You are on page 1of 9

Research Article

pubs.acs.org/acscatalysis

Sulfur and Nitrogen Dual-Doped Molybdenum Phosphide


Nanocrystallites as an Active and Stable Hydrogen Evolution
Reaction Electrocatalyst in Acidic and Alkaline Media
Mohsin Ali Raza Anjum and Jae Sung Lee*
School of Energy and Chemical Engineering, Ulsan National Institute of Science and Technology (UNIST), 50 UNIST-gil, Ulsan
44919, South Korea
*
S Supporting Information
Downloaded via INDIAN INST OF TECH KHARAGPUR on September 27, 2022 at 17:00:08 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Sulfur and nitrogen dual-doped molybdenum


phosphides (MoP/SN) are synthesized via a (thio)urea-
phosphate-assisted strategy in which the reductant (thio)urea
acts as S and N source and phosphoric acid provides the P
atom. The MoP/SN nanoparticles are generated by in situ
phosphidation of indigenously synthesized ammonium phos-
phate-coated P-doped MoSx nanoparticles in a hydrogen
atmosphere. Then, MoP/SN is anchored on graphene to
obtain a hybrid electrocatalyst (MoP/SNG) that exhibits high
activity and stability for electrochemical hydrogen evolution
from water in both acidic and basic electrolytes, outperforming most MoP-based electrocatalysts reported in the literature. The
dual doping and hybridization with graphene enhance electron conductivity of MoP and stabilize small MoP nanoparticles to
increase activity and stability, especially in acid electrolytes.
KEYWORDS: molybdenum phosphide, S and N dual doping, hydrogen evolution reaction, electrocatalysts, urea-phosphate route

T he electrochemical hydrogen evolution reaction (HER)


from water splitting has attracted significant attention
recently for producing sustainable hydrogen with electricity
and protecting P3− in TMP, the least stable oxidation state of P,
from oxidation.
Introduction of more electronegative P atoms into metals
generated from renewable energy sources. The biggest may greatly restrict the electron delocalization in the metal,
challenge in HER research is to replace the most common resulting in lower conductivity.16 However, with an appropriate
and best incumbent Pt catalysts with inexpensive nonprecious atomic ratio of metal and P or doping of other heteroatoms
metals.1 Transition metal carbides, sulfides, borides, nitrides, such as S or N, TMPs can exhibit a metallic character or even
and phosphides have been extensively studied as candidates for superconductivity, especially for the metal-rich phosphides.17
non-Pt electrocatalysts.1−5 Transition metal phosphides Recently, similar strategies have been applied to enhance the
(TMPs) have attracted significant attention over the past few HER activity of MoP: (i) doping with S to form molybdenum
years due to their superior electrical conductivity, mechanical phosphosulfide (MoP|S) on the surface of MoP by a
strength, and chemical stability relative to those of other postsulfidation8 or postphosphidation of MoS2 to form
transition metal compounds.4−6 They are proven to be high- MoS2(1−x)Px solid solution,18 (ii) compounding with carbon
performance electrocatalysts with excellent activity, stability, materials like graphene19 or porous carbons,20 and (iii)
and nearly 100% Faradaic efficiency in acidic, alkaline, and promotion with a TM (Co, Fe, or W).21−24 The TMP
neutral media for HER.6−11 Experimental and theoretical nanostructures are generally synthesized via three common
investigations have revealed that the atomic percentage of P ways: (i) Solution-phase synthesis using organic phosphine like
atoms in the lattice of transition metals (Fe, Co, Ni, Cu, Mo, tri-n-octylphosphine (TOP), triphenylphosphine (TPP) or tri-
and W) plays a crucial role with the more electronegative P n-octylphosphine oxide (TOPO) as a P source in high-boiling-
atom acting as a basic trap of positively charged protons during point solvents (e.g., oleylamine) in an inert atmosphere.17,25
the reaction.6,12−14 Furthermore, an appropriate atomic ratio of (ii) Gas−solid reaction, in which extremely toxic and lethal PH3
metals and P, especially in metal-rich phosphides, offers gas is used as P source directly or produced in situ from
excellent conductivity and more noble metal-like properties hypophosphite;26 in this method, post-treatment is mandatory
relative to those of the parent transition metals.15 Despite the with inert gas to remove residual PH3. (iii) High-temperature
success of TMPs as good electrocatalysts for HER, there are reduction of metal phosphates to form bulk TMP.17
still many remaining challenges to improve their performance
and stability further by tuning their electronic structures, Received: February 19, 2017
maximizing electrical conductivity, doping with other elements, Published: March 16, 2017

© 2017 American Chemical Society 3030 DOI: 10.1021/acscatal.7b00555


ACS Catal. 2017, 7, 3030−3038
ACS Catalysis Research Article

Herein, a simple, economical, and eco-friendly (thio)urea- Scheme 1. Schematic Illustration of (Thio)urea-Phosphate-
phosphate-assisted strategy is demonstrated to synthesize S and Assisted Synthesis of MoP/SN and MoP/SNG
N codoped MoP (MoP/SN) nanocrystallites, which display
excellent electrocatalytic activity and stability for HER in both
acidic and alkaline media. The MoP nanocrystallites show their
best performance when they are grown on S and N dual-doped
graphene (MoP/SNG), displaying one of the best HER
activities among reported non-noble metal electrocatalysts
and still maintaining excellent stability in aqueous acidic and
alkaline solutions. In particular, the stability in acids makes
them a promising practical electrocatalyst because there are
only a few acid-stable nonprecious metal electrocatalysts known
to date.
Distinct from earlier studies, our synthesis relies on an in situ
sulfidation reaction between polyoxometalates (POMs) and
(thio)urea-phosphates to form ammonium phosphate-coated,
reduced P-doped MoSx, followed by the reductive phosphida-
hydrothermal step to obtain MoP/SN NPs supported on
tion in a H2 atmosphere. To the best of our knowledge, this is
graphene (MoP/SNG).
the first systematic report on the fabrication of MoP
H+
electrocatalysts by using phosphoric acid as a P source and [PMo12O40]3 − + (NH 2)2 CS ⎯→
⎯ [PMo12O40]6 − + H 2S + (NH 2)2 CO
reductant (thio)urea as S and N sources while controlling the (1)
phase and size of the nanoparticles. In addition, this method
Δ
induces self-doping of N and S atoms into the skeleton of [PMo12O40]6 − + x H 2S → P − MoSx + y H 2O (2)
phosphides as well as a graphene support that plays crucial roles Δ
in stabilizing the MoP nanoparticles (NPs) and P3− state by (NH 2)2 CO + H3PO4 → (NH 2)2 ·HPO4 (amm. phosphate) + gases
making Mo/P−N and Mo/P−S bonds to increase the electrical (3)
conductivity and provide ample S and N sites for proton These phase changes during the synthesis steps were
adsorption. Avoiding the use of expensive, toxic, and corrosive confirmed by powder X-ray diffraction (XRD) patterns of
chemicals like PH3, hypophosphites, organic phosphine, and MoP/SN electrocatalysts at different reduction temperatures
high-boiling organic solvents is another advantage of this under H2 as shown in Figure 1a. Pure MoP of a hexagonal WC-
synthesis method of MoP nanoparticles. The present synthetic type structure (space group P6̅m2) is obtained above 600 °C,
strategy could also be applied to other transition metal (e.g., and small peaks of molybdenum phosphate are observed in the
Co, Fe, Ni, and W) phosphides for large scale production due samples reduced below 600 °C. Doping of N and S does not
to its simplicity and the use of cheap and environmentally significantly alter the bulk crystal structure of MoP. The
benign precursors.


crystallite size of MoP was obtained by applying the Scherrer
equation to X-ray line broadening, which increases along with
RESULTS AND DISCUSSION the reduction temperature as shown in Figure 1b. The XRD
Urea-Phosphate-Assisted Synthesis for Molybdenum pattern of P-doped MoSx shows P−Mo3S4 as the main phase
Phosphide. Polyoxometalates (POMs) are early transition (Figure S1a), and MoP/N-650 and graphene-supported
metal (Mo, W) polyatomic anionic clusters composed of one catalysts show the same phase behavior as the unsupported
metal oxide and a main group oxyanion (phosphate or silicates) MoP/SN catalysts as shown in Figure S1b and c.
and are good oxidants and versatile inorganic building blocks Scanning electron microscopy (SEM) images of the as-
for the construction of various hybrid materials.27−29 Thiourea synthesized MoP/SN-650 °C and MoP/SNG-20 (with 20 wt %
is a reducing agent that liberates H2S gas and urea, especially in graphene) catalysts show that particles form a closely
the presence of acids (H3PO4 and POM).29 These properties interconnected porous network (Figure S2). No difference in
led us to design a procedure to synthesize S and N codoped morphology is observed after loading the SN-doped MoP
MoP nanocrystallites (MoP/SN) by simply using phosphomo- nanoparticles on graphene (MoP-SN/G-20, Figure S4a and b).
lybdic acid (H3PMo12O40·nH2O) to produce ammonium The surface morphologies of MoP/N-650 (N-doping only) and
phosphate-coated, P-doped MoSx (precursor) via hydrothermal a physical mixture of MoP-SN/G-20 are also shown in Figures
reduction with thiourea followed by hydrogen reduction of the S3 and S4a and b. Elemental mappings by energy-dispersive X-
precursor as illustrated in Scheme 1 and described in detail in ray spectroscopy (EDX)-SEM images of MoP/SN-650 (Figure
the Experimental Section. A Keggin-type H3PMo12O40 is S5) and MoP/SNG-20 (Figure S6) clearly indicate uniform
reduced with thiourea (eq 1) to form a bluish mixed-valence elemental distributions throughout the particles. (EDS)-STEM
reduced species [PMo12O40]6− without the loss of the Keggin mappings of MoP/SNG-20 and MoP/SN-650 in Figures S7a
structure.30 During continuous boiling under autogenous and S8 also demonstrate similar uniform elemental distribu-
pressure, [PMo12O40]6− turns to P-doped MoSx nanoparticles tions. Thus, MoP NPs are homogeneously distributed over
(NPs) by reduction with H2S gas liberated from thiourea in the graphene and uniformly doped with S, N, and C. EDS-STEM
acidic solution (eq 2) with urea-phosphate as a side product. of MoP/SNG-20 (Figure S7b) clearly indicates that there is no
Ammonium o-polyphosphates and biuret encapsulate the P- significant change in the elemental composition after the HER
doped MoSx NPs upon pyrolysis of urea-phosphate (eq 3).31 durability test.
Then, ammonium phosphate-encaged P-doped MoSx NPs are The nature of chemical bonding in as-prepared MoP/SN and
converted to MoP/SN upon reduction by H2. An exfoliated MoP/SNG-20 was investigated by X-ray photoelectron spec-
graphene oxide (GO) suspension is introduced directly into the troscopy (XPS) in Figure 2. The Mo 3d region (Figure 2a) for
3031 DOI: 10.1021/acscatal.7b00555
ACS Catal. 2017, 7, 3030−3038
ACS Catalysis Research Article

Figure 1. XRD patterns of MoP synthesized at different reduction temperatures (a) and their particle sizes calculated by applying the Scherrer
equation to each diffraction peak (b).

Figure 2. High-resolution XPS of MoP/SN-650 and MoP/SNG-20: (a) Mo 3d, (b) P 2p, (c) N 1s, (d) S 2p for MoP/SNG-20, and (e) XPS survey
spectra.

both MoP/SN and MoP/SNG indicates a prominent doublet combined peak of -C−S/-C−SO at 164 eV and a peak of -C−
at 228.2 and 231.3 eV of Moδ+ (0<δ ≤ 4), which can be SO2 at 168.6 eV confirm the successful doping of S2− species
assigned to Mo of MoP.6,8,18 A doublet of Mo4+3d5/2 and into the graphene structure.37,38 Thus, graphene itself is doped
Mo4+3d3/2 with peaks18 at 229.2 and 232.4 eV reveals the with S and N, probably associated with dopant atoms around
successful insertion of the S atom into the structure of MoP, defects such as −C−S-C- and -CS- structures.38 In XPS C 1s
and almost all molybdenum sulfide is converted to MoP during spectra (Figure S10b and c), the main peak of graphitic sp2
the in situ reductive phosphidation reaction. A weak signal of carbon is centered at 284.6 eV, whereas an additional
surface oxides can also be observed at 235.8/232.5 eV,6 which component at 285.6 eV is assigned to C−N and/or CN.
can be assigned to the high oxidation state of Mo (MoO3). The absence of signal at 287.9 eV, typically O−CO, suggests
These surface oxides, invisible in XRD patterns (Figure 1a), can that all oxygenated groups on GO have been removed during
be removed by chemical and electrochemical treatment.32,33 An the reduction reactions. The XPS N 1s spectrum (Figure 2c)
XPS depth profiling of the samples was further performed in could be fitted to Mo−N (394.0 eV), MoS2 (395.3 eV),
Figure S9, which reveals that the oxygen is present only at the pyridinic (398.3), pyrrolic (399.5 eV), graphitic (401.2 eV),
surface of the catalysts. The content of oxygen is greatly and a broad peak of -SCN (397.8 eV).39−41 From the above
reduced with etching time as depicted in Figure S9a and d. XPS study, it is confirmed that more electronegative dopants
The doublet at 129.4 and 130.2 eV in the P 2p (Figure 2b) (S, N, and C) stabilize the P3− state through P−S, P−N/PON,
and one doublet (161.8 and 163.0 eV) in the S 2p (Figure 2d and P−C bonds by withdrawing electron density from P and
and Figure S10a) spectra indicate a similar oxidation state and donating back to vacant d orbitals of Mo, which provides
elemental environment of P and S atoms in MoP/SN and superior conductivity and enhanced stability as shown later.
MoP/SNG-20.8,18 The former confirms the predominant P3+ The morphology of MoP particles was further probed by
species bonded to Mo, and the latter arises from a sulfide high-resolution transmission electron microscopy (HR-TEM).
species (S2−) in both samples.8 Four additional peaks of P−S, Panels a and b in Figure 3 clearly show rod-shaped
P−C, P−O, and P−N/PON at 131.30, 132.8, 133.8, and 134.6 nanocrystallites in the samples MoP/SN-650, MoP/SNG-5,
eV, respectively, are observed in the P 2p spectra, indicating and a physical mixture of MoP-SN/G-20 (Figure S4c and d)
that a P species is likely to be stabilized by sharing electrons embedded in the carbon matrix. These nanorods are converted
with more electronegative S, C, O, and N atoms.8,20,34−36 to nanoparticles (3−7 nm) when an increased amount of GO is
Similarly, in the S 2p spectrum of MoP/SNG-20 (Figure 2d), a added during the synthesis reaction as depicted in Figure 3c
3032 DOI: 10.1021/acscatal.7b00555
ACS Catal. 2017, 7, 3030−3038
ACS Catalysis Research Article

Figure 3. TEM images of (a) MoP/SN-650, (b) MoP/SNG-5, (c) MoP/SNG-20 (inset: particle size distribution), (d) MoP/SNG-30, (e, f) HR-
TEM images of MoP with (100) and (101) planes and corresponding fast Fourier transformation (FFT) patterns, and (g, h) structures of MoP and
MoP/SN.

and d. The lattice spacings of 0.27 and 0.21 nm and codoped MoP/SN-650 sample contains ∼7.9 wt % N and ∼8.5
corresponding fast Fourier transforms (FFT) are assigned to wt % S dopants confirmed by XPS depth profiling in Figure S9f.
(100) and (101) planes for hexagonal MoP in Figure 3e and f. The coupling of N with the S dopants into the structure of
No specific change was observed in lattice parameters of MoP MoP plays a critical role in the high HER activity by
after insertion of S and N atoms in the crystal structure as coordinating with least stable P3−, enhancing electrical
confirmed by the XRD pattern (Figure 1 and Figure S1b and conductivity and generating more adsorption sites for protons.
c). For comparison, N-doped MoP (MoP/N-650) nanoparticles
Electrochemical Hydrogen Evolution Reaction from are also prepared by using urea-phosphate instead of thiourea-
Water in Acidic and Alkaline Media. The HER activities of phosphate by adopting the same procedure. In Figure S11a, η10
MoP/SN electrocatalysts prepared at different reduction and η20 are 164 and 185 mV in acidic media and 154 and 182
temperatures were evaluated in 0.5 M H2SO4 and 1.0 M mV in alkaline media, respectively, highlighting the synergistic
KOH aqueous solutions using a three-electrode electrochemical effect of S and N codoping. The higher HER performances of S
cell (iR-corrected, Figure 4a). As expected, MoP/SN-550 shows and N codoped MoP NPs in acidic media than those of single
little HER activity, whereas performance is greatly enhanced N-doped MoP/N-650 and earlier reported MoS2(1−x)Px solid
with complete reduction of phosphates into MoP (Figure 1a) at solution (η10 ≈ 150 mV)18 clearly indicates the dominant effect
higher temperatures. In particular, MoP/SN-650 shows the of dual doping.
highest activity in 0.5 M H2SO4 with the lowest onset potential The performance of MoP electrocatalysts can be further
of ∼44 mV and overpotentials of 57 and 104 mV to produce enhanced by combining them with a conductive graphene
current densities of 1 and 10 mA/cm2 (η1 and η10), respectively. substrate. Thus, MoP-SN nanoparticles are directly grown on
Interestingly, the HER activity in alkaline solution (Figure 4a) graphene by adding graphene oxide (GO) in the hydrothermal
reaches its maximum at reduction temperatures of 575−650 reactor. During the synthesis process, graphene is also codoped
°C. The optimal reduction temperature of 650 °C is low with S and N as indicated by XPS analysis in Figure 2. The
enough to allow the formation of the tiny NPs (Figure 1b) catalysts are denoted as MoP/SNG-x (x = 5, 10, 20, and 30),
while being high enough to remove all minor surface oxides and where x denotes the wt % of graphene. The performance curves
phosphates on the MoP NP surface.32,33 The overpotentials of in Figure 4b clearly show remarkable improvement in HER
η1 ≈ 17−20 mV and η10 ≈ 90−94 mV are needed to deliver activity when MoP NPs are grown on S and N codoped
current densities of 1 and 10 mA/cm2 for MoP/SN-575 to graphene giving η10 and η100 of 99 and 157 mV, respectively, for
MoP/SN-650. However, further increasing the reduction the MoP/SNG-20 catalyst in 0.5 M H2SO4. It is noted that this
temperature results in decreased HER activity due to sintering activity is substantially higher than those of previously reported
of MoP and removal of S and N species as exhaust gases.38 The MoP-based catalysts in terms of all activity parameters as
3033 DOI: 10.1021/acscatal.7b00555
ACS Catal. 2017, 7, 3030−3038
ACS Catalysis Research Article

Figure 4. HER iR-corrected polarization curves of (a) MoP/SN synthesized at different reduction temperatures and (b) MoP/SNG-x (x = 5, 10, 20,
and 30 wt %) in acid and base electrolytes. Overpotentials required for current densities of 10, 50, and 100 mA/cm2 in (c) 0.5 M H2SO4 and (d) 1.0
M KOH. (e) Tafel slopes in acid and base and (f) measured capacitive currents (ΔJ/2) plotted as a function of scan rate.

summarized in Tables S1 and S2. In 1.0 M KOH, MoP/SNG-x reactions follow the Volmer−Heyrovsky mechanism in acidic
hybrids exhibit excellent HER performances with the lowest media. In 1.0 M KOH, all samples except MoP/SNG-20 exhibit
onset potential close to that of commercial Pt/C and by Tafel slopes in the range of 39−59 mV/dec, suggesting a
reducing η10 to only 63, 52, 49, and 52 mV for catalysts having similar mechanism as in acids. The MoP/SNG-20 displays the
SNG of 5, 10, 20, and 30 wt % with similar mass loadings on highest activity with the most positive onset potential and
the disc electrode. A comparison of overpotentials at current lowest Tafel slope equal to the commercial Pt/C catalyst, i.e.,
densities of 10, 50, and 100 mA/cm2 is presented in Figure 4c ∼31 mV/dec The bare MoP/SN-650 without SNG has a
and d in acid and alkaline media, and MoP/SNG-20 shows the slightly lower activity and a higher Tafel slope than MoP/SNG,
best performance approaching that of Pt/C (Figure 4b). The which reveals that the hybrid structure formed between MoP
MoP-SN and graphene hybrid electrocatalyst was also prepared and SNG reduces the energy required to activate HER by
by reduction of the physical mixture of ammonium phosphate- increasing the conductivity, stabilizing NPs through Mo-P-X-
coated P-doped MoSx (a precursor of MoP/SN) with exfoliated
Mo (X = S, N, and C) bonding (Figure 3g,h) and providing
GO, denoted as MoP-SN/G. As depicted in Figure S11b, the
more proton (H+) adsorption sites. Evidence of such a
physical mixture with 20 wt % graphene (MoP-SN/G-20) does
synergistic coupling effect is already discussed with the
not show any significant enhancement of HER activity in acidic
media at low current densities, but the effect of graphene is electrochemical and spectroscopic data. All electrocatalysts
more pronounced in alkaline solution leading to reduction of display the high exchange current densities (J0), which are
η10 by 20 mV for MoP-SN/G-20. However, it is clear that the expected to be proportional to the catalytic active surface area
direct growth of MoP-SN NPs on graphene yields higher and represent high kinetics of charge transfer reactions over all
performing electrocatalysts than a physical mixture of electrodes. The higher J0 of MoP/SNG-20 (0.15 and 0.46 mA/
presynthesized MoP/SN precursor and graphene. cm2) than that of pristine MoP/SN-650 (0.056 and 0.48 mA/
For understanding the HER mechanism in 0.5 M H2SO4 and cm2) in acid and base, respectively, confirm that the HER
1.0 M KOH, Tafel analysis was conducted in Figure 4e, The kinetics become faster due to electronic and chemical
Tafel slopes of MoP/SN-650 (45 mV/dec), MoP/SNG-5 (51 promotional effects. The summarized electrochemical data in
mV/dec), MoP/SNG-10 (58 mV/dec), MoP/SNG-20 (54 acidic and alkaline media of as-synthesized catalysts are
mV/dec), and MoP/SNG-30 (54 mV/dec) suggest that HER tabulated in Tables S3 and S4.
3034 DOI: 10.1021/acscatal.7b00555
ACS Catal. 2017, 7, 3030−3038
ACS Catalysis Research Article

Figure 5. Electrochemical impedance spectroscopy (EIS) of catalysts at overpotentials of 120 and 125 mV in (a) 0.5 M H2SO4 and (b) 1.0 M KOH.
(c) Equivalent circuit (inset) used for fitting of EIS data. Rs is the overall series resistance; CPEd and Rd are the constant phase element and diffusion
resistance, respectively. CPEdl is the constant phase element, and Rct is the charge transfer resistance of the MoP/electrolyte interface.

Figure 6. Cathodic linear sweeping voltammograms of (a) MoP/SN-650 and (b) MoP/SNG-20 after 2000 CV cycles in 0.5 M H2SO4 and 1.0 M
KOH solutions. (c) Stability comparison after 1000 CV scans for HER and (d) after a 20 h durability test (by chronoamperometry as inset) for
MoP/SNG-20 in acidic and alkaline media.

Electrochemically active surface area (ECSA) is a useful −0.05 and 0.15 VRHE similar to Pt/C and MoP8 were observed
parameter for representing the intrinsic activity of the catalysts in the HER hysteresis for the cathodic- and anodic-going scans,
determined by measuring the double layer capacitance (Cdl) at especially for graphene hybrid electrocatalysts. The anodic
the electrolyte−electrode interface with cyclic voltammetry sweep of the MoP/SN-650 catalyst exhibits higher activity than
(CV). The detailed calculation procedure of ECSA is described the cathodic sweep in both acidic and alkaline media, which
in the Supporting Information. As shown in Figure 4f and may come from possible reduction of surface PO43− species
Table S3, Cdl and ECSA values increase with higher loading of observed by XPS generating more active states during the
SNG. The Cdl and ECSA of MoP/SN-650 (3.96 mF/cm2, 99 cathodic-going sweep.
cm2), MoP/SNG-5 (4.53 mF/cm2, 113.25 cm2), MoP/SNG-10 Electrochemical impedance spectroscopy (EIS) was applied
(8.89 mF/cm2, 222.25 cm2), MoP/SNG-20 (20.5 mF/cm2, for all electrocatalysts in acidic and basic media at different
512.5 cm2), and MoP/SNG-30 (18.58 mF/cm2, 464.5 cm2) overpotentials (Figure S13). The Nyquist plots of MoP/SN-
also confirm that the activity and exchange current density of 650 and MoP/SNG-x (x = 5, 10, 20, and 30) at a given
the electrocatalysts are directly related to ECSA (Table S3). overpotential (η) of 120 mVRHE in 0.5 M H2SO4 and at 125
The cyclic voltammetry (CV) values of MoP/SN-650, MoP/ mVRHE in 1.0 M KOH solutions are compared in Figure 5a and
SNG-10, and MoP/SNG-20 in acidic and alkaline media are b. The EIS Nyquist plots and their equivalent circuit model
compared in Figure S12. The broad redox features between (Figure 5c, inset) show that there are two time constants. The
3035 DOI: 10.1021/acscatal.7b00555
ACS Catal. 2017, 7, 3030−3038
ACS Catalysis Research Article

first one at high frequencies is related to the surface porosity surface area, and improve electron/proton conductivity. As a
(Rd) of the materials and the second one at low frequencies result, S and N doped MoP directly grown on graphene (MoP/
corresponds to the charge transfer (Rct) process at the cathode. SNG-20) demonstrates excellent activity and stability in
The relatively lower Rct and Rd values of MoP/SNG-x than bare electrochemical HER from water in both acidic and alkaline
MoP/SN-650 in all electrolytes could be mainly related to the electrolytes outperforming most MoP-based electrocatalysts
high conductivity of S and N codoped graphene and larger reported to date.
ECSA with better dispersion of smaller nanoparticles. The
diffusion resistance (Warburg impedance) for MoP/SN-650 is
larger than that for MoP/SNG hybrids in both electrolytes
■ EXPERIMENTAL SECTION
Synthesis of S and N Codoped MoP NPs and Their
because of its larger particle sizes.42 Ion diffusion in electrolyte Graphene Composite. All chemicals were utilized as
also becomes slower for small ECSA and large particles.42 The purchased without further purification. Thiourea-phosphate
results of EIS studies are consistent with the performances of was synthesized by mixing thiourea with phosphoric acid.
the electrocatalysts, indicating that charge transfer is the critical Typically, 85% phosphoric acid (5.5 mmol, 634.12 mg, Sigma-
factor determining the HER activity. Aldrich) was dissolved in triply distilled water (20 mL) in a
For the electrochemical durability of MoP-based electro- glass beaker (250 mL) equipped with a magnetic stirring bar,
catalysts to be tested, accelerated cyclic voltammetry (CV) was and an aqueous solution thiourea (5.0 mmol, 380.60 mg,
performed continuously for 2000 cycles at a scan rate of 100 Sigma) was then added. The reaction mixture was heated to 60
mV/s in acidic as well as alkaline media as displayed in Figure 6. °C until a clear solution was obtained and was maintained at
The MoP/SN-650 electrocatalyst exhibits excellent stability in this temperature for 1 h. A transparent yellow solution of
acidic electrolyte with a minor negative shift in overpotential phosphomolybdic acid (PMo12, 0.5 mmol, 913 mg, Sigma-
after the 50th cycle. On the other hand, it displays a small but Aldrich) in water was added slowly. The reaction mixture
continuous loss of activity in alkaline media, indicating became bluish green at once, an indication of the reduction of
continuous corrosion of this catalyst. Similarly, the time PMo12 from [PMo12O40]3− to [PMo12O40]6−. For the synthesis
dependence of current density (chronoamperometry, CA) for of ammonium phosphate-coated P-doped MoSx NPs, the as-
MoP/SNG-20 was also conducted for 20 h as a long-term prepared solution was hydrothermally treated at 180 °C for 15
durability test at a static overpotential in both media (inset of h. The reduced PMo12 was converted to P-doped MoSx (see
Figure 6d). As mentioned, the morphology and atomic XRD in Figure S1a) by internally produced H2S gas, and the
distribution in the electrocatalyst also remains unchanged remaining phosphoric acid reacted with indigenously generated
during the period as shown in EDS-STEM of MoP/SNG-20 ammonia (NH3) gas to form ammonium phosphate. Then, this
(Figure S7b). The performance and stability dramatically freshly prepared suspension of pH 7 was dried at 60 °C to be
increase in both acidic and alkaline electrolytes when these used for MoP synthesis.
NPs are anchored on S and N codoped graphene (MoP/SNG-x The as-prepared ammonium phosphate-coated P-doped
series) as shown in Figure 6b. CV curves (Figure S14) in the MoSx was reduced in a tube furnace under H2 gas with a
range −0.15 and 0.15 V represent the irreversible reduction of ramping rate of 5 °C/min, kept at a required temperature
surface phosphate (PO43−) species to P3−, and the current (550−700 °C) for 3 h, and then naturally cooled to obtain
difference between anodic and cathodic currents ΔJ (Ja − Jc) diamond blackish powders, denoted as MoP/SN-X (X denotes
becomes stabilized after the first 500 CV cycles. The stability of reduction temperature). For comparison, N-doped MoP
MoP/SN-650 with and without graphene was compared in nanoparticles are also prepared by using urea instead of
both electrolytes in Figure 6c. It shows that MoP-SN thiourea following the same procedure.
nanoparticles directly grown on the carbon substrate (MoP/ MoP NPs were grown on S and N codoped graphene (MoP/
SNG) are more stable than the physical mixture MoP-SN/G or SNG) by adding the required amount of exfoliated graphene
bare MoP/SN-650, indicating a close contact between MoP oxide (GO) suspension directly into the hydrothermal step and
and S and N doped graphene is important to take full advantage then reduced following the same conditions. For MoP-SN/G,
of the promotional effect of graphene. Because there are only a the ammonium phosphate-coated P-MoSx powder physically
few nonprecious metal electrocatalysts stable in acids, our mixed with exfoliated GO suspension (20 wt % G of MoP/SN-
MoP/SNG could be a promising candidate for HER in acidic 650) was then dried at 60 °C before reduction in a H2
electrolytes.


atmosphere.
Characterization. Crystallographic information of the as-
CONCLUSIONS synthesized electrocatalysts was investigated by powder X-ray
S and N codoped MoP has been synthesized, adopting a unique diffraction (XRD, PANalytical pw 3040/60 X’pert) with Cu Kα
(thio)urea-phosphate-assisted process in which thiourea acts as radiation. The surface morphologies and structural information
a reducing agent for phosphomolybdic acid as well as an S and were recorded with a field-emission scanning electron micro-
N source and phosphoric acid (H3PO4) as a P source. This scope (FE-SEM, Hitachi, S-4800, 15 kV) and a transmission
method relies on the reduction of in situ-produced ammonium electron microscope (TEM, JEOL, JEM-2100). The chemical
phosphate-coated P-doped MoSx nanoparticles to provide compositions, elemental mapping, and in-depth crystal
uniform doping of S and N at an atomic level into the MoP information on samples were collected by high-resolution
structure. It is also confirmed by XPS data that more transmission electron microscopy (HRTEM, JEOL, JEM-
electronegative dopants (S, N, and C) stabilize P3− through 2100F). X-ray photoelectron spectroscopy (XPS, Thermo-
P−S, P−N/PON, and P−C bonds by withdrawing electron Fisher, Kα) was used to identify the surface atomic composition
density on P and donating back to vacant d orbitals of Mo. The and chemical states.
HER performance and stability of the MoP/SN NPs is further Electrochemical Measurements. The catalyst ink was
improved by loading them onto conductive graphene to reduce prepared by dispersing electrocatalysts (1.0 mg/mL) in a
the particle size of MoP, increase the electrochemical active mixture of 7:3 deionized water/2-propanol and 10 μL of 5%
3036 DOI: 10.1021/acscatal.7b00555
ACS Catal. 2017, 7, 3030−3038
ACS Catalysis Research Article

Nafion solution. The mixture was sonicated for 1 h in a water (7) Tian, J.; Liu, Q.; Asiri, A. M.; Sun, X. J. Am. Chem. Soc. 2014, 136,
bath. The working electrode was prepared by drop casting of 10 7587−7590.
μL ink onto a glassy carbon electrode (0.4−0.5 mg cm−2 (8) Kibsgaard, J.; Jaramillo, T. F. Angew. Chem., Int. Ed. 2014, 53,
loading) followed by dropping 5 μL Nafion solution to fix the 14433−14437.
(9) Chen, Z.; Lv, C.; Chen, Z.; Jin, L.; Wang, J.; Huang, Z. Am. J.
electrocatalyst. The electrode was dried at room temperature
Anal. Chem. 2014, 5, 1200.
before electrochemical measurements. Electrochemical HER (10) Zhu, W.; Tang, C.; Liu, D.; Wang, J.; Asiri, A. M.; Sun, X. J.
activity and stability tests were carried out in a three electrode Mater. Chem. A 2016, 4, 7169−7173.
cell configuration using a rotating disc electrode (RDE, PAR (11) Jiang, P.; Liu, Q.; Liang, Y.; Tian, J.; Asiri, A. M.; Sun, X. Angew.
Model 636 RDE) attached with a potentiostat (Ivium Chem., Int. Ed. 2014, 53, 12855−12859.
Technologies). A Ag/AgCl (3.0 M NaCl) electrode and a Pt (12) Liu, P.; Rodriguez, J. A. J. Am. Chem. Soc. 2005, 127, 14871−
wire were used as reference and counter electrodes, 14878.
respectively. All potentials were referenced to the reversible (13) Pan, Y.; Liu, Y.; Zhao, J.; Yang, K.; Liang, J.; Liu, D.; Hu, W.;
hydrogen electrode (RHE) by the equation ERHE = E(Ag/AgCl) + Liu, D.; Liu, Y.; Liu, C. J. Mater. Chem. A 2015, 3, 1656−1665.
0.059 pH + 0.209. (14) Callejas, J. F.; Read, C. G.; Popczun, E. J.; McEnaney, J. M.;
The HER performance was measured in aqueous 0.5 M Schaak, R. E. Chem. Mater. 2015, 27, 3769−3774.
(15) Blanchard, P. E.; Grosvenor, A. P.; Cavell, R. G.; Mar, A. Chem.
H2SO4 (pH 0.3) and 1.0 M KOH (pH 14) at a scan rate of 5
Mater. 2008, 20, 7081−7088.
mV s−1 at 1600 rpm after 20 cyclic voltammetry (CV) cycles in (16) Carenco, S.; Portehault, D.; Boissiere, C.; Mezailles, N.;
the range of 0.4 to −0.3 VRHE. The electrochemical stability Sanchez, C. Chem. Rev. 2013, 113, 7981−8065.
tests were conducted by performing up to 2000 CV cycles in (17) Shi, Y.; Zhang, B. Chem. Soc. Rev. 2016, 45, 1529−1541.
the above potential ranges. Electrochemical impedance spec- (18) Ye, R.; del Angel-Vicente, P.; Liu, Y.; Arellano-Jimenez, M. J.;
troscopy (EIS) was conducted in the same setup in the Peng, Z.; Wang, T.; Li, Y.; Yakobson, B. I.; Wei, S. H.; Yacaman, M. J.
frequency range of 100 kHz to 1 mHz with a modulation Adv. Mater. 2016, 28, 1427−1432.
amplitude of 10 mV. The EIS spectra were fitted by Z-view (19) Aravind, S. J.; Ramanujachary, K.; Mugweru, A.; Vaden, T. D.
software. For the electrochemical active surface area (ECSA) to Appl. Catal., A 2015, 490, 101−107.
be evaluated, CV was conducted from 0.0 to 0.2 V (in 0.5 M (20) Han, S.; Feng, Y.; Zhang, F.; Yang, C.; Yao, Z.; Zhao, W.; Qiu,
H2SO4) and −0.8 to −0.6 V (in 1.0 M KOH or NaOH) vs Ag/ F.; Yang, L.; Yao, Y.; Zhuang, X. Adv. Funct. Mater. 2015, 25, 3899−
AgCl with different sweep rates between 10 and 100 mV s−1. 3906.


(21) Wang, D.; Zhang, D.; Tang, C.; Zhou, P.; Wu, Z.; Fang, B.
Catal. Sci. Technol. 2016, 6, 1952−1956.
ASSOCIATED CONTENT (22) Liang, X.; Zhang, D.; Wu, Z.; Wang, D. Appl. Catal., A 2016,
*
S Supporting Information 524, 134−138.
The Supporting Information is available free of charge on the (23) Wang, X.-D.; Xu, Y.-F.; Rao, H.-S.; Xu, W.-J.; Chen, H.-Y.;
ACS Publications website at DOI: 10.1021/acscatal.7b00555. Zhang, W.-X.; Kuang, D.-B.; Su, C.-Y. Energy Environ. Sci. 2016, 9,
HER performance comparisons and curves, summary of 1468−1475.
(24) Wang, D.; Zhang, X.; Zhang, D.; Shen, Y.; Wu, Z. Appl. Catal., A
electrochemical data, XRD patterns, surface morpholo-
2016, 511, 11−15.
gies, SEM images, EDX-SEM images, EDS-STEM (25) McEnaney, J. M.; Crompton, J. C.; Callejas, J. F.; Popczun, E. J.;
images, XPS spectra depth profiles, ECSA calculations, Biacchi, A. J.; Lewis, N. S.; Schaak, R. E. Chem. Mater. 2014, 26, 4826−
cyclic voltammograms, EIS data, and CV curves (PDF) 4831.

■ AUTHOR INFORMATION
Corresponding Author
(26) Xing, Z.; Liu, Q.; Asiri, A. M.; Sun, X. Adv. Mater. 2014, 26,
5702−5707.
(27) Zhou, D.; Han, B. H. Adv. Funct. Mater. 2010, 20, 2717−2722.
(28) Chen, Y.-Y.; Zhang, Y.; Jiang, W.-J.; Zhang, X.; Dai, Z.; Wan, L.-
*E-mail: jlee1234@unist.ac.kr.
J.; Hu, J.-S. ACS Nano 2016, 10, 8851−8860.
Notes (29) Shyla, B.; Nagendrappa, G. Spectrochim. Acta, Part A 2011, 78,
The authors declare no competing financial interest.


497−502.
(30) Cotton, F. A.; Wilkinson, G.; Murillo, C. A.; Bochmann, M.
ACKNOWLEDGMENTS Advanced Inorganic Chemistry, 6th ed.; John Wiley and Sons, 1999.
This research was supported by the Basic Science Grant (NRF- (31) McCullough, J. F.; Sheridan, R. C.; Frederick, L. L. J. Agric. Food
2015R1A2A1A10054346), Climate Change Response project Chem. 1978, 26, 670−675.
(32) Wang, T.; Du, K.; Liu, W.; Zhu, Z.; Shao, Y.; Li, M. J. Mater.
(2015M1A2A2074663), the Korea Center for Artificial Photo-
Chem. A 2015, 3, 4368−4373.
synthesis (KCAP, No. 2009-0093880) funded by MSIP, and (33) Cannon, P. J. Inorg. Nucl. Chem. 1959, 11, 124−127.
Project No. 10050509 funded by MOTIE of the Republic of (34) Pelavin, M.; Hendrickson, D.; Hollander, J.; Jolly, W. J. Phys.
Korea.


Chem. 1970, 74, 1116−1121.
(35) Jiang, H.; Zhu, Y.; Su, Y.; Yao, Y.; Liu, Y.; Yang, X.; Li, C. J.
REFERENCES Mater. Chem. A 2015, 3, 12642−12645.
(1) Zou, X.; Zhang, Y. Chem. Soc. Rev. 2015, 44, 5148−5180. (36) Franke, R.; Chassé, T.; Streubel, P.; Meisel, A. J. Electron
(2) Zhong, Y.; Xia, X.; Shi, F.; Zhan, J.; Tu, J.; Fan, H. J. Adv. Sci. Spectrosc. Relat. Phenom. 1991, 56, 381−388.
2016, 3, 1500286. (37) Dong, H.; Liu, C.; Ye, H.; Hu, L.; Fugetsu, B.; Dai, W.; Cao, Y.;
(3) Faber, M. S.; Jin, S. Energy Environ. Sci. 2014, 7, 3519−3542. Qi, X.; Lu, H.; Zhang, X. Sci. Rep. 2015, 5, 17542.
(4) Wang, J.; Cui, W.; Liu, Q.; Xing, Z.; Asiri, A. M.; Sun, X. Adv. (38) Ito, Y.; Cong, W.; Fujita, T.; Tang, Z.; Chen, M. Angew. Chem.,
Mater. 2016, 28, 215−230. Int. Ed. 2015, 54, 2131−2136.
(5) Cui, W.; Cheng, N.; Liu, Q.; Ge, C.; Asiri, A. M.; Sun, X. ACS (39) Sanjinés, R.; Wiemer, C.; Almeida, J.; Levy, F. Thin Solid Films
Catal. 2014, 4, 2658−2661. 1996, 290, 334−338.
(6) Xiao, P.; Sk, M. A.; Thia, L.; Ge, X.; Lim, R. J.; Wang, J.-Y.; Lim, (40) Patterson, T. A.; Carver, J. C.; Leyden, D. E.; Hercules, D. M. J.
K. H.; Wang, X. Energy Environ. Sci. 2014, 7, 2624−2629. Phys. Chem. 1976, 80, 1700−1708.

3037 DOI: 10.1021/acscatal.7b00555


ACS Catal. 2017, 7, 3030−3038
ACS Catalysis Research Article

(41) Duan, J.; Chen, S.; Jaroniec, M.; Qiao, S. Z. ACS Nano 2015, 9,
931−940.
(42) Chen, C.-F.; Mukherjee, P. P. Phys. Chem. Chem. Phys. 2015, 17,
9812−9827.

3038 DOI: 10.1021/acscatal.7b00555


ACS Catal. 2017, 7, 3030−3038

You might also like