You are on page 1of 11

Chemical Engineering Journal 399 (2020) 125826

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Mechanochemical synthesis of InP nanoparticles embedded in hybrid T


conductive matrix for high-performance lithium-ion batteries
⁎ ⁎
Quoc Hai Nguyena,1, Seongjoon Sob,1, Quoc Hanh Nguyenb, Il Tae Kimb, , 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 13120, Republic of Korea

H I GH L IG H T S G R A P H I C A L A B S T R A C T

• mechanochemical
InP@TiO -C was synthesized via a
2
process for the first
time.
• effective
A hybrid TiO -C matrix served as an
2
buffer layer for InP during
cycling.
• Detailed mechanism was thoroughly
investigated with ex situ XRD, TEM,
SEM, and XPS.
• mance
InP@TiO -C showed
2
(~750 mAh g −1
good perfor-
at 1 A g−1
after 800 cycles).

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

Keywords: InP nanoparticles distributed in a TiO2-C hybrid matrix (InP@TiO2-C) are proposed as a promising anode ma-
InP terial for Li-ion batteries. A primary mechanochemical process on the precursor materials (In2O3, Ti, and P) leads
TiO2 to the formation of InP and TiO2 nanocrystal particles. The introduction of carbon during secondary ball milling
Carbon results in the formation of InP nanoparticles that are embedded in the TiO2-C hybrid conductive matrix
Anode
(InP@TiO2-C); this is confirmed by X-ray diffraction (XRD), Raman spectroscopy, scanning electron microscopy
Lithium-ion storage
(SEM), transmission electron microscopy (TEM), X-ray photoelectron spectroscopy (XPS), and energy-dispersive
X-ray characterizations. As an anode material, InP@TiO2-C exhibits impressive electrochemical performance in
both half- and full-cell batteries. Various ex situ analyses including XRD, high-resolution TEM (HRTEM), and XPS
were synergistically used to demonstrate the mechanism of lithium-ion storage for the InP@TiO2-C electrode
during the electrochemical reaction, and to elucidate the roles of active InP particles and the TiO2-C buffering
hybrid matrix. Consequently, as an anode for half-cell batteries, the InP@TiO2-C delivers a high reversible
capacity (~850 mAh g−1 after 120 cycles at 0.1 mA g−1), excellent life span at 0.5 A g−1 (~750 mAh g−1 after
800 cycles), and high rate capability (85% capacity retention at 10 A g−1 compared with that at 0.1 A g−1).
Moreover, as an anode for practical full-cell batteries with LiFePO4@graphite cathodes, it delivers a promising
initial energy density of 214 Wh kg−1 (based on the total mass of the anode and cathode) and a good stability
(retention of 61%) over 150 cycles.


Corresponding authors.
E-mail addresses: itkim@gachon.ac.kr (I.T. Kim), jhhur@gachon.ac.kr (J. Hur).
1
Q. H. Nguyen and S. J. So contributed equally to this study.

https://doi.org/10.1016/j.cej.2020.125826
Received 6 April 2020; Received in revised form 27 May 2020; Accepted 6 June 2020
Available online 11 June 2020
1385-8947/ © 2020 Published by Elsevier B.V.
Q. Hai Nguyen, et al. Chemical Engineering Journal 399 (2020) 125826

1. Introduction reasonable level of theoretical capacity (335 mAh g−1) that is close to
that of graphite [52-58]. Because of these attractive features, TiO2 has
Although Li-ion batteries (LIBs) have been extensively researched been widely used in couple with various active materials such as Si,
over the past two decades, researchers continue to pay more attention SnO2, Fe2O3, and CoO [55,59–61].
to them owing to their high energy densities and output voltages In this study, we firstly propose the solid-state synthesis of InP via a
compared with other energy storage systems [1–3]. Recently, numerous simple, fast, and scalable mechanochemical ball milling process for a
anode materials for Li-ion storage, such as carbon-based, silicon-based, new anode material in lithium-ion batteries (LIBs). Various character-
alloy-based, metal oxide-based, and layer material-based compounds izations (X-ray diffraction (XRD), Raman spectroscopy, X-ray photo-
[4–23] have been explored, with the main focus being overcoming the electron spectroscopy (XPS), scanning electron microscopy (SEM),
drawbacks of LIBs, including low capacity, cyclic instability, poor rate high-resolution transmission electron microscopy (HRTEM), and en-
capability, energy density limitation, and poor safety [24,25]. Although ergy-dispersive spectroscopy (EDS)) reveal that a ball milling of pow-
previous studies have proposed several potential anode materials for ders (In2O3, Ti, P, and C) produces well-defined binary crystals of InP
LIBs, these materials are still far from being satisfactory for next-gen- and TiO2 in amorphous C, which possesses an extremely favorable
eration energy storage systems. Therefore, discovering new anode structural feature for an anode in LIBs. This is because the hybrid TiO2-
materials for Li-ion storage systems continues to be of importance for C matrix not only synergistically buffers the volume change in InP, but
researchers. also allows enhanced mechanical properties and better electrical con-
Owing to their unique physical and chemical properties, metal ductivity. Electrochemical measurements show that InP@TiO2-C has
phosphide (MP) compounds have potential applications in energy sto- excellent performance in terms of the specific discharge capacity (~850
rage, electronics, optoelectronics, and as catalysts [26,27]. Several MP mAh g−1 after 120 cycles at 100 mA g−1), cycle life (~750 mAh g−1
compounds, such as CuP2, Cu3P, GeP, Sn4P3, Co2P, Ni2P, LaP, and InP, after 800 cycles at 500 mA g−1), and rate capability (85% capacity
have been studied for LIBs and sodium ion batteries (SIBs) [28-34]. retention at 10 A g−1 compared with that at 0.1 A g−1) for the half-cell
These MP compounds have been demonstrated to exhibit higher spe- battery test. Moreover, as an anode for practical full-cell batteries, it
cific capacities (440–1750 mAh g−1, Table S1) than the theoretical delivers a promising initial energy density of 131 Wh kg−1 and cyclic
capacity of the commercial graphite electrode (372 mAh g−1 [35]). stability (capacity retention of 61%) over 150 cycles. Therefore, we
Furthermore, they have long life spans and small electrode polariza- provide an in-depth understanding of the reaction principle of the
tions. Therefore, they are considered promising candidates for next- InP@TiO2-C electrode and elucidate the roles of the individual con-
generation electrodes in LIBs. However, they still have shortcomings, stituents in the nanocomposite.
such as particle pulverization and large volume expansion, which re-
main to be resolved.
2. Experimental section
Among the many different MP compounds, InP is a significant
candidate for high-performance Li-ion storage owing to the conversion
2.1. Material synthesis
reaction of Li ions with both In and P components. Its theoretical spe-
cific capacity is estimated to be ~ 735 mAh g−1 based on the overall
InP@TiO2-C was synthesized via two-step mechanochemical ball
electrochemical reaction of InP + 4Li+ + 4e− → LiIn + Li3P, with the
nF milling. In the first milling step, In2O3 (1.48 g, 99.9%, Alfa Aesar), Ti
formula CInP = M (where CInP, n, F, and M denote the theoretical spe- (0.38 g, 99%, Alfa Aesar), and red P (0.33 g, 99.99%, 325 mesh, Alfa
cific capacity of InP, number of electrons involved in the reaction,
Aesar) powders along with zirconium oxide balls (44 g, diameters 3/8
Faraday’s constant, and molecular weight of InP, respectively) [36-38].
and 3/16 in.) were placed in a zirconium oxide vessel reactor (80 ml)
Although InP has been widely studied in various optoelectronics de-
under an Ar atmosphere. The input amounts of each powder were based
vices such as high-speed electronics, lasers, photovoltaics, photo-
on the stoichiometry of the following solid-phase reaction:
detectors, and photo-electrochemical devices because of its exciting
electrical properties [39–48], there are only a few studies on its energy 2In2 O3 + 3Ti + 4P → 4InP + 3TiO2 (ΔGo = −1196kJ/mol) (1)
storage applications. Cui et al. first demonstrated the feasibility of an
The above solid-state reaction is supposed to occur spontaneously
InP thin film fabricated by pulsed laser deposition as an anode material
based on the calculation of change in free energy of reaction (see the
for LIBs with a reversible discharge capacity of ~620 mAh g−1 [36].
detailed calculations in Fig. S1). Ball milling (Pulverisette 5, Fritch) was
Gerngross et al. fabricated a single-crystalline porous InP anode with a
performed at 300 rpm for 10 h. In the second step, C (0.22 g, 99.99%,
maximum possible capacity of ~800 mAh g−1 [37]. However, the cycle
Alfa Aesar) was added to the reactor followed by milling for 10 h to
lives demonstrated in those studies were very short (~10 cycles). In
synthesize InP@TiO2-C. InP@TiO2 (a control sample) was synthesized
2018, Liu et al. improved the electrochemical performance of InP (a
following the same procedure but with only the first milling step. For
specific capacity of ~690 mAh g−1 after 100 cycles) by introducing
comparison, InP@C was also synthesized as another control sample
reduced graphene oxide (rGO) to synthesize InP/rGO [38]. Never-
using two-step ball milling in which InP was first synthesized using In
theless, InP/rGO showed continuous capacity fading from an initial
(1.58 g, 99.99%, Alfa Aesar) and P powders (0.42 g, 98.9%, Alfa Aesar),
discharge capacity of 1872 mAh g−1, and no strategy was presented to
followed by C addition (InP : C = 7 :3, w:w).
overcome this problem. Although all these InP-based composites have
consistently demonstrated the possibility of InP as a potential high-
performance anode, more in-depth studies are required to further im- 2.2. Material characterization
prove its electrochemical performance by the rational design of InP-
based materials. XRD (D/MAX-2200 Rigaku, Japan) was used to determine the
Carbon is one of the most commonly used buffering materials owing formed phases of the as-prepared powders. XPS (Kratos AXIS Nova) and
to its desirable characteristics such as low cost, easy processability, micro Raman spectroscopy (ANDOR Monora500i, 633 nm) were per-
chemical stability, and high conductivity, with reasonably low con- formed to characterize the structural information of as-prepared pow-
tribution to the total capacity (theoretical capacity of 372 mAh g−1 ders. The microscopic morphologies of the powders were observed by
based on the reaction with Li to form LiC6) [4,49–51]. Furthermore, SEM (Hitachi S4700, Japan) and HRTEM (TECNAI G2F30) in con-
TiO2 as a crystalline buffering material has remarkable properties si- junction with EDS. To study the Li-storage mechanism in a more precise
milar to graphite, including low toxicity, low cost, and high safety. It manner, various ex situ analyses (ex situ XRD, XPS, HRTEM, and SEM)
has been widely used as an inactive and active material in LIBs owing to were performed, in which the powders were collected from the elec-
its excellent cyclic stability and small volume change (3–4%) with a trode after a certain number of cycles.

2
Q. Hai Nguyen, et al. Chemical Engineering Journal 399 (2020) 125826

2.3. Electrochemical characterization transfer mechanism; the incident photons can cause a resonant charge
transfer between the semiconductor (ZnS) and adsorbate (graphene),
A coin-type electrochemical half-cell battery (CR2032) was as- forming specific adsorbate–semiconductor complexes, which intensifies
sembled in an Ar-filled glove box, which contained an as-prepared Raman scattering from graphene while relatively reducing scattering
electrode as a working electrode, a lithium foil as a counter electrode, from ZnS. Because the resonance charge transfer between the semi-
polyethylene (PE) as a separator, and an electrolyte solution consisting conductor and carbon as the absorbate can originate from the induced
of 1 M LiPF6 in ethylene carbonate (EC)/diethylene carbonate (DEC) proton, the peak indicating the InP semiconductor is not detected in the
(1:1 by v:v). To prepare a working electrode, a slurry was prepared by InP@TiO2-C nanocomposites [64-66]. To further validate the successful
mixing as-prepared active powders (InP@TiO2-C, InP@TiO2, and formation of InP and TiO2 during the mechanochemical process, XPS
InP@C), acetylene black, and polyvinylidene fluoride (PVDF) at a analysis of InP@TiO2-C was performed (Fig. 1d and Fig. S4). Two peaks
weight ratio of 70:15:15 in an N-methyl-pyrrolidone (NMP) solvent. from In (3d1/2 at 445 eV and 3d3/2 at 452 eV) (Fig. 2d) and one peak
The as-obtained slurry was then casted onto a Cu foil using a doctor from P (2p at 129 eV) (Fig. S4b) represent the presence of InP in the
blade method, followed by subsequent drying in a vacuum oven at InP@TiO2-C composite [67]. In addition, two peaks from Ti (2p1/2 at
70 °C for 12 h, and then punched into disk-shaped (12 mm diameter). A 459 eV and 2p3/2 at 464.8 eV) (Fig. 2d) and one peak from O (2p at
battery cycler (WBCS3000, WonAtech) was used to perform the cy- 531.7 eV) (Fig. S4d) confirm the formation of TiO2 in the as-prepared
clability test in the voltage range of 0.01–3.0 V (versus Li/Li+) and the InP@TiO2-C powder [68].
rate capability test at various current densities (0.1, 0.5, 1, 3, 5, and 10 The SEM image of the InP@TiO2-C composite (Fig. 1e) exhibits a
A g−1). Cyclic voltammetry (CV) and electrochemical impedance uniform particle size (~50 nm), whereas InP@TiO2 and InP@C exhibit
spectroscopy (EIS) measurements were performed using a multichannel a high level of size distribution (probably because of many aggregated
electrochemical workstation (ZIVE MP1, WonAtech). The loading mass particles, Fig. S5). While most of particles are in the range of less
of the active material was ~1 mg cm−2, and all calculations were based than 50 nm for InP@TiO2-C, InP@TiO2 and InP@C show broader size
on the mass of the active material of the anode (the mass of total distribution (Fig. S5c). The EDS analysis of InP@TiO2-C shows a
composite) for half-cell testing. For full-cell testing, a composite of homogeneous distribution of InP, TiO2, and C in the as-obtained In-
LiFePO4-graphite (LFP@G) was prepared as a cathode. LFP@G was P@TiO2-C composite (Fig. S6). The HRTEM (Fig. 1f) and corresponding
prepared via ball milling, in which a mixture of lithium iron phosphate FFT image (Fig. 1g) of InP@TiO2-C demonstrate the presence of InP and
(LFP) (1.4 g, 97%, Sigma-Aldrich)) and graphite (G) (0.6 g, 99.9995%, TiO2 crystal phases, which is consistent with the XRD, Raman, and XPS
Alfa Aesar)) was milled for 12 h. The mass ratios between the anode results. A more detailed microstructure confirmed by the TEM and EDS
(InP@TiO2-C) and cathode (LFP@G) were 1:1.4 and 1:2 to balance the mapping images of InP@TiO2-C indicates the uniformly distributed
capacity contribution from each electrode in a full cell. For both full nanoparticles of InP and TiO2 embedded in amorphous C of the as-
cells, the anode and cathode films were electrochemically pre-dis- prepared InP@TiO2-C composite (Fig. S7).
charged and pre-charged to remove the effects of SEI layer formation The Li-storage behavior of the as-prepared InP@TiO2-C anode was
and allow the batteries to attain a stable condition. The voltage window studied by its galvanostatic discharge/charge profile (Fig. 2a), curren-
applied for the full cell was 0–3.8 V (versus Li/Li+). t–potential curves (CVs) during the first three cycles (Fig. 2b), and
crystal structure of each lithiation/delithiation state (Fig. 2c). The
3. Results and discussion electrochemical reaction mechanism of Li ions with InP follows the
multistep process as shown in Fig. 2c [36,38]. In brief, the first dis-
Fig. 1a illustrates the mechanochemical synthetic scheme of the charge process (stage I in Fig. 2a and the peak at ~0.61 V in Fig. 2b)
InP@TiO2-C composite using two-step ball-milling. The crystalline represents the reaction between InP and Li+ to form In and Li3P
structure of InP@TiO2-C synthesized is consistent with the theoretical (InP + 3Li+ +3e− → In + Li3P; Eq. (2) below), corresponding to the
crystal structures of its crystalline constituents (InP and TiO2) from XRD increasing specific capacity from 0 to ~430 mAh g−1. In stage II (dis-
patterns (Fig. 1b and Fig. S2). The XRD patterns of the synthesized charge from 0.5 to 0.1 V), a strong reduction in the peak is observed at
InP@TiO2-C composite confirm the formation of the cubic phase ~0.2 V, indicating the reaction between In and Li+ to form InLi
structure of InP (PDF#32-0452) and tetragonal rutile phase structure of (In + Li+ +e− → InLi; Eq. (3) below) with a significant capacity
TiO2 (PDF#21-1276) without traces of precursors of In2O3 (PDF#06- contribution up to ~800 mAh g−1. In stage III (discharge from 0.1 to
0416), Ti (PDF#44-1294), and P (PDF#44-0906) in the as-obtained 0.01 V), the specific capacity continuously increases to ~1067 mAh
powder. The crystalline structure of InP@TiO2 without added C is al- g−1, which can be attributed to the formation of the solid electrolyte
most identical to that of InP@TiO2-C because of the amorphous nature interphase (SEI) [38]. Meanwhile, during the first charge process, two
of C (Fig. S3). For InP@C synthesized using two step ball milling of In, major oxidation peaks observed at ~0.72 and ~1.15 V are attributed to
P, and C, the diffraction peaks were all matched with the InP phase (Fig. the two reversible reactions (InLi → In + Li+ + e− and In + Li3P →
S3). All these results confirm the successful synthesis of a target com- InP + 3Li+ + 3e−) as shown in Eqs. (4) and (5). The small anodic
posite material. Raman spectroscopy of InP@TiO2-C in Fig. 1c further peaks observed at 0.5, 0.76, and 1.5 V are related to the multistep
confirms the structural information of the constituents: stretching peaks process of partial delithiation from InP (LixInP → LiyInP + zLi+, x
at ~137, 238, 399, and 594 cm−1 of rutile TiO2 correspond to the = y + z, Fig. S8) where the XRD peak at ~26° is shifted to a higher
symmetries of B1g, multi-photon process, Eg, and A1g, respectively angle due to the Li+ extraction from InP. In the first charge stage (stage
[62,63] and two other characteristic peaks observed at ~1335 and IV, from 0.01 to 0.8 V), the specific capacity increases to ~370 mAh
~1600 cm−1 indicate the presence of carbon. As for the cubic structure g−1, whereas in the second stage of charging (stage V, from 0.8 to
InP, there is no detectable Raman signal in the InP@TiO2-C nano- 3.0 V), the capacity continuously increases to ~800 mAh g−1, in-
composites because of the stronger signal from TiO2. This can be at- dicating full reversion except for the SEI formation. The overall elec-
tributed to either i) different scattering sensitivities between InP and trochemical reaction during the first discharge/charge process can be
TiO2 or ii) surface enhanced Raman scattering (SERS) between InP and written as:
C. Xiaodan et al. reported the Raman quenching of semiconducting ZnS Discharge:
in the presence of TiO2. They attributed this phenomenon to extremely
InP + 3Li+ + 3e− → In + Li3P (2)
sensitive Raman scattering from TiO2, which effectively reduces the

Raman active modes from ZnS [47]. However, Pan et al. demonstrated In + Li +
+e → InLi (3)
the Raman peak quenching of ZnS when mixed with graphene because
of the SERS effect [48]. They attributed this phenomenon to the charge Charge:

3
Q. Hai Nguyen, et al. Chemical Engineering Journal 399 (2020) 125826

Fig. 1. (a) Schematic of the formation of InP@TiO2-C. (b–e) XRD pattern, Raman spectra, deconvoluted In3d XPS profiles, SEM image, high-resolution TEM image,
and FFT image of as-prepared InP@TiO2-C nanocomposite, respectively.

material, two main diffraction peaks of Cu were observed in all samples


InLi → In + Li+ + e− (4)
in the ex situ XRD analysis. At the open circuit potential (Fig. 3a-i), the
In + Li3P → InP + 3Li+ + 3e− (5) three main peaks observed at ~26.3°, 30.43°, and 43.6° correspond to
the (1 1 1), (2 0 0), and (2 2 0) lattice planes of InP (PDF#32–0452),
Notably, there are no significant reduction/oxidation peaks that which is consistent with the characterization of the as-prepared powder
correspond to the intercalation/deintercalation between Li-ions and (Fig. 1b). However, the other peaks reflecting the presence of TiO2
TiO2 or carbon, indicating that TiO2 and C mainly play the roles of could not be observed, which could be attributed to the strong peaks
buffering materials for active InP during cycling. Because TiO2 delivers from Cu, a relatively small amount of TiO2 in InP@TiO2-C, and the
a significantly low specific capacity of ~70 mAh g−1 at a current presence of the PVDF binder that screens the diffraction peaks from
density of 200 mA g−1 (Fig. S9) and small amount of TiO2 (~20 wt%) is TiO2 [71]. The ex situ XRD pattern acquired at 0.5 V of the initial
embedded in the InP@TiO2-C nanocomposite, its contribution to the discharge process (stage I in Fig. 3a-ii) indicates the appearance of In
total specific capacity of the as-prepared anode is very small (PDF#05–0642) and Li3P (PDF#04-0525) phases as well as a reduction
[55,57,69,70]. More quantitatively, we can estimate the actual capacity in peak intensities from InP when compared with those of the pristine
contribution from TiO2 in InP@TiO2-C is as low as 3% using the mea- stage (Fig. 3a-i), which indicates that the reaction occurred according to
sured content (EDX analysis, Table S1) with the theoretical (or mea- Eq. (2). An ex situ HRTEM image for the InP@TiO2-C electrode after
sured) capacity of InP, TiO2, and C, respectively (see the calculation in initial discharge to 0.5 V consistently shows the presence of the In (d-
Table S2). A comparison of the CVs and voltage profiles of InP@TiO2-C spacing of 0.27 nm), InP phases (d-spacing of 0.34 nm), and Li3P (d-
with those of other as-prepared composites (InP@TiO2 and InP@C) spacing of 0.209 nm) (Fig. 3b-i) corresponding to the crystal planes of
indicates a similar behavior (Figs. 2 and S9), demonstrating the similar In (1 0 1), InP (1 1 1), and Li3P (1 0 3) which are consistent with the ex
characteristics of InP despite the different synthesis routes. situ XRD result shown in Fig. 3a-ii. After the discharge process at 0.01 V
For further insightful understanding of the lithiation/delithiation (Fig. 3a-iii), the XRD peaks at ~22.6° ((1 1 1) plane) and ~37° ((2 2 0)
mechanism of the InP@TiO2-C electrode, various ex situ characteriza- plane) confirm the InLi phase (PDF#09–0066), demonstrating that the
tions were performed, including XRD, HRTEM, and XPS. The ex situ In formed in stage I reacted with Li-ions to form the InLi phase (Eq. (3)).
XRD patterns (Fig. 3a), HRTEM images (Fig. 3b), and XPS profiles This reaction is further validated by an HRTEM image (Fig. 3b-ii), in
(Fig. 4) of the InP@TiO2-C electrode material were obtained at various which the crystal plane of the InLi ((2 2 0) plane with a d-spacing of
cutoff potentials (discharges at 0.5 V (D-0.5 V), discharge at 0.01 V (D- 0.24 nm) is observed. In addition, the peak intensity of In strongly re-
0.01 V), charge at 0.75 V (C-0.75 V), and charge at 3.0 V (C-3.0 V)) duced, which is consistent with the reaction in Eq. (3). The residual InP
based on the reduction/oxidation peaks as shown in the CV curve phase in stage II is confirmed by both, the ex situ XRD pattern (Fig. 3a-
(Fig. 2b). Owing to the Cu current collector beneath the electrode iii) and HRTEM image (Fig. 3b-ii), which could be attributed to the

4
Q. Hai Nguyen, et al. Chemical Engineering Journal 399 (2020) 125826

Fig. 2. (a) Initial galvanostatic discharge/charge profile at 100 mA g−1. (b) Initial three cycle current–potential curves at scan rate of 0.2 mV s−1. (c) Crystal
structures of the corresponding lithiation/delithiation stages.

incomplete activation of active InP. To further confirm the highly re- the XRD pattern and HRTEM image, and the In peak in XRD increased
versible conversion of the delithiation process, the ex situ XRD patterns in intensity, as shown in Fig. 3a-iv, indicating the reversible reaction as
(Fig. 3a-(iv and v)) and HRTEM images (Fig. 3b (iii and iv)) of the in Eq. (3). At the full-charging stage at 3.0 V, the ex situ XRD pattern
electrode at 0.75 and 3.0 V of the initial charging cutoff potential were (Fig. 3a-v) and HRTEM image (Fig. 3b-iv) indicate the reappearance of
obtained. After charging to 0.75 V, the InLi phase disappeared in both the InP phase and reduction of the In phase, suggesting the reversible

Fig. 3. (a) Ex situ XRD patterns and (b) ex situ HRTEM of InP@TiO2-C electrode collected at various cutoff potentials.

5
Q. Hai Nguyen, et al. Chemical Engineering Journal 399 (2020) 125826

Fig. 4. Ex situ XPS profiles of InP@TiO2-C electrode collected at various cutoff potentials.

conversion as in Eq. (4). However, the InP peaks in XRD could not fully discharged state (pink color in the P 2p spectrum in the right panel of
recover their intensities after the full-charging process when compared Fig. 4b) when compared with its pristine stage (pink color in the P 2p
with the pristine stage, which could be attributed to the presence of the spectrum in the central panel of Fig. 4a). In addition, Li 1 s spectra
irreversible SEI layer after lithiation. The lithiation mechanism was indicate the Li3P alloy (purple color at 57 eV in the right panel of
further examined using ex situ XPS (Fig. 4). The deconvoluted In 3d Fig. 4b), decomposition of the electrolyte (brown color at 55.5 eV in the
(left panel), P 2p (central panel), and Li 1s (right panel) profiles of the right panel of Fig. 4b–4e), and residues of lithium metal (yellow color at
InP@TiO2-C electrode measured at various cutoff potentials clearly 54 eV in the right panel of Fig. 4b–4e) [74]. These results could be
confirm the lithiation/delithiation mechanism revealed by the CV, ex attributed to the reaction between InP and Li ions to form In and Li3P
situ XRD, and ex situ HRTEM analyses. In brief, in stage I of D-0.5 V (as in Eq. (2)). Fig. 4c exhibits the deconvoluted XPS peaks for D-0.01 V
(Fig. 4b), the XPS signals detected at ~443.7 eV in the In 3d spectra with an increase in the peak intensity of InLi and Li3P (binding energies
(blue color in the left panel of Fig. 4b) indicate the typical state of metal of InP and InLi are the same [75]) as the D-0.5 V case), indicating the
In (In0) [72,73]. The peak detected at 131 eV corresponds to the Li3P continuous reduction of InP to In and Li3P and the reaction of In with Li
phase (purple color in the central panel of Fig. 4b) [74]. Although InP ions to form InLi (as in Eq. (3). Furthermore, the deconvoluted peak of
peaks are still observed in the deconvoluted spectrum in P 2p InP (pink color) at ~129 eV (Fig. 4c) is the residue, implying the in-
(~129 eV), their intensities significantly decrease at the D-0.5 V complete reaction of active InP, as discussed earlier. For C-0.75 V

6
Q. Hai Nguyen, et al. Chemical Engineering Journal 399 (2020) 125826

Fig. 5. (a) Cyclic performances of InP@TiO2-C, InP@TiO2, and InP@C at current density of 100 mA g−1. (b) Long-term cyclic performance of InP@TiO2-C at high
current density of 500 mA g−1.

(Fig. 4d), although the deconvoluted XPS peaks corresponding to In and polarization between two main cathodic and anodic peaks (from 1st to
Li3P phases persist, their intensities decrease, which is consistent with 200th cycle) which enhances the efficiency of charge transfer during
the reaction in Eq. (4). Meanwhile, when charging to 3.0 V (Fig. 4e), the the discharge/charge process (Fig. S11) [80,81]. The long-term cyclic
peak of InP reappears with the disappearance of peaks from InLi and performance of InP@TiO2-C measured at a high current density of
Li3P, which indicates the highly reversible conversion process. 500 mA g−1 for 800 cycles (Fig. 5b) showed a behavior similar to the
The cyclic performance was measured for three different synthe- case of low current density (0.1 A g−1); the specific capacity gradually
sized electrodes (InP@TiO2-C, InP@TiO2, and InP@C) at 100 mA g−1 increased during initial 120 cycles and then remained steady until 800
in the potential range of 0.01–3.0 V (Fig. 5a). The initial coulombic cycles with a capacity of ~750 mAh g−1. We note that the similar
efficiencies (ICEs) of the InP@TiO2-C, InP@TiO2, and InP@C electrodes polarization reduction was observed for the capacity increasing region
were 75%, 84%, and 82% from the initial charge/discharge capacities (470–570 cycle, Fig. S12).[80,81] This result is superior to those for
of 805/1067, 691/821, and 832/942 mAh g−1/mAh g−1, respectively. most other metal phosphide-based anodes in previous works, as sum-
The capacity decrease of initial three cycles for the InP@TiO2-C is marized in Table S3.
pronounced which may be associated with the gradual SEI layer for- The rate capabilities of three electrodes (InP@TiO2-C, InP@TiO2,
mation for hybrid TiO2-C buffering matrix; the coulombic efficiencies of and InP@C) were measured at current densities of 0.1, 0.5, 1, 3, 5, and
InP@TiO2-C were 90 and 93% for 2nd and 3rd cycle, respectively, 10 A g−1 in the voltage range of 0.01–3.0 V as another important cri-
which were lower than those of InP@TiO2 (96% and 97%) and InP@C terion (Fig. 6a). For InP@TiO2, the specific capacities significantly de-
(96 and 97%) as shown in Fig. S10. However, despite higher capacities creased with an increase in current density (from ~600 mAh g−1 at
of InP@TiO2 and InP@C than that of InP@TiO2-C in initial several initial 0.1 A g−1 to ~200 mAh g−1 at 10 mA g−1) and did not fully
cycles, their capacities gradually decreased and became less than 400 reverse after the current density returned to 0.1 A g−1 (~400 mAh
mAh g−1 after 120 cycles. Meanwhile, the InP@TiO2-C electrode g−1). When compared with InP@TiO2, although InP@C showed higher
showed very stable and even slightly increasing capacity until the 250th specific capacities at all current densities (~750, 700, 650, 600, 550,
cycle (807 mAh g−1 with CE of 100% at the 250th cycle). This cyclic and 500 mAh g−1 at 0.1, 0.5, 1, 3, 5, and 10 A g−1, respectively),
stability could be attributed to the presence of the TiO2-C hybrid ma- irreversible capacity recovery was still observed after returning to the
trix, which prevents the aggregation of InP particles and reduces the original current density. For InP@TiO2-C, the rate performance was
stresses/strains of active InP particles, thereby avoiding the particle superior to that of other electrodes at all current densities; capacity
pulverization caused by volume expansion during repeated lithiation/ retentions were 97% (0.5 A g−1), 95% (1.0 A g−1), 88% (3.0 A g−1),
delithiation processes [76]. The increasing capacity with the increase in 87% (5.0 A g−1), and 85% (10.0 A g−1) compared with the capacity at
the number of cycles is associated with the gradual activation of the 0.1 A g−1. Particularly, when the current density was returned to 0.1 A
active material and polymer/gel-like film formation during repeated g−1 from 10 A g−1, the specific capacity rather slightly increased to
cycles [77–79]. This might be also associated with the reduced 720 mA g−1 (120% when compared with the capacity at the initial

7
Q. Hai Nguyen, et al. Chemical Engineering Journal 399 (2020) 125826

Fig. 6. (a) Rate capability and (b) EIS profiles of InP@TiO2-C, InP@TiO2, and InP@C after 35 cycles of rate capability testing. (c) Ex situ SEM images of InP@TiO2-C,
InP@TiO2, and InP@C after rate capability and EIS testing.

current density). This behavior is similar to the cyclic performance its structure without much aggregations during the cycling.
(increasing capacity during 100 cycles in the cyclic performance in Finally, the electrochemical performance of practical full-cell bat-
Fig. 5 resulting from the continuous activation of the InP active mate- teries was measured with LFP@G as a cathode and synthesized
rial and more polymer/gel-like film formation during the repeated cy- InP@TiO2-C as an anode at the current density of 100 mA g−1 (mass
cles). These results could be further understood by comparing Nyquist ratio of 2:1 (cathode:anode)) in the potential range of 1.0–3.8 V
plots and ex situ SEM measurements for the three electrodes after rate (Fig. 7a). The initial three voltage profiles of the full cell indicate a
capability measurements (Fig. 6b and c). The smallest semicircle radius nominal working voltage of ~1.7 V and ICEs of 95%, 97%, and 97%
of InP@TiO2-C when compared with those of InP@TiO2 and InP@C from the charge/discharge capacities of 95/101, 94/97, and 92/95
after rate capability measurement indicates the lowest charge transfer mAh g−1/mAhg−1 for the 1st, 2nd, and 3rd cycles, respectively
resistance, which promotes electron and Li-ion transport in the elec- (Fig. 7b). Fig. 7c shows the cyclic performance of the full cell with mass
trode [82,83]. Clearly, the combination of TiO2 and carbon as a buf- ratios of 1.4:1 and 2:1 between the cathode and anode. The capacity of
fering hybrid matrix allows alleviated volume change and maintenance the full cell was calculated based on the total mass of the anode and
of high conductivity, resulting in the prevention of InP particle pul- cathode. It was found that the full cell with a mass ratio of 2:1 (cath-
verization from the large volume expansion during cycling. In line with ode:anode) showed superior cyclic behavior because of the increased
this, the low charge transfer resistance and high rate capability of In- mass ratio of the cathode, which enabled more efficient contribution to
P@TiO2-C explain the observation in long-term cyclic performance at a the capacity of full cells. In general, the specific capacity balance be-
high capacity (Fig. 6b) where the capacity has been well-maintained at tween cathode and anode (i.e., N/P ratio) is very important for the
0.5 A g−1 relative to the capacity at 0.1 A g−1. Ex situ SEM images of efficient and balanced ion movement in the full cell. In most cases, since
the electrodes after rate capability testing further demonstrate the su- the specific capacity of cathode is less than that of anode, the loading
periority of InP@TiO2-C, in that while many cracks and delaminated amount of cathode should be higher than that of anode to optimize N/P
parts were observed in InP@TiO2 and InP@C, the original morphology ratio. In our particular case, the specific capacity of cathode
was retained for InP@TiO2-C when comparing the images of all samples (LiFePO4@G) was separately measured as ~130 mAh g-1 (Fig. S15),
before/after cycling (Fig. 6c and Fig. S13). Additionally, elemental which is much lower than that of anode (InP@TiO2-C, ~823 mAh g−1
mapped images (Fig. S14) and EDS analysis (Table S4) reveal the uni- in Fig. 5a). Because of this large specific capacity difference between
form distribution of all the components without much change in the cathode and anode, when we applied higher mass ratio of cath-
ratio of all the constituent atoms; indicating the InP@TiO2-C maintain ode:anode (2:1), the full cell showed higher capacities than the case of

8
Q. Hai Nguyen, et al. Chemical Engineering Journal 399 (2020) 125826

Fig. 7. (a) Schematic of full-cell battery. (b) Voltage profile during three cycles. (c) Cyclic performance of InP@TiO2-C/LFP@G full cell at current density of
100 mA g−1.

lower mass ratio of cathode:anode (1.4:1) during the capacity entire interests or personal relationships that could have appeared to influ-
capacity range measured (150 cycles). Specifically, the full cell with a ence the work reported in this paper.
2:1 mass ratio exhibited a capacity of 126 mAh g−1 at the first cycle
(corresponding energy density of 214 Wh kg−1) and 77 mAh g−1 at the Acknowledgments
150th cycle (corresponding energy density of 131 Wh kg−1). This result
suggests that the InP@TiO2-C composite synthesized by the facile and This research was supported by the Basic Science Research Program
scalable mechanical-milling process can be a promising anode material through the National Research Foundation of Korea (NRF) funded by
for the next generation of practical energy storage devices. the Ministry of Education (NRF-2019R1F1A1057709). This research
was also supported by Basic Science Research Capacity Enhancement
4. Conclusions Project through Korea Basic Science Institute (National Research
Facilities and Equipment Center) grant funded by the Ministry of
We demonstrated a novel InP@TiO2-C nanocomposite synthesized Education (2019R1A6C1010016).
using a facile two-step mechanochemical milling process. Various
characterizations (XRD, Raman, SEM, TEM, XPS, and EDX) confirmed Appendix A. Supplementary data
that InP@TiO2-C displayed a beneficial morphology as an anode for
LIBs; this is due to the fact that active InP nanoparticles were effectively Supplementary data to this article can be found online at https://
surrounded by the hybrid crystalline TiO2 and amorphous conductive doi.org/10.1016/j.cej.2020.125826.
C. This hybrid conductive matrix of TiO2-C functioned as an efficient
buffering matrix for active InP when InP@TiO2-C was used as an anode References
in LIBs. InP@TiO2-C exhibited excellent electrochemical performance
for both half- and full-cell batteries. When compared with InP@TiO2 [1] H.S. Moon, J.H. Lee, S. Kwon, I.T. Kim, S.G. Lee, Mechanisms of Na adsorption on
(without adding C) and InP@C (ball milling of commercial InP and C), graphene and graphene oxide: density functional theory approach, Carbon Lett. 16
(2015) 116–120.
the InP@TiO2-C nanocomposite showed superior performance in terms [2] T. Kim, W. Song, D.-Y. Son, L.K. Ono, Y. Qi, Lithium-ion batteries: outlook on
of the specific capacity (~850 mAh g−1 after 120 cycles at present, future, and hybridized technologies, J. Mater. Chem. A 7 (2019)
100 mA g−1), long life span (750 mAh g−1 after 800 cycles at 2942–2964.
[3] H. Kim, M. Kim, Y.H. Yoon, Q.H. Nguyen, I.T. Kim, J. Hur, S.G. Lee, Sb2Te3-TiC-C
500 mA g−1), and rate performance (~97%, 95%, 88%, 87%, 85%, and nanocomposites for the high-performance anode in lithium-ion batteries,
120% at 0.5, 1, 3, 5, and 10 A g−1, and returned 0.1 A g−1 of current Electrochim Acta (2018).
density, respectively, as compared with the capacity at an initial current [4] D. Saikia, T.H. Wang, C.J. Chou, J. Fang, L.D. Tsai, H.M. Kao, A comparative study
of ordered mesoporous carbons with different pore structures as anode materials for
density of 0.1 A g−1). With regard to the practicality of the full cell, the lithium-ion batteries, RSC Adv. 5 (2015) 42922–42930.
InP@TiO2-C anode delivered a promising initial energy density of 214 [5] E. Yoo, J. Kim, E. Hosono, H. Zhou, T. Kudo, I. Honma, Large reversible Li storage of
Wh kg−1 and a stability of as high as 150 cycles during cyclic perfor- graphene nanosheet families for use in rechargeable lithium ion batteries, Nano
Lett. 8 (2008) 2277–2282.
mance. [6] Q.H. Nguyen, I.T. Kim, J. Hur, Core-shell Si@ c-PAN particles deposited on graphite
as promising anode for lithium-ion batteries, Electrochim Acta 297 (2019)
Declaration of Competing Interest 355–364.
[7] C. Liu, X.G. Liu, J. Tan, Q.F. Wang, H. Wen, C.H. Zhang, Nitrogen-doped graphene
by all-solid-state ball-milling graphite with urea as a high-power lithium ion battery
The authors declare that they have no known competing financial anode, J. Power Sources 342 (2017) 157–164.

9
Q. Hai Nguyen, et al. Chemical Engineering Journal 399 (2020) 125826

[8] Q.H. Nguyen, N.T. Hung, S.J. Park, I.T. Kim, J. Hur, Enhanced performance of 159 (2012) A1941–A1948.
carbon-free intermetallic zinc titanium alloy (Zn-ZnxTiy) anode for lithium-ion [38] S. Liu, W. Wei, X. He, Indium phosphide/reduced graphene oxide composites as
batteries, Electrochim Acta (2019). high-performance anodes in lithium-ion batteries, ChemElectroChem 5 (2018)
[9] T.A. Nguyen, I.T. Kim, S.W. Lee, Chitosan-tethered iron oxide composites as an 3315–3322.
antisintering porous structure for high-performance li-ion battery anodes, J. Am. [39] S. Tamang, C. Lincheneau, Y. Hermans, S. Jeong, P. Reiss, Chemistry of InP na-
Ceram. Soc. 99 (2016) 2720–2728. nocrystal syntheses, Chem. Mater. 28 (2016) 2491–2506.
[10] N.Q. Hai, J.S. Choi, Y.-C. Lee, I.T. Kim, J. Hur, 3D hierarchical structure of MoS2@ [40] M.-D. Gerngross, J. Carstensen, H. Föll, Single-Crystalline membranes in indium
G-CNT combined with post-film annealing for enhanced lithium-ion storage, J. Ind. phosphide: fabrication process and characterization using FFT impedance Analysis,
Eng. Chem. (2018). J. Electrochem. Soc. 159 (2012) H857–H863.
[11] Q.H. Nguyen, J. Hur, MoS2–TiC–C nanocomposites as new anode materials for high- [41] E. Aharon-Shalom, A. Heller, Efficient p-InP (Rh-H alloy) and p-InP (Re-H alloy)
performance lithium-ion batteries, J. Nanosci. Nanotechnol. 19 (2019) 996–1000. hydrogen evolving photocathodes, J. Electrochem. Soc. 129 (1982) 2865–2866.
[12] I.T. Kim, J.C. Knight, H. Celio, A. Manthiram, Enhanced electrochemical perfor- [42] X. Duan, Y. Huang, Y. Cui, J. Wang, C.M. Lieber, Indium phosphide nanowires as
mances of Li-rich layered oxides by surface modification with reduced graphene building blocks for nanoscale electronic and optoelectronic devices, Nature 409
oxide/AlPO4 hybrid coating, J. Mater. Chem. A 2 (2014) 8696–8704. (2001) 66.
[13] Y.S. Mun, Y.N. Pham, V.K.H. Bui, S.T. Tanaji, H.U. Lee, G.W. Lee, J.S. Choi, [43] Q. Lin, D. Sarkar, Y. Lin, M. Yeung, L. Blankemeier, J. Hazra, W. Wang, S. Niu,
I.T. Kim, Y.C. Lee, Tin oxide evolution by heat-treatment with tin-aminoclay (SnAC) J. Ravichandran, Z. Fan, Scalable indium phosphide thin-film nanophotonics plat-
under argon condition for lithium-ion battery (LIB) anode applications, J. Power form for photovoltaic and photoelectrochemical devices, ACS Nano 11 (2017)
Sources 437 (2019). 5113–5119.
[14] I.T. Kim, A. Magasinski, K. Jacob, G. Yushin, R. Tannenbaum, Synthesis and elec- [44] P. Lobaccaro, A. Raygani, A. Oriani, N. Miani, A. Piotto, R. Kapadia, M. Zheng,
trochemical performance of reduced graphene oxide/maghemite composite anode Z. Yu, L. Magagnin, D.C. Chrzan, Electrodeposition of high-purity indium thin films
for lithium ion batteries, Carbon 52 (2013) 56–64. and its application to indium phosphide solar cells, J. Electrochem. Soc. 161 (2014)
[15] T.L. Nguyen, D. Park, I.T. Kim, FexSnyOz Composites as Anode Materials for D794–D800.
Lithium-Ion Storage, J. Nanosci. Nanotechnol. 19 (2019) 6636–6640. [45] J. Wallentin, N. Anttu, D. Asoli, M. Huffman, I. Åberg, M.H. Magnusson, G. Siefer,
[16] Y.S. Mun, D. Kim, I.T. Kim, Electrochemical performance of FeSb2-P@C composites P. Fuss-Kailuweit, F. Dimroth, B. Witzigmann, InP nanowire array solar cells
as anode materials for lithium-ion storage, J. Nanosci. Nanotechnol. 18 (2018) achieving 13.8% efficiency by exceeding the ray optics limit, Science 339 (2013)
1343–1346. 1057–1060.
[17] T.P. Nguyen, J.H. Kim, I.T. Kim, Electrochemical performance of Sn/SnO/Ni3Sn [46] J. Wang, M.S. Gudiksen, X. Duan, Y. Cui, C.M. Lieber, Highly polarized photo-
composite anodes for lithium-ion batteries, J. Nanosci. Nanotechnol. 19 (2019) luminescence and photodetection from single indium phosphide nanowires, Science
1001–1005. 293 (2001) 1455–1457.
[18] T.N. Pham, S.T. Tanaji, J.S. Choi, H.U. Lee, I.T. Kim, Y.C. Lee, Preparation of Sn- [47] N. Miao, B. Xu, N.C. Bristowe, J. Zhou, Z. Sun, Tunable magnetism and extra-
aminoclay (SnAC)-templated Fe3O4 nanoparticles as an anode material for lithium- ordinary sunlight absorbance in indium triphosphide monolayer, J. Am. Chem. Soc.
ion batteries, RSC Adv. 9 (2019) 10536–10545. 139 (2017) 11125–11131.
[19] S.Y. Son, S.A. Hong, S.Y. Oh, Y.C. Lee, G.W. Lee, J.W. Kang, Y.S. Huh, I.T. Kim, [48] M.S. Gudiksen, J. Wang, C.M. Lieber, Size-dependent photoluminescence from
Crab-shell biotemplated SnO2 composite anodes for lithium-ion batteries, J. single indium phosphide nanowires, J. Phys. Chem. B 106 (2002) 4036–4039.
Nanosci. Nanotechnol. 18 (2018) 6463–6468. [49] N.Q. Hai, H. Kim, I.S. Yoo, J. Hur, Facile and scalable preparation of a MoS2/carbon
[20] S. Batool, M. Idrees, J. Kong, J. Zhang, S. Kong, M. Dong, H. Hou, J. Fan, H. Wei, nanotube nanocomposite anode for high-performance lithium-ion batteries: effects
Z. Guo, Assessment of the electrochemical behaviour of silicon@carbon nano- of carbon nanotube content, J. Nanosci. Nanotechnol. 19 (2019) 1494–1499.
composite anode for lithium-ion batteries, J. Alloy. Compd. 832 (2020) 154644. [50] N.Q. Hai, H. Kim, I.S. Yoo, J.H. Kim, J. Hur, Comparative study of mechanically
[21] M. Idrees, L. Liu, S. Batool, H. Luo, J. Liang, B. Xu, S. Wang, J.J.E.S. Kong, Cobalt- milled MoS2 and MoSe2 in graphite matrix as anode materials for high-performance
doping enhancing electrochemical performance of silicon/carbon nanocomposite as lithium-ion batteries, J. Nanosci. Nanotechnol. 18 (2018) 6469–6474.
highly efficient anode materials in lithium-ion, Batteries (2019). [51] D. Gangaraju, V. Sridhar, I. Lee, H. Park, Graphene - carbon nanotube - Mn3O4
[22] B.M. Li, Y. Yan, C.T. Shen, Y. Yu, Q.H. Wang, M.K. Liu, Extraordinary lithium ion mesoporous nano-alloys as high capacity anodes for lithium-ion batteries, J. Alloy.
storage capability achieved by SnO2 nanocrystals with exposed 221 facets, Compd. 699 (2017) 106–111.
Nanoscale 10 (2018) 16217–16230. [52] S. El-Deen, A. Hashem, A.A. Ghany, S. Indris, H. Ehrenberg, A. Mauger, C. Julien,
[23] R.J. Li, X.Z. Zhu, Q.F. Fu, G.S. Liang, Y.J. Chen, L.J. Luo, M.Y. Dong, Q. Shao, Anatase TiO 2 nanoparticles for lithium-ion batteries, Ionics 24 (2018) 2925–2934.
C.F. Lin, R.B. Wei, Z.H. Guo, Nanosheet-based Nb12O29 hierarchical microspheres [53] A.R. Armstrong, G. Armstrong, J. Canales, P.G. Bruce, TiO2–B nanowires as nega-
for enhanced lithium storage, Chem. Commun. 55 (2019) 2493–2496. tive electrodes for rechargeable lithium batteries, J. Power Sources 146 (2005)
[24] W. Qi, J.G. Shapter, Q. Wu, T. Yin, G. Gao, D. Cui, Nanostructured anode materials 501–506.
for lithium-ion batteries: principle, recent progress and future perspectives, J. [54] B. Qiu, M. Xing, J. Zhang, Mesoporous TiO2 nanocrystals grown in situ on graphene
Mater. Chem. A 5 (2017) 19521–19540. aerogels for high photocatalysis and lithium-ion batteries, J. Am. Chem. Soc. 136
[25] S. Goriparti, E. Miele, F. De Angelis, E. Di Fabrizio, R.P. Zaccaria, C. Capiglia, (2014) 5852–5855.
Review on recent progress of nanostructured anode materials for Li-ion batteries, J. [55] Y. Yang, Y. Bai, S. Zhao, Q. Chang, W. Zhang, Electrochemical performances of Si/
Power Sources 257 (2014) 421–443. TiO2 composite synthesized by hydrothermal method, J. Alloy. Compd. 579 (2013)
[26] X. He, R. Wang, M.C. Stan, E. Paillard, J. Wang, H. Frielinghaus, J. Li, In situ in- 7–11.
vestigations on the structural and morphological changes of metal phosphides as [56] J.S. Chen, L.A. Archer, X.W.D. Lou, SnO2 hollow structures and TiO2 nanosheets for
anode materials in lithium-ion batteries, Adv. Mater. Interfaces 4 (2017) 1601047. lithium-ion batteries, J. Mater. Chem. 21 (2011) 9912–9924.
[27] Y. Pei, Y. Cheng, J. Chen, W. Smith, P. Dong, P.M. Ajayan, M. Ye, J. Shen, Recent [57] S. Huang, L. Kavan, I. Exnar, M. Grätzel, Rocking chair lithium battery based on
developments of transition metal phosphides as catalysts in the energy conversion nanocrystalline TiO2 (anatase), J. Electrochem. Soc. 142 (1995) L142–L144.
field, J. Mater. Chem. A 6 (2018) 23220–23243. [58] N. Li, G. Liu, C. Zhen, F. Li, L. Zhang, H.M. Cheng, Battery performance and pho-
[28] S.-O. Kim, A. Manthiram, Phosphorus-rich CuP2 embedded in carbon matrix as a tocatalytic activity of mesoporous anatase TiO2 nanospheres/graphene composites
high-performance anode for lithium-ion batteries, ACS Appl. Mater. Interfaces 9 by template-free self-assembly, Adv. Funct. Mater. 21 (2011) 1717–1722.
(2017) 16221–16227. [59] M. Madian, A. Eychmuller, L. Giebeler, Current Advances in TiO2-Based
[29] S. Liu, X. He, J. Zhu, L. Xu, J. Tong, Cu 3 p/RGO nanocomposite as a new anode for Nanostructure Electrodes for High Performance Lithium Ion Batteries, Batteries-
lithium-ion batteries, Sci. Rep. 6 (2016) 35189. Basel 4 (2018).
[30] W. Li, X. Li, J. Yu, J. Liao, B. Zhao, L. Huang, A. Abdelhafiz, H. Zhang, J.H. Wang, [60] Y. Zhang, G.W. Wen, S. Fan, W.H. Ma, S.H. Li, T. Wu, Z.C. Yu, B.R. Zhao, Alcoholic
Z. Guo, A self-healing layered GeP anode for high-performance Li-ion batteries hydroxyl functionalized partially reduced graphene oxides for symmetric super-
enabled by low formation energy, Nano Energy 61 (2019) 594–603. capacitors with long-term cycle stability, Electrochim Acta 313 (2019) 59–69.
[31] J. Choi, W.-S. Kim, K.-H. Kim, S.-H. Hong, Sn 4 P 3–C nanospheres as high capa- [61] W.W. Zhong, J.D. Huang, S.Q. Liang, J. Liu, Y.J. Li, G.M. Cai, Y. Jiang, J. Liu, New
citive and ultra-stable anodes for sodium ion and lithium ion batteries, J. Mater. prelithiated V2O5 superstructure for lithium-ion batteries with long cycle life and
Chem. A 6 (2018) 17437–17443. high power, Acs Energy Lett 5 (2020) 31–38.
[32] A. Lu, X. Zhang, Y. Chen, Q. Xie, Q. Qi, Y. Ma, D.-L. Peng, Synthesis of Co2P/ [62] S. Challagulla, K. Tarafder, R. Ganesan, S. Roy, Structure sensitive photocatalytic
graphene nanocomposites and their enhanced properties as anode materials for reduction of nitroarenes over TiO 2, Sci. Rep. 7 (2017) 8783.
lithium ion batteries, J. Power Sources 295 (2015) 329–335. [63] H.L. Ma, J.Y. Yang, Y. Dai, Y.B. Zhang, B. Lu, G.H. Ma, Raman study of phase
[33] H. Guo, H. Cai, W. Li, C. Chen, K. Chen, Y. Zhang, Y. Li, M. Wang, Y. Wang, Tailored transformation of TiO2 rutile single crystal irradiated by infrared femtosecond
Ni2P nanoparticles supported on N-doped carbon as a superior anode material for laser, Appl. Surf. Sci. 253 (2007) 7497–7500.
Li-ion batteries, Inorganic Chemistry, Frontiers (2019). [64] Y.C. Chen, R.J. Young, J.V. Macpherson, N.R. Wilson, Single-walled carbon nano-
[34] H. Usui, Y. Domi, R. Yamagami, K. Fujiwara, H. Nishida, H. Sakaguchi, Sodiation- tube networks decorated with silver nanoparticles: A novel graded SERS substrate, J
desodiation reactions of various binary phosphides as novel anode materials of na- Phys Chem C 111 (2007) 16167–16173.
ion battery, ACS Appl. Energy Mater. 1 (2018) 306–311. [65] K.N. Kudin, B. Ozbas, H.C. Schniepp, R.K. Prud'homme, I.A. Aksay, R. Car, Raman
[35] N.Q. Hai, S.H. Kwon, H. Kim, I.T. Kim, S.G. Lee, J. Hur, High-performance MoS2- spectra of graphite oxide and functionalized graphene sheets, Nano Lett. 8 (2008)
based nanocomposite anode prepared by high-energy mechanical milling: The ef- 36–41.
fect of carbonaceous matrix on MoS2, Electrochim Acta 260 (2018) 129–138. [66] Y.Y. Wang, Z.H. Ni, T. Yu, Z.X. Shen, H.M. Wang, Y.H. Wu, W. Chen, A.T.S. Wee,
[36] Y.-H. Cui, M.-Z. Xue, X.-L. Wang, K. Hu, Z.-W. Fu, InP as new anode material for Raman studies of monolayer graphene: The substrate effect, J Phys Chem C 112
lithium ion batteries, Electrochem. Commun. 11 (2009) 1045–1047. (2008) 10637–10640.
[37] M.-D. Gerngross, E. Quiroga-Gonzalez, J. Carstensen, H. Föll, Single-crystalline [67] J. Thomas, G. Kaganowicz, J. Robinson, X-ray photoelectron spectroscopy analysis
porous indium phosphide as anode material for li-ion batteries, J. Electrochem. Soc. of changes in InP and InGaAs SURFACES EXPOSED TO VARIOUS PLASMA

10
Q. Hai Nguyen, et al. Chemical Engineering Journal 399 (2020) 125826

ENVIRONMENTs, J. Electrochem. Soc. 135 (1988) 1201–1206. with half-cell and full-cell applications as promising anode material for li-ion bat-
[68] H. Zhou, M. Wang, H. Ding, G. Du, Preparation and characterization of barite/TiO2 teries, Appl. Surf. Sci. 144718 (2019).
composite particles, Adv. Mater. Sci. Eng. 2015 (2015). [77] W.P. Kang, Y.B. Tang, W.Y. Li, X. Yang, H.T. Xue, Q.D. Yang, C.S. Lee, High in-
[69] S. Liu, H. Jia, L. Han, J. Wang, P. Gao, D. Xu, J. Yang, S. Che, Nanosheet-constructed terfacial storage capability of porous NiMn2O4/C hierarchical tremella-like na-
porous TiO2–B for advanced lithium ion batteries, Adv. Mater. 24 (2012) nostructures as the lithium ion battery anode, Nanoscale 7 (2015) 225–231.
3201–3204. [78] Z.P. Zeng, H.L. Zhao, P.P. Lv, Z.J. Zhang, J. Wang, Q. Xia, Electrochemical prop-
[70] P. Kubiak, M. Pfanzelt, J. Geserick, U. Hörmann, N. Hüsing, U. Kaiser, erties of iron oxides/carbon nanotubes as anode material for lithium ion batteries,
M. Wohlfahrt-Mehrens, Electrochemical evaluation of rutile TiO2 nanoparticles as J. Power Sources 274 (2015) 1091–1099.
negative electrode for Li-ion batteries, J. Power Sources 194 (2009) 1099–1104. [79] X.J. Zhu, Z.P. Guo, P. Zhang, G.D. Du, R. Zeng, Z.X. Chen, S. Li, H.K. Liu, Highly
[71] Q.H. Nguyen, H. Kim, I.T. Kim, W. Choi, J. Hur, Few-layer NbSe2@ graphene porous reticular tin-cobalt oxide composite thin film anodes for lithium ion bat-
heterostructures as anodes in lithium-ion half-and full-cell batteries, Chem. Eng. J. teries, J. Mater. Chem. 19 (2009) 8360–8365.
122981 (2019). [80] S.X. Liu, H.B. Yin, H.B. Wang, J.C. He, Electrochemical performance of WO2
[72] W.-J. Li, Y.-N. Zhou, Z.-W. Fu, Fabrication and lithium electrochemistry of InSe thin modified LiFePO4/C cathode material for lithium-ion batteries, J. Alloy. Compd.
film, Appl. Surf. Sci. 257 (2011) 2881–2885. 561 (2013) 129–134.
[73] Z.M. Detweiler, S.M. Wulfsberg, M.G. Frith, A.B. Bocarsly, S.L. Bernasek, The oxi- [81] B. Zhang, G. Chen, P. Xu, Z. Lv, Effect of ultrasonic irradiation on the structure and
dation and surface speciation of indium and indium oxides exposed to atmospheric electrochemical properties of cathode material LiNi0.5Mn0.5O2 for lithium bat-
oxidants, Surf. Sci. 648 (2016) 188–195. teries, Solid State Ionics 178 (2007) 1230–1234.
[74] K.N. Wood, K.X. Steirer, S.E. Hafner, C. Ban, S. Santhanagopalan, S.-H. Lee, [82] K. Sun, J.N. Dong, Z.X. Wang, Z.Y. Wang, G.H. Fan, Q. Hou, L.Q. An, M.Y. Dong,
G. Teeter, Operando X-ray photoelectron spectroscopy of solid electrolyte inter- R.H. Fan, Z.H. Guo, Tunable Negative Permittivity in Flexible Graphene/PDMS
phase formation and evolution in Li 2 SP 2 S 5 solid-state electrolytes, Nat. Metacomposites, J Phys Chem C 123 (2019) 23635–23642.
Commun. 9 (2018) 2490. [83] Z.Y. Wang, K. Sun, P.T. Xie, Y. Liu, Q.L. Gu, R.H. Fan, Permittivity transition from
[75] W.-J. Li, Y.-N. Zhou, Z.-W. Fu, Nanocomposite Cr2O3–InP as a Storage Lithium positive to negative in acrylic polyurethane-aluminum composites, Compos. Sci.
Material, J. Electrochem. Soc. 157 (2010) A957–A961. Technol. 188 (2020).
[76] Q.H. Nguyen, Q.H. Nguyen, J. Hur, High-performance ZnTe-TiO2-C nanocomposite

11

You might also like