Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Applied Surface Science 509 (2020) 144718

Contents lists available at ScienceDirect

Applied Surface Science


journal homepage: www.elsevier.com/locate/apsusc

Full Length Article

High-performance ZnTe-TiO2-C nanocomposite with half-cell and full-cell T


applications as promising anode material for Li-Ion batteries

Quoc Hanh Nguyena,1, Quoc Hai Nguyenb,1, Jaehyun Hura,
a
Department of Chemical and Biological Engineering, Gachon University, Seongnam, Gyeonggi 461-701, Republic of Korea
b
Faculty of Applied Sciences, Ton Duc Thang University, Ho Chi Minh City 700000, Viet Nam

A R T I C LE I N FO A B S T R A C T

Keywords: C-coated ZnTe-TiO2 nanocomposite is successfully synthesized via annealing followed by high-energy me-
Zinc telluride chanical milling. ZnTe-TiO2-C nanocomposite is comprised of intermetallic ZnTe nanoalloy homogeneously
Titanium oxide distributed in a hybrid matrix of rutile TiO2 and conductive amorphous C. The presence of rutie TiO2 and C
Anode significantly improves the electrochemical performance of ZnTe; that is, highly stable cyclic performance
High-energy mechanical milling
(611 mAh g−1 at 0.1 A g−1 with ~95% capacity retention over 300 cycles and 492 mAh g−1 at 0.5 A g−1 after
Li-ion batteries
500 cycles) and good rate capability (76% capacity retention at 10 A g−1 compared with the capacity at
0.1 A g−1). The improved electrochemical performance is primarily associated with the TiO2-C hybrid matrix
that efficiently surrounds the ZnTe nanocrystallites, thereby providing high electric conductivity and at the same
time functioning as a buffering medium that alleviates the large volume change during extended cycling.
Furthemore, the practical application of ZnTe-TiO2-C as an anode is investigated with full cell consisting of a
ZnTe-TiO2-C anode and a LiFePO4-graphite (LFP@G) cathode. The ZnTe-TiO2-C//LFP@G full cell exhibits good
cyclic stability with 79% capacity retention at 0.1 A g−1 after 200 cycles. These results suggest the ZnTe-TiO2-C
nanocomposite is a promising candidate for a high-performance anode in next-generation LIBs.

1. Introduction anode material because it is naturally abundant, eco-friendly and in-


expensive. Additionally, Zn possesses such advantages as an extremely
Recently, the rapid advances in portable electronics and electric high volumetric capacity (2927 mAh cm−3) and a moderate working
vehicles have necessitated the development of next-generation energy potential (~0.4 V vs. Li/Li+). However, the practical application of Zn
storage devices with high energy densities and enhanced safety. In this alone still causes the huge volume expansions and the pulverization of
regard, Li-ion batteries (LIBs) are one of the most promising recharge- active particles during Li alloying/dealloying, resulting in poor elec-
able battery systems because of their high energy and power densities, trochemical performance, as with other Li-alloying materials [22–27].
long cycle lives, and excellent rate capabilities [1–4]. Despite the low One of the effective approaches for resolving this problem is the in-
cost and good cyclic stability of commercial graphite, it has a limited troduction of appropriate mechanical buffer to the active composites,
energy density owing to its relatively low theoretical gravime- which can efficiently absorb the strain generated from cycling and al-
triccapacity of 372 mAh g−1, low volumetric capacity resulting from leviate particle agglomeration. Titanium oxides have been considered
low tap density (< 1 g cm−3), and poor rate capability, as well as the as supplemental materials for preventing the pulverization of the alloy
safety issues (Li plating on graphite surfaces). Thus, enormous attempts anode in LIBs owing to their structural stability [28–32]. Furthermore,
have been implemented to develop alternative anode materials to re- rutile TiO2 can react with Li reversibly as an active material [33–39].
place graphite anodes [5–10]. The lithiation/delithiation reaction of TiO2 can be summarized as fol-
In recent years, Li-alloying materials have been proposed as new lows.
high-performance LIB anodes because of their high theoretical capa-
TiO2 + xLi+ + xe− ↔ Lix TiO2 (0 ≤ x ≤ 1) (1)
cities resulting from a large number of Li+ ions during cycling and their
high reaction potentials, which mitigate safety issue associated with Li In addition, conductive C coating has been regarded as a strategy to
plating [11–21]. Zn has been extensively investigated as a potential improve the electrochemical performance of LIBs.


Corresponding author.
E-mail address: jhhur@gachon.ac.kr (J. Hur).
1
Q.H. Nguyen and Q.H. Nguyen contributed equally to this work.

https://doi.org/10.1016/j.apsusc.2019.144718
Received 19 September 2019; Received in revised form 11 November 2019; Accepted 15 November 2019
Available online 28 November 2019
0169-4332/ © 2019 Elsevier B.V. All rights reserved.
Q.H. Nguyen, et al. Applied Surface Science 509 (2020) 144718

Another strategy to prevent the high volume expansion of Zn during nanocomposite from the metastable ZnO-Te-Ti alloy was completed,
cycling is the combination of zinc with other metallic components. without impurities. The overall chemical reaction is expressed as fol-
Telluride is a great candidate in alloying with Zn because it can react lows.
with Li+ to form Li2Te (~1.9 V), with a reaction potential far from that Annealing
of Zn (~0.4 V), implying that progressive electrochemical reactions ZnO + Te + Ti → ZnTe + TiO2 + metastable ZnO − Te − Ti
occur during cycling [14,19]. Therefore, the incorporation of Te into HEMM
(+C )
Zn, forming intermetallic ZnTe, can be an effective strategy for alle- → ZnTe − TiO2 − C (2)
viating the abrupt volume expansion upon prolonged cycling. Fur-
thermore, it has an extremely high theoretical volumetric capacity of
~2621 mAh cm−3 and a higher electric conductivity (2 × 10- 2.2. Material characterization
4
MS m−1) than other elements in the chalcogen group (Se, 1 × 10-
10
MS m−1; S, 5 × 10-22 MS m−1) which can enhance the electro- The crystalline structures of the as-synthesized powders were
chemical performance of Zn-based anode in LIBs. characterized using an X-ray diffraction (XRD; D/MAX-2200 Rigaku,
In this study, we synthesize a ZnTe-TiO2-C composite via a combi- Japan) with Cu Kα radiation (λ = 1.5406 nm) (scan rate: 2° min−1, 2θ
nation of annealing and high-energy mechanical milling (HEMM). The range: 20°–80°). To understand the reaction mechanism of ZnTe-TiO2-C
bimetallic ZnTe alloy nanocrystallites were embedded in a conductive anode, ex -situ XRD analysis was performed for ZnTe-TiO2-C electrodes
hybrid matrix of TiO2 and C, which yielded excellent electrochemical at pristine, fully-discharged, and fully charged states. X-ray photoelec-
performance (reversible and stable cycle life and good rate capability). tron spectroscopy (XPS, Kratos AXIS Nova) was performed to in-
The merits of ZnTe-TiO2-C nanocomposite as a high-performance anode vestigate the chemical and bonding states of the composite. The mi-
are regarded as follows: (i) the distinct reaction potentials between zinc croscopic morphologies of the as-prepared powder samples were
and telluride with Li+ allows the sequential electrochemical lithiation/ analyzed using scanning electron microscopy (SEM, Hitachi S4700,
delithiation and reduces the abrupt volume expansion upon cycling; (ii) Japan) and high-resolution transmission electron microscopy (HRTEM,
the introduction of Te and the conductive C for Zn improves the electric JEOL JEM-2100F) analyses. The elemental compositions of the as-
conductivity; and (iii) the conductive TiO2-C hybrid matrix acts as a synthesized samples were analyzed using scanning transmission elec-
buffering medium that suppresses the huge volume expansion of active tron microscopy (STEM) equipped with energy-dispersive X-ray spec-
ZnTe and prevents particle agglomeration during cycling. troscopy (EDS). Ex-situ SEM analysis was conducted to examine the
different morphologies of the electrode materials for the disassembled
2. Experimental section cell after cycling.

2.1. Material preparation 2.3. Electrochemical measurements

To prepare the ZnTe-TiO2-C nanocomposite, ZnO powder CR2032 coin cells were assembled in a glove box using an anode as
(< 50 nm, > 97%, Sigma–Aldrich), Te powder (200 mesh, 99.8%, a working electrode, a metallic Li foil as a counter electrode, 1 M LiPF6
Sigma-Aldrich), and Ti powder (325 mesh, 99.99%, Alfa Aesar) were solution dissolved in ethylene carbonate/diethylene carbonate (EC/
mixed in a mortar with a molar ratio of 2:2:1. Then, the resulting DEC, 1:1, v:v) as an electrolyte, and polyethylene (PE) as a separator.
mixtures were annealed under a flowing Ar atmosphere at 600 °C for The working electrodes were prepared by mixing the as-prepared
8 h. Subsequently, acetylene carbon black powder was added in such a powders (70 wt%), conductive Super P carbon black (15 wt%), and a
way that the mass ratio of carbon black was 30 wt% of the total weight binder (15 wt% polyvinylidene fluoride (PVDF)) in an N-methyl-pyr-
to the annealed ZnO-Te-Ti alloy powder with and manually ground. rolidone (NMP) solvent. These slurry mixtures were then coated onto
These powder mixtures were then transferred in a ZrO2 bowl (80 cm3) Cu foils via the doctor-blade technique and vacuum-dried overnight at
in which two different sized hardened ZrO2 balls (diameters of 3/8 and 70 °C. The electrochemical performance was measured within the vol-
3/16 in.) were placed along with powder at the mass ratio of 1:20 tage range of 0.01–2.5 V vs. Li/Li+ galvanostatically at a current
(powder:balls). The bowl was tightly sealed under an Ar atmosphere in density of 0.1 A g−1 with a WBCS3000 battery cycler (WonATech).
a glove box, and HEMM was operated at 300 rpm for 20 h utilizing a Cyclic voltammetry (CV) was carried out at a scan rate of 0.2 mV s−1
planetary mechanical milling machine (Pulverisette 5, Fritch). For using a ZIVE MP1 (WonATech) machine. The measurement of rate
comparison, a ZnTe-TiO2 sample was separately synthesized via the capability was performed at a current density range of 0.1–10 10 A g−1.
same procedure without the addition of carbon black at the milling Electrochemical impedance spectroscopy (EIS) was measured in the
step. Another control sample (ZnTe-C) was synthesized using Zn, Te, frequency range of 100 kHz–100 mHz using a ZIVE MP1 analyzer
and C powders, with a weight ratio of 7:3 (ZnTe:C). The overall syn- (WonATech). For full-cell electrochemical measurements, LiFePO4-
thetic procedure for producing the ZnTe-TiO2-C composite is illustrated graphite (LFP@G) electrode was synthesized as a cathode material
in Fig. 1. During the first annealing step, the powder mixtures of ZnO, using one-step HEMM; The powders of LiFePO4 (< 5 µm, > 97%,
Te, and Ti were mostly converted into the ZnTe alloy and the TiO2 Sigma-Aldrich) and graphite (200 mesh, 99.9999%, Alfa Aesar) were
phase, along with the metastable ZnO-Te-Ti alloy. In the second HEMM first mixed and manually ground at a mass ratio of 7:3. Subsequently,
process with C addition, the formation of the ZnTe-TiO2-C the HEMM was applied for the mixture under an Ar atmosphere for 10 h

Fig. 1. Overall synthesis process for the ZnTe-TiO2-C composite.

2
Q.H. Nguyen, et al. Applied Surface Science 509 (2020) 144718

the NMP solvent). To enhance the stability of the full cell, the cells were
first pre-charged for cathode and pre-discharged for anode to stabilize
the solid–electrolyte interphase (SEI) layer. GCD cycling measurements
were performed with a WBC3000 battery cycler at 0.1 A g−1 in the
potential range of 1.0–3.8 V. The specific capacity of full cell was de-
termined based on the weight including both active anode and cathode
material, with their weight ratio of ~1:2 (anode:cathode).

3. Results and discussions

3.1. Characterization of synthesized samples

The XRD patterns of the ZnTe-TiO2-C, ZnTe-TiO2, and ZnTe-C


composites are shown in Fig. 2. The XRD patterns of all these samples
exhibited the peaks of crystalline ZnTe, which well match the reference
peaks (PDF#15-0746), suggesting the successful formation of the ZnTe
phase. For the ZnTe-TiO2-C and ZnTe-TiO2 nanocomposites, additional
peaks corresponding to the crystalline structure of rutile TiO2 (PDF#21-
1276) were observed at ~27.6°, 36.1°, 54.3°, and 76.5°, which are as-
signed to the (1 1 0), (1 0 1), (2 1 1), and (2 0 2) planes, respectively.
Although small peaks corresponding to ZnO, Ti, and Te phases were
observed after the first heat treatment (Fig. S1), these phases were
completely transformed into ZnTe and TiO2 phases after the HEMM
(Fig. 2). In all cases, no other peaks were detected, suggesting that there
are no impurities formed during the milling process. The Raman spectra
shown in Fig. S2 verify the presence of C in the ZnTe-TiO2-C and ZnTe-C
composites. The first peak at ~1329 cm−1 is related to the Dl.-band of
the disordered C, and another peak at ~1584 cm−1 corresponds to the
G-band of the graphitized C [40]. For ZnTe-TiO2-C and ZnTe-TiO2, the
Fig. 2. XRD patterns of the ZnTe-C, ZnTe-TiO2, and ZnTe-TiO2-C composites. characteristic peaks of rutile TiO2 were observed at ~144, 445, and
610 cm−1, which are designated to the peaks of B1g, Eg, and A1g, re-
spectively. Additionally, the broad peak at ~235 cm−1 is ascribed to
at 300 rpm. A coin-type (CR2032) full cells were prepared using ZnTe-
the second-order or disorder-induced scattering [41].
TiO2-C electrode (anode) and LFP@G electrode (cathode). PE was used
XPS measurement was performed to investigate the chemical
as the separator while the electrolyte was identical to that in half-cell.
bonding state and elemental composition of the ZnTe-TiO2-C nano-
The slurry for the cathode material was prepared by mixing 80 wt% of
composite (Fig. 3). In the Zn 2p spectrum, Zn 2p1/2 and Zn 2p3/2 peaks
LFP@G, 10 wt% of Super P, and 10 wt% of binder (PVDF dissolved in
were observed at ~1044.8 and 1021.7 eV, respectively (Fig. 3a). The

Fig. 3. XPS spectra of the ZnTe-TiO2-C composite.

3
Q.H. Nguyen, et al. Applied Surface Science 509 (2020) 144718

Fig. 4. (a) SEM image, (b, c) HRTEM images, and (d) STEM image with EDS mapping images of the ZnTe-TiO2-C composite.

XPS peaks at ~583.9 and 573.1 eV in Fig. 3b are attributed to Te 3d5/2 discharge and charge specific capacities of 839 and 558 mAh g−1, re-
and Te 3d3/2, respectively, confirming the reduction of ZnO to metallic spectively, with the corresponding initial coulombic efficiency (CE) of
Zn and the successful formation of the ZnTe alloy after the HEMM [42]. ~66.5%. It is thought that the irreversible capacity loss can be ascribed
The XPS peaks at ~464.4 and 457.9 eV can be indexed to Ti 2p1/2 and to the electrolyte decomposition and the SEI layer formation in the first
Ti 2p3/2, respectively (Fig. 3c), and the O 1s peak at ~532.1 eV is as- lithiation process. However, the ZnTe-TiO2-C electrode, after the first
cribed to the O 1s electron binding energy for TiO2 (Fig. 3d) [43]. In the cycle, exhibited excellent capacity retention, with a gradual increase in
C 1s spectrum, the peaks at ~288.9, 285.4, and 284.4 eV, corre- the CE (94.8%, 98.1%, and 99.11% for the 2nd, 10th, and 100th cycles,
sponding to the bonds of C]O, CeO, and CeC, respectively (Fig. 3e), respectively). To investigate the Li-ion storage mechanism, CV was
which can be explained by the partial transformation of the CeC bonds measured on the ZnTe-TiO2-C nanocomposite for the initial three cycles
in the amorphous C after the HEMM [44]. The full XPS spectra in Fig. 3f from 0.01 V to 2.5 V at a scanning rate of 0.1 mV s−1 (Fig. 5b). During
confirm the presence of all the elements (Zn, Te, Ti, O, and C) in the the initial discharge process, the first reduction peak at ~1.1 V was
ZnTe-TiO2-C composite, indicating the successful formation of ZnTe, detected, indicating the alloying reaction of rutile TiO2 to form the
TiO2, and C after the HEMM. Li0.5TiO2 phase, and the broad peak at ~0.9 V is assigned to the li-
The microscopic morphology of the ZnTe-TiO2-C nanocomposite thiation reaction of Te, forming the Li2Te phase [45,46]. The next two
were characterized via SEM, HRTEM, and elemental distribution peaks at ~0.6 and < 0.4 V can be ascribed to the formation of the SEI
through STEM together with the EDS mapping. Fig. 4a and S3 show film on the surface of the electrode and the progressive lithiations of Zn
SEM images of the ZnTe-TiO2-C, ZnTe-TiO2, and ZnTe-C composites. On that sequentially form various phases of Li-Zn (Zn → LiZn 4→LiZn) ,
average, the particle sizes of as-prepared samples were in the range of respectively [20,22]. Upon the first charging cycle, a series of small
hundreds of nanometers –a few micrometers that are much smaller than peaks were observed below 0.7 V, which are ascribed to the series of
pristine powder sizes (owing to the continuous collision and fracture delithiation reactions from LiZn to metallic Zn
during the HEMM. The HRTEM images presented in Fig. 4b and c in- (LiZn → Li2 Zn3 → LiZn2 → LiZn 4→Zn) [7,20]. The subsequent oxida-
dicate the presence of nanosized ZnTe and TiO2 crystallites (darker tion peak detected at ~1.25 V corresponds to the transformation of
areas with sizes of 5–10 nm) dispersed into a carbon matrix without Li2Te into the ZnTe phase, and the broad peak at ~2.2 V is associated
aggregation. The interplanar distance of 0.352 nm is assigned to the with the delithiation reaction of Li0.5TiO2, which formed the TiO2 phase
(1 1 1) plane of ZnTe, and the two interplanar spacings of 0.248 and [45,46]. In the subsequent cycles, while the SEI peaks at ~0.6 V dis-
0.324 nm can be indexed to the (1 0 1) and (1 1 0) planes, respectively, appeared, all the lithiation/delithiation peaks in the 2nd and 3rd cycles
of the TiO2 phase. These results agree well with the foregoing XRD remained nearly unchanged, suggesting the good reversibility of the
analysis. The selected-area electron diffraction (SAED) patterns (insets ZnTe-TiO2-C electrode. The overall electrochemical reaction me-
in Fig. 4b and c) confirm the presence of crystalline ZnTe (1 1 1) and chanism of the ZnTe-TiO2-C electrode can be written as follows.
TiO2 ((1 0 1) and (1 1 0)), according to the interplanar spacings and During the discharge process:
lattice planes. EDS mapping images in Fig. 4d reveal the uniform dis-
TiO2 + 0.5Li+ + 0.5e− → Li 0.5TiO2 ( 1.1 V vs. Li/Li + ) (3)
persion of ZnTe and TiO2 within the C matrix in the composite. Clearly,
the overlapped signal from Zn and Te as well as that from Ti and O ZnTe + 2Li+ + 2e− → Zn + Li2 Te ( 0.9 V vs. Li/Li + ) (4)
further confirms the successful formation of the ZnTe alloy and TiO2
phase. All these structural characterizations evidence that the ZnTe- Zn → LiZn 4 → LiZn ( < 0.4 V vs. Li/Li + ) (5)
TiO2-C is composed of ZnTe and TiO2 nanocrystals well-distributed in a During the charge process:
conductive C matrix which provides a strong buffering effect by sup-
LiZn → Li2 Zn3 → LiZn2 → LiZn 4 → Zn ( < 0.7 V vs. Li/Li + ) (6)
pressing the large stress from ZnTe crystallites, and increase the con-
ductivity of the composite. Zn + Li2 Te → ZnTe + 2Li+ + 2e− ( 1.25 V vs. Li/Li + ) (7)

Li 0.5TiO2 → TiO2 + 0.5Li+ + 0.5e− ( 2.2 V vs. Li/Li + ) (8)


3.2. Electrochemical performance of ZnTe-TiO2-C half-cell
To clarify the alloying/dealloying reaction mechanisms of the ZnTe-
Fig. 5a presents the charge/discharge voltage profiles for the 1st, TiO2-C electrode, ex-situ XRD experiments were carried out, as shown in
2nd, 10th, and 100th cycles of the ZnTe-TiO2-C nanocomposite elec- Fig. 5c. The electrochemical cells were disassembled at three potential
trode from the GCD measurement at 0.1 A g−1 within the potential points, as marked by the arrows in Fig. 5b, corresponding to the as-
range of 0.01–2.5 V. The ZnTe-TiO2-C electrode showed the initial prepared, completely discharged, and completely charged states. For

4
Q.H. Nguyen, et al. Applied Surface Science 509 (2020) 144718

Fig. 5. (a) Voltage profiles; (b) CV curves; (c) ex situ XRD pattern of the ZnTe-TiO2-C electrode for the first cycle; (d) schematic illustration of the lithiation/
delithiation reaction mechanism.

the pristine electrode, the crystalline phases of ZnTe and TiO2, along 611 mAh g−1 over 300 cycles, corresponding to 95% capacity retention
with a prominent peak corresponding to Cu foil, were observed, which (compared with its capacity at the second cycle). This was mainly due
agrees well with the XRD results in Fig. 2. At fully discharged state to the homogeneous dispersion of nanocrystalline ZnTe in the efficient
(0.01 V), several new peaks corresponding to Li0.5TiO2 (at ~24°, 26°, buffering hybrid matrix of the TiO2 and the amorphous C matrix. A
37°, and 44°), Li2Te (at ~23°, 27°, 39°, and 46°) and LiZn (at ~41° and slight increase in capacity was observed with the increasing cycle
48°) were observed, while the peaks corresponding to ZnTe and TiO2 number, which is ascribed to the late activation of embedded active
disappeared, indicating the complete conversion of ZnTe and TiO2. At particles in the composites [31,32]. The TiO2 functioned not only as an
completely charged state (2.5 V), all the peaks corresponding to LiZn, active phase but also as a supplementary buffering material [28]. As
Li2Te, and Li0.5TiO2 vanished, and new ZnTe and TiO2 peaks were shown in Fig. 6b, the long-term cyclic performance of these electrodes
observed, indicating the complete reversibility of the active ZnTe and measured at 0.5 A g−1 exhibited a similar trend to the cyclic perfor-
TiO2 without structural changes upon the extended cycling process. The mance measured at 0.1 A g−1 although the capacities were generally
detailed electrochemical mechanism is schematically presented in reduced. The ZnTe-TiO2-C electrode displayed the best performance; a
Fig. 5d. specific capacity of 492 mAh g−1 at 500 cycles, which was much higher
Fig. 6a compares the cyclic performance of the ZnTe-TiO2-C, ZnTe- than those of the ZnTe-TiO2 and ZnTe-C electrodes (139 and
TiO2, and ZnTe-C electrodes at 0.1 A g−1 current density, and Fig. 6b 382 mAh g−1, respectively). The superior cyclic performance of ZnTe-
illustrates the long-term cyclic performance of these electrodes mea- TiO2-C is ascribed to the exisitence of the TiO2-C hybrid matrix into the
sured at a high current density (0.5 A g−1). As shown in Fig. 6a, among ZnTe alloy, which reduced the stresses and strains of ZnTe particles and
the electrodes, ZnTe-TiO2 exhibited the lowest specific capacity of prevented the aggregation of active ZnTe during continuous electro-
~123 mAh g−1 after 300 cycles, which can be attributed to the large chemical reactions. Additionally, the excellent structural stability of
volume changes of the active particles during the lithiation/delithiation TiO2 with very small volume changes (~3%) during the lithiation/de-
processes, which resulted in the particles pulverization and the loss of lithiation processes contributed to the cyclic stability of the composite.
electrical contact between the current collector and the active particles The rate capabilities at different current densities from 0.1 A g−1 to
upon extended cycling. The ZnTe-C electrode exhibited a significantly 10 A g−1 are shown for three electrodes, as presented in Fig. 6c. The
higher capacity than ZnTe-TiO2 in the early stage of cycling ZnTe-TiO2 and ZnTe-C electrodes exhibited relatively poor rate cap-
(472 mAh g−1 until 70 cycles). This can be interpreted by the effect of ability, delivering low capacities of 134 and 302 mAh g−1, respectively,
the C-matrix addition to the ZnTe alloy, which facilitated the lithiation/ at 10 A g−1. By contrast, the ZnTe-TiO2-C electrode presented an ex-
delithiation processes. However, its capacity decreased significantly to cellent rate performance with the higher specific capacities at all cur-
311 mAh g−1 over 200 cycles. The introduction of the hybrid buffering rent densities, delivering specific charge capacities of ~539, ~501,
matrix of TiO2 and C to the ZnTe composite resulted in considerable ~487, ~455, and ~435 mAh g−1, respectively, at 0.1, 0.5, 1, 3, and
enhancements in the specific capacity and cyclic stability. Specifically, 5 A g−1. In addition, ZnTe-TiO2-C retained a high specific capacity of
the ZnTe-TiO2-C electrode delivered a highly stable specific capacity of ~406 mAh g−1 even at high current density of 10 A g−1, corresponding

5
Q.H. Nguyen, et al. Applied Surface Science 509 (2020) 144718

Fig. 6. (a) Cyclic performances at 0.1 A g−1; (b) long-term cyclic performance at 0.5 A g−1; (c) rate capability; (d) EIS spectra of the ZnTe-TiO2-C, ZnTe-TiO2, and
ZnTe-C electrodes after 50 cycles.

to ~76% of its initial capacity at the current density of 0.1 A g−1. This reaction potential between cathode and anode. The specific capacity of
good rate performance suggests the facile Li+ ion transport in the ZnTe- the ZnTe-TiO2-C//LFP@G full cell was evaluated based on the sum-
TiO2-C electrode, which are associated with the homogeneous dis- mation of cathode and anode mass. The specific capacity of the full cell
tribution of nanosized ZnTe in the conductive TiO2-C matrix preventing was measured to be as high as ~81 mAh g−1 even after 200 cycles
the aggregation of ZnTe particles under fast charge/discharge cycling. (Fig. 7b). Remarkably, after the first cycle, the CEs of the full cell were
To evaluate the enhanced electrical conductivity of ZnTe-TiO2-C com- ~100%, without significant capacity fading until 200 cycles. The rate
pared with the ZnTe-TiO2 and ZnTe-C electrodes, EIS measurements performance measurement showed the specific capacities of the ZnTe-
were conducted for these three samples after 50 cycles at 0.1 A g−1, as TiO2-C//LFP@G full cell as ~97, 65, 53, and 38 mAh g−1 at 0.1, 0.5, 1,
shown in Fig. 6d. Among the three electrodes, ZnTe-TiO2-C exhibited and 3 A g−1, respectively (Fig. 7c). Upon a series of different current
the smallest semicircles and the lowest charge transfer resistance (Rct) density and resuming to the original current density of 0.1 A g−1, the
after 50th cycle (detailed data are presented in Table S1). These results recovery of capacity retention was as high as ~84% (~81 mAh g−1),
suggest that the TiO2-C introduction as a hybrid buffering matrix pro- representing the excellent reversibility and good rate performance of
vided efficient electron-transport pathways for the active particles and the ZnTe-TiO2-C//LFP@G full cell. The energy densities of the full cell
stabilized the thin SEI layer, enhancing the electric conductivity of the for ZnTe-TiO2-C//LFP@G estimated using the operational voltage of
composite. 2.5 V at different current rates were ~242, 163, 132, and 95 Wh kg−1,
respectively, at various current densities of 0.1, 0.5, 1, and 3 A g−1
(Fig. 7d). Notably, the energy density of the full cell (ZnTe-TiO2-C//
3.3. Electrochemical performance of ZnTe-TiO2-C//LiFePO4-G full cell LFP@G) measured at 0.1 A g−1 exceeded that of a previously reported
Li4Ti5O12//LiFePO4 full cell (energy density of ~140 Wh kg−1) [47].
To investigate the practical application of the ZnTe-TiO2-C nano-
composites in a LIB system, we constructed a coin-type full-cell LIB
using LFP@G and ZnTe-TiO2-C as the cathode and anode, respectively, 4. Conclusion
as shown in Fig. 7. The electrochemical performance of the full cell was
measured in the potential range of 1.0–3.8 V at 0.1 A g−1. The mass We demonstrated a ZnTe-TiO2-C nanocomposite prepared via a
ratio between the cathode and anode was controlled to be ~1:2, con- simple and low-cost HEMM as a high-performance anode for LIBs. The
sidering their capacity balance. Fig. S4 shows a schematic principle of ZnTe-TiO2-C nanocomposite comprising of intermetallic ZnTe uni-
ZnTe-TiO2-C//LFP@G full-cell LIB operation. Briefly, Li+ ions transfer formly distributed in a multifunctional hybrid matrix of rutile TiO2 and
from ZnTe-TiO2-C anode to LFP@G cathode during the discharge pro- C exhibited great electrochemical performance because of many bene-
cess, while Li+ ions move back from LFP@G to ZnTe-TiO2-C upon ficial features including microstructure that is efficient to Li+ ion
charge process. As shown in Fig. 7a, the charge/discharge potential transport, high electronic conductivity, and the hybrid buffering effect
profile of the ZnTe-TiO2-C//LFP@G full cell displayed an apparent cell of the rutile TiO2 and amorphous C matrix for active ZnTe during cy-
potential at ~2.5 V, corresponding to the difference in the main cling. Compared with ZnTe-TiO2 and ZnTe-C electrodes, the ZnTe-TiO2-

6
Q.H. Nguyen, et al. Applied Surface Science 509 (2020) 144718

Fig. 7. Full-cell performance obtained using ZnTe-TiO2-C as an anode and LFP@G as a cathode: (a) GCD profiles at 0.1 A g−1 over the voltage range of 1–3.8 V; (b)
cyclic performance at 0.1 A g−1; (c) rate capability; and (d) energy density of the ZnTe-TiO2-C//LFP@G full cell at different current rates.

C electrode showed a highly reversible capacity of 611 mAh g−1 until Appendix A. Supplementary material
300 cycles, with an excellent capacity retention of 95%, and an superior
rate performance at high current density of 10 A g−1 (76% capacity Supplementary data to this article can be found online at https://
retention of its capacity at 0.1 A g−1). Furthermore, the ZnTe-TiO2-C doi.org/10.1016/j.apsusc.2019.144718.
electrode delivered a good specific capacity of ~492 mAh g−1 over 500
cycles even at 0.5 A g−1. Furthemore, the electrochemical reaction References
mechanism revealed by ex situ XRD analysis ensured the progressive
lithiation/delithiation processes of the active ZnTe and TiO2 phase [1] M. Armand, J.M. Tarascon, Building better batteries, Nature 451 (2008) 652–657.
during cycling. Overall, ZnTe-TiO2-C nanocomposite can be an alter- [2] J.M. Tarascon, M. Armand, Issues and challenges facing rechargeable lithium bat-
teries, Nature 414 (2011) 171–179.
native anode material for high-performance rechargeable LIBs. [3] S. Goriparti, E. Miele, F. De Angelis, E. Di Fabrizio, R.P. Zaccaria, C. Capiglia,
Review on recent progress of nanostructured anode materials for Li-ion batteries, J.
Power Sources 257 (2014) 421–443.
CRediT authorship contribution statement [4] V. Etacheri, R. Marom, R. Elazari, G. Salitra, D. Aurbach, Challenges in the devel-
opment of advanced Li-ion batteries: a review, Energy Environ. Sci. 4 (2011)
Quoc Hanh Nguyen: Methodology, Investigation, Data curation, 3243–3262.
[5] N.T. Hung, J. Bae, J.H. Kim, H.B. Son, I.T. Kim, J. Hur, Facile preparation of a zinc-
Writing - original draft. Quoc Hai Nguyen: Conceptualization, based alloy composite as a novel anode material for rechargeable lithium-ion bat-
Investigation, Visualization. Jaehyun Hur: Conceptualization, Writing teries, Appl. Surf. Sci. 429 (2018) 210–217.
- review & editing, Supervision, Project administration, Funding ac- [6] N.Q. Hai, S.H. Kwon, H. Kim, I.T. Kim, S.G. Lee, J. Hur, High-performance MoS2-
based nanocomposite anode prepared by high-energy mechanical milling: The ef-
quisition. fect of carbonaceous matrix on MoS2, Electrochim. Acta 260 (2018) 129–138.
[7] Q.H. Nguyen, N.T. Hung, S.J. Park, I.T. Kim, J. Hur, Enhanced performance of
carbon-free intermetallic zinc titanium alloy (Zn-ZnxTiy) anode for lithium-ion
Declaration of Competing Interest batteries, Electrochim. Acta 301 (2019) 229–239.
[8] Q.H. Nguyen, J.S. Choi, Y.-C. Lee, I.T. Kim, J. Hur, 3D hierarchical structure of
The authors declare that they have no known competing financial MoS2@ G-CNT combined with post-film annealing for enhanced lithium-ion sto-
rage, J. Ind. Eng. Chem. 69 (2019) 116–126.
interests or personal relationships that could have appeared to influ-
[9] N.Q. Hai, H. Kim, I.S. Yoo, J.H. Kim, J. Hur, Comparative study of mechanically
ence the work reported in this paper. milled MoS2 and MoSe2 in graphite matrix as anode materials for high-performance
lithium-ion batteries, J. Nanosci. Nanotechnol. 18 (2018) 6469–6474.
[10] N.Q. Hai, H. Kim, I.S. Yoo, J. Hur, Facile and scalable preparation of a MoS2/carbon
Acknowledgments nanotube nanocomposite anode for high-performance lithium-ion batteries: effects
of carbon nanotube content, J. Nanosci. Nanotechnol. 19 (2019) 1494–1499.
[11] W.J. Zhang, A review of the electrochemical performance of alloy anodes for li-
This work was supported by the Korea Institute of Energy
thium-ion batteries, J. Power Sources 196 (2011) 13–24.
Technology Evaluation and Planning (KETEP) and the Ministry of [12] M. Obrovac, V. Chevrier, Alloy negative electrodes for Li-ion batteries, Chem. Rev.
Trade, Industry & Energy (MOTIE) of the Republic of Korea (No. 114 (2014) 11444–11502.
20194030202290). It was also supported by the Korea Electric Power [13] J. Leibowitz, E. Allcorn, A. Manthiram, SnSb–TiC–C nanocomposite alloy anodes for
lithium-ion batteries, J. Power Sources 279 (2015) 549–554.
Corporation (grant number: R18XA02).

7
Q.H. Nguyen, et al. Applied Surface Science 509 (2020) 144718

[14] S.Y. Son, J. Hur, K.H. Kim, H.B. Son, S.G. Lee, I.T. Kim, SnTe-TiC-C composites as reaction mechanism of Zn-TiOx-C nanocomposite anode materials for Li secondary
high-performance anodes for Li-ion batteries, J. Power Sources 365 (2017) batteries, J. Electrochem. Soc. 164 (2017) A2683–A2688.
372–379. [33] Y.S. Hu, L. Kienle, Y.G. Guo, J. Maier, High lithium electroactivity of nanometer-
[15] Q.H. Nguyen, J. Hur, MoS2-TiC-C nanocomposites as new anode materials for high- sized rutile TiO2, Adv. Mater. 18 (2006) 1421–1426.
performance lithium-ion batteries, J. Nanosci. Nanotechnol. 19 (2019) 996–1000. [34] D. Wang, D. Choi, Z. Yang, V.V. Viswanathan, Z. Nie, C. Wang, Y. Song, J.G. Zhang,
[16] C.M. Park, J.H. Kim, H. Kim, H.J. Sohn, Li-alloy based anode materials for Li sec- J. Liu, Synthesis and Li-ion insertion properties of highly crystalline mesoporous
ondary batteries, Chem. Soc. Rev. 39 (2010) 3115–3141. rutile TiO2, Chem. Mater. 20 (2008) 3435–3442.
[17] I.T. Kim, S.-O. Kim, A. Manthiram, Effect of TiC addition on SnSb–C composite [35] E. Baudrin, S. Cassaignon, M. Koelsch, J.P. Jolivet, L. Dupont, J.M. Tarascon,
anodes for sodium-ion batteries, J. Power Sources 269 (2014) 848–854. Structural evolution during the reaction of Li with nano-sized rutile type TiO2 at
[18] E. Allcorn, A. Manthiram, High-rate, high-density FeSb–TiC–C nanocomposite an- room temperature, Electrochem. Commun. 9 (2007) 337–342.
odes for lithium-ion batteries, J. Mater. Chem. A 3 (2015) 3891–3900. [36] M.A. Reddy, M.S. Kishore, V. Pralong, V. Caignaert, U. Varadaraju, B. Raveau,
[19] H. Kim, M. Kim, Y.H. Yoon, Q.H. Nguyen, I.T. Kim, J. Hur, S.G. Lee, Sb2Te3-TiC-C Room temperature synthesis and Li insertion into nanocrystalline rutile TiO2,
nanocomposites for the high-performance anode in lithium-ion batteries, Electrochem. Commun. 8 (2006) 1299–1303.
Electrochim. Acta 293 (2019) 8–18. [37] C. Jiang, I. Honma, T. Kudo, H. Zhou, Nanocrystalline rutile TiO2 electrode for high-
[20] S.O. Kim, A. Manthiram, High-performance Zn-TiC-C nanocomposite alloy anode capacity and high-rate lithium storage, Electrochem. Solid State Lett. 10 (2007)
with exceptional cycle life for lithium-ion batteries, ACS Appl. Mater. Interfaces 7 A127–A129.
(2015) 14801–14807. [38] C.M. Park, W.S. Chang, H. Jung, J.H. Kim, H.J. Sohn, Nanostructured Sn/TiO2/C
[21] D. Applestone, A. Manthiram, Cu6Sn5–TiC–C nanocomposite alloy anodes with high composite as a high-performance anode for Li-ion batteries, Electrochem. Commun.
volumetric capacity for lithium ion batteries, RSC Adv. 2 (2012) 5411–5417. 11 (2009) 2165–2168.
[22] Y. Hwa, J.H. Sung, B. Wang, C.-M. Park, H.-J. Sohn, Nanostructured Zn-based [39] X. Yang, Z. Wen, X. Xu, B. Lin, S. Huang, Nanosized silicon-based composite derived
composite anodes for rechargeable Li-ion batteries, J. Mater. Chem. 22 (2012) by in situ mechanochemical reduction for lithium ion batteries, J. Power Sources
12767–12773. 164 (2007) 880–884.
[23] H.T. Kwon, C.M. Park, Electrochemical characteristics of ZnSe and its nanos- [40] Q.H. Nguyen, I.T. Kim, J. Hur, Core-shell Si@ c-PAN particles deposited on graphite
tructured composite for rechargeable Li-ion batteries, J. Power Sources 251 (2014) as promising anode for lithium-ion batteries, Electrochim. Acta 297 (2018)
319–324. 355–364.
[24] S. Saadat, Y.Y. Tay, J. Zhu, P.F. Teh, S. Maleksaeedi, M.M. Shahjamali, [41] L. Kernazhitsky, V. Shymanovska, T. Gavrilko, V. Naumov, L. Fedorenko,
M. Shakerzadeh, M. Srinivasan, B.Y. Tay, H.H. Hng, Template-free electrochemical V. Kshnyakin, J. Baran, Laser-excited excitonic luminescence of nanocrystalline
deposition of interconnected ZnSb nanoflakes for li-ion battery anodes, Chem. TiO2 powder, Ukr. J. Phys. 59 (2014) 248–255.
Mater. 23 (2011) 1032–1038. [42] Y.J. Jang, J.W. Jang, J. Lee, J.H. Kim, H. Kumagai, J. Lee, T. Minegishi, J. Kubota,
[25] Z. Zhang, Y. Fu, X. Yang, Y. Qu, Q. Li, Nanostructured ZnSe anchored on graphene K. Domen, J.S. Lee, Selective CO production by Au coupled ZnTe/ZnO in the
nanosheets with superior electrochemical properties for lithium ion batteries, photoelectrochemical CO2 reduction system, Energy Environ. Sci. 8 (2015)
Electrochim. Acta 168 (2015) 285–291. 3597–3604.
[26] C.M. Park, H.J. Sohn, Quasi-intercalation and facile amorphization in layered ZnSb [43] L. Zhu, Q. Lu, L. Lv, Y. Wang, Y. Hu, Z. Deng, Z. Lou, Y. Hou, F. Teng, Ligand-free
for li-ion batteries, Adv. Mater. 22 (2010) 47–52. rutile and anatase TiO2 nanocrystals as electron extraction layers for high perfor-
[27] M.G. Park, C.K. Lee, C.M. Park, Amorphized ZnSb-based composite anodes for high- mance inverted polymer solar cells, RSC Adv. 7 (2017) 20084–20092.
performance Li-ion batteries, RSC Adv. 4 (2014) 5830–5833. [44] K.-H. Liao, A. Mittal, S. Bose, C. Leighton, K.A. Mkhoyan, C.W. Macosko, Aqueous
[28] T. Song, U. Paik, TiO2 as an active or supplemental material for lithium batteries, J. only route toward graphene from graphite oxide, ACS Nano 5 (2011) 1253–1258.
Mater. Chem. A 4 (2016) 14–31. [45] T. Zeng, P. Ji, X. Hu, G. Li, Nano-Sn doped carbon-coated rutile TiO2 spheres as a
[29] L. Liu, X. Chen, Titanium dioxide nanomaterials: self-structural modifications, high capacity anode for Li-ion battery, RSC Adv. 6 (2016) 48530–48536.
Chem. Rev. 114 (2014) 9890–9918. [46] J.-U. Seo, C.-M. Park, ZnTe and ZnTe/C nanocomposite: a new electrode material
[30] G.N. Zhu, Y.G. Wang, Y.Y. Xia, Ti-based compounds as anode materials for Li-ion for high-performance rechargeable Li-ion batteries, J. Mater. Chem. A 2 (2014)
batteries, Energy Environ. Sci. 5 (2012) 6652–6667. 20075–20082.
[31] H. Zhao, W. Qi, X. Li, H. Zeng, Y. Wu, J. Xiang, S. Zhang, B. Li, Y. Huang, SnSb/ [47] M. Qin, Y. Li, X.-J. Lv, Preparation of Ce-and La-doped Li4Ti5O12 nanosheets and
TiO2/C nanocomposite fabricated by high energy ball milling for high-performance their electrochemical performance in Li half cell and Li4Ti5O12/LiFePO4 full cell
lithium-ion batteries, RSC Adv. 6 (2016) 32462–32466. batteries, Nanomaterials 7 (2017) 150.
[32] H. Choi, K. Kim, W. Cho, C.M. Park, J.H. Kim, Synthesis and electrochemical

You might also like