials have been extensively studied for
vehicles because of the high capacity and long cycle life. Never- LIBs, their application to NIBs has been less explored so far.
theless, since it is envisioned that the lithium supply may run LTO nanorods, carbon coated LTO nanowires, and LTO mate-
out due to the increasing demand, sodium-ion batteries (NIBs) rials with hierarchically porous structures have recently been
as a promising candidate have again aroused intensive interest synthesized and demonstrated to deliver good sodium ion
due to more abundant and inexpensive sodium resources.[2–4] storage performance.[8,25,26] In spite of these improvements,
Despite efforts in developing the suitable new electrode they still are insufficient in practice and the major problems of
NIBs are fast capacity fading and poor high-rate performance.
In this respect, rational nanoscale surface engineering of elec-
G. B. Xu, Prof. L. W. Yang, Dr. X. L. Wei, troactive nanoarchitectures is a viable means to enhance the
Prof. J. W. Ding, Prof. J. X. Zhong properties.[27]
Hunan Key Laboratory of Micro-Nano
Energy Materials and Devices 2D nanosheets less than 10 nm are promising in high-per-
School of Physics and Optoelectronics formance batteries because of the short paths enabling fast
Xiangtan University ion diffusion, large exposed surface, as well as abundant ion
Hunan 411105, China insertion channels.[28,29] Ultrathin LTO nanosheets with a larger
E-mail: ylwxtu@xtu.edu.cn
specific surface area often have a larger discharge capacity
Prof. L. W. Yang, Prof. P. K. Chu
Department of Physics and Materials Science
and improved rate capability due to unprecedented pseudo-
City University of Hong Kong capacitive effect as the electrochemical interaction takes place
Tat Chee Avenue on the surface[30,31] or increased storage site resulting from the
Kowloon, Hong Kong, China cooccupation at the 8a and 16c sites at the surface region of
E-mail: paul.chu@cityu.edu.hk spinel structure.[14,32,33] Recently, 0D/2D heterostructure mate-
DOI: 10.1002/adfm.201505435 rials have attracted much attention in photomodulation and

Adv. Funct. Mater. 2016, 26, 3349–3358 © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 3349
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER

under vacuum, the LTO/MoS2 composite is


prepared. In our experiments, the loading
amounts of MoS2 QDs have important influ-
ence on electrochemical performance of
the LTO/MoS2 heterostructure composite.
As shown in the inset in Figure 2 and
Figure S1a (Supporting Information),
the LTO/MoS2 heterostructure composite
becomes darker with more MoS2 QDs,
implying that the MoS2 QDs are interspersed
homogeneously in the LTO nanosheets. The
cycling performances of the LTO/MoS2 heter-
ostructure composite with different amount
of MoS2 QDs as shown in Figure S1b (Sup-
porting Information) reveal that the sample
prepared with 2.5 wt% MoS2 QDs in NMP
Figure 1. Schematic showing the synthesis of the heterostructure composite composed of (designated as LTO-MoS2-2) delivers the best
ultrasmall MoS2 QDs and ultrathin LTO nanosheets. The vacuum-filtration assisted assembly electrochemical performance.
process enables facile and controllable integration of ultrathin LTO nanosheets and ultrasmall
MoS2 QDs. “D” and “S” represent the diffusion-controlled and enhanced surface-controlled
The phase structures of the materials are
lithium/sodium storage behavior, respectively. characterized by powder X-ray diffraction
(XRD). As depicted in Figure 2, the diffrac-
photodetection.[34] For example, in the MoS2 monolayers doped tion peaks of the as-prepared LTO nanosheets are well indexed
with graphene quantum dots (QDs), abnormal photolumines- to cubic spinel LTO with space group of Fd-3m (JCPDS No.
cence (PL) tuning was observed due to the competition between 49-0207). The sharp and well-defined peaks suggest high crys-
the photogenerated exciton and charge exciton induced by tallinity and no other peaks such as TiO2 can be detected indi-
charge transfer at the 0D/2D multilayer interface.[34] However, cating high purity as well. Compared to the LTO nanosheets,
the application of the heterostructures formed from 0D/2D three weak peaks at 14.3°, 39.4° and 49.7° originating from
materials to energy storage has been rarely reported, although the ultrasmall MoS2 QDs, which are well consistent with
MoS2 QDs have value-adding functions for solar cells, hydrogen hexagonal MoS2 (JCPDS No. 77-1716), can be observed. The
evolution, intracellular microRNA detection, and multiphoton crystal structure of the ultrasmall MoS2 QDs, LTO nanosheets,
bioimaging as a result of quantum confinement and edge and LTO-MoS2-2 composite is further characterized by trans-
effects.[35] Herein, the heterostructured composite of MoS2 mission electron microscopy (TEM). The typical TEM image
QDs and LTO nanosheets is prepared by self-assembly and the in Figure 3a shows that the as-made MoS2 QDs are well-dis-
Li/Na storage properties are assessed. The LTO nanosheets persed and have a size of about 3 nm. The characteristic (101)
homogeneously anchored with ultrasmall MoS2 QDs possess lattice plane of hexagonal MoS2 (PDF No. 37-1492) with a lat-
enhance surface-controlled Li/Na storage behavior, rendering tice spacing of 0.27 nm can be well resolved in high resolu-
the resulting heterostructure composite with outstanding Li/Na tion TEM image (HRTEM) (Figure 3b), suggesting high crys-
storage properties such as high capacity and high-rate capability tallinity of the MoS2 QDs. The ultrasmall MoS2 QDs solution
as well as excellent cycling stability.

2. Results and Discussion


As shown in Figure 1, the synthesis of LTO/MoS2 heterostruc-
tured composite involves the preparation and vacuum-filtration
assisted assembly of MoS2 QDs and LTO nanosheets. In brief,
a simple and mild oxidant(H2O2)-triggered spontaneous exfolia-
tion process is employed to prepare the ultrasmall MoS2 QDs
in N-Methyl-2-pyrrolidinone(NMP).[35] Then highly-crystalline
ultrathin LTO nanosheets prepared via topotactic transfor-
mation of layered ultrathin Li1.81H0.19Ti2O5·xH2O (H-LTO)
nanosheets with a minute of gadolinium ions at 700 °C[14]
are added into the NMP solution with ultrasmall MoS2 QDs.
Once being introduced, the MoS2 QDs tend to reside on the
surface of energetically more stable LTO nanosheets by van der
Waals force to make the total free energy of solution system
being minimized. After being stirred for about 1 h, the solu-
tion is vacuum-filtered with a microporous filtering mem- Figure 2. XRD patterns of the LTO nanosheets and LTO-MoS2-2 com-
brane with pore size of 0.15 µm and after being dried at 50 °C posite and the related photographs.

3350 wileyonlinelibrary.com © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2016, 26, 3349–3358
www.afm-journal.de
www.MaterialsViews.com

FULL PAPER
Figure 3. a) TEM image of MoS2 QDs with the size distribution in the inset; b) HR-TEM image of MoS2 QDs; c) Photographs of the MoS2 QDs solu-
tion taken under visible light (left) and 365 nm UV light (right) illumination; d,f) Typical TEM and HR-TEM images of the LTO nanosheets as well as
corresponding e) SAED patterns; g,h) typical TEM and HR-TEM images of the LTO-MoS2-2 composite.

is very stable without any noticeable precipitation after several be corroborated by Raman scattering and XPS to be discussed
weeks at room temperature. The solution under illumination later.
by a 365 nm lamp shows intense blue fluorescence (Figure 3c) Figure 4a depicts the Raman spectra acquired from the LTO-
consistent with the strong quantum confinement effect.[35] MoS2-2 composite. The Raman peaks at 670, 431, 348, 262, and
Figure 3d shows typical TEM image of the LTO nanosheets, 234 cm−1 are five characteristic phonon modes (A1g+Eg+3F2g)
demonstrating that the sample retains the sheet-like feature of of cubic spinel LTO, which are consistent with those from LTO
the H-LTO precursor with a thickness of 10 nm, which is con- nanosheets. The low-frequency Raman bands (F2g) at 340, 268,
sistent with our previous results.[24] The representative SAED and 234 cm−1 originate from the vibration of lithium octahe-
patterns (Figure 3e) and HRTEM image (Figure 3f) suggest drally coordinated by oxygen. The band at 431 cm−1 (Eg) is the
single crystallinity of individual LTO nanosheet. Well-defined stretching mode of Li-O in the LiO4 tetrahedra and the one at
lattice fringes with a separation of 0.48 nm in the HRTEM 670 cm−1 (A1g) is the stretching mode of Ti-O covalent bonding
image correspond to the spacing of the (111) atomic planes in the TiO6 octahedra. Besides, these Raman modes from cubic
of spinel LTO. Figure 3g shows the typical TEM image of the spinel LTO, two characteristic peaks at 383.8 and 409.4 cm−1
LTO-MoS2-2 composite, indicating that the ultrasmall MoS2 ascribed to the in-plane E12g and out-of-plane A1g modes of
QDs are dispersed well on the surface of LTO nanosheets. MoS2,[35] respectively, are observed from the LTO-MoS2-2 com-
Figure 3h is the corresponding HRTEM image disclosing that posite. Furthermore, the A1g phonons soften (redshift) while the
besides the lattice fringes of the LTO nanosheets, a few black A1g phonons stiffen (blueshift) from the LTO-MoS2-2 composite
spots with a size of about 3 nm originating from the ultrasmall compared to bulk MoS2. The results imply that the thickness of
MoS2 QDs can be observed. Although it is rather difficult to MoS2 QDs is thinned to fewer layers.[36] The surface electronic
obtain clear lattice fringes of the MoS2 QDs due to the influ- states and chemical composition of the LTO-MoS2-2 composite
ence of the LTO nanosheets, the presence of MoS2 QDs can are determined by X-ray photoelectron spectroscopy (XPS).

Adv. Funct. Mater. 2016, 26, 3349–3358 © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 3351
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER

Figure 4. a) Raman spectra of LTO-MoS2-2 composite and LTO nanosheets; b) survey XPS spectrum of the LTO-MoS2-2 composite; c,d) high-resolution
Mo 3d and S 2p XPS spectra of the LTO-MoS2-2 composite.

The results presented in Figure 4, Figures S2 and S3 (Sup- (i) at every sweep rate (v) compared to the LTO nanosheets elec-
porting Information), reveal the existence of Li, Ti, O, Mo, and trode (see inset of Figure 5a). The broadening of the CV peaks
S with a stoichiometric Mo/S atomic ratio of about 1:2. Quan- is different from those observed in graphene QDs coated VO2
titative XPS reveals that the detectable Mo content of ≈1 wt% arrays[27] and CuO nanowires[39] as well as LTO nanosheets
is below the 2.5 wt% of the MoS2 QDs in NMP. The two peaks modified with Ag nanocrystals,[24] in which the main role of the
at 228.6 and 231.8 eV attributed to Mo 3d5/2 and Mo 3d3/2, graphene QDs and Ag nanocrystals is to reduce the electrode
respectively and due to Mo4+ in pure MoS2[35,37] are observed polarization of electrode materials. The results imply another
from high-resolution Mo 3d spectrum(see Figure 4c). The high- lithium storage process in the LTO-MoS2-2 electrode. Each of
resolution S 2p peaks at 161.3 and 169.0 eV represent sulfide. broad anodic or cathodic peaks includes two overlapped peaks
It should be pointed out that the binding energy of Mo4+ in the (or a shoulder) in LTO-MoS2-2 electrode, which can be decon-
LTO-MoS2-2 composite shifts toward low-energy direction com- voluted into two pairs of redox peaks. Compared with these of
pared with that previously reported,[37] suggesting strong elec- LTO nanosheets, for example, at the sweep rate of 1 mV s−1 as
tron interaction between the LTO nanosheets and coupled MoS2 shown in Figure S4 (Supporting Information), the redox peaks
QDs.[38] Based on aforementioned results, ultrasmall MoS2 at 1.74 and 1.41 eV are mainly attributed to the redox reaction
QDs with good crystallinity are successfully coupled with LTO of Ti4+/Ti3+, respectively, associated with lithium insertion/
nanosheets and the highly dispersed MoS2 QDs on the LTO extraction in spinel LTO lattice, which are consistent with those
nanosheets benefit the formation of LTO/MoS2 heterojunctions reported for other LTO-based materials.[18,33] The redox peaks
which are expected to exhibit unique Li/Na storage behavior, at 1.81 and 1.36 eV cannot be attributed to lithium insertion/
thereby enhancing energy storage performance of the electrode extraction into MoS2 because lithium extraction from MoS2
materials. would exhibit an anodic peak located at around 2.3 V due to
For exploring the benefits granted by nanoscale LTO/MoS2 the oxidation of Mo into MoS2.[40] Therefore, this pair of redox
heterojunctions, the obtained composites as LIB and NIB peaks should originate from the coupling of MoS2 QDs and
anode materials are investigated using CR2032- type coin cells. LTO nanosheets. The results also indicate that the coupled
Figure 5a shows the CV curves of the LTO-MoS2-2 electrode MoS2 QDs speed up lithium diffusion and increase lithium
at different sweeping rates in the potential range between 1.0 storage capacity. Differing from the results for LIBs, the peaks
and 2.5 V (vs Li/Li+). The well-defined and symmetrical redox corresponding to sodium insertion/extraction of LTO exhibit no
peaks are observed, indicating excellent electrode kinetics of the obvious broadening in the CV curve of the LTO-MoS2-2 com-
LTO-MoS2-2 electrode. Interestingly, the LTO-MoS2-2 electrode posite (see Figure 5b). This likely originates from the broad and
exhibits much broader redox peaks with higher peak currents inconspicuous redox peaks of the LTO for NIBs.[8] In general,

3352 wileyonlinelibrary.com © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2016, 26, 3349–3358
www.afm-journal.de
www.MaterialsViews.com

FULL PAPER
(a) (b)

(c) (d)

(e) (f)

Figure 5. a,b) CV curves of the LTO-MoS2-2 composite and LTO nanosheets (see inset) at various scanning rates for LIBs and NIBs; c,d) determina-
tion of the b-value of the LTO-MoS2-2 composite and LTO nanosheets electrodes for LIBs and NIBs using the relationship between the peak current
(i) and scanning rate (v) in (a) and (b) according to i = avb; e,f) Nyquist plots of the LTO-MoS2-2 composite and LTO nanosheets for LIBs and NIBs
after 100 discharging and charging cycles.

the currents (i) in the CV curves of active electrode materials and cycling stability.[5] The surface-controlled capacitive contri-
response to the sweeping rates (v) obeying a power law rela- bution is also determined using another analysis, in which the
tionship of i = avb, where a and b are adjustable values.[5,41] current at a particular voltage is considered to contain contribu-
A b-value of 0.5 indicates that the current is a typical battery tions from capacitive (k1) and diffusion (k2) controlled processes
behavior controlled by semi-infinite linear diffusion, and the b according to following equation,[42] i = k1v + k2v1/2. The amount
value of 1 implies that the current is proportional to the sweep of charge stored due to the capacitive controlled process at a
rate, revealing surface-controlled charge storage processes. typical scan rate is presented in Figure S5 (Supporting Informa-
Figure 5c,d presents the log(v)-log(i) plots for the LTO- tion), which further confirm the enhanced surface-controlled
MoS2-2 composite and LTO nanosheets as electrode materials Li/Na storage in the LTO-MoS2-2 composite compare to LTO
in LIBs and NIBs. The fitted b values for cathodic and anodic nanosheets. Second, the enhanced surface-controlled storage
peak currents in the LTO-MoS2-2 composite (0.70 and 0.71 behavior of the H-LTO-MoS2 composite is attributed to the
for LIBs, 0.70/0.65 for NIBs) are larger than those of the LTO hetero-interface effect between the LTO nanosheets and MoS2
nanosheets (0.58 and 0.59 for LIBs, 0.44/0.60 for NIBs). The QDs. It is proposed that the change of structural environments
unique phenomena observed from the LTO-MoS2-2 composite in the near-surface region of LTO results in a distribution of
suggest two points. First, there is more contribution from sur- redox potentials in the near-surface area due to surface-related
face-controlled Li/Na storage that behaves similarly to pseudo- lithium storage or the cooccupation at the 8a and 16c sites at
capacitive effect in supercapacitor materials for the LTO-MoS2-2 the surface region.[43] This unique effect becomes increasing
composite and it is expected to ameliorate high-rate capability dominated when the size is reduced to 10 nm.[9] In our case, on

Adv. Funct. Mater. 2016, 26, 3349–3358 © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 3353
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER

(a) (b)

(c) (d)

(e) (f)

Figure 6. a,b) Charging–discharging voltage profiles of the LTO-MoS2-2 composite and LTO nanosheets as LIBs electrodes for the 1st, 500th, and
1000th cycles at a current rate of 10 C; c,d) charging–discharging voltage profiles of the LTO-MoS2-2 composite and LTO nanosheets as LIBs elec-
trodes at different current rates; e) corresponding rate capabilities at different current rates from 1 C to 30 C for the LTO-MoS2-2 composite and
LTO nanosheets; f) corresponding long-cycling performance and Coulombic efficiency versus cycle number of the LTO-MoS2-2 composite and LTO
nanosheets at a rate of 10 C.

account of band alignment and potential difference resulting eradicating the structural and interfacial instability of the LTO
from intrinsic band gap properties of MoS2 and LTO, the strong due to the surface reconstruction or mechanical failure associ-
hetero-interfacial effect between the LTO nanosheets and cou- ated with too high surface Li/Na storage[43] and enhancing the
pled MoS2 QDs produces the space charge layer near the LTO/ cycling stability. In addition, the nanocavities between the cou-
MoS2 interface to seriously change the structural environ- pled MoS2 QDs offer favorable transport routes for Li/Na ions
ments in the near-surface region of LTO nanosheets, leading and simultaneously provide a nontrivial contribution to the
to a distribution of redox potentials in the near-surface area capacity due to physical adsorption of Li/Na ions between the
accompanying enhanced surface-controlled Li/Na storage. The MoS2 QDs.
hetero-interfacial effect likely also enhances the relative current Figure 6a,b shows the galvanostatic charging–discharging
associated with high surface electrochemical activity within the profiles of the LTO nanosheets and LTO-MoS2-2 composite
surface reaction region. In other words, this hetero-interfacial electrodes for the 1st, 500th, and 1000th cycles at a current rate
effect makes the gradients of the Li/Na concentrations during of 10 C in the voltage range from 1.0 to 2.5 V versus Li/Li+.
Li/Na insertion/extraction being suppressed notably within the Both electrodes exhibit similar charging and discharging flat
space charge layer near the LTO/MoS2 interface, and subse- plateaus at around 1.5 V, confirming characteristic two-phase
quently the diffusion distance for Li/Na ions is notably short- lithium insertion/extraction reaction mechanism of spinel
ened in the LTO/MoS2 heterostructure to allow Li/Na ions to LTO.[22,44] However, the charging/discharging voltage profiles
migrate more quickly throughout the LTO nanosheets, thereby of the LTO-MoS2-2 composite electrode have a flatter profile

3354 wileyonlinelibrary.com © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2016, 26, 3349–3358
www.afm-journal.de
www.MaterialsViews.com

FULL PAPER
than those of the reference LTO nanosheets for every cycle, Owing to the aforementioned advantages of MoS2 QDs cou-
suggesting smaller polarization (Δ E) between the charging pling, we further examine the sodium storage performance
and discharging plateaus. The lower electrochemical polariza- of the LTO/MoS2 heterostructure composite. As shown in
tion implies improved reaction kinetics due to higher surface Figure 5b, the CVs between 0.5 and 2.5 V acquired from both
electron conductivity of unique LTO/MoS2 heterostructure. the LTO-MoS2-2 composite and LTO nanosheets show one
On the one hand, an extended solid solution region for Li+ anodic peak at 1.09 V and cathodic peak at 0.66 V representing
insertion into LTO is observed from the LTO-MoS2-2 com- sodium insertion-extraction in the LTO crystal lattice. A pair of
posite electrode in the charging–discharging voltage profiles weak anodic and cathodic peaks is observed at around 0.9 V
from ≈2.3 to ≈1.5 V, similar to those reported previously from implying that the Na+ extraction process follows a more com-
small particles such as TiO2[45,46] and Si,[5] in which an addi- plicated path.[8] Figure 7a,b depict the charging–discharging
tional surface lithium storage mechanism exists. On the other voltage profiles of the LTO nanosheets and LTO-MoS2-2 com-
hand, a smaller slope below 1.5 V that is a typical feature of posite electrodes cycled at different current rates from 0.1 to
a supercapacitor is observed from the LTO-MoS2-2 composite 5 C between 0.5 and 2.5 V. Both electrodes exhibit the same
electrode, implying higher capacitance associated with lithium features in the curves: a slow voltage decay from 1.5 to 0.8 V
storage by surface control of the LTO/MoS2 heterostructure. and a flat plateau from 0.8 to 0.5 V consistent with the CV
The rate capability of the LTO-MoS2-2 composite as an anode results. The region of the slow voltage decay from 1.5 to 0.8 V
for LIBs is investigated by charging–discharging at various is extended for the LTO-MoS2-2 composite electrode indicating
current rates from 1 to 30 C (see Figure 6c,d). The Δ E of the an enhanced surface Na storage behavior. The polarization
LTO-MoS2-2 composite electrode is depressed compared to the between the charging and discharging plateaus of the LTO-
LTO nanosheet electrode for every C-rate, suggesting better rate MoS2-2 composite electrode is smaller than that of the refer-
capability. At 1 C, the LTO-MoS2-2 composite electrode has a ence LTO nanosheets suggesting a better rate capability as well.
discharge capacity of 189 mAhg−1, which is beyond the theoret- For example, as shown in Figure 7c, at rates of 2 and 5 C, the
ical value of 175 mAh g−1 suggesting that there exist additional capacities of the LTO-MoS2-2 composite electrode are 118 and
lithium storage sites in the LTO-MoS2-2 composite electrode.[47] 91 mAh/g and greater than those of the LTO nanosheets elec-
As the rates is increased from 1 to 5 ,10, 20, and 30 C, the dis- trode of 73 and 49 mAh/g, respectively. The cycling stability is
charge capacities are 177, 172, 166, and 161 mAhg−1 with the investigated at 2 C for 200 cycles and Figure 7d shows that the
capacity retention of 93.6% , 91% , 88%, and 85.1%, respec- reversible capacity of the LTO-MoS2-2 composite electrode can
tively. In comparison, the LTO nanosheets electrode cannot be retained at 101 mAhg−1 after 200 cycles, which is superior to
match such high capacity retention at every rate. For example, that (60 mAhg−1) of the LTO nanosheets. These results demon-
the capacity of the LTO nanosheets at 30 C is only 120 mAhg−1, strate that the LTO/MoS2 heterostructure composite is suitable
which is ≈68.8% of that at 1 C. More importantly, as shown in for high-power NIBs due to excellent stable capacity and high
Figure 6e, the relative increase (Δ C) of discharge capacity is rate capacity.
larger at higher rates and the Coulombic efficiencies at large The superior Li/Na storage performance of the LTO/MoS2
rates approach 100% for each cycle, implying ultrafast lithium heterostructure composite can be attributed to some enhance-
diffusion and high reversibility of electrochemical reaction due ments of transport kinetics in electrode. To understand the
to heterostructure effect in the LTO-MoS2-2 composite elec- transport kinetics in the LTO-MoS2-2 composite electrode, elec-
trode, thereby leading to excellent rate capability. In addition, trochemical impedance spectra (EIS) are performed after the
a stable capacity of 178 mAhg−1 with negligible loss for subse- 100th cycling test. The obtained Nyquist plots are presented
quent cycling can be retained when the current rate is returned in Figure 5e,f, revealing a purely resistive response at high
to 1 C after the rate capability test, suggesting superior sta- frequency region represented by the ohmic resistance (Rs) of
bility of the LTO/MoS2 heterostructure. Figure 6f displays the the electrode and electrolyte, a semicircle arc due to the charge-
long-term cycling performance of the LTO-MoS2-2 composite transfer resistance (Rct) on electrode–electrolyte interface in
electrode at a high current density of 10 C, demonstrating a high-to-middle frequency region and an inclined straight line
stable capacity of 170 mAhg−1 with the Coulombic efficiency ascribed to the Warburg impedance (Rw) in low frequency
approaching 100% after 20 cycles. A capacity of 160 mAh g−1 region. The Warburg-like response manifests larger slopes for
is still retained after 1000 cycles with a capacity loss of only the LTO-MoS2-2 composite electrodes for both LIBs and NIBs
5.9%. The cycling performance of the LTO-MoS2-2 composite suggesting higher Li/Na mobility[19] and enhanced pseudoca-
electrode is superior to that of the LTO nanosheets with a pacitive properties compared to the LTO nanosheets.[49] The
capacity loss of 15.8% after 1000 cycles and the cycling stability semicircle arcs of the LTO-MoS2-2 composite electrode for
(0.0059% decay per cycle) of the LTO-MoS2-2 composite elec- both LIBs and NIBs are smaller than those of reference LTO
trode is among the best observed from LTO-based anodes so nanosheets, implying smaller Rct values at the electrode–elec-
far. The high-rate and long-term cycling performance of the trolyte interface,[22,47] which are consistent with the simulation
LTO-MoS2-2 composite electrode are comparable to the results results (see Table S1 in the Supporting Information) using
obtained from LTO nanowire arrays,[20] Cu-doped LTO-TiO2 modified Randles equivalent circuit (see Figure S6 in the Sup-
nanosheets,[48] self-supported LTO nanosheets arrays,[32] LTO-C porting Information). The results indicate that the LTO/MoS2
nanotube arrays,[18] carbon-encapsulated F-doped LTO,[19] LTO heterostructure facilitates rapid charge transfer at the electrode/
tailored by cathodically induced graphene[22] (see Table S1 in electrolyte interface in the LTO-MoS2-2 composite electrode,
the Supporting Information), and so on, confirming the posi- thereby leading to better Li/Na storage capability. Further-
tive effects rendered by the LTO/MoS2 heterostructure. more, the indicative exchange current density can be analyzed

Adv. Funct. Mater. 2016, 26, 3349–3358 © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 3355
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER

(a) (b)

(c) (d)

Figure 7. a,b) Charging–discharging voltage profiles of the LTO-MoS2-2 composite and LTO nanosheets as NIBs electrodes at different current rates;
c) corresponding rate capabilities at different current rates from 0.1 C to 5 C for the LTO-MoS2-2 composite and LTO nanosheets electrodes; d) long
cycling performance of the LTO-MoS2-2 composite and LTO nanosheets as NIBs electrode at a rate of 2 C.

according to the equation of i0 = RT/nFRct (see Table 1).[19] The high surface Li/Na storage behaviors.[43] The homogeneously
enhanced exchange current density of the LTO-MoS2-2 com- coupled MoS2 QDs can also prevent dissolution of the LTO by
posite electrode for NIBs is not as obvious as that of LIBs due decreasing the exposed surface area of the LTO. In addition to
to slow Na+ diffusion in the LTO host. However, the exchange promoting Li/Na insertion–extraction accompanying enhanced
current density of the LTO-MoS2-2 composite electrode for LIBs surface-controlled Li/Na storage during the electrochemical
after 100 cycles is nearly double that of the LTO nanosheets reactions, the “sensitization” effect of the coupled ultrasmall
electrode (i0 = 0.84 versus 0.41 mAcm−2). The enhancement in MoS2 QDs likely makes the LTO nanosheets surface lipophilic
the charge-transfer kinetics in the LTO-MoS2-2 composite elec- thereby facilitating electrolyte penetration and ion transfer to
trode results from the special structure of the LTO/MoS2 het- improve the reaction kinetics.[27] Consequently, outstanding
erostructure. First, the heterostructure combination of the LTO Li/Na storage properties such as excellent rate capability and
nanosheets and MoS2 QDs promotes Li/Na insertion–extrac- remarkable cycling stability are achieved from the LTO/MoS2
tion due to the shorter Li+/Na+ diffusion distance and higher composite.
surface electrochemical activity associated with the coupled het-
erointerfacial effect between LTO nanosheets and MoS2 QDs as
well as avoiding the drawback of grain boundaries benefiting 3. Conclusion
from the singe-cryastal nature of LTO nanosheets. Second,
the structure stability is better since the formation of covalent High capacity, high rate, and durable Li/Na storage are achieved
bonds between the MoS2 QDs and LTO during the electro- from the heterostructure composite composed of single-crystal
chemical reactions can act as an effective buffer to eradicate LTO nanosheets homogeneously anchored with ultrasmall
the structural and interfacial instability of the LTO due to the MoS2 QDs. The strong hetero-interfacial effect between the LTO
surface reconstruction or mechanical failure associated with too nanosheets and MoS2 QDs endows the 0D/2D heterostructure

Table 1. Rs and Rct values of the LTO nanosheets and LTO-MoS2-2 composite for LIBs and NIBs after 100 cycles in half-cells according to Figure 5e,f
as well as exchange current density (i0) calculated from to i0 = RT/nFRct, where n and F are constants.

Samples LIBs NIBs


Rs [Ω] Rct [Ω] i0 [mA cm−2] Rs [Ω] Rct [Ω] i0 [mA cm−2]
LTO 5.13 62.1 0.41 5.01 202 0.123
LTO-MoS2-2 4.21 30.2 0.845 4.66 161 0.154

3356 wileyonlinelibrary.com © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2016, 26, 3349–3358
www.afm-journal.de
www.MaterialsViews.com

in the voltage range from 1 to 2.5 V (vs Li/Li+) and 0.5 to 2.5 V (vs Na/

FULL PAPER
with enhanced surface-controlled Li/Na storage. The MoS2
QDs may work as a surface sensitizer and protector. As a Na+). Cyclic voltammetry (CV) was carried out on an electrochemical
result, the LTO/MoS2 heterostructure composite possesses workstation (CHI660D) in the voltage range between 1 and 2.5 V (vs Li/
Li+) as well as 0.5 and 2.5 V (vs Na/Na+). Electrochemical impedance
outstanding Li/Na storage properties. With regard to lithia- spectroscopy (EIS) was performed between 0.01 Hz and 100 kHz on the
tion, the LTO/MoS2 heterostructure composite shows a stable CHI660E electrochemical workstation. All the tests were conducted out
reversible specific capacity of 170 mAhg−1 after 20 cycles and it at room temperature.
is maintained at 160 mAhg−1 after 1000 cycles at a high rate of
10 C with 94.1% retention. During sodiation, a high capacity of
91 mAh/g at a rate of 5 C and superior capacity retention with
101 mAhg−1 after 200 cycles at the rate of 2 C are achieved. The
Supporting Information
results provide an effective strategy to improve the capacity, rate Supporting Information is available from the Wiley Online Library or
capability, and cycling stability of LTO-based anode material for from the author.
LIBs and NIBs, thus pushing LTO as a cost-effective alternative
for next-generation high-performance electrochemical energy
storage devices. Acknowledgements
This work was financially supported by the National Natural Science
Foundation of China (Grant Nos. 11474242, 51272220, 11374252, and
51472209), Program for Changjiang Scholars and Innovative Research
4. Experimental Section Team in University (Grants No. IRT13093), as well as City University of
Synthesis of Ultrathin LTO Nanosheets and Ultrasmall MoS2 QDs: The Hong Kong Applied Research Grant (ARG) No. 9667104.
ultrathin LTO nanosheets were prepared by a modified hydrothermal
method followed by annealing treatment.[24] The ultrasmall MoS2 Received: December 16, 2015
QDs were prepared by a spontaneous mild oxidant (H2O2)-triggered Revised: January 27, 2016
exfoliation process.[35] Typically, 200 mg of pristine MoS2 flake were Published online: March 2, 2016
dispersed in a solution of H2O2 (30 wt% aqueous solution) and NMP
(v/v = 1:1) and stirred at 35 °C for 10 h. The mixture was centrifuged for
30 min at 5000 rpm and the upper half of the supernatant was collected.
After centrifugation for two to four times, the mixture became colorless [1] M. Armand, J.-M. Tarascon, Nature 2008, 451, 652.
and the ultrasmall MoS2 QDs dispersed in NMP were obtained. [2] X. Xiang, K. Zhang, J. Chen, Adv. Mater. 2015, 27, 5343.
Synthesis of the LTO/MoS2 Composite: Since it was rather difficult [3] N. Yabuuchi, K. Kubota, M. Dahbi, S. Komaba, Chem. Rev. 2014,
to collect the ultrasmall MoS2 QDs powder, the MoS2 QDs dispersed 114, 11636.
in NMP were used to prepare the LTO/MoS2 composite directly. [4] H. Pan, Y.-S. Hu, L. Chen, Energy Environ. Sci. 2013, 6, 2338.
First, the ultrathin LTO nanosheets with an appropriate content were [5] B. Wang, X. Li, B. Luo, L. Hao, M. Zhou, X. Zhang, Z. Fan, L. Zhi,
added slowly to the NMP solution with ultrasmall MoS2 QDs. After
Adv. Mater. 2015, 27, 1526.
being stirred for about 1 h, the solution was vacuum-filtered through
[6] X. Chen, C. Li, M. Graetzel, R. Kostecki, S. S. Mao, Chem. Soc. Rev.
a microporous filtering membrane with a pore size of 0.15 µm. The
2012, 41, 7909.
collected samples were dried at 50 °C for 10 h in vacuum to remove
residual NMP molecules to obtain the LTO-MoS2 composite. To adjust [7] C. Wu, P. Kopold, Y.-L. Ding, P. A. van Aken, J. Maier, Y. Yu, ACS
the content of MoS2 QDs in the composite, the NMP solutions with Nano 2015, 9, 6610.
different concentrations of MoS2 QDs (1, 2.5, 5, and 10 wt%) were [8] G. Hasegawa, K. Kanamori, T. Kiyomura, H. Kurata, K. Nakanishi,
prepared by evaporating the NMP content. These samples were T. Abe, Adv. Energy Mater. 2015, 5, 1400730.
designated as LTO-MoS2-1, LTO-MoS2-2, LTO-MoS2-3, and LTO-MoS2-4, [9] S. Ganapathy, M. Wagemaker, ACS Nano 2012, 6, 8702.
respectively. [10] T.-F. Yi, S.-Y. Yang, Y. Xie, J. Mater. Chem. A 2015, 3, 5750.
Materials Characterization: The morphology, microstructure, and [11] Y. Sun, L. Zhao, H. Pan, X. Lu, L. Gu, Y.-S. Hu, H. Li, M. Armand,
composition of the samples were studied by SEM (Hitachi, S4800), TEM Y. Ikuhara, L. Chen, Nature Commun. 2013, 4, 1870.
(JEOL-2100F) equipped with SAED, and XRD (Rigaku, D/MAX 2500) [12] X. Yu, H. Pan, W. Wan, C. Ma, J. Bai, Q. Meng, S. N. Ehrlich,
using a copper Kα radiation source (λ = 0.154 nm), Raman scattering Y.-S. Hu, X.-Q. Yang, Nano Lett. 2013, 13, 4721.
(Renishaw InVia system at an excitation wavelength of 532 nm), and [13] L. Yu, H. B. Wu, X. W. D. Lou, Adv. Mater. 2013, 25, 2296.
XPS (Kratos Analytical Ltd., UK) using an Al Kα source. [14] G. Xu, L. Yang, X. Wei, J. Ding, J. Zhong, P. Chu, J. Power Sources
Electrochemical Characterization: The electrochemical tests were 2015, 295, 305.
conducted using two-electrode CR2032 type coin cells. The working [15] B. Li, C. Han, Y.-B. He, C. Yang, H. Du, Q.-H. Yang, F. Kang, Energy
electrodes in the lithium and sodium cells were prepared by pasting the Environ. Sci. 2012, 5, 9595.
mixture of the active materials, carbon black, and polyvinylidene fluoride [16] N. Li, Z. Chen, W. Ren, F. Li, H.-M. Cheng, Proc. Natl. Acad. Sci.
(PVDF) at a weight ratio of 8:1:1 onto a Cu foil. After being dried at 80 °C
U.S.A. 2012, 109, 17360.
for 6 h in air and then at 120 °C in vacuum for 12 h, the electrodes
[17] N. Li, G. Zhou, F. Li, L. Wen, H. M. Cheng, Adv. Funct. Mater. 2013,
were punched into round disks with a diameter of 1.2 cm and the mass
23, 5429.
of the active materials ranging from 1 and 2 mg. Metallic lithium and
sodium foils were used as the counter electrodes, and Celgard 2400 [18] J. Liu, K. Song, P. A. van Aken, J. Maier, Y. Yu, Nano Lett. 2014, 14,
polypropylene and glass fiber (Whatman GF/D) as separators for the 2597.
lithium and sodium cells, respectively. The electrolytes were 1 M LiPF6 [19] Y. Ma, B. Ding, G. Ji, J. Y. Lee, ACS Nano 2013, 7, 10870.
in a 1:1 (V:V) mixture of ethylene carbonate/dimethyl carbonate and 1 M [20] L. Shen, E. Uchaker, X. Zhang, G. Cao, Adv. Mater. 2012, 24, 6502.
NaClO4 in 1:1 (V:V) mixture of ethylene carbonate/propylene carbonate [21] L. Shen, X. Zhang, E. Uchaker, C. Yuan, G. Cao, Adv. Energy Mater.
for lithium and sodium cells, respectively. All coin cells were assembled 2012, 2, 691.
in an Ar-filled glove box. The galvanostatic charging/discharging was [22] Y. Yang, B. Qiao, X. Yang, L. Fang, C. Pan, W. Song, H. Hou, X. Ji,
performed using a multichannel battery test system (NEWARE BTS-610) Adv. Funct.Mater. 2014, 24, 4349.

Adv. Funct. Mater. 2016, 26, 3349–3358 © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 3357
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER

[23] L. Zhao, Y. S. Hu, H. Li, Z. Wang, L. Chen, Adv. Mater. 2011, 23, 1385. [38] Y. Hou, A. B. Laursen, J. Zhang, G. Zhang, Y. Zhu, X. Wang, S. Dahl,
[24] G. Xu, W. Li, L. Yang, X. Wei, J. Ding, J. Zhong, P. K. Chu, J. Power I. Chorkendorff, Angew. Chem. Int. Ed. 2013, 52, 3621.
Sources 2015, 276, 247. [39] C. Zhu, D. Chao, J. Sun, I. M. Bacho, Z. Fan, C. F. Ng, X. Xia,
[25] K.-T. Kim, C.-Y. Yu, C. S. Yoon, S.-J. Kim, Y.-K. Sun, S.-T. Myung, H. Huang, H. Zhang, Z. X. Shen, Adv. Mater. Interfaces 2015, 2,
Nano Energy 2015, 12, 725. 1400499.
[26] Q. Zhou, L. Liu, J. Tan, Z. Yan, Z. Huang, X. Wang, J. Power Sources [40] C. Zhu, X. Mu, P. A. van Aken, J. Maier, Y. Yu, Adv. Energy Mater.
2015, 283, 243. 2015, 5, 1401170.
[27] D. Chao, C. Zhu, X. Xia, J. Liu, X. Zhang, J. Wang, P. Liang, J. Lin, [41] V. Augustyn, J. Come, M. A. Lowe, J. W. Kim, P.-L. Taberna,
H. Zhang, Z. X. Shen, H. Fan, Nano Lett. 2014, 15, 565. S. H. Tolbert, H. D. Abruña, P. Simon, B. Dunn, Nature Mater.
[28] F. Bonaccorso, L. Colombo, G. Yu, M. Stoller, V. Tozzini, 2013, 12, 518.
A. C. Ferrari, R. S. Ruoff, V. Pellegrini, Science 2015, 347, 1246501. [42] G. A. Muller, J. B. Cook, H.-S. Kim, S. H. Tolbert, B. Dunn, Nano
[29] Y. Sun, S. Gao, Y. Xie, Chem. Soc. Rev. 2014, 43, 530. Lett. 2015, 15, 1911.
[30] J. Chen, L. Yang, S. Fang, Y. Tang, Electrochim. Acta 2010, 55, 6596. [43] W. Borghols, M. Wagemaker, U. Lafont, E. Kelder, F. Mulder, J. Am.
[31] C. Lai, Y. Dou, X. Li, X. Gao, J. Power Sources 2010, 195, 3676. Chem. Soc. 2009, 131, 17786.
[32] S. Chen, Y. Xin, Y. Zhou, Y. Ma, H. Zhou, L. Qi, Energy Environ. Sci. [44] J. M. Feckl, K. Fominykh, M. Döblinger, D. Fattakhova-Rohlfing,
2014, 7, 1924. T. Bein, Angew. Chem. Int. Ed. 2012, 51, 7459.
[33] Y.-Q. Wang, L. Gu, Y.-G. Guo, H. Li, X.-Q. He, S. Tsukimoto, [45] C. Chen, Y. Wen, X. Hu, X. Ji, M. Yan, L. Mai, P. Hu, B. Shan,
Y. Ikuhara, L.-J. Wan, J. Am. Chem. Soc. 2012, 134, 7874. Y. Huang, Nature Commun. 2015, 6, 6929.
[34] Z. Li, R. Ye, R. Feng, Y. Kang, X. Zhu, J. M. Tour, Z. Fang, Adv. Mater. [46] J. Wang, J. Polleux, J. Lim, B. Dunn, J. Phys. Chem. C 2007, 111,
2015, 27, 5235. 14925.
[35] L. Dong, S. Lin, L. Yang, J. Zhang, C. Yang, D. Yang, H. Lu, Chem. [47] S. Yu, L. Yang, Y. Tian, P. Yang, F. Jiang, S. Hu, X. Wei, J. Zhong, J.
Commun. 2014, 50, 15936. Mater. Chem. A 2013, 1, 12750.
[36] C. Lee, H. Yan, L. E. Brus, T. F. Heinz, J. Hone, S. Ryu, ACS Nano [48] C. Chen, Y. Huang, C. An, H. Zhang, Y. Wang, L. Jiao, H. Yuan,
2010, 4, 2695. ChemSusChem 2015, 8, 114.
[37] J. Kibsgaard, Z. Chen, B. N. Reinecke, T. F. Jaramillo, Nature Mater. [49] V. Augustyn, P. Simon, B. Dunn, Energy Environ. Sci. 2014, 7,
2012, 11, 963. 1597.

3358 wileyonlinelibrary.com © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2016, 26, 3349–3358

Вам также может понравиться