You are on page 1of 7

Applied Surface Science 534 (2020) 147679

Contents lists available at ScienceDirect

Applied Surface Science


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

Efficient TiC-C hybrid conductive matrix for ZnTe anode in Lithium-ion T


storage
Quoc Hai Nguyena,1, , Quoc Hanh Nguyenb,1, Seongjoon Sob, Jaehyun Hurb,
⁎ ⁎

a
Biomaterials and Nanotechnology Research Group, Faculty of Applied Sciences, Ton Duc Thang University, Ho Chi Minh City 700000, Viet Nam
b
Department of Chemical and Biological Engineering, Gachon University, Seongnam, Gyeonggi 461-701, Republic of Korea

ARTICLE INFO ABSTRACT

Keywords: ZnTe alloy is simply prepared by the primary annealing step, along with the formation of TiC-C hybrid con-
ZnTe ductive matrix in the secondary ball-milling process. As-prepared ZnTe@TiC-C nanocomposite consists of active
TiC ZnTe nanocrystallites embedded in a TiC-C buffering material, which is adopted as a potential anode material for
Anode Lithium-ion storage. In comparison with stand-alone carbon matrix as a buffering material, the presence of the
Ball milling
TiC-C hybrid matrix improves the electrochemical performances of ZnTe active material in terms of capacity,
LIBs
stability, reversibility, and rate capability performances. Also, the optimum content of TiC in the ZnTe@TiC-C
composites is evaluated based on the electrochemical performances, in that the ZnTe@TiC(20%)-C composite
exhibits the highest reversible capacities of ~547 mAh g−1 after 300 cycles at 0.1 A g−1, the good rate capability
(84% capacity retention at 10 A g−1 when compared with the capacity at 0.1 A g−1), and the cycle life at 1 A
g−1 (~700 mAh cm−3 after 500 cycles). The enhanced electrochemical performance of the ZnTe active material
with the presence of the TiC-C hybrid matrix can be attributed to the effective mitigation against the large
volume change in the ZnTe and the high conductivity, thereby facilitating electron transport during prolonged
cycling.

1. Introduction number of Li+ ions coordinated with electrode materials during the
alloying/dealloying reactions as well as the reduced safety concerns
Owing to the outstanding properties such as high capacities, high associated with their high reaction potentials. Among various Li-al-
power densities, long cycle life, low self-discharge, and no memory loying-based materials, zinc-based materials have been studied as po-
effect, lithium-ion batteries (LIBs) are still attracting much attention as tential anodes in LIBs due to the cost-effectiveness, eco-friendliness, and
one of the most important rechargeable batteries in various commercial the natural abundance of Zn. Zn not only exhibits a very high volu-
applications [1–11]. To enhance the electrochemical performance of metric capacity (~1511 mAh cm−3 corresponding to the theoretical
LIBs that could meet the requirements for the next generation batteries, gravimetric capacity of 410 mAh g−1) but also easily form various Zn-
the anode materials have been significantly improved with main fo- metal binary alloying compounds such as Zn-Sb, Zn-S, Zn-P, Zn-Se, and
cuses on overcoming some drawbacks including low capacity, cyclic Zn-Te, among others [37–48]. However, Zn-based anode materials de-
instability, and poor rate capability. Although graphite has been widely monstrate poor electrochemical performances, as they undergo sig-
used as a stable and low-cost anode material for LIBs, this material nificant changes in volume and particle agglomeration during electro-
presents major drawbacks, which include a limited theoretical specific chemical cycling [20,21]. Moreover, in several studies, it was reported
capacity (~372 mAh g−1), low rate capability, and safety issues (Li that Te, which is a chalcogen group element, can alloy with Li to form
plating) [12–15]. To overcome these limitations, most researches have Li2Te with a theoretical capacity of 420 mAh g−1 [49–51]. Even though
been directed toward the exploration of new materials for the realiza- Te exhibits a relative lower gravimetric capacity when compared with
tion of high-performance anodes [16–21]. the atoms of other chalcogen group elements, e.g., S (1675 mAh g−1)
Recently, Li-alloying materials such as Li-Sn, [22–24] Li-Si, [25–27] and Se (678 mAh g−1), it has an high theoretical volumetric capacity of
Li-Sb, [28–31] Li-P, [32–35], and Li-Ge [36] have been proposed as ~2621 mAh cm−3 due to its high density (6.24 g cm−3), which is al-
potential anode materials for the next generation LIBs due to a large most higher than that of S by a factor of 3. Furthermore, Te exhibits a


Corresponding authors.
E-mail addresses: nguyenquochai@tdtu.edu.vn (Q.H. Nguyen), 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.2020.147679
Received 8 June 2020; Received in revised form 13 August 2020; Accepted 24 August 2020
Available online 26 August 2020
0169-4332/ © 2020 Elsevier B.V. All rights reserved.
Q.H. Nguyen, et al. Applied Surface Science 534 (2020) 147679

higher electronic conductivity (2 × 10−4 MS m−1) than those of Se 2.2. Materials characterization
(1 × 10−10 MS m−1) and S (5 × 10−22 MS m−1) [50]. Therefore, the
combination of Zn and Te with respect to the formation of a bimetallic The crystal structure of the as-prepared samples was examined
alloying compound (ZnTe) is an effective strategy for the realization of using X-ray diffraction (XRD; D/MAX-2200 Rigaku, Japan), where 2θ
a potential anode for LIBs, thus utilizing their high capacities (theore- Bragg angles were scanned over a range of 20–80° at a scan speed of 2°
tical capacity of 416 mAh g−1) and conductivities. In addition, the min−1. Scanning electron microscopy (SEM, Hitachi S4700, Japan) and
distinct reaction potential between Zn and Te in ZnTe allows for se- high-resolution transmission electron microscopy (HRTEM, JEOL JEM-
quential lithiation, which mitigates the large and abrupt change in 2100F) were used to characterize the morphology of as-prepared
volume during cycling (theoretical volume change ratio of ~83%) powders. The compositions of the powders were analyzed using scan-
[52,53]. Furthermore, the use of the standalone ZnTe-based alloys is ning transmission electron microscopy (STEM), and the corresponding
impractical, given that they may be subject to a degree of particle energy-dispersive X-ray spectroscopy (EDS) mapping images were ob-
pulverization due to the volume expansion during the repeated lithia- tained for each constituent element. The ex-situ XRD and ex-situ SEM
tion/delithiation processes, which results in a low rate capability and measurements were performed in which the powders was collected
severe capacity fading. from the electrode after a certain number of cycling.
Hence, the application of an appropriate buffering matrix to the
active material may effectively enhance the electrochemical perfor- 2.3. Electrochemical measurements
mances. One of the most promising reinforcements that have been
verified in several recent studies is the TiC-C hybrid conductive matrix, The working electrodes were prepared using the following proce-
which has remarkable properties based on the good mechanical prop- dures, to evaluate the electrochemical performances of the nano-
erties of TiC (high stiffness, electrochemical stability, and high con- composites. The slurries, which included the active materials, carbon
ductivity) and the synergetic buffering roles between TiC and C during black, and poly(vinylidene fluoride) (PVDF) with a weight ratio of
cycling [54,55]. Therefore, this hybrid matrix can act as a structural 70:15:15 were mixed into N-methyl-pyrrolidone (NMP) as a solvent.
barrier and prevent the aggregation of active particles as well as par- After casting those slurries on copper foils using a doctor blade, the
ticle pulverization, which improves the electrochemical characteriza- films were then dried at 70 °C overnight in a vacuum oven. The coin-
tions of various Li-alloying active materials [22,24,30,31,37]. type cell (CR2032) include an as-prepared electrode as a working
We herein report an effective strategy to achieve good electro- electrode, a metallic lithium foil as a counter electrode, polyethylene
chemical performances by the introduction of intermetallic ZnTe na- (PE) as a separator, and an electrolyte consisting of 1 M LiPF6 in
noparticles into the multifunctional TiC-C hybrid conductive matrix, ethylene carbonate (EC)/diethylene carbonate (DEC) (1:1 by v/v).
which results in the preparation of the ZnTe@TiC-C composite using Cyclic voltammetry (CV) was carried out using a ZIVE MP1 (WonAtech)
the primary annealing followed by the secondary ball-milling process. machine in the voltage range of 0.01–2.5 V at a scanning rate of
The good electrochemical performance of the ZnTe@TiC-C electrode 0.1 mV s−1. A battery cycler (WBCS3000, WonAtech) was operated in
based on the following advantages of this nanocomposite is demon- the potential range of 0.01–2.5 V to test galvanostatic charge/discharge
strated: (i) the formation of intermetallic ZnTe contributes to the mi- measurements (at a constant current density of 100 mA g−1) and rate
tigation of the change in volume via the progressive electrochemical capabilities (at different current densities of 0.1 A g−1, 0.5 A g−1, 1 A
reactions during cycling, and enhances the conductivity; and (ii) the g−1, 3 A g−1, 5 A g−1, and 10 A g−1). Thereafter, electrochemical
presence of the TiC-C hybrid conductive matrix can suppress the vo- impedance spectroscopy (EIS) measurements were carried out within a
lume expansion of active ZnTe and prevent the particle disruption frequency range of 0.1–100,000 Hz at 10 mV AC amplitude using a
during repeated cycling. Furthermore, this efficient hybrid matrix can ZIVE MP1 (WonAtech) analyzer. To estimate the volumetric capacity,
provide enhanced electrical conductivity, which can shorten the Li+ the tap density of as-synthesized powder was measured using powder
ion diffusion distance and facilitate electron transport. density measurement equipment (MAT-7000, Seishin, Japan).

3. Results and discussions


2. Experimental section
Fig. 1 presents the schematic of the overall synthesis of the com-
2.1. Material fabrication
posites using a facile annealing step followed by a ball-milling process.
In the initial annealing step, a metastable Zn-Te-Ti alloy was formed
The ZnTe-based nanocomposites were synthesized via annealing
including a new phase of ZnTe and three unreacted continents of Zn, Te,
followed by ball milling. In brief, powders of Zn (< 10 µm, ≥ 98%,
and Ti, which are confirmed by XRD result in Fig. S1. In the subsequent
Aldrich), Te (200 mesh, 99.8%, Aldrich), and Ti (325 mesh, 99.99%,
ball-milling step, with the presence of amorphous carbon, the TiC phase
Alfa Aesar) were used as raw-materials with four different molar ratios
was synthesized along with the complete formation of ZnTe phase from
(1:1:0.46, 1:1:1.07, 1:1:1.93, and 1:1:3.22). The primary annealing in
unreacted continents of Zn and Te. The as-prepared composite includes
argon atmosphere was applied at 600 °C for 8 h at the heating rate of 5
ZnTe, TiC, and amorphous carbon (denoted as ZnTe@TiC-C composite).
°C min−1 for the mixed-powder of Zn-Te-Ti. Subsequently, the heat-
The chemical reaction for these two steps can be summarized as fol-
treated powder of Zn-Te-Ti and an appropriate amount of carbon black
lows.
(99.99%, Alfa Aesar) were added into a ZrO2 bowl (80 cm3) with
hardened ZrO2 balls in an Ar atmosphere, such that the compositions of Annealing ball milling
Zn + Te + Ti ZnTe + Zn + Te + Ti ZnTe@TiC C
the TiC phase were estimated to be 10 wt%, 20 wt%, 30 wt%, and 40 wt
(1)
%, respectively, with 20 wt% of excess carbon black. The secondary
ball-milling was operated at a speed of 300 rpm for 40 h using a pla- The XRD patterns (Fig. 2a) of ZnTe-based composites with various
netary ball mill (Pulverisette 5, Fritch). The resulting products were estimated TiC content (0 wt%, 10 wt%, 20 wt%, 30 wt%, and 40 wt%)
denoted as ZnTe@TiC(10%)-C, ZnTe@TiC(20%)-C, ZnTe@TiC(30%)-C, confirm the successful formation of the cubic phase structure of ZnTe
and ZnTe@TiC(40%)-C. For comparison, ZnTe@C composites without (PDF#15–0746) in all samples. In detail, the peaks at 2θ = 25.3°, 29.2°,
TiC was prepared by the same procedure without Ti at the initial mixing 41.8°, and 49.5° observed for all samples correspond to the (1 1 1),
stage. Another control sample of TiC was prepared by ball milling at a (2 0 0), (2 2 0), and (3 1 1) planes of crystalline ZnTe with the interplanar
molar ratio of 1:1 (Ti:C). spacing of ~3.5 A°, ~3.1 A°, ~2.16 A°, and 1.84 A°, respectively. In
addition, the successful formation of the conductive TiC phase during

2
Q.H. Nguyen, et al. Applied Surface Science 534 (2020) 147679

Fig. 1. Schematic of overall synthesis of ZnTe@TiC-C composite.

SEM. As shown in Fig. 2c, the average particle sizes of the ZnTe@TiC-C
nanocomposite were generally in the range of sub-micrometers to
several micrometers. From a comparison with the ZnTe and ZnTe@C
composites, the ZnTe@TiC-C nanocomposite exhibited a greater
number of particles with sizes less than 100 nm (up to 86% in Fig. S4).
This can be attributed to the effect of the mechanically strong TiC
present in the ZnTe@TiC-C composite, in which the fragmentation of
ZnTe particles was more pronounced during the ball-milling process.
Furthermore, the TEM image (Fig. S5a) and the STEM images with the
corresponding elemental mapping (Fig. S5b) revealed the homogeneous
distribution of ZnTe, TiC, and C within the composite. These structural
features of ZnTe@TiC-C are considered as beneficial for anode mate-
rials in LIBs.
Fig. 3 presents a comparison between the first-cycle discharge/
charge profiles of ZnTe@C and those of various ZnTe@TiC-C elec-
trodes, as obtained from the galvanostatic charge and discharge (GCD)
measurements at a current density of 100 mA g−1, over a voltage range
of 0.01–2.5 V (vs. Li/Li+) (the detailed data are summarized in Table
Fig. 2. (a) XRD patterns of ZnTe-based composites; (b) HRTEM image (inset:
S2). As seen clearly, the initial discharge and charge capacities gen-
FFT pattern) and (c) SEM image of ZnTe@TiC-C composite.
erally decrease with the increase of TiC content from 0 wt% to 40 wt%,
given that the contribution of TiC to the capacity was lower (~125 mAh
ball milling in the presence of Ti and C was demonstrated by the ad- g−1 in Fig. S6) [56,57] than that of ZnTe. In detail, the initial dis-
ditional peaks at ~35.9°, 41.7°, 60.4°, and 76.1° corresponding to the charge/charge capacities of of ZnTe@-C, ZnTe@TiC(10%)-C,
(1 1 1), (2 0 0), (2 2 0), and (3 1 1) planes of crystalline TiC ZnTe@TiC(20%)-C, ZnTe@TiC(30%)-C, and ZnTe@TiC(40%)-C were
(PDF#32–1383), which was in good agreement with the findings of 926/634, 859/598, 667/446, 544/334, and 596/291 mAh g−1/mAh
previous studies [22,24,29–31]. With increasing TiC content, the TiC
peaks were more distinct (dotted marks at 2θ of ~35.9°), and the ZnTe
peaks reduced their intensities. This can be attributed to the increased
content of TiC, given that the sequential collision and deformation of
ZnTe against the mechanically hard and strong TiC phase reduced the
crystallinity of the ZnTe during the milling process [54,55]. Although a
portion of the pure Zn, Ti, and Te phases remained after the annealing
(dotted circles in Fig. S1), they were all primarily converted into the
well-defined ZnTe and TiC phases after the secondary ball-milling
process (Fig. 2a) except for a trace amount of pure Ti (1.1 wt% as
calculated from Table S1 with Fig. S2 and S3) in ZnTe@TiC(20%)-C.
From the Table S1, the actual contents of ZnTe/C in ZnTe@C and
ZnTe@TiC(20%)-C were 81%/19% and 60.6%/19.6%, respectively,
which were not much deviated from the theoretical contents (80%/20%
and 60%/20% of ZnTe/C in ZnTe@C and ZnTe@TiC(20%)-C, respec-
tively). From the HRTEM image (Fig. 2b), the presence of ZnTe and the
TiC nanocrystallite phases could be confirmed with interplanar dis-
tances of 0.216 nm (2 2 0) phase of ZnTe and 0.249 nm (1 1 1) phase of
TiC, which were in good agreement with the XRD results discussed
above. Also, the presence of crystalline ZnTe (2 2 0) and TiC (1 1 1)
could be confirmed from the corresponding FFT image by indexing the
lattice planes and interplanar spacings (inset in Fig. 2b). The powder
Fig. 3. Initial charge/discharge profiles of ZnTe@C and various ZnTe@TiC-C
morphology of the ZnTe@TiC-C nanocomposite was characterized by
electrodes at 100 mA g−1.

3
Q.H. Nguyen, et al. Applied Surface Science 534 (2020) 147679

Fig. 4. CV curves of (a) ZnTe-TiC(10%)-C; (b) Ex-situ XRD patterns of the ZnTe@TiC-C electrode at three different cut-off potentials and (c) Schematic illustration for
the reaction mechanism.

g−1, corresponding to their initial coulombic efficiencies (ICEs) of cycle, due to the reduced electrode polarization during the discharge
~68%, 67%, 67%, 64%, 51%, respectively. The ICEs exhibited a similar processes, whereas all the oxidation peaks were remained at the same
trend to the initial discharge/charge capacities, in that the increase in potentials [24]. Secondly, the alloying/dealloying reaction mechanism
the TiC content led to a decrease in the ICEs. This could be attributed to for ZnTe@TiC-C during cycling was further confirmed by carrying out
the lower crystallinity of ZnTe with a higher TiC content, which pre- ex-situ XRD analysis on the ZnTe@TiC-C electrode based on three cut-
vents Li+ ion diffusion during the initial cycling process and induces off potentials as (Fig. 4b) indicated using arrows in the CV plots
more side reactions, thus resulting in the formation of an SEI layer over (Fig. 4a). In the pristine state, all the observed ZnTe peaks were con-
a wide range of potential [24]. As a result, distinct plateaus associated sistent with the XRD results (Fig. 2a). When the electrode was dis-
with SEI layer formation at ~0.7 V were observed for the charged to 0.01 V vs. Li/Li+ (full discharge state), the additional peaks
ZnTe@TiC(30%)-C and ZnTe@TiC(40%)-C electrodes, which were re- of Li2Te (~23°, ~27°, and ~39°) were detected, and the ZnTe peaks
latively unclear for the lower TiC content cases (ZnTe@C, disappeared, which indicates the complete conversion reaction of ZnTe.
ZnTe@TiC(10%)-C, and ZnTe@TiC(20%)-C electrodes). We note the peaks corresponding to LiZn were not clearly detected after
To better understand the electrochemical reaction mechanism, CVs full discharge. After charging to 2.5 V (fully charge state), the phases of
(Fig. 4a and Fig. S7) and ex-situ XRD measurements (Fig. 4b) were Li2Te and LiZn were not observed, whereas crystalline ZnTe peaks re-
carried out on the ZnTe@TiC-C composites. First, CV plots in all cases of appeared, indicating the reversible lithiation/delithiation process
ZnTe@TiC-C electrodes exhibit similar behaviors, in which all the peak without structural changes. In addition, the peaks from TiC were not
positions were consistent with the voltage plateaus observed in GCD clearly detected due to the low signal level in the ex-situ XRD mea-
voltage profiles (Fig. 3). However, in comparison with the cases of TiC surements [43]. To further investigate the reaction mechanism, the ex-
contents of 10% and 20%, the ZnTe-based electrodes with 30% and situ HRTEM was performed after full discharge and charge (Fig. S8); the
40% of TiC content showed generally broader voltage ranges, which results showed the lattice spacings of 0.376 nm and 0.326 nm which
could be caused by the reduced crystallinity of ZnTe when the TiC correspond to the (1 1 1) and (2 0 0) planes of Li2Te and the lattice
content was high. In the initial lithiation process, a reduction peak distances of 0.220 nm and 0.188 nm which indicate the (2 2 0) and
observed at ~1.7 V in all cases of ZnTe-based electrodes related to the (3 1 1) planes of LiZn, respectively, after full discharge (Fig. S8a).
formation of Li2Te and Zn phase from the alloying reaction between Te Meanwhile, at fully charged state (Fig. S8b), the crystalline phases of
and Li+ ions (ZnTe + 2Li+ → Li2Te + Zn at 1.7 V). Two reduction ZnTe (0.352 nm d-spacing from (1 1 1) plane) and TiC (0.216 nm d-
peaks at 0.9 and 0.7 V indicate the formation of Li2Te and LixZn from spacing from (2 0 0) plane) were observed, indicating the reversible
ZnTe (ZnTe + (2 + x)Li+→ Li2Te + LixZn at 0.9 and 0.7 V). In ad- reaction of ZnTe. Based on the CV results and ex-situ XRD patterns, the
dition, the various small peaks observed in the voltage range of detailed electrochemical mechanism of the ZnTe@TiC-C electrode
0.4–0.01 V are associated with the successive formation of a LiZn phase during the first lithiation/delithiation can be schematically illustrated
from the reaction between Zn and Li-ions (LixZn + (1-x)Li+ → LiZn at as Fig. 4c and expressed as follows:
0.4–0.01 V) along with a solid-electrolyte interphase (SEI) layer for- During the lithiation process:
mation on the surface of the electrode material [37,43]. In the initial
ZnTe → Li2Te + Zn → Li2Te + LixZn → Li2Te + LiZn (2)
delithiation process, the sequential reversible reaction of LiZn to Zn
occurred when the potential increased from 0.01 V to 1.25 V (LiZn → During the de-lithiation process:
LixZn → xLi+ + Zn at 1.25 V). Meanwhile, an oxidation peak observed
at ~2.0 V represents the dealloying reaction of Li2Te phase Li2Te + LiZn → Li2Te + LixZn → Li2Te + Zn (3)
(Li2Te + Zn → ZnTe + 2Li+ at 2.0 V) [43]. Moreover, the reduction → ZnTe
peak at ~0.9 V was shifted toward a higher potential after the first
Fig. 5a presents a comparison of the cyclic performances and

4
Q.H. Nguyen, et al. Applied Surface Science 534 (2020) 147679

Fig. 5. (a) Cyclic performances (gravi-


metric capacity) of ZnTe@C and
ZnTe@TiC-C electrodes at 100 mA g−1;
(b) Long-term cyclic performance
(gravimetric and volumetric capacity)
of ZnTe@TiC-C electrodes at 1 A g−1;
(c) Rate capability of ZnTe@C and
ZnTe@TiC-C electrodes; (d) Normalized
capacity retention of ZnTe@C and
ZnTe@TiC-C electrodes at different
current densities.

coulombic efficiencies between ZnTe@TiC-C (20 wt% TiC) and at a high current density of 1 A g−1 (Fig. 5b) demonstrates the stable
ZnTe@C (0 wt% TiC) electrodes at a current density of 100 mA g−1. In specific capacity of ~480 mAh g−1 after 500 cycles, corresponding to
comparison with ZnTe@TiC-C, the ZnTe@C electrode exhibited a the capacity retention of 106% (relative to the capacity at 2nd cycle).
higher initial discharge capacity (926 mAh g−1 as compared to 667 The result also showed the very high volumetric capacity of ~700 mAh
mAh g−1), as it contained higher ZnTe active material content (80 wt% cm−3 after 500 cycles, which was converted from the specific capacity
as compared to 60 wt%) in the composite. However, there was a rapid with the measured tap density (1.46 g cm−3). It is of note that the
decrease in the capacity during the 150 cycles (~298 mAh g−1 with a volumetric capacity can be calculated by multiplying the measured
corresponding capacity retention of 55%). Meanwhile, the introduction specific capacity by the tap density. This value (~700 mAh cm−3) is
of the TiC-C buffering matrix into the ZnTe alloy resulted in a sig- greater than the theoretical volumetric capacity of the commercial
nificant improvement in the cyclic stability, which exhibited a capacity graphite anode (~385 mAh cm−3 with low tap density of 0.96 g cm−3)
of ~547 mAh g−1 after 300 cycles (capacity retention of 108% relative by a factor of 1.8 [28,30], which confirms that the ZnTe@TiC-C is
to the capacity at 2nd cycle). The gradually increasing capacity with the applicable to practical energy storage systems.
increasin in cycle number may be attributed to the gradual activation of To further evaluate the effect of the TiC-C matrix on the composites,
the electrode and polymer/gel-like film formation [58–60]. Further- the rate capabilities (Fig. 5c) and the normalized capacity retentions
more, the optimum TiC content in the ZnTe@TiC-C composites was (Fig. 5d) of the ZnTe@C and ZnTe@TiC-C electrodes were evaluated at
evaluated based on the comparison of the cyclic performances, as current densities ranging from 0.1 to 10 A g−1. As shown in Fig. 5c, the
shown in Fig. S9a. Among these electrodes, the ZnTe-TiC(10%)-C ex- ZnTe@TiC-C electrode retained a high specific capacity of ~432 mAh
hibited the highest capacity after 100 cycles (535 mAh g−1 as com- g−1 at a very high current density of 10 A g−1, which corresponded to a
pared to ~524, 330, and 330 mAh g−1 for ZnTe@TiC(20%)-C, capacity retention of ~84% compared to the capacity at 0.1 A g−1;
ZnTe@TiC(30%)-C and ZnTe@TiC(40%)-C, respectively). However, the whereas the ZnTe@C electrode exhibited capacities within the ap-
ZnTe-TiC(10%)-C showed the instable cycling behavior after 100 cycles proximate range of ~325 mAh g−1 at 10 A g−1, corresponding to
where the capacity decreased to ~328 mAh g−1 after 300 cycles. This ~67% of capacity retention. Moreover, when the current density was
could be attributed to the insufficient amount of TiC in the composite returned from 10 A g−1 to its initial value of 0.1 A g−1, the capacity
that could not sufficiently prevent the ZnTe active particles from the retention of the ZnTe@TiC-C electrode was slightly higher (~545 mAh
volume expansion during repeated cycling. Although the ZnTe-based g−1) than the initial capacity (514 mAh g−1), which indicates an in-
electrodes with 30 wt% and 40 wt% of TiC content exhibited much creased activation of the electrode. This better performance can be at-
improved cyclic stabilities, they showed relatively low reversible dis- tributed to the facile Li+ ion transport property of the active particles,
charge capacities due to their low ZnTe contents (~351 and 357 mAh in addition to the uniform incorporation of the conductive TiC-C dis-
g−1 for ZnTe@TiC(30%)-C and ZnTe@TiC(40%)-C, respectively). persing matrix that confers the high conductivity for the composites
When the TiC content was optimized to be 20 wt% (ZnTe@TiC(20%)- [22,24,30,31,37]. The normalized capacity retentions, as shown in
C), both high capacity and cyclic stability were realized. In particular, it Fig. 5d, further confirm the relatively higher capacity retention of
exhibited a capacity of ~547 mAh g−1 after 300 cycles (capacity re- ZnTe@TiC-C electrode than that of the ZnTe@C electrode. Further-
tention of 82%), which can be attributed to the introduction of an ap- more, the comparison of rate capability between various TiC content in
propriate amount of TiC to the TiC-C hybrid matrix, which effectively ZnTe@TiC-C electrode was presented in Fig. S9b, which demonstrated
suppresses the aggregation of active ZnTe and mitigates the stresses and the best performance of the ZnTe@TiC(20%)-C as compared to the
strains in the ZnTe during the extended cycling. In addition, the long- others.
term cyclic performance of the ZnTe@TiC(20%)-C electrode measured Finally, EIS measurements (Fig. 6a and Table S3) and ex-situ SEM

5
Q.H. Nguyen, et al. Applied Surface Science 534 (2020) 147679

Fig. 6. (a) EIS spectra of ZnTe@C and ZnTe@TiC-C electrodes after 50 cycles; Ex-situ SEM images of (b) ZnTe@C and (c) ZnTe@TiC(20%)-C electrodes after 50
cycles.

(Fig. 6b and c) imaging were conducted after 50 cycles of cyclic per- potential as a promising high-performance anode material for re-
formance at 100 mA g−1 to evaluate the changes in the conductivities chargeable Li-ion storage.
and morphology of ZnTe@TiC-C and ZnTe@C electrodes during cy-
cling. As a result, the ZnTe@TiC-C electrode exhibited relatively lower CRediT authorship contribution statement
charge-transfer resistances (smaller semi-circles) when compared with
that of the ZnTe@C electrode after 50 cycles. Also, the cross-sectional Quoc Hai Nguyen: Data curation, Writing - original draft,
image of ZnTe-C (Fig. 6b) revealed a poor morphology with distinct Investigation. Quoc Hanh Nguyen: Data curation, Writing - original
cracks, pulverization, and aggregation; whereas the ZnTe@TiC-C elec- draft, Investigation. Seongjoon So: Data curation, Investigation.
trode retained its original morphology with a smooth surface, and no Jaehyun Hur: Conceptualization, Methodology, Writing - review &
cracks or peeling were observed (Fig. 6c). These results confirm that the editing, Supervision, Funding acquisition.
addition of an appropriate amount of TiC to the TiC-C hybrid matrix
allows for the development of good electron transport pathways be- Declaration of Competing Interest
tween active ZnTe particles, thereby suppressing the electrode pulver-
ization and enhancing the electronic conductivity of the composites, The authors declare that they have no known competing financial
which further verify the effectiveness of the addition of the TiC-C ma- interests or personal relationships that could have appeared to influ-
trix into ZnTe for the enhancement of the electrochemical performance. ence the work reported in this paper.

4. Conclusion Acknowledgments

In summary, ZnTe@TiC-C nanocomposites with various TiC con- This work was supported by the Korea Institute of Energy
tents were successfully synthesized using a facile two-step procedure of Technology Evaluation and Planning (KETEP) and the Ministry of
the primary annealing followed by the secondary ball milling. The as- Trade, Industry & Energy (MOTIE) of the Republic of Korea (No.
prepared ZnTe@TiC-C nanocomposite includes the ZnTe nanocrys- 20194030202290) and by the Basic Science Research Program through
tallites uniformly distributed in the well-mixed TiC-C hybrid conductive the National Research Foundation of Korea (NRF) funded by the
matrix, which resulted in the superior electrochemical performance Ministry of Education (NRF-2019R1F1A1057709).
with respect to the reversible capacity, life-span, and high rate cap-
ability. Among the various ZnTe-based electrodes, the ZnTe-TiC(20%)- Appendix A. Supplementary material
C electrode demonstrated the best performance, exhibiting a reversible
capacity of 547 mAh g−1 at 100 mA g−1 after 300 cycles, and an good Supplementary data to this article can be found online at https://
rate capability with a capacity retention of 84% at 10 A g−1. Moreover, doi.org/10.1016/j.apsusc.2020.147679.
the ZnTe-TiC(20%)-C exhibited a stable reversible capacity of ~480
mAh g−1 until 500 cycles at a high current density of 1 A g−1, which References
corresponds to the volumetric capacity of ~700 mAh cm−3. These good
[1] M. Armand, J.-M. Tarascon, Building better batteries, Nature 451 (2008) 652.
characteristics of the ZnTe@TiC-C nanocomposite confirm its high

6
Q.H. Nguyen, et al. Applied Surface Science 534 (2020) 147679

[2] J.-M. Tarascon, M. Armand, Issues and challenges facing rechargeable lithium Electrochim. Acta 293 (2019) 8–18.
batteries, Materials For Sustainable Energy: A Collection of Peer-Reviewed Research [32] M.-P. Bichat, J.-L. Pascal, F. Gillot, F. Favier, Electrochemical lithium insertion in
and Review Articles from Nature Publishing Group, World Scientific, 2011, pp. Zn3P2 zinc phosphide, Chem. Mater. 17 (2005) 6761–6771.
171–179. [33] H. Hwang, M.G. Kim, Y. Kim, S.W. Martin, J. Cho, The electrochemical lithium
[3] J.B. Goodenough, Y. Kim, Challenges for rechargeable Li batteries, Chem. Mater. 22 reactions of monoclinic ZnP2 material, J. Mater. Chem. 17 (2007) 3161–3166.
(2009) 587–603. [34] W. Li, L. Gan, K. Guo, L. Ke, Y. Wei, H. Li, G. Shen, T. Zhai, Self-supported Zn3P2
[4] V. Etacheri, R. Marom, R. Elazari, G. Salitra, D. Aurbach, Challenges in the devel- nanowire arrays grafted on carbon fabrics as an advanced integrated anode for
opment of advanced Li-ion batteries: a review, Energy Environ. Sci. 4 (2011) flexible lithium ion batteries, Nanoscale 8 (2016) 8666–8672.
3243–3262. [35] M. Bichat, L. Monconduit, J. Pascal, F. Favier, Anode materials for lithium ion
[5] Y.S. Mun, D. Kim, I.T. Kim, Electrochemical Performance of FeSb2-P@C Composites batteries in the Li-Zn-P system, Ionics 11 (2005) 66–75.
as Anode Materials for Lithium-Ion Storage, J. Nanosci. Nanotechnol. 18 (2018) [36] Y.J. Cho, H.S. Im, Y. Myung, C.H. Kim, H.S. Kim, S.H. Back, Y.R. Lim, C.S. Jung,
1343–1346. D.M. Jang, J. Park, Germanium sulfide (II and IV) nanoparticles for enhanced
[6] Y.S. Mun, Y.N. Pham, V.K.H. Bui, S.T. Tanaji, H.U. Lee, G.W. Lee, J.S. Choi, performance of lithium ion batteries, Chem. Commun. 49 (2013) 4661–4663.
I.T. Kim, Y.C. Lee, Tin oxide evolution by heat-treatment with tin-aminoclay (SnAC) [37] S.O. Kim, A. Manthiram, High-Performance Zn-TiC-C Nanocomposite Alloy Anode
under argon condition for lithium-ion battery (LIB) anode applications, J. Power with Exceptional Cycle Life for Lithium-Ion Batteries, ACS Appl. Mater. Inter. 7
Sources 437 (2019). (2015) 14801–14807.
[7] T.A. Nguyen, I.T. Kim, S.W. Lee, Chitosan-Tethered Iron Oxide Composites as an [38] H. Choi, K. Kim, W. Cho, C.-M. Park, J.-H. Kim, Synthesis and Electrochemical
Antisintering Porous Structure for High-Performance Li-Ion Battery Anodes, J. Am. Reaction Mechanism of Zn-TiOx-C Nanocomposite Anode Materials for Li Secondary
Ceram. Soc. 99 (2016) 2720–2728. Batteries, J. Electrochem. Soc. 164 (2017) A2683–A2688.
[8] T.L. Nguyen, D. Park, I.T. Kim, FexSnyOz Composites as Anode Materials for [39] L. He, X.-Z. Liao, K. Yang, Y.-S. He, W. Wen, Z.-F. Ma, Electrochemical character-
Lithium-Ion Storage, J. Nanosci. Nanotechnol. 19 (2019) 6636–6640. istics and intercalation mechanism of ZnS/C composite as anode active material for
[9] T.P. Nguyen, J.H. Kim, I.T. Kim, Electrochemical Performance of Sn/SnO/Ni3Sn lithium-ion batteries, Electrochim. Acta 56 (2011) 1213–1218.
Composite Anodes for Lithium-Ion Batteries, J. Nanosci. Nanotechnol. 19 (2019) [40] A.-R. Park, K.-J. Jeon, C.-M. Park, Electrochemical mechanism of Li insertion/ex-
1001–1005. traction in ZnS and ZnS/C anodes for Li-ion batteries, Electrochim. Acta 265 (2018)
[10] T.N. Pham, S.T. Tanaji, J.S. Choi, H.U. Lee, I.T. Kim, Y.C. Lee, Preparation of Sn- 107–114.
aminoclay (SnAC)-templated Fe3O4 nanoparticles as an anode material for lithium- [41] S. Saadat, Y.Y. Tay, J. Zhu, P.F. Teh, S. Maleksaeedi, M.M. Shahjamali,
ion batteries, RSC Adv. 9 (2019) 10536–10545. M. Shakerzadeh, M. Srinivasan, B.Y. Tay, H.H. Hng, Template-free electrochemical
[11] S.Y. Son, S.A. Hong, S.Y. Oh, Y.C. Lee, G.W. Lee, J.W. Kang, Y.S. Huh, I.T. Kim, deposition of interconnected ZnSb nanoflakes for li-ion battery anodes, Chem.
Crab-Shell Biotemplated SnO2 Composite Anodes for Lithium-Ion Batteries, J. Mater. 23 (2011) 1032–1038.
Nanosci. Nanotechnol. 18 (2018) 6463–6468. [42] C.M. Park, H.J. Sohn, Quasi-Intercalation and Facile Amorphization in Layered
[12] S. Goriparti, E. Miele, F. De Angelis, E. Di Fabrizio, R.P. Zaccaria, C. Capiglia, ZnSb for Li-Ion Batteries, Adv. Mater. 22 (2010) 47–52.
Review on recent progress of nanostructured anode materials for Li-ion batteries, J. [43] J.U. Seo, C.M. Park, ZnTe and ZnTe/C nanocomposite: a new electrode material for
Power Sources 257 (2014) 421–443. high-performance rechargeable Li-ion batteries, J. Mater. Chem. A 2 (2014)
[13] W.-J. Zhang, A review of the electrochemical performance of alloy anodes for li- 20075–20082.
thium-ion batteries, J. Power Sources 196 (2011) 13–24. [44] Y. Hwa, J.H. Sung, B. Wang, C.-M. Park, H.-J. Sohn, Nanostructured Zn-based
[14] H. Yang, S. Amiruddin, H.J. Bang, Y.-K. Sun, J. Prakash, A review of Li-ion cell composite anodes for rechargeable Li-ion batteries, J. Mater. Chem. 22 (2012)
chemistries and their potential use in hybrid electric vehicles, J. Ind. Eng. Chem. 12 12767–12773.
(2006) 12–38. [45] H.-T. Kwon, C.-M. Park, Electrochemical characteristics of ZnSe and its nanos-
[15] L. Lu, X. Han, J. Li, J. Hua, M. Ouyang, A review on the key issues for lithium-ion tructured composite for rechargeable Li-ion batteries, J. Power Sources 251 (2014)
battery management in electric vehicles, J. Power Sources 226 (2013) 272–288. 319–324.
[16] N.Q. Hai, H. Kim, I.S. Yoo, J.H. Kim, J. Hur, Comparative Study of Mechanically [46] Z. Zhang, Y. Fu, X. Yang, Y. Qu, Q. Li, Nanostructured ZnSe anchored on graphene
Milled MoS2 and MoSe2 in Graphite Matrix as Anode Materials for High- nanosheets with superior electrochemical properties for lithium ion batteries,
Performance Lithium-Ion Batteries, J. Nanosci. Nanotechnol. 18 (2018) 6469–6474. Electrochim. Acta 168 (2015) 285–291.
[17] N.T. Hung, S.-H. Park, J. Bae, Y.S. Yoon, J.H. Kim, H.B. Son, D. Lee, I.T. Kim, J. Hur, [47] C.-M. Park, H.-J. Sohn, Tetragonal zinc diphosphide and its nanocomposite as an
Sb-AlxCy-C nanocomposite alloy anodes for lithium-ion batteries, Electrochim. Acta anode for lithium secondary batteries, Chem. Mater. 20 (2008) 6319–6324.
210 (2016) 567–574. [48] N.T. Hung, J. Bae, J.H. Kim, H.B. Son, I.T. Kim, J. Hur, Facile preparation of a zinc-
[18] Q.H. Nguyen, J.S. Choi, Y.-C. Lee, I.T. Kim, J. Hur, 3D hierarchical structure of based alloy composite as a novel anode material for rechargeable lithium-ion bat-
MoS2@ G-CNT combined with post-film annealing for enhanced lithium-ion sto- teries, Appl. Surf. Sci. 429 (2018) 210–217.
rage, J. Ind. Eng. Chem. 69 (2019) 116–126. [49] A.-R. Park, C.-M. Park, Cubic Crystal-Structured SnTe for Superior Li-and Na-Ion
[19] Q.H. Nguyen, I.T. Kim, J. Hur, Core-shell Si@ c-PAN particles deposited on graphite Battery Anodes, ACS Nano 11 (2017) 6074–6084.
as promising anode for lithium-ion batteries, Electrochim. Acta (2018). [50] J.-U. Seo, G.-K. Seong, C.-M. Park, Te/C nanocomposites for Li-Te secondary bat-
[20] C.-M. Park, J.-H. Kim, H. Kim, H.-J. Sohn, Li-alloy based anode materials for Li teries, Sci. Rep. 5 (2015) 7969.
secondary batteries, Chem. Soc. Rev. 39 (2010) 3115–3141. [51] Y. Liu, J. Wang, Y. Xu, Y. Zhu, D. Bigio, C. Wang, Lithium–tellurium batteries based
[21] M. Obrovac, V. Chevrier, Alloy negative electrodes for Li-ion batteries, Chem. Rev. on tellurium/porous carbon composite, J. Mater. Chem. A 2 (2014) 12201–12207.
114 (2014) 11444–11502. [52] J. Besenhard, M. Hess, P. Komenda, Dimensionally stable Li-alloy electrodes for
[22] J. Leibowitz, E. Allcorn, A. Manthiram, SnSb–TiC–C nanocomposite alloy anodes for secondary batteries, Solid State Ion. 40–41 (1990) 525–529.
lithium-ion batteries, J. Power Sources 279 (2015) 549–554. [53] N. Nitta, F. Wu, J.T. Lee, G. Yushin, Li-ion battery materials: present and future,
[23] H. Zhao, W. Qi, X. Li, H. Zeng, Y. Wu, J. Xiang, S. Zhang, B. Li, Y. Huang, SnSb/ Mater. Today 18 (2015) 252–264.
TiO2/C nanocomposite fabricated by high energy ball milling for high-performance [54] A. Mani, P. Aubert, F. Mercier, H. Khodja, C. Berthier, P. Houdy, Effects of residual
lithium-ion batteries, RSC Adv. 6 (2016) 32462–32466. stress on the mechanical and structural properties of TiC thin films grown by RF
[24] S.Y. Son, J. Hur, K.H. Kim, H.B. Son, S.G. Lee, I.T. Kim, SnTe-TiC-C composites as sputtering, Surf. Coat. Technol. 194 (2005) 190–195.
high-performance anodes for Li-ion batteries, J. Power Sources 365 (2017) [55] H. Jia, Z. Zhang, Z. Qi, G. Liu, X. Bian, Formation of nanocrystalline TiC from
372–379. titanium and different carbon sources by mechanical alloying, J. Alloy. Compd. 472
[25] M. Obrovac, L. Christensen, Structural changes in silicon anodes during lithium (2009) 97–103.
insertion/extraction, Electrochem. Solid-State Lett. 7 (2004) A93–A96. [56] S. Cao, Z. Xue, C. Yang, J. Qin, L. Zhang, P. Yu, S. Wang, Y. Zhao, X. Zhang, R. Liu,
[26] N. Liu, Z. Lu, J. Zhao, M.T. McDowell, H.-W. Lee, W. Zhao, Y. Cui, A pomegranate- Insights into the Li+ storage mechanism of TiC@C-TiO2 core-shell nanostructures as
inspired nanoscale design for large-volume-change lithium battery anodes, Nat. high performance anodes, Nano Energy 50 (2018) 25–34.
Nanotechnol. 9 (2014) 187. [57] C.-M. Park, H.-J. Sohn, Electrochemical characteristics of TiSb2 and Sb/TiC/C na-
[27] S.-O. Kim, A. Manthiram, A facile, low-cost synthesis of high-performance silicon- nocomposites as anodes for rechargeable Li-ion batteries, J. Electrochem. Soc. 157
based composite anodes with high tap density for lithium-ion batteries, J. Mater. (2010) A46–A49.
Chem. A 3 (2015) 2399–2406. [58] W.P. Kang, Y.B. Tang, W.Y. Li, X. Yang, H.T. Xue, Q.D. Yang, C.S. Lee, High in-
[28] E. Allcorn, A. Manthiram, NiSb–Al2O3–C nanocomposite anodes with long cycle life terfacial storage capability of porous NiMn2O4/C hierarchical tremella-like nanos-
for Li-ion batteries, J. Phys. Chem. C 118 (2014) 811–822. tructures as the lithium ion battery anode, Nanoscale 7 (2015) 225–231.
[29] I.T. Kim, S.-O. Kim, A. Manthiram, Effect of TiC addition on SnSb–C composite [59] Z.P. Zeng, H.L. Zhao, P.P. Lv, Z.J. Zhang, J. Wang, Q. Xia, Electrochemical prop-
anodes for sodium-ion batteries, J. Power Sources 269 (2014) 848–854. erties of iron oxides/carbon nanotubes as anode material for lithium ion batteries,
[30] E. Allcorn, A. Manthiram, High-rate, high-density FeSb–TiC–C nanocomposite an- J. Power Sources 274 (2015) 1091–1099.
odes for lithium-ion batteries, J. Mater. Chem. A 3 (2015) 3891–3900. [60] X.J. Zhu, Z.P. Guo, P. Zhang, G.D. Du, R. Zeng, Z.X. Chen, S. Li, H.K. Liu, Highly
[31] H. Kim, M. Kim, Y.H. Yoon, Q.H. Nguyen, I.T. Kim, J. Hur, S.G. Lee, Sb2Te3-TiC-C porous reticular tin-cobalt oxide composite thin film anodes for lithium ion bat-
nanocomposites for the high-performance anode in lithium-ion batteries, teries, J. Mater. Chem. 19 (2009) 8360–8365.

You might also like