You are on page 1of 15

Nano Energy 82 (2021) 105750

Contents lists available at ScienceDirect

Nano Energy
journal homepage: http://www.elsevier.com/locate/nanoen

Full paper

Hierarchical Co and Nb dual-doped MoS2 nanosheets shelled micro-TiO2


hollow spheres as effective multifunctional electrocatalysts for HER, OER,
and ORR
Dinh Chuong Nguyen a, Thi Luu Luyen Doan a, Sampath Prabhakaran a, Duy Thanh Tran a,
Do Hwan Kim b, Joong Hee Lee a, c, *, Nam Hoon Kim a, **
a
Department of Nano Convergence Engineering, Jeonbuk National University, Jeonju, 54896 Jeonbuk, Republic of Korea
b
Division of Science Education, Jeonbuk National University, Jeonju, 54896 Jeonbuk, Republic of Korea
c
Carbon Composite Research Center, Department of Polymer-Nano Science and Technology, Jeonbuk National University, Jeonju, 54896 Jeonbuk, Republic of Korea

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

Keywords: Heteroatom doping engineering has emerged as an intriguing strategy to enhance electrocatalytic efficiency.
Co,Nb dual-doping effects Herein, a multiple transition metal doping approach is developed through the incorporation of both Co and Nb
Hollow core-shell spheres into hierarchical MoS2 ultrathin nanosheets directly grown on micro-TiO2 hollow spheres (Co,Nb-MoS2/TiO2
Multifunctional catalyst
HSs) to boost the hydrogen evolution (HER), oxygen evolution reaction (OER), and oxygen reduction reaction
Water splitting
Oxygen reduction reaction
(ORR). The Co and Nb dual-doping effects modify the electronic structure of the host MoS2 towards maximizing
the HER, OER, and ORR performance. Additionally, the unique hollow spherical structure and heterostructured
synergistic effects between the TiO2 core and MoS2 shell provide effective channels for electron transfer and large
surface area with abundant exposed void spaces for ion diffusion/penetration. Therefore, the Co,Nb-MoS2/TiO2
HSs catalyst demonstrates extraordinary activity and stability with small overpotentials of 58.8 and 260.0 mV at
10 mA cm− 2 for the HER and OER, respectively. When employed as both cathodic and anodic electrode in an
electrolyzer, the catalyst requires an operating voltage of 1.57 V to achieve 10 mA cm− 2. The catalyst also ex­
hibits great potential for the ORR with high onset potential of + 0.96 V and half-wave potential of +0.87 V, as
well as direct four-electron transfer process.

1. Introduction Accordingly, the design of alternative multifunctional electrocatalysts


that meet the requirements of cost-effectiveness, high catalytic effi­
In recent years, many different research efforts have been under­ ciency, and good stability for the HER, OER, and ORR is an important
taken in the energy field, especially for renewable energy conversion task, but so far remains a great challenge.
and storage techniques, such as water splitting, metal-air batteries, and In recent years, layered molybdenum dichalcogenide (MoS2) has
fuel cells, to meet the urgent demand in the near future for green energy. received considerable attention in the catalysis field by reason of its
These technologies are primarily realized via the three major HER, OER, great advantages, such as unique two-dimensional layered graphene-
and ORR reactions [1,2]. Unfortunately, the sluggish kinetics and high like structure, and excellent electrochemical and mechanical proper­
overpotential of these reactions result in the practical efficiency of the ties [7–11]. Previous studies have proven that MoS2 not only has high
related energy systems becoming much inferior, as compared with the electrocatalytic activity for the HER [8,11], but also serves as a prom­
theoretical efficiency [3,4], and thus efficient electrocatalysts are ising low-cost electrocatalyst for both the ORR and OER [12,13]. In
needed. Even though noble metals (e.g. Pt) or transition metal oxides (e. general, natural semiconducting material with low electrical conduc­
g. RuO2 or IrO2)-based catalysts have demonstrated excellent catalytic tivity and poor exposed active edge sites owing to the irregular reag­
activity, their high cost, insufficient stability, and poor electrical con­ gregation and agglomeration of MoS2 induces unsatisfactory
ductivity are critical barriers for large-scale practical application [5,6]. performance to meet the practical requirements [14–16]. Therefore, a

* Corresponding author at: Department of Nano Convergence Engineering, Jeonbuk National University, Jeonju, 54896 Jeonbuk, Republic of Korea.
** Corresponding author.
E-mail addresses: jhl@jbnu.ac.kr (J.H. Lee), nhk@jbnu.ac.kr (N.H. Kim).

https://doi.org/10.1016/j.nanoen.2021.105750
Received 7 November 2020; Received in revised form 18 December 2020; Accepted 4 January 2021
Available online 7 January 2021
2211-2855/© 2021 Elsevier Ltd. All rights reserved.
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

unique approach is needed to seek an effective route to optimize the activity and enhanced the electrochemically active surface area (ECSA).
catalytic activity of MoS2-based catalysts towards replacing noble metal Moreover, the strong electronic interactions between the inner TiO2
catalysts. hollow sphere and outer Co,Nb-MoS2 layer at the interface significantly
Among the tremendous efforts that have been made, transition metal improved electrical conductivity, thereby providing favorable kinetic
doping has been considered an effective pathway to promote the elec­ for lower overpotential values of electrochemical reactions. Therefore,
trochemical properties of MoS2, due to its ability to adjust the electronic the Co,Nb-MoS2/TiO2 HSs has the ability to simultaneously catalyze the
structure of MoS2 and optimize the electron density of Mo and S, HER, OER, and ORR with high electrocatalytic performance. The cata­
resulting in the increase of conductivity and new active sites [17,18]. Shi lyst only required overpotential (η) of 58.8 mV for the HER, and 260.0
et al. reported that the doping of transition metals dramatically mV for the OER, to achieve 10 mA cm− 2, which is superior to most of the
enhanced the HER activity of MoS2 through exposing more active sites efficient HER and OER catalysts reported recently. In addition, the fast
and modifying the 3d electron configuration, leading to a positive shift and favorable kinetics were demonstrated through small Tafel slopes of
of about − 0.13 V vs. RHE in the onset overpotential, and far higher 40.2 and 65.0 mV dec− 1 for HER and OER, respectively. When applied
turnover (15.44 s− 1 at 300 mV) than that of the pure MoS2 reported as two electrodes of the electrochemical water splitting device, the Co,
elsewhere [19]. In another study, Xiong et al. reported significant in­ Nb-MoS2/TiO2 HSs was capable of driving 10 and 50 mA cm− 2 at small
crease in the HER and OER performances of the MoS2 towards efficient operating voltage values of 1.57 and 1.88 V, respectively, and achieved
water splitting through Co-doping engineering, demonstrating that stable gaseous production for 60 h with a retention of 89.2%. The pos­
effective Co-doping resulted in a minimized bandgap of Co-doped MoS2 itive onset potential (Eonset) and half-wave potential (E1/2) of + 0.96 and
(~ 0 eV), which is beneficial to catalyze both the HER and OER reactions + 0.87 V, respectively, with desirable 4-electron transfer pathway, were
[20]. Noticeably, the doping of secondary elements, such as Co, Fe, and also found for the Co,Nb-MoS2/TiO2 HSs towards the ORR. Furthermore,
Nb, into MoS2 not only effectively regulates the electronic structure the catalyst showed good stability under the ORR condition with a
towards HER efficiency, but also provides additional OER active sites to current density retention of 96.65% after 60,000 s along with excellent
boost OER performance [21,22]. In addition, Wang et al. have per­ alcohol tolerance, which even outperformed the commercial Pt/C
formed computational research on the substitutional doping of different catalyst.
transition metals for MoS2, demonstrating that the atoms doping can
strongly interact with the S-vacancy, as well as adjust the electronic 2. Experimental section
structure of MoS2, leading to improvement of the ORR activity of MoS2
[23]. Considering the above-mentioned reasons, it is reasonable to 2.1. Chemical reagents
expect the development of multifunctional catalysts via transition
metal-doped MoS2 materials. Sodium lauryl sulfate (CH3(CH2)11OSO3Na, ≥ 98.5%), potassium
In addition to improving intrinsic activity, the development of het­ persulfate (K2S2O8, ≥ 99%), methanol (CH3OH, ≥ 99.9%), styrene
erostructures based on MoS2 and a conductive/stable backbone is also (C8H8CH=CH2, ≥ 99%), ethanol (CH3CH2OH, ≥ 99.8%), hexadecyl­
emerging as another effective way to improve catalytic performance. amine (CH3(CH2)15NH2, 98%), ammonium hydroxide (NH4OH, 28.0%
Evidence has been reported that the contactable surface area of the MoS2 NH3 in H2O), titanium isopropoxide (Ti[OCH(CH3)2]4, 99.9%), cobalt
for reactants/electrolyte ions could be extended through hetero­ nitrate hexahydrate (Co(NO3)2.6H2O, ≥ 98%), ruthenium oxide (RuO2,
structural form by alleviating the bulky aggregates. Titanium dioxide 99.9%), niobium chloride (NbCl5, 99%), ammonium tetrathiomolybdate
(TiO2), with the excellent merits of cheap price, richness, nontoxicity, ((NH4)2MoS4, 99.97%), potassium hydroxide (KOH, ≥ 85%), nafion (5
environmental friendliness, superior chemical stability/durability, and wt%), platinum on carbon (Pt/C, 20 wt%), and nickel foam were pur­
excellent physicochemical properties [24–26], has been extensively chased from Sigma Co. (USA).
utilized in many different electrochemical applications, especially as a
popular catalyst support [27–29]. Liang et al. has reported that 2.2. Synthesis of Co,Nb-MoS2/TiO2 hollow spheres
three-dimensional (3D) heterostructured catalysts containing TiO2 and
MoS2 nanosheets (NSs), as core and shell parts, respectively, showed Firstly, Co and Nb dual-doped-MoS2 NSs were directly deposited on
dramatic HER performance, in comparison with the pure MoS2 NSs [30]. the PS@TiO2 spherical surface via a hydrothermal reaction, which can
This study highlighted that the very rough TiO2 layer core not only be described as follows. Firstly, 5 mL of 0.05 M (NH4)2MoS4 solution was
provided a large amount of favorable sites for the growth of MoS2 mixed with 20 mL ethanol solution containing 0.06 g of PS@TiO2
without agglomerated form, but during the HER process also acted as an spheres (see Supporting information) under magnetic stirring for 1 h.
electron transfer layer for smooth charge transfer. Among various Next, 10 mL ethanol solution of Co(NO3)2.6H2O (6 mM) and NbCl5 (3
morphologies of TiO2 support materials, hollow TiO2 sphere has a large mM) was added into that solution under stirring for 3 h. The obtained
and curved growth surface to benefit MoS2 loading, and offers an solution was then poured into 50 mL Teflon-lined stainless steel auto­
effective transport path to improve the mass transfer property of the clave. The hydrothermal reaction was performed at 200 ◦ C for 6 h. After
material [31,32]. As compared to solid particles, the porous hollow that, the centrifugation method was used to collect the Co,Nb-MoS2/
structures possess lower density and mass requirements, and these fea­ PS@TiO2 spheres, followed by rinsing with ethanol/distilled water, and
tures are significant for cost reduction [33]. Although there have various finally drying at 60 ◦ C overnight.
efforts to improve the electrocatalytic activities of MoS2, no data has To convert the Co,Nb-MoS2/PS@TiO2 spheres to Co,Nb-MoS2/TiO2
been available on inducing multifunctionality to MoS2 via combining HSs, the Co, Nb-MoS2/PS@TiO2 precursor was annealed at 650 ◦ C with
co-doping engineering and constructing MoS2-based heterostructures heating rate of 1 ◦ C min− 1 in Ar atmosphere (200 sccm) for 3 h (Scheme
towards the HER, OER, and ORR applications. 1).
Herein, we have developed a novel multi-shelled hollow structure Individual TiO2 hollow spheres (TiO2 HSs), Co-doped MoS2/TiO2
containing Co and Nb dual-doped MoS2 NSs directly grown on TiO2 hollow spheres (Co-MoS2/TiO2 HSs), Nb-doped MoS2/TiO2 hollow
hollow microspheres (denoted as Co,Nb-MoS2/TiO2 HSs) as highly spheres (Nb-MoS2/TiO2 HSs), MoS2/TiO2 hollow spheres (MoS2/TiO2
active and cost-effective multifunctional electrocatalysts. The spherical HSs), and Co,Nb-MoS2 NSs materials were also fabricated via the
hollow structure with ultrathin NSs attached on the surface can enlarge method described above for comparison.
the contact area for electrolyte/reactant, promote the electrolyte/reac­
tant–electrode diffusion/penetration process, and enhance the charge 2.3. Material characterization
transfer ability. In addition, the Co, Nb-dual doping adjusted the unique
electronic structure of the MoS2 towards improving intrinsic catalytic The morphological and structural characteristics of materials were

2
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

Scheme 1. Schematic showing the synthesis procedure of Co,Nb-MoS2/TiO2 HSs.

investigated using field-emission scanning electron microscopy (FE- equations [34]:


SEM) and energy-dispersive X-ray analysis (EDAX) through a Supra 40
EiR− = Eraw − EiR (3)
VP instrument (Zeiss Co., Germany). Transmission electron microscopy corrected

(TEM) and high-resolution TEM (HR-TEM) were carried out by H-7650


EiR = J × Rs (4)
instrument (Hitachi Ltd., Japan) in the Jeonju center of KBSI, to examine
the structure of materials in detail. X-ray diffraction (XRD) analysis was where Rs is series resistance (Ω).
used to identify the crystalline structure and chemical components using Cycle voltammetry (CV) was performed at various scan rates of 5, 10,
a D/Max 2500V/PC (Rigaku Co., Japan) in the 2θ range (10− 90)◦ , with 15, 20, 25, and 30 mV s− 1, in the potential region 0.92–1.01 V vs. RHE,
Cu Kα radiation (λ = 0.154 nm) at the Center for University-Wide to calculate the double layer capacitance (Cdl). Charge–transfer resis­
Research Facilities (CURF) in Jeonbuk National University, Korea The tance (Rct) of materials was identified by electrochemical impedance
elemental composition and valence state of elements on the surface of spectroscopy (EIS) at an open circuit potential in the frequency range
materials were observed via X-ray photoelectron spectroscopy (XPS) 105 to 10− 2 Hz with an amplitude of 5 mV. The long-term stability of
analysis on a Theta Probe instrument (Thermo Fisher Scientific Inc., materials was estimated through chronoamperometric measurement at
USA). The specific surface area of materials was found via the Bru­ current responses of 10 and 50 mA cm− 2 for both the HER and OER for
nauer–Emmett–Teller (BET) method through an ASAP 2020 Plus system 30 h. Multistep chronoamperometry was also carried out at potential
(Micromeritics Instrument Corp., USA). range of 0.058-0.198 V for the HER, and 1.49–1.59 V for the OER, with
increment of 0.01 V per 500 s.
2.4. Electrochemical characterization To evaluate the catalytic performance towards overall water split­
ting, a hand-made electrolyzer was assembled using the as-synthesized
Electrochemical properties of the materials were assessed on an catalyst-loaded NF as cathode and anode (mass loading of
electrochemical workstation (CH660D, CH Instruments Inc., USA) in­ 5 mg cm− 2). The LSV and chronoamperometric measurements were run
tegrated into a standard three-electrode configuration, in which in N2-saturated 1.0 M KOH solution. The amount of generated gaseous
graphite rod and Ag/AgCl were used as counter electrode and reference during the experiment process was collected via a drainage collection
electrode, respectively. To fabricate the working electrode, 5 mg of process. The Faradaic efficiency (FE) for both the HER and OER was
catalyst was dissolved in 500 µL of isopropanol solution. After adding calculated according to the following equation [30]:
20 µL of Nafion (5 wt%), the mixed solution was sonicated for 30 min to
n × F × ngas
obtain a homogeneous catalyst ink, followed by pipetting on a piece of FE = (5)
J×t
nickel foam (NF, 1 cm × 1 cm) for the HER, OER, and overall water
splitting test, or on a polished RDE surface (diameter ~ 3 mm) for the where, n is the transferred electron number, F is the Faraday constant
ORR test. All potential values were calibrated to reversible hydrogen (96,485 cmol− 1), ngas is the amount of O2 or H2 produced (mmol), and t
electrode (RHE), according to the Nernst equation [33]: is the time for the constant oxidation/reduction current (s).
In the case of ORR investigation, CV measurement was conducted in
ERHE = EAg/AgCl + 0.059∙pH + E0Ag/Agcl (1)
N2- and O2-saturated 0.1 M KOH, while LSV was performed in only O2-
saturated 0.1 KOH medium. A 10 mV s− 1 scan rate was applied for both
where E0Ag/AgCl = 0.198 V (vs. NHE). Additionally, the reaction kinetics
CV and LSV test. In addition, LSV was performed by varying the rotating
were evaluated from Tafel plots with the Tafel equation:
speed as 400, 800, 1200, 1600, and 2000 rpm to clarify the ORR
η = blog|J| + a (2) mechanism. The electron transfer number (n) was determined based on
the Koutechy–Levich equation, as reported elsewhere [35,36]. Chro­
where η is the overpotential (mV), b reflects the Tafel slope (mV dec− 1), noamperometry was also used to probe the stability and methanol
J is the current density (mA cm− 2), and a reflects the exchange current tolerance of the materials in ORR operation. Moreover, rotating
density. ring-disk electrode (RRDE) measurement also conducted for the Co,
For the HER and OER experiments, the electrolyte was nitrogen (N2)- Nb-MoS2/TiO2 HSs and Pt/C. Based on the RRDE result, the n and
saturated 1.0 KOH solution. Linear sweep voltammetry (LSV) was per­ percentage of produced hydrogen peroxide (H2O2) consistent with ORR
formed at 5 mV s− 1 with iR-correction according to the following

3
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

reaction were determined via the following equations [37,38]: layer (~ 20 nm) on the PS microsphere surface. In addition to C and O
elements, Ti element is also observed in the EDAX spectrum of the
Jd
n=4 × (6) PS@TiO2 in Fig. S1f, further confirming the formation of core–shell
Jd + JNr
PS@TiO2 microspheres. Interestingly, the PS@TiO2 surface is much
Jr rougher than that of its PS counterpart, which can provide abundant
%H2 O2 = 2 × N
(7) nano-voids as chambers for high MoS2 NSs loading with full edge/active
Jd + JNr sites exposure [15]. The low-magnification SEM image of the Co,
Nb-MoS2/PS@TiO2 (Fig. S2a) indicates the good preservation of the
where N reflects the collection efficiency of the RRDE, taken as 0.42; Jd
spherical structure after hydrothermal treatment. In addition, it also
and Jr reflect the disk and ring current, respectively.
exhibits the presence of dense NSs on the sphere surface. A magnified
SEM image (Fig. S2b) clearly shows that all PS@TiO2 spheres were fully
3. Results and discussion covered with uniform and thin Co,Nb-MoS2 NSs, forming a hierarchical
core–shell structure. Moreover, the average diameter of the spheres is
3.1. Morphological and structural characterization found to be around ~ 480 nm, which is about 90 nm higher than that of
the pure PS@TiO2, effectively demonstrating the formation of the outer
FE-SEM and energy-dispersive X-ray (EDAX) analyses were used to layer. The EDAX pattern of the Co,Nb-MoS2/PS@TiO2 in Fig. S2c also
verify the morphology and elemental composition of the materials. confirms the presence of Co, Nb, Mo, S, Ti, O, and C elements. After
Fig. S1a and b show SEM images of PS at different magnifications, which annealing, the Co,Nb-MoS2/TiO2 HSs (Fig. 1a and b) is similar to the Co,
reveal the spherical structure, along with the average diameter of ~ Nb-MoS2/PS@TiO2 in terms of the unique spherical structure and
350 nm, together with very smooth surface. A typical EDAX pattern of diameter. Nevertheless, the C element was not found in the EDAX
the PS (Fig. S1c) indicates existence of C and O elements in the material, spectrum (Fig. 1c), proving the complete conversion of PS@TiO2 into
demonstrating that the obtained microspheres are PS. The SEM images TiO2 HSs. Growing Co,Nb-MoS2 NSs on the entire TiO2 skeleton is ex­
of PS@TiO2 (Fig. S1d and e) show that the initial spherical structure is pected to maximum the surface contact of Co,Nb-MoS2 NSs with the
almost intact; however, the diameter of PS@TiO2 has increased to ~ electrolyte/reactant, as well as promote the possible utilization of
390 nm, providing favorable evidence of the coating of thin TiO2 shell

Fig. 1. (a) and (b) SEM images, and (c) EDAX spectrum of the Co,Nb-MoS2/TiO2 HSs; (d) TEM image of the TiO2 HSs; (e) and (f) TEM images of the Co,Nb-MoS2/
TiO2 HSs at low and high magnifications, respectively; (g) HR-TEM image of the Co,Nb-MoS2 shell (Inset: SAED pattern of the Co,Nb-MoS2 shell); and (h–o) EDS
elemental mapping images of the Co,Nb-MoS2/TiO2 HSs.

4
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

catalytic active sites, leading to the increase of electrocatalytic perfor­ temperature during calcination treatment. Likewise, the bright/dark
mance. Furthermore, such unique core–shell structure can offer an contrast characteristics of the Co,Nb-MoS2/TiO2 HSs material (Fig. 1e)
effective pathway for electron transportation from the TiO2 core to the prove its structured hollow spherical form. In addition, TEM images of
surrounding Co,Nb-MoS2 NSs, thus effectively boosting the electrode the Co,Nb-MoS2/TiO2 HSs also indicate the ultrathin secondary NSs
kinetics of electrochemical reactions. Through the EDAX analysis, the uniformly grown on the surface of TiO2 HSs platform, demonstrating the
atomic percentage of particular elements contained in the Co, successful formation of unique core–shell Co,Nb-MoS2/TiO2 HSs. The
Nb-MoS2/TiO2 HSs material was found to be 12.5, 27.1, 7.35, 15.07, average diameter was found to be ~ 480 nm for the Co,Nb-MoS2/TiO2
2.91, and 1.42 at% for Ti, O, Mo, S, Nb, and Co, respectively. SEM HSs with an ~ 390 nm core of TiO2 and ~ 45 nm shell of Co,Nb-MoS2
analysis was also conducted for the Co,Nb-MoS2 NSs for comparison layer. A closer observation of the shell layer (Fig. 1f) reveals the very
(Fig. S3), and exhibits the aggregation of numerous NSs into the form of rough and wrinkled characteristics of the Co,Nb-MoS2 NSs, with only
compact/stacked layered structure, suggesting deficiency in the surface single and few-layer NSs (around 1–4 layers) attached to the surface of
site exposure [16]. the TiO2 HSs substrate. The HR-TEM image taken from the shell region
TEM analysis was then conducted for the as-synthesized materials to (Fig. 1g) identifies lattice fringes with interlayer spacing of 0.27 and
examine their morphological characteristics in detail. The TEM images 0.64 nm, which are in good agreement with the (100), and (002) planes
of the PS and PS@TiO2 microspheres (Fig. S4) show their solid feature of MoS2, respectively [39]. In addition, the selective area electron
with average diameter of ~ 350 and ~ 390 nm, respectively; thus, the diffraction (SAED) pattern of the layer shell (Inset Fig. 1g) displays
thickness of outer TiO2 layer is found to be ~ 20 nm, which agrees well several bright circular rings that are assigned to the (100), (103), and
with SEM observation. Fig. 1d shows the TEM image of the TiO2 HSs, (110) planes of MoS2, suggesting that the outer part contains MoS2
which distinctly exhibits its hollow structure with a uniform size and phase [39]. It is notable that no signal related to the compounds of Co or
thickness behavior. The formation of such hollow structure can be Nb is observed from the HR-TEM image and SAED pattern, implying that
attributed to the decomposition of the inner PS sphere under high both Co and Nb may act only as effective dopants to boost the

Fig. 2. (a) N2 adsorption–desorption curves of the Co,Nb-MoS2/TiO2 HSs, Co,Nb-MoS2 NSs, and TiO2 HSs, and the corresponding pore-size distribution curves
(inset); (b) XRD patterns of the materials; (c) XPS survey spectra of the materials [(1) MoS2/TiO2 HSs, (2) Co-MoS2/TiO2 HSs, (3) Nb-MoS2/TiO2 HSs, and (4) Co,Nb-
MoS2/TiO2 HSs]; and (d–i) High-resolution XPS spectra of the Ti2p, O1s, Mo3d, S2p, Nb3d, and Co2p for the MoS2/TiO2 HSs, Co-MoS2/TiO2 HSs, Nb-MoS2/TiO2 HSs,
and Co,Nb-MoS2/TiO2 HSs, respectively.

5
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

electrochemical property of MoS2. Fig. 1h–o show the results of the MoS2/TiO2 HSs to further investigate its structural characteristics
Energy dispersive X-ray spectroscopy (EDS) mapping that was also (Fig. S7). As seen, two Raman peaks at 382.1 and 404.8 cm− 1 assigned
carried out for an individual Co,Nb-MoS2/TiO2 HSs. The Co, Nb, Mo, properly with the E12g in-plane and A1g out-of-plane vibrational modes
and S elements are mainly found in the shell area, in which the Mo and S of the MoS2 are clearly observed from Raman spectrum [45], as
possess much stronger signals than those of the Co and Nb, which is compelling evidence for the formation of MoS2 phase in the Co,
attributed to the lower concentration of both Co and Nb associated with Nb-MoS2/TiO2 HSs structure.
their doping role. High densities of Ti and O2 elements are observed at XPS analysis was also conducted for the materials to further clarify
the center of the hollow sphere, confirming the core–shell architecture their structural details and chemical composition. The survey XPS
with Ti and O in the core, and Co, Nb, Mo, and S elements in the shell spectra of the Co,Nb-MoS2/TiO2 HSs, Co-MoS2/TiO2 HSs, Nb-MoS2/
area. On the other hand, EDS line scan profiles (Fig. S5a–h) depict the TiO2 HSs, and MoS2/TiO2 HSs materials (Fig. 2c) exhibit the predomi­
strong signals of the Co, Nb, Mo, and S elements in the outer area, nant existence of peaks at ca. 161.8, 228.9, 458.8, and 530.4 eV, which
whereas the signals of Ti and O elements are mainly observed in the index well to the S2p Mo3d, Ti2p, and O1s orbitals, respectively,
inner area, reconfirming the unique core–shell configuration of the Co, reconfirming the formation of pure phase MoS2 and TiO2 in our work.
Nb-MoS2/TiO2 HSs material. Additionally, the appearance of minor peaks at ca. 207.0 and 782.6 eV,
To reflect the potential of the Co,Nb-MoS2/TiO2 HSs for electro­ which are assigned to the Nb3d and Co2p orbitals, in the Co,Nb-MoS2/
catalytic applications, we evaluated its specific surface area using Bru­ TiO2 HSs, Co-MoS2/TiO2 HSs, and Nb-MoS2/TiO2 HSs proves the suc­
nauer–Emmett–Teller (BET) gas sorptometry analysis. Fig. 2a shows that cessful incorporation of Co and Nb into the MoS2 structure. Fig. 2d–i
the N2 adsorption–desorption isotherm of the Co,Nb-MoS2/TiO2 HSs provide the high-resolution XPS spectra of the Ti2p, O1s, Mo3d, S2p,
possesses the full features of the type IV isotherm with a pronounced Co2p, and Nb3d core-level, respectively. The Ti2p region (Fig. 2d) can
hysteresis loop, suggesting the presence of mesoporous structure. The be deconvoluted into two peaks of Ti2p3/2 and Ti2p1/2 at ca. 458.8 and
Co,Nb-MoS2/TiO2 HSs shows a large specific surface area of 198.8 m2 464.6 eV, respectively, which is a characteristic of the Ti 4+ oxidation
g− 1, which is beneficial for the catalysis process, along with pore size state in TiO2 [30]. In addition, the appearance of the O1s peaks (Fig. 2e)
distribution located in the range (1.9–2.5) nm (Inset of Fig. 2a), indi­ at ca. 530.3 and 531.7 eV corresponding to the metal–oxygen bonding
cating its excellent mesoporous nature. Such mesoporous structure has and the chemisorbed oxygen, such as the adsorbed OH- group, respec­
the advantage of high void channel exposure to facilitate electrolyte ion tively, confirms the formation of TiO2 phase by using the present syn­
diffusion/penetration, as well as gaseous product detachment, thereby thesis approach [46]. As compared to the Ti2p in pure TiO2 (Fig. S8a),
improving the reaction rate during the catalysis process [40,41]. the binding energies of the Ti2p in the Co,Nb-MoS2/TiO2 HSs show a
Remarkably, the specific surface area of both the TiO2 HSs (162.0 m2 slightly positive shift of ca. 0.28 eV, representing the strong electronic
g− 1) and Co,Nb-MoS2 NSs (71.2 m2 g− 1) counterparts are significantly interactions at the interface of TiO2-core and Co,Nb-MoS2-shell, which
lower than the Co,Nb-MoS2/TiO2 HSs (Fig. 2a), implying the important are advantageous to electron transfer from TiO2 to Co,Nb-MoS2 for the
roles of the core–shell structure in enhancing the surface area. This is catalysis process [47,48]. For the Mo3d spectrum (Fig. 2f), three peaks
because of the agglomeration phenomenon of the pure Co,Nb-MoS2 NSs located at ca. 225.6, 228.2, and 231.4 eV, assigned to Mo–S, Mo4+
was restricted via growth on the TiO2 HSs backbone, as seen in the SEM 3d5/2, and Mo4+ 3d3/2, respectively, demonstrate the presence of
and TEM imagery. Furthermore, ultrathin, very rough, and wrinkled MoS2 phase in the as-synthesized material [49]. Meanwhile, the
features of the Co,Nb-MoS2 shell layer significantly improve the surface remaining fitted peaks at ca. 232.5 and 235.6 eV match the Mo6+ 3d5/2
area of the Co,Nb-MoS2/TiO2 HSs material. XRD measurements were and Mo6+ 3d3/2 from the MoO3 phase, which might originate from an
conducted for the PS, PS@TiO2, TiO2 HSs, MoS2/TiO2 HSs, Co-MoS2/­ oxidized surface [50]. All the Mo–S, Mo4+ 3d5/2, and Mo4+ 3d3/2 are
TiO2 HSs, Nb-MoS2/TiO2 HSs, and Co,Nb-MoS2/TiO2 HSs materials. negatively shifted by ca. 0.4 eV, in contrast to those of the single Co,
Fig. S6 shows the XRD patterns of the PS, PS@TiO2 microsphere, and Nb-MoS2 NSs (Fig. S8c), further evidencing that the formation of the
TiO2 HSs. The broad diffraction peak at 2θ of 20◦ of both the PS and Co,Nb–MoS2/TiO2 HSs core–shell structure enhances the electrical
PS@TiO2 sample can be assigned to the amorphous structures of the PS conductivities. In the case of the high-resolution XPS spectrum of the
spheres. In addition, a series of additional peaks at 2θ of 25.2, 37.8, 48.0, S2p (Fig. 2g), binding energy values at ca. 161.8 and 163.4 eV are
53.9, 55.0, 62.6, 68.7, 70.3, and 75.1◦ , which agree with the (101), consistent with the S2- 2p3/2 and S2- 2p1/2 of the Mo–S bonds,
(004), (200), (105), (211), (204), (116), (220), and (215) crystalline respectively [51]. The binding energies of the Mo3d and S2p peaks for
planes of TiO2 (PDF#021-1272) [42], are also found for the PS@TiO2, the Co-MoS2/TiO2 HSs clearly show negative shifts of about 0.4 and
proving the successful formation of the outer TiO2 layer. After anneal­ 0.35 eV, respectively, as compared with the undoped MoS2/TiO2 HSs.
ing, only the diffraction XRD peaks of TiO2 phase are observed in the Likewise, the binding energies of the Mo3d and S2p peaks for the
TiO2 HSs, implying that under high thermal condition, the inner PS Nb-MoS2/TiO2 HSs are 0.6 and 0.42 eV lower than that of the undoped
spheres are fully decomposed, agreeing well with the TEM observations. MoS2/TiO2 HSs, respectively. This result proves the increase in elec­
For the case of the MoS2/TiO2 HSs, Co-MoS2/TiO2 HSs, Nb-MoS2/TiO2 tronic density of the host MoS2 after being doping by the Co or Nb [52].
HSs, and Co,Nb-MoS2/TiO2 HSs materials (Fig. 2b), besides the char­ More importantly, the Mo3d and S2p peaks of the Co,Nb-MoS2/TiO2
acteristic peaks of the TiO2 phase, the other peaks located at 2θ of 14.4, HSs have more negative shifts than those of the Co-MoS2/TiO2 HSs
32.6, 33.5, 39.5, 49.7, 58.3, 60.1, and 72.7◦ for all materials can be and Nb-MoS2/TiO2 HSs counterparts, thereby proving that the dual Co
attributed to the (002), (100), (101), (103), (105), (110), (008), and and Nb doping can further enrich the electronic density of MoS2. In the
(203) crystalline planes of the hexagonal MoS2 phase (PDF#037-1492) Nb3d region of the XPS spectra for Nb-MoS2/TiO2 HSs and Co,
[43], meaning that the majority of phase composites are TiO2 and MoS2. Nb-MoS2/TiO2 HSs (Fig. 2h), the peaks located at ca. 203.6 and
No obvious peaks related to the possible crystal phases of Co- or 204.9 eV should be assigned to the Nb1+ 3d5/2 and Nb1+ 3d3/2,
Nb-based compounds are observed from the Co-MoS2/TiO2 HSs, respectively [53], while the other fitted peaks at 208.5 and 211.2 eV are
Nb-MoS2/TiO2 HSs, and Co,Nb-MoS2/TiO2 HSs, which correspond with well consistent with high oxidation states of Nb5+ 3d5/2 and Nb5+
the SAED and HR-TEM results, reconfirming the doping roles of Co and 3d3/2, respectively [54], indicating the coexistence of Nb1+ and Nb5+
Nb into the MoS2 structure [18,22]. Furthermore, the weak (002) oxidation states. Fig. 2i provides the high-resolution XPS spectrum of
diffraction peak indicates that the as-prepared MoS2 using the current Co2p, which displays two major sets of binding energies at ca. 778.7 and
synthesis strategy has an ultrathin two-dimensional (2D) layered 794.4 eV, corresponding to the Co3+ 2p3/2 and Co3+ 2p1/2, respec­
structure [44], similar to graphene structure, as demonstrated in the tively, and ca. 781.6 and 797.4) eV, which match well with the Co2+
SEM and TEM observations. 2p3/2 and Co2+ 2p1/2, respectively [54]. Therefore, these results
We conducted Raman spectroscopy measurement for the Co,Nb- demonstrate that both Co and Nb are successfully doped into MoS2

6
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

structure, resulting in an increase of electronic densities in MoS2, thus This can be probably attributed to the low density of the active edge/site
inducing the decrease of empty d-orbitals density towards weakening exposure related to the agglomerated structure of the Co,Nb-MoS2 NSs,
the adsorption ability of the OH group and H atom [55–57]. Such as observed in the SEM result. In contrast, the rationally designed hi­
weakened OH adsorption and H binding can expedite OH/H desorption, erarchical Co,Nb-MoS2/TiO2 HSs core–shell structure led to significantly
as well as promote the release process of products from the catalysis enhanced catalytic activities. Excellent HER behavior was found for the
surface, and thus intrinsically improve the catalytic activities towards Co,Nb-MoS2/TiO2 HSs with very small η10 and η50 of 58.8 and 111.6 mV,
the HER, OER, and ORR [58,59]. respectively. On the other hand, the Co,Nb-MoS2/TiO2 HSs shows
impressively enhanced catalytic activities, when compared with the
3.2. Electrocatalytic performance MoS2/TiO2 HSs (η10 = 222.0 and η50 = 322.2 mV), indicating that the
Co and Nb doping is favorable for the HER. In particular, the signifi­
The electrocatalytic activities of the Co,Nb-MoS2/TiO2 HSs were cantly positive shifts of η for the Co, Nb dual-doping compared with
initially investigated for the HER in alkaline condition. Fig. 3a and b, those of the single Co-MoS2/TiO2 HSs (η10 = 123.2 and η50
respectively, show the iR-corrected polarization curves and comparison = 196.5 mV), and Nb-MoS2/TiO2 HSs (η10 = 200.0 and η50
of η at 10 and 50 mA cm− 2 (denoted as η10 and η50, respectively) for the = 300.2 mV) demonstrate the further enhanced catalytic ability of the
Co,Nb-MoS2/TiO2 HSs HSs, Co-MoS2/TiO2 HSs, Nb-MoS2/TiO2 HSs, MoS2/TiO2 HSs via both Co and Nb incorporation. To the best of our
MoS2/TiO2 HSs, Co,Nb-MoS2 NSs, TiO2 HSs, and the commercial Pt/C knowledge, the excellent HER activities of the Co,Nb-MoS2/TiO2 HSs are
catalyst. The LSV result shows that the pure TiO2 HSs exhibits poor HER superior to most of the recently non-noble metal HER catalysts reported
behavior with the high η10 and η50 of 272.0 and 404.1 mV, respectively, elsewhere in term of η10 (Fig. 3c and Table S1). To further understand
suggesting the major role of the Co,Nb-MoS2 NSs in constructing the the HER kinetics, Tafel slopes of the materials were calculated according
electrocatalytic behavior of the Co,Nb-MoS2/TiO2 HSs. Although the to the Tafel Eq. (1), along with the linear region of Tafel plots (Fig. 3d).
catalytic activities of the Co,Nb-MoS2 NSs (η10 = 170.0 mV and η50 The Tafel slope of 40.2 mV dec− 1 was found for the Co,Nb-MoS2/TiO2
= 265.5 mV) for the HER are higher than those of the TiO2 HSs, its HSs, suggesting that the material follows a Volmer–Heyrovsky mecha­
catalytic efficiency is still insufficient to provide desirable performance. nism, in which the reaction rate is limited by the Heyrovsky step [60,

Fig. 3. (a) iR-corrected polarization curves obtained at a scan rate of 5 mV s− 1 for the Co,Nb-MoS2/TiO2 HSs, Co-MoS2/TiO2 HSs, Nb-MoS2/TiO2 HSs, MoS2/TiO2
HSs, TiO2 HSs, Co,Nb-MoS2 NSs, and Pt/C in the 1.0 M KOH electrolyte; (b) Comparison of η10 and η50 for the as-synthesized samples; (c) Comparison of η10 for the
Co,Nb-MoS2/TiO2 HSs and other reported HER catalysts; (d) Tafel plots of the as-synthesized samples; (e) Chronoamperometric curves of the Co,Nb-MoS2/TiO2 HSs
at applied potentials of 58.8 and 111.6 mV for 30 h; (f) Polarization curves before and after chronoamperometric test; (g) The SEM and TEM (inset) images of the
post-HER Co,Nb-MoS2/TiO2 HSs material; and (h–o) Corresponding EDS elemental mapping of the post-HER Co,Nb-MoS2/TiO2 HSs.

7
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

61]: superior stability towards HER. There is no obvious change in the po­
larization curve after 30 h stability test at high operating potential,

H2 O + e →H + ∗
OH −
(8)
when compared with the initial one (Fig. 3f), suggesting that during
long-term operation, its catalytic behavior is retained well. Also, η10 and
H∗ + H2 O + e− →H2 + OH − (9)
η50 of the Co,Nb-MoS2/TiO2 HSs (Inset of Fig. 3f) show only slightly
The Tafel slopes for the Co-MoS2/TiO2 HSs, Nb-MoS2/TiO2 HSs, Co, positive shifts of about 6.4 and 20.9 mV, respectively, after the chro­
Nb-MoS2 NSs, and MoS2/TiO2 HSs of 81.2, 134.9, 107.8, and noamperometry test over 30 h, further demonstrating the high dura­
116.5 mV dec− 1, respectively, also confirm the Volmer–Heyrovsky bility of the catalyst for the HER.
mechanism while using these catalysts for the HER reaction, and the The SEM and TEM images of the post-HER Co,Nb-MoS2/TiO2 HSs
electrochemical H2 gas desorption may be a rate-determining step. For (Fig. 3g) confirm that Co,Nb-MoS2/TiO2 HSs still has spherical structure
the case of the TiO2 HSs, the Tafel slope of 135.0 mV dec− 1 suggests the with little re-aggregation of NSs on the TiO2 surface. The SEM image
Volmer step limits the rate of reaction, leading to inferior HER behavior (Fig. 3h) and EDS elemental mapping (Fig. 3i–o) for the post-HER Co,
in this material [62]. Even though the value of Tafel slope for the Co, Nb-MoS2/TiO2 HSs clearly reveal the uniform distribution of inner Ti
Nb-MoS2/TiO2 HSs is slightly higher in comparison with that of the Pt/C and O elements and outer Co, Nb, Mo, and S elements throughout the
(38.9 mV dec− 1), it looks to be relatively smaller than those of recently hollow sphere region, further validating the good structural stability of
reported efficient HER catalysts (Table S1). The lowest Tafel slope value the catalyst for the HER. The high-resolution XPS spectra of the post-
of the Co,Nb-MoS2/TiO2 HSs reveals its more favorable and faster HER HER catalyst (Fig. S9a–f) are almost unchanged after stability test,
kinetics, thereby resulting in better H2 production. In addition to the implying the retained composition of the Co,Nb-MoS2/TiO2 HSs during
excellent catalytic HER activities, the Co,Nb-MoS2/TiO2 HSs also ex­ the HER. To evaluate the HER stability of the Co,Nb-MoS2/TiO2 HSs at
hibits good durability and stability. Chronoamperometric curves for the different operating potentials, we performed multi-step chro­
Co,Nb-MoS2/TiO2 HSs (Fig. 3e) show a current density retention of ⁓ noamperometry with operating potential ranging (58.8–198.4) mV.
97.3% at a small applied η of 58.8 mV, and ⁓ 85.2% at a high applied η Fig. S10 shows that the current density driven by the Co,Nb-MoS2/TiO2
of 111.6 mV, during 30 h of continuous operation, demonstrating its HSs remains the same value during 500 s at each applied potential

Fig. 4. (a) iR-corrected polarization curves obtained at a scan rate of 5 mV s− 1 for Co,Nb-MoS2/TiO2 HSs, Co-MoS2/TiO2 HSs, Nb-MoS2/TiO2 HSs, MoS2/TiO2 HSs,
TiO2 HSs, Co,Nb-MoS2 NSs, and RuO2 in 1.0 M KOH electrolyte; (b) Comparison of η10 and η50 for the as-synthesized samples; (c) Comparison of η10 for Co,Nb-MoS2/
TiO2 HSs and other recently reported efficient OER catalysts; (d) Tafel plots of the as-synthesized samples; (e) Chronoamperometric curves of Co,Nb-MoS2/TiO2 HSs
at applied potentials of 1.49 V and 1.59 V for 30 h each; (f) Polarization curves before and after the chronoamperometric test; (g) SEM and TEM (inset) images of the
post-OER Co,Nb-MoS2/TiO2 HSs; and (h–o) Corresponding EDS elemental mapping of the post-OER Co,Nb-MoS2/TiO2 HSs.

8
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

throughout the test, meaning that the Co,Nb-MoS2/TiO2 HSs is highly phases. For the case of the Mo3d, the Mo4+ 3d5/2 and Mo4+ 3d3/2 peaks
stable under continuous change of applied potential during the HER. show decrease in their intensities after the OER test, whereas the peak
This test reconfirms the good durability and stability of the Co,Nb-MoS2/ intensity of the Mo6+ increases, suggesting that MoS2 phase is partially
TiO2 HSs towards the HER, which is extremely important for industrial transformed into MoO3, due to the oxidation process that occurs during
application. the OER [63]. Moreover, after the OER stability test, the core-level
The OER performance of the Co,Nb-MoS2/TiO2 HSs was assessed by metal–oxygen peak of the O1s also increases, along with the reduction
LSV measurement at 5 mV s− 1 in 1.0 KOH medium. For comparison, the of the S2p peak, implying the MoS2 oxidation under the anodic condi­
OER activity of Co-MoS2/TiO2 HSs, Nb-MoS2/TiO2 HSs, MoS2/TiO2 HSs, tion in alkaline media. Furthermore, the robust stability of the Co,
Co,Nb-MoS2 NSs, TiO2 HSs, and RuO2 catalyst were also evaluated Nb-MoS2/TiO2 HSs for the OER was verified through the multi-step
under the same conditions. Fig. 4a and b show that the Co,Nb-MoS2/ chronoamperometric measurement at operating potential range be­
TiO2 HSs reveals the best OER activity with small η10 and η50 of 260.1 tween 1.49 V and 1.65 V. There is no significant change in the current
and 361.0 mV, respectively, which are even lower than those of the density of the Co,Nb-MoS2/TiO2 HSs at 1.49 V in the first 500 s, and the
state-of-art RuO2 catalyst (η10 = 293.2 mV and η50 = 394.5 mV). Both same chronoamperometric responses are observed in the remaining
Co,Nb-MoS2 NSs (η10 = 319.5 mV and η50 = 447.0 mV) and TiO2 HSs steps of the test (Fig. S12), indicating its OER activities are retained well
(η10 = 501.6 mV and η50 = 616.5 mV) possess poorer OER activities under the continuous change of the applied potentials. Thus, the ach­
than the Co,Nb-MoS2/TiO2 HSs. Moreover, the OER catalytic activities ieved result further demonstrates the great potential of the material to
of the undoped MoS2/TiO2 HSs (η10 = 447.5 mV and η50 = 556.1 mV), be employed as efficient and stable OER catalyst for industrial
the Co-MoS2/TiO2 HSs (η10 = 317.5 mV and η50 = 375.7 mV), and the application.
Nb-MoS2/TiO2 HSs (η10 = 407.9 mV and η50 = 510.4 mV) are defi­ Fig. 5a shows the lab-made water splitting device utilizing the Co,
nitely inferior in comparison with the Co,Nb-MoS2/TiO2 HSs. Thus, the Nb-MoS2/TiO2 HSs-loaded NF as both cathodic and anodic electrodes
LSV results demonstrate that the chemical coupling effect of TiO2 HSs that the excellent activity and superior stability of the Co,Nb–MoS2/
core and Co,Nb-MoS2 NSs shell, together with the Co and Nb co-doping TiO2 HSs for HER and OER inspired us to construct. A reference elec­
effects, contribute to the remarkable OER activities of the Co,Nb-MoS2/ trolyzer device assembled by Pt/C-loaded NF cathode and RuO2-loaded
TiO2 HSs. More interesting, the OER activities of the Co,Nb-MoS2/TiO2 NF anode electrodes was also prepared for comparison. The polarization
HSs almost outperform those of the recently reported OER catalysts in curves in Fig. 5b show that the catalytic performance of the bifunctional
terms of η10 and η50, as seen in Fig. 4c. The OER kinetics of the catalysts Co,Nb-MoS2/TiO2 HSs catalyst for overall water splitting is comparable
are analyzed through Tafel plot measurement, in which the Tafel slope to that of the Pt/C//RuO2 electrolyzer. The Co,Nb-MoS2/TiO2 HSs-based
of the Co,Nb-MoS2/TiO2 HSs is found to be 65.0 mV dec− 1, lower than device requires cell voltages of 1.59 V and 1.88 V to reach current
that of the Co-MoS2/TiO2 HSs, Nb-MoS2/TiO2 HSs, MoS2/TiO2 HSs, Co, density of (10 and 50 mA cm− 2, which is only 0.06 V and 0.09 V higher
Nb-MoS2 NSs, and TiO2 HSs of 125.9, 165.8, 175.5, 147.6, and than those of the Pt/C//RuO2-based device, respectively (Fig. 5c)),
187.9 mV dec− 1, respectively. The smallest Tafel slope of the Co,Nb- clearly demonstrating the good performance of the Co,Nb-MoS2/TiO2
MoS2/TiO2 HSs implies its fastest OER rate compared with its counter­ HSs-based electrolyzer. Such exciting performance is superior to that of
parts, consistent with the higher OER activities of this material (Fig. 4d). many efficient bifunctional electrocatalysts for water splitting that have
This is further evidenced when the Tafel slope of the Co,Nb-MoS2/TiO2 previously been developed (Fig. 5d and Table S3). In addition, the Co,
HSs is compared to that of other OER catalysts recently reported else­ Nb-MoS2/TiO2 HSs-based device exhibits a rapid reaction kinetic, and
where (Table S2), implying its potential as a superior OER catalyst in only requires an operating voltage of 2.16 V to achieve 100 mA cm− 2,
practical OER application. The long-term durability and stability have which is even smaller in contrast to that of the Pt/C//RuO2 cell of
been recognized as necessary parameters for the OER catalyst in 2.25 V. The long-term stability of the Co,Nb-MoS2/TiO2 HSs for overall
commercialization. To evaluate the stability of the Co,Nb-MoS2/TiO2 water splitting was examined via chronoamperometry at high operating
HSs for OER, chronoamperometric measurement was adopted at low voltage of 1.59 V. After 60 h of continuous electrolysis, the current
and high operating voltages of 1.49 and 1.59 V, respectively. The density driven by the Co,Nb-MoS2/TiO2 HSs is almost stable with a
chronoamperometric curves (Fig. 4e) exhibit stable current density after current retention of ⁓89.2%, better than that of the Pt/C//RuO2-based
30 h of continuous catalysis with 95.7% and 87.6% initial current electrolyzer of ⁓65.8% at 30 h (Fig. 5e), evidencing the good stability of
density retention at low and high applied potentials, respectively, the Co,Nb-MoS2/TiO2 HSs under water splitting condition. Water
evidencing the robust stability of the catalyst towards the OER. displacement method was used to collect gaseous species produced from
Furthermore, a negligible change for the polarization curves of the Co, the Co,Nb-MoS2/TiO2 HSs-based electrolyzer, and the results show that
Nb-MoS2/TiO2 HSs after 30 h stability test at high operating potential in the molar ratio of H2 to O2 agrees well with 2–1. More interestingly, both
comparison with the initial one (Fig. 4f) reconfirms the excellent sta­ experimental amounts of H2 and O2 gas fit well with the theoretical
bility of the Co,Nb-MoS2/TiO2 HSs for OER. After operating under amount during full water splitting (Fig. 5f), thereby generating high
chronoamperometric condition, positive shifts of about 4.62 and faradaic efficiencies of 99.6% and 99.8% for the HER and OER,
11.8 mV are found for η10 and η50 of the Co,Nb-MoS2/TiO2 HSs (Inset of respectively.
Fig. 4f), respectively, implying the good retention of OER activity during Fig. 5g provides the EIS spectra of the EIS measurement that was
long-term continuous OER. The structural stability of the Co,Nb-MoS2/ conducted for different materials to clarify the origin of the excellent
TiO2 HSs for OER is demonstrated by SEM, TEM, and XPS analyses. electrocatalytic property of the Co,Nb–MoS2/TiO2 HSs. The Nyquist
Fig. 4g distinctly shows that the post-OER Co,Nb-MoS2/TiO2 HSs display plots of the EIS spectra were fitted with an equivalent circuit model
the maintenance of the initial hollow sphere structure with well-ordered (Inset in Fig. 5g) to calculate Rct. The Rct of the Co,Nb-MoS2/TiO2 HSs is
ultrathin Co,Nb-MoS2 shell after OER catalysis. SEM image and EDS found to be only 0.55 Ω, which is significantly lower than that of the Co,
elemental mapping of the post-OER Co,Nb-MoS2/TiO2 HSs (Fig. 4h–o) Nb-MoS2 NSs (0.73 Ω), and TiO2 HSs (0.95 Ω), suggesting that the
indicate the uniform distribution of the Co, Nb, Mo, S, Ti, and O ele­ core–shell structure with strong electronic interactions between TiO2
ments over the selected region, in which Co, Nb, Mo, and S are still HSs-core and Co,Nb-MoS2-shell creates a lower resistance connection to
located in the shell part, while the core part contains Ti and O elements, accelerate fast ion/electron transfer ability. In comparison with the
further verifying the intact core–shell architecture. Fig. S11 provides the pristine MoS2/TiO2 HSs (0.85 Ω), Co-MoS2/TiO2 HSs (0.60 Ω), and Nb-
high-resolution XPS spectra of the Co2p, Nb3d, Mo3d, S2p, Ti2p, and MoS2/TiO2 HSs (0.77 Ω), the smaller Rct of the Co,Nb-MoS2/TiO2 HSs
O1s for the post-OER Co,Nb-MoS2/TiO2 HSs. The core-level peaks of the also indicates the important roles of both Co and Nb incorporation for
Ti2p, Co2p, and Nb3d after the OER test are almost similar to those of promoting the charge transfer process at the catalyst–electrolyte inter­
the as-synthesized one, demonstrating the retained Co, Nb, and TiO2 face, consistent with its enhanced catalytic performance.

9
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

Fig. 5. (a) The optical image of a lab-made electrolyzer with Co,Nb-MoS2/TiO2 HSs as cathode and anode electrodes; (b) Polarization curves of the Co,Nb-MoS2/TiO2
HSs(+/− ) and Pt/C (+) // RuO2 (− ) cell; (c) Comparison of cell voltage at 10 and 50 mA cm− 2 for the Co,Nb-MoS2/TiO2 HSs (+/− ) and Pt/C (+) // RuO2 (− ) cell;
(d) Comparison of cell voltage at 10 mA cm− 2 for the Co,Nb-MoS2/TiO2 HSs (+/− ) and other efficient bifunctional electrocatalysts reported recently; (e) Chro­
noamperometric curves of the Co,Nb-MoS2/TiO2 HSs (+/− ) and Pt/C(+)//RuO2(− ) electrolyzers at voltages of 1.59 and 1.53 V, for 60 h each, respectively; (f) The
theoretically calculated and experimentally measured O2 and H2 with time at 100 mA cm− 2; (g) EIS results of the as-synthesized samples with an equivalent electrical
circuit (inset); (h) Comparison of the CV curves at 10 mV s− 1 of the samples; (i) Cdl values of the samples (1) Co,Nb-MoS2/TiO2 HSs, (2) Co-MoS2/TiO2 HSs, (3) Nb-
MoS2/TiO2 HSs, (4) MoS2/TiO2 HSs, (5) Co,Nb-MoS2 NSs, and (6) TiO2 HSs.

In addition, the ECSA was determined to further explain the high catalytic activities of the Co,Nb-MoS2/TiO2 HSs towards the ORR.
enhanced electrocatalytic performance of the Co,Nb-MoS2/TiO2 HSs. In The LSV measurements was then conducted for the materials. Fig. 6b
this work, the ECSA value was deduced from Cdl calculated by the CV and c show that the Co,Nb-MoS2/TiO2 HSs demonstrates exceptional
measurement (Fig. S13). Fig. 5h compares the CV cycle for the materials ORR activities with an Eonset and E1/2 at + 0.96 and + 0.86 V, far higher
at 10 mV s− 1. It can be seen that the CV cycle of Co,Nb-MoS2/TiO2 HSs than those of the Co-MoS2/TiO2 HSs (0.92 V and 0.81 V), Nb-MoS2/TiO2
has a larger area than its counterparts, indicating the exposure of a HSs (0.76 V and 0.67 V), MoS2/TiO2 HSs (0.74 V and 0.63 V), Co,Nb-
better active area in the Co,Nb-MoS2/TiO2 HSs material. Based on the MoS2 NSs (0.81 V and 0.71 V), and TiO2 HSs (0.71 V and 0.62 V). These
calculation of line slope in Fig. 5i, the Cdl values for the Co,Nb-MoS2/ results suggest that the formation of unique core–shell architecture and
TiO2 HSs, Co-MoS2/TiO2 HSs, Co,Nb-MoS2 NSs, Nb-MoS2/TiO2 HSs, the successful incorporation of both Co and Nb are the primary factors
MoS2/TiO2 HSs, and TiO2 HSs are found to be 23.7, 22.4, 15.2, 6.4, 4.5, for dramatically expediting the ORR performance. Although the Eonset
and 2.2 mF cm− 2, respectively. In this regard, the Co,Nb-MoS2/TiO2 and E1/2 of the Co,Nb-MoS2/TiO2 HSs are slightly lower than those of Pt/
HSs possesses the highest Cdl value, and thus produces the largest ECSA, C of 0.97 V and 0.87 V, it is noteworthy that the ORR activities of the Co,
suggesting the enhancement of catalytic active sites provided by the Nb-MoS2/TiO2 HSs can be favorably compared with those of recently
unique structure of the Co,Nb-MoS2/TiO2 HSs and the doping effects of reported ORR catalysts (Table S4), further demonstrating the competi­
Co and Nb dopants. tive activities of the Co,Nb-MoS2/TiO2 for the ORR. The Tafel plots
In other regards, the ORR performance of the Co,Nb-MoS2/TiO2 HSs (Fig. 6d) demonstrate that the Co,Nb-MoS2/TiO2 HSs possesses the small
was also examined in alkaline condition using a basic three-electrode Tafel slope value of 56.1 mV dec− 1, which is only 1.6 mV dec− 1 higher
system, in which the Co,Nb-MoS2/TiO2 HSs-loaded RDE acted as a than that of the Pt/C (54.5 mV dec− 1), suggesting that the advantages of
working electrode. In the CV measurements (Fig. 6a), the Co,Nb-MoS2/ the morphological and electronic structural properties of the Co,Nb-
TiO2 HSs clearly shows an oxygen reduction peak starting at ~ 0.96 V in MoS2/TiO2 HSs not only improve the ORR activities, but also enhance
O2-saturated 0.1 M KOH solution, while no reduction peak can be the ORR kinetics. The superior kinetics of the Co,Nb-MoS2/TiO2 HSs is
observed from the CV cycle in N2-saturated electrolyte, reflecting the further demonstrated, since it is compared to those of the Co-MoS2/TiO2

10
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

Fig. 6. (a) CV curves of the Co,Nb-MoS2/TiO2 HSs in N2- and O2-saturated 0.1 M KOH solution; (b) Polarization curves of the materials; (c) Comparison of the Eonset
and E1/2 of the materials; (d) Tafel plots of the materials; (e) Polarization curves of the Co,Nb-MoS2/TiO2 HSs at various rotating speeds; (f) Koutecky–Levich plots of
the Co,Nb-MoS2/TiO2 HSs and n value (inset); (g) Comparison of the polarization curves of the Co,Nb-MoS2/TiO2 HSs before and after 1000 CV cycles; (h) Chro­
noamperometric curves of the Co,Nb-MoS2/TiO2 HSs and Pt/C catalyst at operating potentials of 0.86 and 0.87 V, for 60,000 s each, respectively; (i) Methanol
crossover tests for the Co,Nb-MoS2/TiO2 HSs and Pt/C catalyst under the influence of 1.0 M methanol addition. [(1) Co,Nb-MoS2/TiO2 HSs, (2) Co-MoS2/TiO2 HSs,
(3) Co,Nb-MoS2 NSs, (4) Nb-MoS2/TiO2 HSs, (5) MoS2/TiO2 HSs, (6) TiO2 HSs, and (7) Pt/C)].

HSs (76.5 mV dec− 1), Nb-MoS2/TiO2 HSs (93.5 mV dec− 1), MoS2/TiO2 results. Moreover, the long-term durability and stability of the Co,Nb-
HSs (100.1 mV dec− 1), Co, Nb-MoS2 HSs (86.4 mV dec− 1), and TiO2 MoS2/TiO2 HSs towards ORR is investigated by cycling stability test in
HSs (113.9 mV dec− 1). The ORR mechanism of the Co,Nb-MoS2/TiO2 O2-saturated 0.1 M KOH solution. After cycling test, the E1/2 value is
HSs was further investigated through LSV measurement at various only 24 mV lower than that of the initial one, indicating no obvious
rotating rates of 400, 800, 1200, 1600, and 2000 rpm. The LSV curves in change in the ORR performance of the Co,Nb-MoS2/TiO2 HSs before and
Fig. 6e show a quick increase in the limited diffusion current density of after 1000 cycles (Fig. 6g). This is favorable evidence for the good
the Co,Nb-MoS2/TiO2 HSs since the rotation speed increases, implying durability and stability of the Co,Nb-MoS2/TiO2 HSs towards ORR. In
an effective diffusion process during the ORR catalysis. Based on the addition, chronoamperometry was also carried out for the Co,Nb-MoS2/
Koutecky–Levich (K–L) equation, the K–L plots obtained from the Co, TiO2 HSs to further probe its stability (Fig. 6h). The Co,Nb-MoS2/TiO2
Nb-MoS2/TiO2 HSs (Fig. 6f) indicates n ≈ 3.96 in the potential range of HSs maintains 96.65% of its initial current density after 60,000 s of
(+ 0.31 to + 0.66) V, implying a direct 4e- pathway occurring during continuous ORR at operating potential of 0.86 V, superior to that of the
the ORR in 0.1 M KOH. In order to further investigate the ORR mech­ Pt/C catalyst (a retention of only 88.78% activity at 30,000 s). The SEM
anism of the Co,Nb-MoS2/TiO2 HSs, we carried out RRDE test of the Co, images of the post-ORR Co,Nb-MoS2/TiO2 HSs (Fig. S15a and b) show
Nb-MoS2/TiO2 HSs along with the Pt/C for comparison. The RRDE po­ that its initial morphology is still well retained after 60,000 s stability
larization curves of the Co,Nb-MoS2/TiO2 HSs and Pt/C are given in test. The EDAX pattern of the post-ORR Co,Nb-MoS2/TiO2 HSs still in­
Fig. S14a, which clearly exhibits that the disk current is far higher than dicates the presence of the required elements, such as Co, Nb, Mo, S, Ti,
that of the corresponding ring current. Based on the RRDE test result and and O (Inset of Fig. S15b).
the Eqs. (6) and (7), the n and H2O2 percentage of Co,Nb-MoS2/TiO2 HSs The TEM image in Fig. S15c displays the good preservation of the
are found to be between 3.89 and 3.95 and 9.2–11.9% in the 4.0–6.0 V hierarchical core–shell structure with hollow spherical feature of the
potential range, respectively, which is almost comparable to the Pt/C post-ORR Co,Nb-MoS2/TiO2 HSs. A set of lattice fringes with interplanar
(3.91–3.99 and 4.56–5.84%) (Fig. S14b and c). Therefore, the obtained distances of 0.64 and 0.27 nm are observed from the HR-TEM image
results corroborate that the ORR process mainly occurs through a direct (Fig. S15d), which correspond with the (002), and (100) planes of the
4e- pathway for the Co,Nb-MoS2/TiO2 HSs, consistent with the RDE test MoS2, respectively, proving that after a continuous ORR operation in

11
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

alkaline medium, the MoS2 phase is still stable. In addition, the EDS application of Perdew–Burke–Ernzerhof functional for exchange-
elemental mapping images (Fig. S15e–l) of the post-ORR Co,Nb-MoS2/ correlation energy. Fig. 7a and b show the top view and side view of
TiO2 HSs display the uniform distribution of the Ti and O elements at the the atomic structure for the pristine MoS2 layer, and Co and Nb dual-
core region, while the Mo, S, Co, and Nb elements are distributed at the doped MoS2 layer, respectively, along with some important parame­
shell area, further verifying its core–shell structure is retained well. ters, such as formation energy, and atomic distance. Electronic band
Thus, the SEM, TEM, and EDS analyses clearly show the good structure of the Co,Nb-doped MoS2 and the sets of its counterparts, as
morphological stability of the material under long-term operation. The well as their interaction ability with adsorbents were identify by density
excellent alcohol tolerance of the Co,Nb-MoS2/TiO2 HSs was assessed by of states (DOS) calculations [66,67]. Fig. 7c compares DOS result of the
adding 1.0 M methanol to 0.1 M O2-saturated KOH solution during the Co,Nb-doped MoS2 with Co-doped MoS2, Nb-doped MoS2, and single
ORR process. The chronoamperometric test (Fig. 6i) indicates that after MoS2. As seen, the electronic states of the Co,Nb-doped MoS2 are close to
the methanol addition into electrolyte, the current response of the Pt/C Fermi level, and are significantly higher than for its counterparts,
catalyst immediately increases, due to effective methanol oxidation at implying faster charge transfer and lower activation energy for inter­
its surface, while the Co,Nb-MoS2/TiO2 HSs presents no clear current mediate adsorptions of Co,Nb-doped MoS2 [17]. The increased elec­
change, indicating the great potential of this catalyst as cathodic catalyst tronic states around the Fermi level is mainly due to the
in direct methanol fuel cell application. co-incorporation of Co and Nb, causing the enhanced electron transfer
In order to exclude the NF influence on the electrocatalytic perfor­ and intermediate adsorption ability, and thus the catalytic activity.
mance, we adhered the as-synthesized Co,Nb-MoS2/TiO2 HSs sample Moreover, it is well-known that adsorption free energies of hydrogen
onto glassy carbon (GC) electrode (label as Co,Nb-MoS2/TiO2 HSs/GC) (∆GH*) is a key parameter in reflecting HER activity; in general, an
instead of the NF and conducted LSV measurement under electro­ excellent HER electrocatalyst demonstrates an optimal ∆GH* value
catalytic HER and OER conditions for the obtained electrode along with around 0 eV [68]. Accordingly, to further understand the synergistic
pure GC and pristine NF for comparison (Fig. S16a and b) [64,65]. As effects deriving from MoS2 and Co, Nb dual doping for the HER, ∆GH* on
seen, the pure GC and pristine NF exhibit very poor HER and OER Mo, S, Co, and Nb centers in MoS2 and Co,Nb-doped MoS2 were calcu­
performance, suggesting the key contribution of the host Co, lated and compared in Fig. 7d. The pure MoS2 possesses a high ∆GH*
Nb-MoS2/TiO2 HSs in achieving extraordinary catalytic activity. It is values of 2.525 and 1.733 eV for Mo and S sites, respectively. Whereas,
well known that onset overpotential is an important descriptor reflected after incorporation of Co and Nb into MoS2 structure, the ∆GH* values of
properly the catalytic activity of materials, and thus to evaluate Mo and S centers are significantly reduced to 0.055 and 1.687 eV,
correctly the catalytic activity of the Co,Nb-MoS2/TiO2 HSs, we deter­ respectively, demonstrating remarkable enhancement of HER activity.
mined onset overpotential values of the Co,Nb-MoS2/TiO2 HSs/GC and In particularly, the adsorption free energy of H on Mo center in Co,
Co,Nb-MoS2/TiO2 HSs/NF samples based on start point of the Tafel Nb-doped MoS2 is very close to 0 eV, consistent with its good HER
slope. Fig. S16c and d show that the onset overpotential of Co, performance. Noticeably, the ∆GH* values of the Co (0.969 eV) and Nb
Nb-MoS2/TiO2 HSs/GC is found to be 44.2 and 248.3 mV for the HER (0.303 eV) sites in Co,Nb-doped MoS2 are also close to 0 eV, suggesting
and OER, respectively, which is almost comparable with the Co, that Co and Nb may act as additional active sites in this structure. Thus,
Nb-MoS2/TiO2 HSs/NF (29.7 and 228.6 mV), corroborating the excel­ the computational studies corroborate a much better HER activity of the
lent catalytic activity of Co,Nb-MoS2/TiO2 HSs - in other words, the NF Co,Nb-doped MoS2 than that of the pure MoS2, which are consistent well
substrate has little influence on the electrocatalytic properties when with experimental results.
compared with the host Co,Nb-MoS2/TiO2 HSs. Furthermore, the CV
measurement at different scan rates was also performed for the NF 4. Conclusions
(Fig. S16e) to evaluate its ECSA through determination of Cdl. The result
in Fig. S16f reveals that the NF reaches Cdl of 1.52 mF cm− 2, which is In summary, we have developed a novel Co,Nb-MoS2/TiO2 HSs
15.6 time lower than that of the Co,Nb-MoS2/TiO2 HSs/NF, suggesting hybrid material with mesoporous nanosheet-shelled heterostructure via
the much smaller ECSA of NF than that of the Co,Nb-MoS2/TiO2 a simple and effective strategy. Benefiting from the synergistic effects of
HSs/NF. This also implies that the enhanced exposure of accessible valuable factors, including the unique hollow core–shell structure con­
active sites for reactants and/or electrolyte originates mainly from the taining strong electronic interactions between the core part and shell
host Co,Nb-MoS2/TiO2 HSs; the NF substrate thus possesses negligible layer, the Co and Nb dual-doping effects, and the ultrathin and meso­
contribution to the total catalytic performance. porous characteristics of the MoS2 NSs, the catalytic performance of the
In addition, we also examined the HER, OER, and ORR performance Co,Nb-MoS2/TiO2 HSs towards the HER, OER, and ORR reactions is
of Co,Nb-MoS2/TiO2 HSs prepared under different molar ratio of the significant improved. As a result, the Co,Nb-MoS2/TiO2 HSs displays
reactant Co(NO3)2 and NbCl5, including 2:1, 1:1, and 1:2 (denoted as 1- good activities for the simultaneous HER and OER in alkaline condition
Co,Nb-MoS2/TiO2 HSs, 2-Co,Nb-MoS2/TiO2 HSs, and 3-Co,Nb-MoS2/ with the HER and OER η10 values of 58.8 and 260.0 mV, respectively.
TiO2 HSs, respectively) to investigate the influence of Co and Nb doping The electrolyzer with symmetric cell based on the Co,Nb-MoS2/TiO2 HSs
content on the catalytic performance. The HER polarization curves of as both cathode and anode electrodes can drive 10 mA cm− 2 at an
these samples in Fig. S17a show that the 1-Co,Nb-MoS2/TiO2 HSs operating voltage of 1.59 V, which is superior to that of the recently
sample possesses an optimal HER catalytic activity along with smaller η reported bifunctional electrocatalysts. In addition, the Co,Nb-MoS2/
and Tafel slope value than for the others (Fig. S17b and c). Likewise, TiO2 HSs demonstrates efficient electrocatalytic activities for the ORR
Fig. S17d–f reveal the better OER catalytic activity of the 1-Co,Nb- with high Eonset value of + 0.96 V, E1/2 of + 0.86 V, and a direct 4e-
MoS2/TiO2 HSs sample as compared to the others. Furthermore, the 1- pathway in 0.1 M KOH media. Our work paves a way for the effective
Co,Nb-MoS2/TiO2 HSs also demonstrates the best ORR catalytic activity improvement of the intrinsic catalytic activities of MoS2-based nano­
with a lowest Eonset and E1/2 value (Fig. S17g–i). Therefore, the above- material towards HER, OER, and ORR via the doping effect to upgrade
obtained results clearly indicate dependence of the electrocatalytic ac­ its advantaged properties, thus leading to it being suitable for multi­
tivity of Co,Nb-MoS2/TiO2 HSs on the Co and Nb doping content; the Co, functional applications in the forthcoming critically needed green en­
Nb-MoS2/TiO2 HSs with the molar ratio of Co:Nb = 2:1 shows the ergy conversion and storage technologies.
optimal electrocatalytic performance for all HER, OER, and ORR.
Furthermore, to provide deep insight into the origin of the catalytic CRediT authorship contribution statement
enhancement, we performed theoretical calculations based on density
functional theory (DFT). In this work, the DFT calculations were carried Dinh Chuong Nguyen: Methodology, writing-original draft, vali­
out with Vienna ab initio simulation package (VASP), together with the dation, validation visualization. Thi Luu Luyen Doan: Methodology,

12
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

Fig. 7. (a and b) Schematic illustration for the atomic structure of MoS2, H-Mo (MoS2), Co,Nb-MoS2, and H-Mo (Co,Nb-MoS2) with top and side view as well as their
corresponding structural parameters; (c) Calculated DOS curves for MoS2, Co-MoS2, Nb-MoS2 and Co,Nb-MoS2; and (d) The Gibbs free energy of H* (ΔGH*) for (1) H-
Co(Co,Nb-MoS2), (2) H-Nb(Co,Nb-MoS2), (3) H-Mo(Co,Nb-MoS2), (4) H-S(Co,Nb-MoS2), (5) H-Mo(MoS2), and (6) H-S(MoS2).

investigation, formal analysis. Sampath Prabhakaran: Density- [2] Z. Qian, Y. Chen, Z. Tang, Z. Liu, X. Wang, Y. Tian, W. Gao, Hollow nanocages of
NixCo1− xSe for efficient zinc–air batteries and overall water splitting, Nanomicro
functional theory calculation. Duy Thanh Tran: Writing- reviewing
Lett. 11 (2019) 1–17, https://doi.org/10.1007/s40820-019-0258-0.
and editing. Do Hwan Kim: Density-functional theory calculation, [3] T. Van Tam, S.G. Kang, M.H. Kim, S.G. Lee, S.H. Hur, J.S. Chung, W.M. Choi, Novel
writing- reviewing and editing. Joong Hee Lee: Conceptualization, data graphene hydrogel/B-doped graphene quantum dots composites as trifunctional
curation, writing- reviewing and editing, supervision. Nam Hoon Kim: electrocatalysts for Zn− air batteries and overall water splitting, Adv. Energy Mater.
9 (2019), 1900945, https://doi.org/10.1002/aenm.201900945.
Conceptualization, writing- reviewing and editing, supervision, project [4] W. Liu, L. Yu, R. Yin, X. Xu, J. Feng, X. Jiang, D. Zheng, X. Gao, X. Gao, W. Que,
administration. P. Ruan, F. Wu, W. Shi, X. Cao, Non-3d metal modulation of a 2D Ni–Co
heterostructure array as multifunctional electrocatalyst for portable overall water
splitting, Small 16 (2020) 1–8, https://doi.org/10.1002/smll.201906775.
Declaration of Competing Interest [5] D.C. Nguyen, D.T. Tran, T.L.L. Doan, D.H. Kim, N.H. Kim, J.H. Lee, Rational design
of core@shell structured CoSx@Cu2MoS4 hybridized MoS2/N,S-codoped graphene
The authors declare that they have no known competing financial as advanced electrocatalyst for water splitting and Zn-air battery, Adv. Energy
Mater. 10 (2020), 1903289, https://doi.org/10.1002/aenm.201903289.
interests or personal relationships that could have appeared to influence [6] Y. Guo, P. Yuan, J. Zhang, H. Xia, F. Cheng, M. Zhou, J. Li, Y. Qiao, S. Mu, Q. Xu,
the work reported in this paper. Co2P–CoN double active centers confined in N-doped carbon nanotube:
heterostructural engineering for trifunctional catalysis toward HER, ORR, OER,
and Zn–Air batteries driven water splitting, Adv. Funct. Mater. 28 (2018) 1–9,
Acknowledgments https://doi.org/10.1002/adfm.201805641.
[7] Q. Xu, Y. Liu, H. Jiang, Y. Hu, H. Liu, C. Li, Unsaturated sulfur edge engineering of
This research was supported by the Basic Science Research Program strongly coupled MoS2 nanosheet–carbon macroporous hybrid catalyst for
enhanced hydrogen generation, Adv. Energy Mater. 9 (2019), 1802553, https://
(2019R1A2C1004983) and the Regional Leading Research Center Pro­ doi.org/10.1002/aenm.201802553.
gram (2019R1A5A8080326) through the National Research Foundation [8] W. Zhou, M. Chen, M. Guo, A. Hong, T. Yu, X. Luo, C. Yuan, W. Lei, S. Wang,
funded by the Ministry of Science and ICT of Republic of Korea. Magnetic enhancement for hydrogen evolution reaction on ferromagnetic MoS2
catalyst, Nano Lett. 20 (2020) 2923–2930, https://doi.org/10.1021/acs.
nanolett.0c00845.
Appendix A. Supporting information [9] Y. Zhang, Y. Gao, S. Yao, S. Li, H. Asakura, K. Teramura, H. Wang, D. Ma,
Sublimation-induced sulfur vacancies in MoS2 catalyst for one-pot synthesis of
secondary amines, ACS Catal. 9 (2019) 7967–7975, https://doi.org/10.1021/
Supplementary data associated with this article can be found in the
acscatal.9b01429.
online version at doi:10.1016/j.nanoen.2021.105750. [10] S. Sun, Q. An, M. Watanabe, J. Cheng, H. Ho Kim, T. Akbay, A. Takagaki,
T. Ishihara, Highly correlation of CO2 reduction selectivity and surface electron
References accumulation: a case study of Au-MoS2 and Ag-MoS2 catalyst, Appl. Catal. B
Environ. 271 (2020), 118931, https://doi.org/10.1016/j.apcatb.2020.118931.
[11] Q. Li, X. Bai, C. Ling, Q. Zhou, S. Yuan, Q. Chen, J. Wang, Forming atom–vacancy
[1] J. Wang, W. Cui, Q. Liu, Z. Xing, A.M. Asiri, X. Sun, Recent progress in cobalt-based interface on the MoS2 catalyst for efficient hydrodeoxygenation reactions, Small
heterogeneous catalysts for electrochemical water splitting, Adv. Mater. 28 (2016) Methods 3 (2019) 1–7, https://doi.org/10.1002/smtd.201800315.
215–230, https://doi.org/10.1002/adma.201502696.

13
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

[12] H. Huang, X. Feng, C. Du, S. Wu, W. Song, Incorporated oxygen in MoS2 ultrathin splitting: revisiting activity parameters with a critical assessment, Energy Environ.
nanosheets for efficient ORR catalysis, J. Mater. Chem. A 3 (2015) 16050–16056, Sci. 11 (2018) 744–771, https://doi.org/10.1039/c7ee03457a.
https://doi.org/10.1039/c5ta01600b. [35] C. Zhu, Z. Yin, W. Lai, Y. Sun, L. Liu, X. Zhang, Y. Chen, S.-L. Chou, Fe-Ni-Mo
[13] K. Yan, L. Yiran, Direct growth of MoS2 microspheres on Ni foam as a hybrid nitride porous nanotubes for full water splitting and Zn-air batteries, Adv. Energy
nanocomposite efficient for oxygen evolution reaction, Small 12 (2016) Mater. 8 (2018), 1802327, https://doi.org/10.1002/aenm.201802327.
2975–2981, https://doi.org/10.1002/smll.201600332. [36] Y. Guo, P. Yuan, J. Zhang, Y. Hu, I.S. Amiinu, X. Wang, J. Zhou, H. Xia, Z. Song,
[14] J. Xiao, D. Choi, L. Cosimbescu, P. Koech, J. Liu, J.P. Lemmon, Exfoliated MoS2 Q. Xu, S. Mu, Carbon nanosheets containing discrete Co-Nx-By-C active sites for
nanocomposite as an anode material for lithium ion batteries, Chem. Mater. 22 efficient oxygen electrocatalysis and rechargeable zn-air batteries, ACS Nano 12
(2010) 4522–4524, https://doi.org/10.1021/cm101254j. (2018) 1894–1901, https://doi.org/10.1021/acsnano.7b08721.
[15] J. Pei, H. Geng, E.H. Ang, L. Zhang, X. Cao, J. Zheng, H. Gu, Controlled synthesis of [37] D.C. Nguyen, D.T. Tran, T.L. Luyen Doan, N.H. Kim, J.H. Lee, Constructing MoPx@
hollow C@TiO2@MoS2 hierarchical nanospheres for high-performance lithium-ion MnPy heteronanoparticle-supported mesoporous N,P-codoped graphene for
batteries, Nanoscale 10 (2018) 17327–17334, https://doi.org/10.1039/ boosting oxygen reduction and oxygen evolution reaction, Chem. Mater. 31 (2019)
c8nr05451g. 2892–2904, https://doi.org/10.1021/acs.chemmater.9b00071.
[16] M. Yu, S. Zhao, H. Feng, L. Hu, X. Zhang, Y. Zeng, Y. Tong, X. Lu, Engineering thin [38] R. Zhou, Y. Zheng, M. Jaroniec, S.-Z. Qiao, Determination of the electron transfer
MoS2 nanosheets on TiN nanorods: advanced electrochemical capacitor electrode number for the oxygen reduction reaction: from theory to experiment, ACS Catal. 6
and hydrogen evolution electrocatalyst, ACS Energy Lett. 2 (2017) 1862–1868, (2016) 4720–4728, https://doi.org/10.1021/acscatal.6b01581.
https://doi.org/10.1021/acsenergylett.7b00602. [39] Y. Zhang, W. Sun, X. Rui, B. Li, H.T. Tan, G. Guo, S. Madhavi, Y. Zong, Q. Yan, One-
[17] J. Deng, H. Li, J. Xiao, Y. Tu, D. Deng, H. Yang, H. Tian, J. Li, P. Ren, X. Bao, pot synthesis of tunable crystalline Ni3S4@Amorphous MoS2 core/shell
Triggering the electrocatalytic hydrogen evolution activity of the inert two- nanospheres for high-performance supercapacitors, Small 11 (2015) 3694–3702,
dimensional MoS2 surface via single-atom metal doping, Energy Environ. Sci. 8 https://doi.org/10.1002/smll.201403772.
(2015) 1594–1601, https://doi.org/10.1039/c5ee00751h. [40] L. Yu, H. Zhou, J. Sun, F. Qin, F. Yu, J. Bao, Y. Yu, S. Chen, Z. Ren, Cu nanowires
[18] J. Deng, H. Li, S. Wang, D. Ding, M. Chen, C. Liu, Z. Tian, K.S. Novoselov, C. Ma, shelled with NiFe layered double hydroxide nanosheets as bifunctional
D. Deng, X. Bao, Multiscale structural and electronic control of molybdenum electrocatalysts for overall water splitting, Energy Environ. Sci. 10 (2017)
disulfide foam for highly efficient hydrogen production, Nat. Commun. 8 (2017) 1820–1827, https://doi.org/10.1039/c7ee01571b.
14430, https://doi.org/10.1038/ncomms14430. [41] Q. Zhou, T.T. Li, J. Qian, Y. Hu, F. Guo, Y.Q. Zheng, Self-supported hierarchical
[19] Y. Shi, Y. Zhou, D.R. Yang, W.X. Xu, C. Wang, F. Bin Wang, J.J. Xu, X.H. Xia, H. CuOx@Co3O4 heterostructures as efficient bifunctional electrocatalysts for water
Y. Chen, Energy level engineering of MoS2 by transition-metal doping for splitting, J. Mater. Chem. A 6 (2018) 14431–14439, https://doi.org/10.1039/
accelerating hydrogen evolution reaction, J. Am. Chem. Soc. 139 (2017) c8ta03120g.
15479–15485, https://doi.org/10.1021/jacs.7b08881. [42] H.B. Jiang, Q. Cuan, C.Z. Wen, J. Xing, D. Wu, X.-Q. Gong, C. Li, H.G. Yang,
[20] Q. Xiong, Y. Wang, P.F. Liu, L.R. Zheng, G. Wang, H.G. Yang, P.K. Wong, H. Zhang, Anatase TiO2 crystals with exposed high-index facets, Angew. Chem. Int. Ed. 123
H. Zhao, Cobalt covalent doping in MoS2 to induce bifunctionality of overall water (2011) 3848–3852, https://doi.org/10.1002/ange.201007771.
splitting, Adv. Mater. 30 (2018) 1–7, https://doi.org/10.1002/adma.201801450. [43] H. Liu, X. Chen, L. Deng, M. Ding, J. Li, X. He, Perpendicular growth of few-layered
[21] J. Staszak-Jirkovský, C.D. Malliakas, P.P. Lopes, N. Danilovic, S.S. Kota, K. MoS2 nanosheets on MoO3 nanowires fabricated by direct anion exchange
C. Chang, B. Genorio, D. Strmcnik, V.R. Stamenkovic, M.G. Kanatzidis, N. reactions for high-performance lithium-ion batteries, J. Mater. Chem. A 4 (2016)
M. Markovic, Design of active and stable Co-Mo-Sx chalcogels as pH-universal 17764–17772, https://doi.org/10.1039/c6ta06741g.
catalysts for the hydrogen evolution reaction, Nat. Mater. 15 (2016) 197–203, [44] M. Xu, F.L. Yi, Y. Niu, J. Xie, J. Hou, S. Liu, W.H. Hu, Y. Li, C.M. Li, Solvent-
https://doi.org/10.1038/nmat4481. mediated directionally self-assembling MoS2 nanosheets into a novel worm-like
[22] Q. Xiong, X. Zhang, H. Wang, G. Liu, G. Wang, H. Zhang, H. Zhao, One-step structure and its application in sodium batteries, J. Mater. Chem. A 3 (2015)
synthesis of cobalt-doped MoS2 nanosheets as bifunctional electrocatalysts for 9932–9937, https://doi.org/10.1039/c5ta00315f.
overall water splitting under both acidic and alkaline conditions, Chem. Commun. [45] N.D. Chuong, T.D. Thanh, N.H. Kim, J.H. Lee, Hierarchical heterostructures of
54 (2018) 3859–3862, https://doi.org/10.1039/c8cc00766g. ultrasmall Fe2O3-encapsulated MoS2/N-graphene as an effective catalyst for
[23] Z. Wang, J. Zhao, Q. Cai, F. Li, Computational screening for high-activity MoS2 oxygen reduction reaction, ACS Appl. Mater. Interfaces 10 (2018) 24523–24532,
monolayer-based catalysts for the oxygen reduction reaction via substitutional https://doi.org/10.1021/acsami.8b06485.
doping with transition metal, J. Mater. Chem. A 5 (2017) 9842–9851, https://doi. [46] Z. Chen, Y.J. Xu, Ultrathin TiO2 layer coated-CdS spheres core-shell nanocomposite
org/10.1039/c7ta00577f. with enhanced visible-light photoactivity, ACS Appl. Mater. Interfaces 5 (2013)
[24] S.W. Kim, T.H. Han, J. Kim, H. Gwon, H.S. Moon, S.W. Kang, S.O. Kim, K. Kang, 13353–13363, https://doi.org/10.1021/am4043068.
Fabrication and electrochemical characterization of TiO2 three-dimensional [47] L. An, Y. Li, M. Luo, J. Yin, Y.Q. Zhao, C. Xu, F. Cheng, Y. Yang, P. Xi, S. Guo,
nanonetwork based on peptide assembly, ACS Nano 3 (2009) 1085–1090, https:// Atomic-level coupled interfaces and lattice distortion on CuS/NiS2 nanocrystals
doi.org/10.1021/nn900062q. boost oxygen catalysis for flexible Zn-air batteries, Adv. Funct. Mater. 27 (2017)
[25] Z. Xiao, Z. Yang, L. Wang, H. Nie, M. Zhong, Q. Lai, X. Xu, L. Zhang, S. Huang, 1–9, https://doi.org/10.1002/adfm.201703779.
A lightweight TiO2/Graphene interlayer, applied as a highly effective polysulfide [48] Y. Liu, S. Jiang, S. Li, L. Zhou, Z. Li, J. Li, M. Shao, Interface engineering of (Ni, Fe)
absorbent for fast, long-life lithium-sulfur batteries, Adv. Mater. 27 (2015) S2@MoS2 heterostructures for synergetic electrochemical water splitting, Appl.
2891–2898, https://doi.org/10.1002/adma.201405637. Catal. B 247 (2019) 107–114, https://doi.org/10.1016/j.apcatb.2019.01.094.
[26] E. Borgarello, J. Kiwi, E. Pelizzetti, M. Visca, M. Grätzel, Photochemical cleavage of [49] H. Zhu, J. Zhang, R. Yanzhang, M. Du, Q. Wang, G. Gao, J. Wu, G. Wu, M. Zhang,
water by photocatalysis, Nature 289 (1981) 158–160, https://doi.org/10.1038/ B. Liu, J. Yao, X. Zhang, When cubic cobalt sulfide meets layered molybdenum
289158a0. disulfide: a core-shell system toward synergetic electrocatalytic water splitting,
[27] S. Nong, W. Dong, J. Yin, B. Dong, Y. Lu, X. Yuan, X. Wang, K. Bu, M. Chen, Adv. Mater. 27 (2015) 4752–4759, https://doi.org/10.1002/adma.201501969.
S. Jiang, L.M. Liu, M. Sui, F. Huang, Well-dispersed ruthenium in mesoporous [50] Y. Guo, J. Tang, J. Henzie, B. Jiang, W. Xia, T. Chen, Y. Bando, Y.M. Kang, M.S.
crystal TiO2 as an advanced electrocatalyst for hydrogen evolution reaction, J. Am. A. Hossain, Y. Sugahara, Y. Yamauchi, Mesoporous iron-doped MoS2/CoMo2S4
Chem. Soc. 140 (2018) 5719–5727, https://doi.org/10.1021/jacs.7b13736. heterostructures through organic-metal cooperative interactions on spherical
[28] J.X. Feng, H. Xu, Y.T. Dong, X.F. Lu, Y.X. Tong, G.R. Li, Efficient hydrogen micelles for electrochemical water splitting, ACS Nano 14 (2020) 4141–4152,
evolution electrocatalysis using cobalt nanotubes decorated with titanium dioxide https://doi.org/10.1021/acsnano.9b08904.
nanodots, Angew. Chem. Int. Ed. 56 (2017) 2960–2964, https://doi.org/10.1002/ [51] Q. Qin, L. Chen, T. Wei, X. Liu, MoS2/NiS yolk–shell microsphere-based electrodes
anie.201611767. for overall water splitting and asymmetric supercapacitor, Small 15 (2019) 1–13,
[29] H.A. Hamedani, N.K. Allam, M.A. El-Sayed, M.A. Khaleel, H. Garmestani, F. https://doi.org/10.1002/smll.201803639.
M. Alamgir, An experimental insight into the structural and electronic [52] Y. Shi, Y. Zhou, D.R. Yang, W.X. Xu, C. Wang, F. Bin Wang, J.J. Xu, X.H. Xia, H.
characteristics of strontium-doped titanium dioxide nanotube arrays, Adv. Funct. Y. Chen, Energy level engineering of MoS2 by transition-metal doping for
Mater. 24 (2014) 6783–6796, https://doi.org/10.1002/adfm.201401760. accelerating hydrogen evolution reaction, J. Am. Chem. Soc. 139 (2017)
[30] J. Liang, C. Wang, P. Zhao, Y. Wang, L. Ma, G. Zhu, Y. Hu, Z. Lu, Z. Xu, Y. Ma, 15479–15485, https://doi.org/10.1021/jacs.7b08881.
T. Chen, Z. Tie, J. Liu, Z. Jin, Interface engineering of anchored ultrathin TiO2/ [53] F.L. Deepak, H. Cohen, S. Cohen, Y. Feldman, R. Popovitz-Biro, D. Azulay, O. Millo,
MoS2 heterolayers for highly-efficient electrochemical hydrogen production, ACS R. Tenne, Fullerene-like (IF) NbxMo1− xS2 nanoparticles, J. Am. Chem. Soc. 129
Appl. Mater. Interfaces 10 (2018) 6084–6089, https://doi.org/10.1021/ (2007) 12549–12562, https://doi.org/10.1021/ja074081b.
acsami.7b19009. [54] X.J. Chua, J. Luxa, A.Y.S. Eng, S.M. Tan, Z. Sofer, M. Pumera, Negative
[31] Y. Wang, R. Zhang, Y. chao Pang, X. Chen, J. Lang, J. Xu, C. Xiao, H. Li, K. Xi, S. electrocatalytic effects of p-doping niobium and tantalum on MoS2 and WS2 for the
Ding, Carbon@titanium nitride dual shell nanospheres as multi-functional hosts for hydrogen evolution reaction and oxygen reduction reaction, ACS Catal. 6 (2016)
lithium sulfur batteries, Energy Storage Mater. 16 (2019) 228–235, https://doi. 5724–5734, https://doi.org/10.1021/acscatal.6b01593.
org/10.1016/j.ensm.2018.05.019. [55] Z. Shi, K. Nie, Z.J. Shao, B. Gao, H. Lin, H. Zhang, B. Liu, Y. Wang, Y. Zhang, X. Sun,
[32] H. Ren, R. Yu, J. Wang, Q. Jin, M. Yang, D. Mao, D. Kisailus, H. Zhao, D. Wang, X.M. Cao, P. Hu, Q. Gao, Y. Tang, Phosphorus-Mo2C@carbon nanowires toward
Multishelled TiO2 hollow microspheres as anodes with superior reversible capacity efficient electrochemical hydrogen evolution: composition, structural and
for lithium ion batteries, Nano Lett. 14 (2014) 6679–6684, https://doi.org/ electronic regulation, Energy Environ. Sci. 10 (2017) 1262–1271, https://doi.org/
10.1021/nl503378a. 10.1039/c7ee00388a.
[33] X. Lai, J.E. Halpert, D. Wang, Recent advances in micro-/nano-structured hollow [56] Q. Liu, Q. Fang, W. Chu, Y. Wan, X. Li, W. Xu, M. Habib, S. Tao, Y. Zhou, D. Liu,
spheres for energy applications: from simple to complex systems, Energy Environ. T. Xiang, A. Khalil, X. Wu, M. Chhowalla, P.M. Ajayan, L. Song, Electron-doped 1T-
Sci. 5 (2012) 5604–5618, https://doi.org/10.1039/c1ee02426d. MoS2 via interface engineering for enhanced electrocatalytic hydrogen evolution,
[34] S. Anantharaj, S.R. Ede, K. Karthick, S. Sam Sankar, K. Sangeetha, P.E. Karthik, Chem. Mater. 29 (2017) 4738–4744, https://doi.org/10.1021/acs.
S. Kundu, Precision and correctness in the evaluation of electrocatalytic water chemmater.7b00446.

14
D.C. Nguyen et al. Nano Energy 82 (2021) 105750

[57] D. Voiry, R. Fullon, J. Yang, C. De Carvalho Castro, E. Silva, R. Kappera, I. Bozkurt, [63] D. Xiong, Q. Zhang, W. Li, J. Li, X. Fu, M.F. Cerqueira, P. Alpuim, L. Liu, Atomic-
D. Kaplan, M.J. Lagos, P.E. Batson, G. Gupta, A.D. Mohite, L. Dong, D. Er, V. layer-deposited ultrafine MoS2 nanocrystals on cobalt foam for efficient and stable
B. Shenoy, T. Asefa, M. Chhowalla, The role of electronic coupling between electrochemical oxygen evolution, Nanoscale 9 (2017) 2711–2717, https://doi.
substrate and 2D MoS2 nanosheets in electrocatalytic production of hydrogen, Nat. org/10.1039/c7nr00140a.
Mater. 15 (2016) 1003–1009, https://doi.org/10.1038/nmat4660. [64] M.Q. Yu, Y.H. Li, S. Yang, P.F. Liu, L.F. Pan, L. Zhang, H.G. Yang, Mn3O4 nano-
[58] X. Chia, A. Ambrosi, D. Sedmidubský, Z. Sofer, M. Pumera, Precise tuning of the octahedrons on Ni foam as an efficient three-dimensional oxygen evolution
charge transfer kinetics and catalytic properties of MoS2 materials via electrocatalyst, J. Mater. Chem. A 3 (2015) 14101–14104, https://doi.org/
electrochemical methods, Chem. Eur. J. 20 (2014) 17426–17432, https://doi.org/ 10.1039/c5ta02988k.
10.1002/chem.201404832. [65] W. Zhou, X.J. Wu, X. Cao, X. Huang, C. Tan, J. Tian, H. Liu, J. Wang, H. Zhang,
[59] B.B. Xiao, P. Zhang, L.P. Han, Z. Wen, Functional MoS2 by the Co/Ni doping as the Ni3S2 nanorods/Ni foam composite electrode with low overpotential for
catalyst for oxygen reduction reaction, Appl. Surf. Sci. 354 (2015) 221–228, electrocatalytic oxygen evolution, Energy Environ. Sci. 6 (2013) 2921–2924,
https://doi.org/10.1016/j.apsusc.2014.12.134. https://doi.org/10.1039/c3ee41572d.
[60] Z.H. Ibupoto, A. Tahira, P.Y. Tang, X. Liu, J.R. Morante, M. Fahlman, J. Arbiol, [66] C. Yang, M.Y. Gao, Q.B. Zhang, J.R. Zeng, X.T. Li, A.P. Abbott, In-situ activation of
M. Vagin, A. Vomiero, MoSx@NiO composite nanostructures: an advanced self-supported 3D hierarchically porous Ni3S2 films grown on nanoporous copper
nonprecious catalyst for hydrogen evolution reaction in alkaline media, Adv. as excellent pH-universal electrocatalysts for hydrogen evolution reaction, Nano
Funct. Mater. 29 (2019) 1–10, https://doi.org/10.1002/adfm.201807562. Energy 36 (2017) 85–94, https://doi.org/10.1016/j.nanoen.2017.04.032.
[61] Y.-F. Xu, M.-R. Gao, Y.-R. Zheng, J. Jiang, S.-H. Yu, Nickel/Nickel(II) oxide [67] J.S. Moon, J.H. Jang, E.G. Kim, Y.H. Chung, S.J. Yoo, Y.K. Lee, The nature of active
nanoparticles anchored onto cobalt(IV) diselenide nanobelts for the sites of Ni2P electrocatalyst for hydrogen evolution reaction, J. Catal. 326 (2015)
electrochemical production of hydrogen, Angew. Chem. Int. Ed. 125 (2013) 92–99, https://doi.org/10.1016/j.jcat.2015.03.012.
8708–8712, https://doi.org/10.1002/ange.201303495. [68] J. Su, Y. Yang, G. Xia, J. Chen, P. Jiang, Q. Chen, Ruthenium-cobalt nanoalloys
[62] M. Gong, W. Zhou, M.C. Tsai, J. Zhou, M. Guan, M.C. Lin, B. Zhang, Y. Hu, D. encapsulated in nitrogen-doped graphene as active electrocatalysts for producing
Y. Wang, J. Yang, S.J. Pennycook, B.J. Hwang, H. Dai, Nanoscale nickel oxide/ hydrogen in alkaline media, Nat. Commun. 8 (2017) 1–10, https://doi.org/
nickel heterostructures for active hydrogen evolution electrocatalysis, Nat. 10.1038/ncomms14969.
Commun. 5 (2014) 1–6, https://doi.org/10.1038/ncomms5695.

15

You might also like