Benck 2012

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

Research Article

pubs.acs.org/acscatalysis

Amorphous Molybdenum Sulfide Catalysts for Electrochemical


Hydrogen Production: Insights into the Origin of their Catalytic
Activity
Jesse D. Benck, Zhebo Chen, Leah Y. Kuritzky, Arnold J. Forman, and Thomas F. Jaramillo*
Department of Chemical Engineering, Stanford University, 381 North-South Mall, Stanford, California 94305, United States
*
S Supporting Information

ABSTRACT: We present a scalable wet chemical synthesis for a


catalytically active nanostructured amorphous molybdenum sulfide
material. The catalyst film is one of the most active nonprecious metal
materials for electrochemical hydrogen evolution, drawing 10 mA/cm2 at
∼200 mV overpotential. To identify the active phase of the material, we
perform X-ray photoelectron spectroscopy after testing under a variety of
conditions. As deposited, the catalyst resembles amorphous MoS3, but
domains resembling MoS2 in composition and chemical state are created
under reaction conditions and may contribute to this material’s high
electrochemical activity. The activity scales with electrochemically active surface area, suggesting that the rough, nanostructured
catalyst morphology also contributes substantially to the film’s high activity. Electrochemical stability tests indicate that the
catalyst remains highly active throughout prolonged operation. The overpotential required to attain a current density of 10 mA/
cm2 increases by only 57 mV after 10 000 reductive potential cycles. Our enhanced understanding of this highly active
amorphous molybdenum sulfide hydrogen evolution catalyst may facilitate the development of economical electrochemical
hydrogen production systems.
KEYWORDS: electrocatalysis, molybdenum sulfide, hydrogen evolution, wet chemical synthesis, nanostructured film

■ INTRODUCTION
Hydrogen, produced today mainly from natural gas or coal, is
Molybdenum sulfide materials have shown high activity and
stability for the HER in acidic environments,24−31 but previous
an extremely important chemical utilized at the global scale for procedures for synthesizing these catalysts have typically
major industrial processes, including petroleum refining and involved ultrahigh-vacuum processing,32 high-temperature
ammonia synthesis.1,2 Hydrogen has also been proposed as a treatment,27,28 sulfidization using H2S gas,27,32 or electro-
future energy carrier that could be used to power electronic deposition,29 which could limit the range of potential
devices, vehicles, and homes.3,4 Developing methods to applications. Recent work has revealed scalable procedures for
economically produce hydrogen from renewable energy synthesizing highly active amorphous molybdenum sulfide
resources could lead to substantial reductions in fossil fuel catalysts.11,28,33,34 However, much remains to be learned about
consumption and lower global CO2 emissions. the properties of amorphous molybdenum sulfide and the
A great deal of attention has recently been devoted to origins of its catalytic activity.
producing hydrogen from water through sustainable electro- Herein, we report a facile wet chemical synthesis for a
chemical processes such as photoelectrochemical water splitting nanostructured amorphous molybdenum sulfide catalyst and
or electrolysis driven by an external renewable source of aim to understand the origin of its catalytic activity. The
electricity.4−7 The hydrogen evolution reaction (HER), 2H+ + synthesis technique is based on readily available precursors;
2e− → H2, is a fundamental component of water splitting. avoids the need for high-vacuum processing, high-temperature
Achieving high energetic efficiency for water splitting requires treatment, or a separate sulfidization step; and enables
the use of a catalyst to minimize the overpotential necessary to straightforward catalyst deposition onto a wide range of
drive the HER.5,8 substrates. Physical and chemical characterization suggests
The best known catalysts for the HER are precious metals, that the resulting catalyst is similar to other recently reported
such as platinum, ruthenium, and iridium, but the scarcity and amorphous MoS3 materials in structure and composition, but
high cost of these materials prohibit their wide-scale deploy- this material currently has the highest reported activity of any
ment.5,9−11 Nickel alloy catalysts show high activity for the molybdenum sulfide catalyst produced through a room
HER in alkaline electrolytes, but often degrade in acidic
solutions.12−19 The development of a scalable, environmentally Received: February 26, 2012
friendly synthesis for an inexpensive, highly active, acid-stable Revised: July 25, 2012
HER catalyst remains a major challenge.5,20−23 Published: August 10, 2012

© 2012 American Chemical Society 1916 dx.doi.org/10.1021/cs300451q | ACS Catal. 2012, 2, 1916−1923
ACS Catalysis Research Article

temperature wet chemical process. To understand the origin of spectroscopy (XPS) was performed using a Phi VersaProbe
this material’s activity, we investigated changes in the catalyst spectrometer with an Al Kα source, and binding energies were
morphology, composition, and activity during catalysis. These calibrated to the adventitious C 1s peak at 284.6 eV. Raman
results suggest that the high activity of this material likely arises spectra were collected using a Cobolt Blues 473 nm blue diode
from both the inherently favorable surface properties of the pumped solid state laser at ∼1 mW and an Andor charge-
active molybdenum sulfide phase and the rough, nano- coupled detector held at −50 °C with a thermoelectric cooling
structured catalyst film morphology. Having enhanced the plate.
understanding of this material’s properties, we propose Electrochemical Characterization. Electrochemical meas-
strategies for further improving its performance. urements were performed in a three-electrode electrochemical

■ EXPERIMENTAL SECTION
Catalyst Synthesis. The molybdenum sulfide catalyst was
cell in a rotating disk electrode (RDE) configuration with a Bio-
Logic potentiostat (VMP3). All tests were performed in 80−
100 mL of 0.5 M sulfuric acid (H2SO4) electrolyte prepared in
prepared via a room temperature wet chemical synthesis. All Millipore water (18 MΩ cm) with a Hg/Hg2SO4 in saturated
reagents were purchased and used as received. K2SO4 reference electrode (Hach) and a graphite rod counter
A molybdenum precursor solution was prepared by adding electrode (Ted Pella). The molybdenum sulfide catalyst film on
0.60 g of ammonium heptamolybdate (Mo7O24(NH4)6·4H2O, a glassy carbon disk was used as the working electrode. The
Sigma-Aldrich) to 12 mL of 0.2 M sulfuric acid (H2SO4, reversible hydrogen electrode (RHE) was calibrated to between
Aldrich) in Millipore water (18 MΩ cm). In a second −0.700 V and −0.706 V vs the Hg/Hg2SO4 reference electrode
container, a sulfur precursor solution was prepared by adding as measured using platinum mesh working and counter
0.075 g of sodium sulfide (Na2S, Alfa Aesar) to 12 mL of electrodes under a H2 gas purge. The potential scale was
Millipore water. These two solutions were mixed, yielding a 24 calibrated to RHE after each experiment to prevent
mL solution of 0.02 M ammonium heptamolybdate, 0.04 M contamination of the cell with platinum ions dissolved from
sodium sulfide, and 0.1 M sulfuric acid in water. Upon mixing the electrodes during this calibration step.
the solutions, a suspension of nanoparticles was immediately Prior to further characterization, the catalyst was electro-
formed. The suspension was centrifuged at 12 000 rpm (relative chemically activated by performing approximately three
centrifugal force 17 400) for 30 min to collect the nanoparticles. potential cycles between 0.10 and −0.25 V vs RHE at 20
After discarding the supernatant, the solid centrifuge pellet was mV/s. After the initial stabilization of the activity, the
rinsed once with 15 mL of isopropyl alcohol, and the rinsing performance of the catalyst for the hydrogen evolution reaction
liquid was discarded. The pellet was then redispersed in an was measured using a linear sweep voltammogram beginning at
additional 15 mL of isopropyl alcohol via sonication for 10 min the open circuit potential of ∼0.30 V vs RHE and ending at
to give the final catalyst solution used for drop-casting. −0.30 V vs RHE with a scan rate of 5 mV/s. Throughout these
The catalyst samples with typical mass loading used for the measurements, the electrolyte was purged with H2 gas, and the
scanning electron microscopy, X-ray photoelectron spectrosco- working electrode was rotated at 1600 rpm in the RDE to
py, Raman spectroscopy, and electrochemical testing were remove hydrogen gas bubbles formed at the catalyst surface.
prepared by drop-casting 10 μL of the redispersed suspension To measure electrochemical capacitance, the potential was
onto 5-mm-diameter glassy carbon disks (Sigradur G, HTW swept from 0.15 to 0.35 V and back to 0.15 V three times at
Hochtemperatur-Werkstoffe GmbH). The sample used for X- each of six different scan rates (10, 20, 40, 80, 160, and 320
ray diffraction analysis was prepared by drop-casting 2 mL of mV/s). The working electrode was held at each potential vertex
the redispersed suspension onto a 50 × 75 mm glass for 20 s before beginning the next sweep. These scans were
microscope slide. The sample used for transmission electron performed on the molybdenum sulfide catalyst while purging
microscopy was prepared by diluting the sonicated nanoparticle with N2 gas and rotating at 1600 rpm in the RDE. The rate of
suspension by 400× in isopropyl alcohol and drop-casting 2 μL rotation was observed to have no effect on the measured
of this mixture onto a holey carbon support (Ted Pella capacitive current. The same set of scans was also performed on
Catalogue no. 01824). A series of samples with mass loadings a flat standard MoS2 sample (preparation details provided in
ranging from 400% to 25% of the typical preparation were the Supporting Information) in a compression cell with a N2
synthesized by redispersing the solid pellet formed after gas purge and no rotation.
centrifugation in 3.75 mL of isopropyl alcohol and using a Extended potential cycling was performed to investigate
1:1 dilution series in isopropyl alcohol to make five solutions changes in composition during catalysis and to evaluate the
with 400%, 200%, 100%, 50%, and 25% of the typical catalyst catalyst’s durability. These tests were performed by taking linear
concentration. A 10 μL aliquot of each solution was then sweep voltammograms after repeatedly cycling the potential
deposited onto a 5 mm glassy carbon disk. All samples were between 0.10 and −0.25 V vs RHE at 50 mV/s. The first 20
dried under low vacuum (∼25 Torr) for several hours prior to potential cycles were performed at 0 rpm, and the next 5 cycles
characterization. were performed at 1600 rpm. This pattern was repeated three
Physical and Chemical Characterization. Scanning additional times. Performing the majority of the cycles at 0 rpm
electron microscopy (SEM) was performed before and after prevented the RDE apparatus from overheating due to
electrochemical characterization using an FEI Magellan XHR prolonged continuous rotation, and periodic rotation still
microscope operated with a beam voltage of 5.0 kV and current enabled removal of H2 bubbles. After every 100 potential cycles,
of 25 pA. Transmission electron microscopy (TEM) of the the potential was swept from 0.10 to −0.30 V vs RHE at 5
catalyst was collected using an FEI Titan 80-300 microscope mV/s and 1600 rpm. The slow scans were used to minimize
operated at 300 kV. contributions from capacitance current and obtain a more
X-ray diffraction (XRD) spectra were collected using a accurate measure of the electrochemical activity. For electro-
PANalytical X’Pert Pro diffractometer with a Cu Kα X-ray chemical stability testing, this program was repeated 100 times
source operated at 40 kV and 45 mA. X-ray photoelectron for 10 000 total potential cycles at 50 mV/s (approximately
1917 dx.doi.org/10.1021/cs300451q | ACS Catal. 2012, 2, 1916−1923
ACS Catalysis Research Article

48 h). The samples prepared for X-ray photoelectron of the catalyst particle is ∼60 nm, which corresponds to the size
spectroscopy measurements were cycled using this program of the features observed using SEM after the particles
for 4 h. A continual H2 gas purge was used for the duration of agglomerate into the rough, nanostructured film morphology.
these tests. As appropriate, these samples were rinsed in The absence of lattice or Moiré fringes observed in the bright
Millipore water to remove any residual sulfuric acid prior to field TEM image suggests that the material is amorphous,
spectroscopic measurements. which was further confirmed by the lack of any XRD peaks that

■ RESULTS AND DISCUSSION


The morphology of the catalyst film was characterized using
would suggest a crystalline catalyst film (see Supporting
Information Figure S1). Raman spectroscopy revealed the
absence of the distinctive A1g and E12g vibrational peaks expected
scanning electron microscopy, as shown in Figure 1. The film for crystalline MoS2 (see Supporting Information Figure S2),
which suggests that crystalline MoS2 is not present in significant
quantities in the as-deposited catalyst.37−39
X-ray photoelectron spectra collected before electrochemical
testing (Figure 3) show that the as-deposited (“no testing”)

Figure 1. SEM images showing (a) the catalyst film surface


morphology and (b) a thickness cross section of a segment of the
catalyst film that delaminated from the substrate surface.

has a rough, nanostructured surface with feature sizes ranging


from 50 to 100 nm. This structure arises from the
agglomeration of the catalyst nanoparticles upon drop-casting
and is favorable for catalysis because it provides a high surface
area interface between the catalyst and the electrolyte, which is
conducive to high overall electrochemical activity. An image of
a portion of the catalyst film that delaminated from the
substrate shows that the film is ∼5−7 μm thick, assuming an
observation angle between 0° and 45° with respect to the film
surface. The amount of catalyst on the substrate is greater than
the typical loading for precious metal electrocatalysts, which are
often deposited as isolated nanoparticles rather than as a
porous film.5,35,36 However, due to the low material cost and
straightforward synthetic procedure of the molybdenum sulfide
catalyst, this film structure could potentially be practical for
Figure 3. X-ray photoelectron spectra of the catalyst surface before
wide-scale deployment. testing, after one reductive potential cycle, and after 4 h of continuous
The crystallinity of the film was investigated before reductive cycling. The data are scaled to normalize the Mo 3d peak
electrochemical characterization using TEM and XRD. TEM areas. Before electrochemical testing, the shape and binding energy of
of an isolated nanoparticle (Figure 2) shows that the diameter the S 2p peak are indicative of amorphous MoS3. The Mo 3d peaks
show that the majority of the molybdenum occurs in chemical state
Mo(A)Sx. The peak positions and shapes begin to change after the
initial reductive potential cycle, corresponding to the catalyst activation
observed during electrochemical testing. These changes are accen-
tuated after extended electrochemical testing. In the tested samples,
the S 2p peak shape and binding energy more closely resemble MoS2,
and the majority of the molybdenum occurs in chemical state Mo(B)Sx,
with binding energies near the Mo 4+ peaks observed in MoS2.

material resembles molybdenum trisulfide (MoS3). The sulfur


2p region shows a peak at a binding energy of 163.0 eV, with a
broad full width at half-maximum (fwhm) of 2.2 eV and no
evidence of the expected spin−orbit splitting doublet
corresponding to the 2p1/2 and 2p3/2 lines, which suggests
that the sulfur atoms near the surface exist in multiple oxidation
states. These observations are consistent with previous reports
of MoS3 materials, which contain sulfur in a combination of
Figure 2. Transmission electron micrograph of an isolated particle S22− and S2− groups.11,29,34,40,41 A small peak observed near 169
deposited from a diluted molybdenum sulfide catalyst suspension. eV corresponds to the binding energy of sulfur in a sulfate
1918 dx.doi.org/10.1021/cs300451q | ACS Catal. 2012, 2, 1916−1923
ACS Catalysis Research Article

group42,43 and likely arises from residual SO42− from the active phase, the nature of which is discussed below. A similar
sulfuric acid used in the catalyst synthesis. The presence of reductive activation was also observed in other recent studies of
small nitrogen 1s and sodium 1s peaks indicates that there may amorphous molybdenum sulfide HER catalyst materials.29,34
also be some residual Na+ and NH4+ from the catalyst synthesis, A polarization curve demonstrating the HER activity of the
but these peaks disappear after the catalyst has been immersed molybdenum sulfide catalyst on a glassy carbon electrode is
in water or sulfuric acid (see Supporting Information Figure compared with a bare glassy carbon control in Figure 5a. The
S3).
The molybdenum 3d region reveals a mixture of Mo
oxidation states. Peak fitting shows that ∼ 27% of the Mo signal
corresponds to a 6+ oxidation state. XPS spectra of an untested
catalyst sample aged in air show that this feature arises primarily
from oxidation to form a thin layer of Mo(6+)O3 (see
Supporting Information Figure S4). The remaining 73% of
the Mo signal arises from the molybdenum sulfide. This Mo
exists in a chemical state labeled here as Mo(A)Sx, which likely
corresponds to a lower oxidation state of either 4+ or 5+. The
amorphous structure of this material prohibits the definitive
determination of the formal oxidation state, as the observed
binding energies are likely influenced by a number of factors,
including the oxidation state, sulfur coordination, and local
bonding structure. To quantify the Mo and S components of
the molybdenum sulfide film, the peak areas pertaining to the
Mo 3d and the sulfide/persulfide S 2p lines were measured and
calibrated versus a reference scan of a natural MoS2 crystal,
which served as a standard (see Supporting Information Figure
S5). This analysis reveals that the composition near the catalyst
surface (∼3 nm probe depth) is 22% Mo and 78% S as-
synthesized (disregarding the S pertaining to sulfate as well as
all other elements). These values closely match the 25% Mo
and 75% S expected for molybdenum trisulfide. Thus, the
atomic composition and binding energies measured by XPS and Figure 5. Electrochemical activity of the molybdenum sulfide catalyst.
the amorphous structure observed by TEM and XRD suggest Polarization curves (a) show that the catalyst exhibits high activity for
that the as-deposited material is predominantly composed of the HER. A Tafel plot (b) shows the electrochemical activity of the
amorphous MoS3.40,41,44−53 wet-chemical-synthesized amorphous molybdenum sulfide catalyst
We investigated the electrochemical properties of the along with digitized data of HER measurements of several other
molybdenum sulfide catalyst using a three-electrode electro- materials for comparison, including MoS2 nanotriangles in pH 0.24
chemical cell in a rotating disk electrode configuration. Catalyst H2SO4 (orange trace),24 MoO3/MoS2 core−shell nanowires in 0.5 M
H2SO4 (blue trace),27 electrodeposited amorphous MoS3 in pH 0
activation is observed upon the first reductive potential cycle electrolyte (green trace),29 wet chemical synthesized amorphous
(Figure 4). Initially, the onset of reductive current is observed MoS3/multiwalled carbon nanotube composite in 1 M H2SO4 (purple
near −200 mV vs RHE, but after completing the first potential trace),34 and MoS2 on reduced graphene oxide in 0.5 M H2SO4 (red
cycle, the onset of reductive current shifts to approximately trace).28
−150 mV vs RHE and remains stable in subsequent cycles. The
current observed on the first reductive cycle is likely a
combination of hydrogen evolution and electrochemical catalyst shows high activity for the HER with ∼200 mV
reduction of the catalyst material to form the catalytically overpotential necessary to achieve 10 mA/cm2 hydrogen
evolution current density, which is a useful metric for
comparing catalysts for solar hydrogen production.5,54 Meas-
urements performed on four duplicate samples show consistent
activity, with the overpotentials necessary to achieve 10 mA/
cm2 current density ranging from 198 to 204 mV (see
Supporting Information Figure S6).
The activity of the wet-chemical-synthesized amorphous
molybdenum sulfide catalyst is compared with several other
MoS2 and MoS3 materials in the Tafel plot shown in Figure
5b.27−29,32,34 This material shows the highest activity of any
reported molybdenum sulfide catalyst synthesized using a room
temperature wet chemical procedure. The activity is within ∼50
mV of the best reported molybdenum sulfide catalyst
synthesized using any technique.28 The average Tafel slope
Figure 4. Cyclic voltammograms indicate that the molybdenum sulfide for this catalyst is 60 mV/decade, with individual slopes ranging
catalyst is activated during the cathodic sweep of the first cycle (solid from 53 to 65 mV/decade for the catalyst films shown in
red line). Enhanced activity is observed on the anodic sweep of the Figures 5 and Supporting Information S6. Although the Tafel
first cycle (dotted red line) and in subsequent cycles (blue lines). slope alone is insufficient to determine the specific mechanism
1919 dx.doi.org/10.1021/cs300451q | ACS Catal. 2012, 2, 1916−1923
ACS Catalysis Research Article

of the HER,27,55,56 it does match several earlier reports for sample’s composition has changed from 22% Mo and 78% S
molybdenum sulfide catalysts, which also exhibited slopes of (consistent with molybdenum trisulfide) to a composition of
∼60 mV/decade.27,32 34% Mo and 66% S, closely matching the 33% Mo and 67% S
To understand the changes in the material during electro- expected in MoS2. After four hours of reductive cycling, the
chemical testing and identify the active form of the catalyst, we composition appears to have continued to change, albeit
performed XPS after catalyst activation via the first reductive slightly, to 39% Mo and 61% S. These values, which are also
potential cycle and after 4 h of continuous reductive cycling within experimental error of MoS2, could suggest a subtle
(Figure 3). XPS spectra collected after electrochemical testing depletion of sulfur during catalysis. All of these observations
reveal pronounced changes in the material, both in the regarding the quantification and chemical state of Mo and S
elemental ratio and in the chemical states of the Mo and S. suggest that the MoS3 material resulting from the wet chemical
The S 2p peak corresponding to the catalyst material after the synthesis could be considered a “precatalyst.” This material
initial activation cycle shifts from 163.0 to 162.4 eV and begins undergoes major changes during operation to create the active
to resemble a single spin−orbit splitting doublet. After 4 h of phase, which more closely resembles MoS2, consistent with the
reductive cycling, the S 2p region displays a distinct spin−orbit trends observed in previous reports of amorphous MoS3
splitting doublet with the S 2p3/2 peak at a binding energy of materials.29,34 However, Raman spectra collected after catalysis
162.2 eV, very close to the value observed in crystalline MoS2 (see Supporting Information Figure S2) confirm that the
(see Supporting Information Figure S5). The same can be said sample still does not contain crystalline MoS2 in a detectable
for the S 2s line, which shifts from 227.4 eV before testing to quantity. Therefore, the reduced phase created during catalysis
226.2 eV after testing, the binding energy observed in is likely an amorphous MoS2.
crystalline MoS2. This trend suggests that the sulfur in the Because of the amorphous and nanostructured nature of this
sample is partially reduced during catalysis and the dominant catalyst, there is likely significant surface site heterogeneity in
form of sulfur present after testing exists in a chemical this material. It is therefore difficult to accurately determine the
environment similar to MoS2. structure and relative concentration of various types of surface
Significant changes are also observed in the Mo 3d region sites, which makes the identification of the active sites of this
after electrochemical testing. First, the amount of Mo(6+)O3 has material particularly challenging. Previous work on many
decreased substantially to a negligible quantity that obviates the different types of molybdenum sulfide materials shows that
need for fitted features in the deconvoluted spectra of the “after this general class of catalyst can exhibit high activity for
testing” samples in Figure 3; thus, for the sake of clarity, those hydrogen evolution, despite wide variations in atomic structure
particular features are omitted. This significant decrease in and chemical composition.11,23,25,27−32,34,58 The catalyst ma-
Mo(6+)O3 was also observed in catalyst samples allowed to rest terial presented in this work likely exhibits many types of
in water or 0.5 M H2SO4 with no applied potential, so we surface sites due to its amorphous, nanostructured nature, some
attribute this difference to the chemical dissolution of the native of which could resemble the active sites previously described in
surface oxide (see Supporting Information Figure S7). This the studies cited above. It is therefore possible that many
observation is consistent with the expectation that any MoO3 different types of surface sites participate in the reaction. The
would dissolve in the electrolyte used during catalyst results of this study provide initial indications that domains
testing.27,57 The second noticeable change is a significant shift resembling MoS2 are created during catalysis and may
in the binding energies of the Mo peaks corresponding to the contribute to this material’s high electrochemical activity, but
molybdenum sulfide. Prior to testing, the majority of the further characterization is necessary to confirm this hypothesis.
molybdenum exists as Mo(A)Sx. After the initial activation and Synchrotron techniques such as X-ray absorption spectroscopy
after 4 h of reductive cycling, the majority of the near-surface could yield greater insights about the catalyst composition and
molybdenum is found in a slightly more reduced state labeled structure.11,49
here as Mo(B)Sx, with 3d3/2 and 3d5/2 peaks near 232.6 and The intrinsic per-site activity of a catalyst is an important
229.5 eV, respectively. Approximately one-third of the Mo metric necessary to compare catalyst materials and to guide
remains as Mo(A)Sx in both cases. A small amount of Mo(6+)O3 catalyst development. To estimate this key figure of merit, we
may be hidden within the Mo(A)Sx peaks, causing them to used electrochemical capacitance surface area measurements to
appear at a slightly higher binding energy than observed before determine the active surface area of the catalyst film.59,60 When
catalytic testing. The shift from Mo (A) S x to Mo (B) S x combined with a simplified model of the surface structure
corresponds to a change in chemical environment, and the (details provided in the Supporting Information), this
direction of this shift is consistent with a reduction of the technique enables independent estimation of the density of
molybdenum, as expected based on the reductive potentials electrochemically active sites and the average activity of each
applied during catalyst activation and HER catalysis. The Mo site, reported as a per-site turnover frequency (TOF). Although
3d3/2 and 3d5/2 binding energies in the Mo(B)Sx are close to the the analysis shown here utilizes known properties for MoS2,
values observed in crystalline MoS 2 (see Supporting which may not perfectly reflect MoS3 or the reduced phase
Information Figure S5), which has a Mo oxidation state of created during electrochemical testing, it nevertheless provides
4+. The small remaining difference in the binding energies may useful insights for future improvements to catalyst design.
arise from the different sulfur coordination or local bonding To find the electrochemically active surface area, as shown in
structure in this catalyst. All the changes in the molybdenum Figure 6a, we measured the non-faradic capacitive current
spectra measured after catalysis show that the active form of the associated with electrochemical double layer charging upon
catalyst resembles MoS2, although it is likely not a pure phase, repeated potential cycling. This double layer charging current,
since the Mo still exists in multiple chemical states. ic, is proportional to both the scan rate, ν, and the
Finally, quantification of the Mo and sulfide/persulfide S electrochemically active surface area of the electrode, Aechem:61
signals after catalysis reveals changes in the stoichiometry of the
near-surface region. After the initial reductive cycle, the ic ∝ ν × Aechem

1920 dx.doi.org/10.1021/cs300451q | ACS Catal. 2012, 2, 1916−1923


ACS Catalysis Research Article

capacitance associated with double layer charging is expected


to be similar (i.e., within an order of magnitude) for many
metallic and semiconducting materials in the same aqueous
electrolyte,62 the resulting surface area estimate for our work
presented here is accurate to within an order of magnitude or
better.
The RF was used to estimate the density of electrochemically
accessible sites on the catalyst surface, ∼1017/cm2 of geometric
electrode area (calculation details are provided in the
Supporting Information). Using this surface site density, we
calculated a turnover frequency (TOF) of 0.3 H2/s per surface
site at 200 mV overpotential (∼10 mA/cm2 of current density,
calculated per geometric electrode area). Conservatively,
because of the order of magnitude inaccuracy of these
calculations, the TOF may fall between 0.03 H2/s and 3 H2/
s per surface site. This range is comparable to the TOFs
reported for other highly active molybdenum sulfide catalysts at
similar current densities.27,29 However, previous reports of well-
defined MoS2 nanoparticles indicate that the intrinsic TOF for
a crystalline MoS2 edge site is ∼3 orders of magnitude
higher.27,32 As discussed previously, there is likely significant
surface site heterogeneity in this amorphous molybdenum
sulfide material, which raises the possibility that only a fraction
of the surface sites may be active for hydrogen evolution.
Figure 6. Electrochemical capacitance measurements for determi- Therefore, the true TOF of the most active sites could be
nation of the molybdenum sulfide catalyst surface area. Cyclic orders of magnitude greater than the average value of 0.3 H2/s.
voltammograms (a) were taken in a potential range where no faradic In any event, the high surface area nanostructured
processes were observed to measure the capacitive current from
double layer charging. The capacitive current measured at 0.30 V vs
morphology of the catalyst film clearly plays an important
RHE was plotted as a function of scan rate (b) for the wet-chemical- role in providing a significant number of active sites. To further
synthesized amorphous molybdenum sulfide and the MoS2 flat investigate the effect of electrochemically active surface area on
standard. The ratio of the capacitive currents for the molybdenum the observed activity, we synthesized samples with mass
sulfide catalyst and the flat standard was used to determine the relative loadings varying from 25% to 400% of the typical catalyst
roughness factor. preparation. As shown in Figure 7, we found that the activity of

The capacitive currents for the molybdenum sulfide catalyst


and for a flat MoS2 standard (preparation details provided in
the Supporting Information) were measured as a function of
scan rate, as shown in Figure 6b. A potential range of 0.15−0.35
V vs RHE was selected for the capacitance measurements
because no obvious electrochemical features corresponding to
faradic current were observed in this region. Furthermore, the
dependence of the current on the scan rate in this region is
linear for both materials, which is consistent with capacitive
charging behavior. Current arising from a faradic process would
yield a square root dependence with respect to scan rate as a
result of mass transfer (reactant diffusion) limitations.62 The
Figure 7. Electrochemical activity of catalyst samples as a function of
ratio of the currents for the catalyst and the flat MoS2 standard roughness factor. The catalyst activity increases linearly with the log of
was taken as the relative roughness factor, RF: the roughness factor.
ic(catalyst) catalyst active surface area
RF = = these samples, as measured by the overpotential needed to
ic(flat standard) substrate geometric surface area
reach 1 mA/cm2 on a geometric area basis, varies linearly with
From this analysis, we measured a RF value of 90. This log(RF). This shows that increasing surface area translates to
calculation relies on the assumption that the intrinsic, surface- improved activity in a predictable manner based on standard
area-normalized capacitance of the molybdenum sulfide catalyst models of electrochemical kinetics. Note, however, that the
and the flat MoS2 standard are the same. Although the chemical slope of the regression line, 38 mV/decade, is slightly lower
and physical characteristics of MoS2 differ from the amorphous than the expected value of 60 mV/decade based on the Tafel
molybdenum sulfide catalyst, we chose to use a MoS2 flat analysis described earlier. This difference may result from
standard due to the difficulty of fabricating a perfectly flat, well- limitations from mass transport or charge transport for thicker,
defined analogue of the amorphous molybdenum sulfide higher surface area films. Ultimately, the linearity of the data
catalyst. The capacitance measured for the flat MoS2, provides additional evidence that the high surface area
approximately 60 μF/cm2, is consistent with expectations for morphology of the catalyst film contributes significantly to its
a flat electrode.59,62 Because the surface-area-normalized high activity and may explain, in part, why this catalyst has
1921 dx.doi.org/10.1021/cs300451q | ACS Catal. 2012, 2, 1916−1923
ACS Catalysis


Research Article

activity superior to similar materials reported previously29,34 CONCLUSIONS


(additional discussion is presented in the Supporting We have developed a scalable wet chemical synthesis to
Information along with further electrochemical data in Figure produce a highly active, stable, amorphous molybdenum sulfide
S8 and SEM images in Figure S9). This result suggests that HER electrocatalyst. This synthetic technique requires no high-
high catalytic current densities could be achieved from a much temperature processing or secondary sulfidization step and
lower catalyst loading by depositing the molybdenum sulfide allows for direct catalyst deposition onto many types of
catalyst onto a high surface area support.27 Similarly, the substrates. We performed extensive spectroscopic and electro-
activity of this catalyst could be further improved by increasing chemical characterization to understand changes in this material
the roughness factor beyond 90. during catalysis and investigate the origin of the catalytic
The long-term stability of a catalyst is another important activity of amorphous molybdenum sulfide catalysts, for which
metric to consider for commercial applications. The stability of there have been several recent literature reports. Domains
this catalyst film was assessed by repeated potential cycling for resembling MoS2 in both composition and chemical state are
more than 10 000 cycles to replicate diurnal cycling experienced created during catalysis and may contribute to the high HER
by a HER catalyst for solar water splitting. The lower bound of activity of this material. The high density of active sites that
the potential cycles, −0.25 V vs RHE, was chosen to reach a results from the rough, nanostructured surface morphology also
current density in excess of 10 mA/cm2, which is a useful contributes to the high geometric current densities. The
metric for comparing catalysts for solar hydrogen production. catalyst’s stability was also ascertained; it was found to retain its
The results of the stability testing are displayed in Figure 8, activity throughout extended reductive potential cycling. The
catalyst activity and stability could be further improved by
designing a substrate structure to increase the total surface area
and prevent catalyst delamination during operation. This highly
active, stable HER catalyst is a promising candidate material
that could help to enable the widespread deployment of cost-
effective systems for electrochemical hydrogen production.


*
ASSOCIATED CONTENT
S Supporting Information
Experimental details, additional XPS spectra, XRD spectra,
Raman spectra, electrochemical activity of duplicate samples,
additional SEM images of the catalyst films, electrochemical
Figure 8. Electrochemical stability of the molybdenum sulfide catalyst.
activity variation with mass loading, and TOF calculation
The overpotential required to reach a current density of 10 mA/cm2 details. This material is available free of charge via the Internet
at http://pubs.acs.org.


increases by only 57 mV after 10 000 reductive potential cycles,
indicating that the catalyst remains highly active.
AUTHOR INFORMATION
which shows that the overpotential required to attain 10 mA/ Corresponding Author
cm2 current density increases by only 57 mV after 10 000 *Phone: 650-498-6879. E-mail: jaramillo@stanford.edu.
potential cycles. Although not quite as stable as some other Notes
molybdenum sulfide materials,27,28 the absolute overpotential The authors declare no competing financial interest.


required to drive 10 mA/cm2 remains low.
Several factors may contribute to the slight loss in activity.
ACKNOWLEDGMENTS
One hypothesis is that surface adsorbates, potentially from
impurities in the electrolyte, reference electrode, or within the This work was supported as part of the Center on
film itself, may poison the active sites of the catalyst over the Nanostructuring for Efficient Energy Conversion (CNEEC)
course of the stability testing. We tested this hypothesis by at Stanford University, an Energy Frontier Research Center
pausing the stability testing after the 10 000th cycle to replace funded by the U.S. Department of Energy, Office of Science,
the electrolyte solution, after which the catalyst immediately Office of Basic Energy Sciences under Award No. DE-
recovered 30 mV of activity during subsequent measurements. SC0001060. J.D.B. acknowledges support from the National
Delamination of the catalyst film from the substrate is another Science Foundation Graduate Research Fellowship Program
likely cause of activity loss, an effect confirmed by scanning and a Stanford Graduate Fellowship. A.J.F. was supported by
electron microscope imaging performed after electrochemical the U.S. Department of Energy, Office of Energy Efficiency and
testing (see Supporting Information Figure S10). This loss of Renewable Energy, under Subcontract NFT-9-88567-01 under
catalyst loading could potentially be remedied by using a prime Contract No. DE-AC36-08GO28308. The authors thank
conductive binder to secure the catalyst to the glassy carbon Benjamin Reinecke for performing the TEM imaging, Andrey
substrate or by depositing this material upon a different Malkovskiy for help with Raman spectroscopy, and Yelena
substrate altogether. Although further study is required to detail Gorlin for helpful discussions about electrochemical stability
the specific mechanisms of the catalyst degradation, our measurements.
preliminary efforts indicate that factors other than the inherent
properties of the catalyst material are responsible for the
observed decrease in activity. Overall, this highly active material
■ REFERENCES
(1) Häussinger, P.; Lohmüller, R.; Watson, A. M. Hydrogen, 6. Uses.
remains an excellent HER catalyst throughout the course of the In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag
rigorous accelerated durability test. GmbH & Co. KGaA: Weinheim, 2000.

1922 dx.doi.org/10.1021/cs300451q | ACS Catal. 2012, 2, 1916−1923


ACS Catalysis Research Article

(2) U.S. Energy Information Administration. The Impact of (34) Vrubel, H.; Merki, D.; Hu, X. Energy Environ. Sci. 2012, 5 (3),
Increased Use of Hydrogen on Petroleum Consumption and Carbon 6136−6144.
Dioxide Emissions. http://www.eia.gov/oiaf/servicerpt/hydro/pdf/ (35) Gilliam, R.; Kirk, D.; Thorpe, S. Electrocatalysis 2012, 3, 1−7.
oiafcneaf%2808%2904.pdf (accessed August 5, 2012). (36) Gasteiger, H. A.; Kocha, S. S.; Sompalli, B.; Wagner, F. T. Appl.
(3) Turner, J. A. Science 2004, 305 (5686), 972−974. Catal., B 2005, 56 (1−2), 9−35.
(4) Nowotny, J.; Sorrell, C. C.; Sheppard, L. R.; Bak, T. Int. J. (37) Eda, G.; Yamaguchi, H.; Voiry, D.; Fujita, T.; Chen, M.;
Hydrogen Energy 2005, 30 (5), 521−544. Chhowalla, M. Nano Lett. 2011, 11 (12), 5111−5116.
(5) Walter, M. G.; Warren, E. L.; McKone, J. R.; Boettcher, S. W.; Mi, (38) Lee, C.; Yan, H.; Brus, L. E.; Heinz, T. F.; Hone, J.; Ryu, S. ACS
Q. X.; Santori, E. A.; Lewis, N. S. Chem. Rev. 2010, 110 (11), 6446− Nano 2010, 4 (5), 2695−2700.
6473. (39) Wieting, T. J.; Verble, J. L. Phys. Rev. B 1971, 3 (12), 4286−
(6) Cook, T. R.; Dogutan, D. K.; Reece, S. Y.; Surendranath, Y.; 4292.
Teets, T. S.; Nocera, D. G. Chem. Rev. 2010, 110 (11), 6474−6502. (40) Weber, T.; Muijsers, J. C.; Niemantsverdriet, J. W. J. Phys. Chem.
(7) Turner, J.; Sverdrup, G.; Mann, M. K.; Maness, P. C.; Kroposki, 1995, 99 (22), 9194−9200.
B.; Ghirardi, M.; Evans, R. J.; Blake, D. Int. J. Energy Res. 2008, 32 (5), (41) Iranmahboob, J.; Gardner, S. D.; Toghiani, H.; Hill, D. O. J.
379−407. Colloid Interface Sci. 2004, 270 (1), 123−126.
(8) Bak, T.; Nowotny, J.; Rekas, M.; Sorrell, C. C. Int. J. Hydrogen (42) Wagner, C. D.; Taylor, J. A. J. Electron Spectrosc. Relat. Phenom.
Energy 2002, 27 (10), 991−1022. 1980, 20 (1−2), 83−93.
(9) Kibler, L. A. ChemPhysChem 2006, 7 (5), 985−991. (43) Siriwardane, R. V.; Cook, J. M. J. Colloid Interface Sci. 1986, 114
(10) Subbaraman, R.; Tripkovic, D.; Strmcnik, D.; Chang, K.-C.; (2), 525−535.
Uchimura, M.; Paulikas, A. P.; Stamenkovic, V.; Markovic, N. M. (44) Jacobson, A. J.; Chianelli, R. R.; Rich, S. M.; Whittingham, M. S.
Science 2011, 334 (6060), 1256−1260. Mater. Res. Bull. 1979, 14 (11), 1437−1448.
(11) Tang, M. L.; Grauer, D. C.; Lassalle-Kaiser, B.; Yachandra, V. K.; (45) Liang, K. S.; Cramer, S. P.; Johnston, D. C.; Chang, C. H.;
Amirav, L.; Long, J. R.; Yano, J.; Alivisatos, A. P. Angew. Chem., Int. Ed. Jacobson, A. J.; Deneufville, J. P.; Chianelli, R. R. J. Non-Cryst. Solids
2011, 50 (43), 10203−10207. 1980, 42 (1−3), 345−356.
(12) Krstajic, N. V.; Jovic, V. D.; Gajic-Krstajic, L.; Jovic, B. M.; (46) Chang, C. H.; Chan, S. S. J. Catal. 1981, 72 (1), 139−148.
Antozzi, A. L.; Martelli, G. N. Int. J. Hydrogen Energy 2008, 33 (14), (47) Bhattacharya, R. N.; Lee, C. Y.; Pollak, F. H.; Schleich, D. M. J.
3676−3687. Non-Cryst. Solids 1987, 91 (2), 235−242.
(13) Lasia, A.; Rami, A. J. Electroanal. Chem. 1990, 294 (1−2), 123− (48) Belanger, D.; Laperriere, G.; Marsan, B. J. Electroanal. Chem.
141. 1993, 347 (1−2), 165−183.
(14) Rami, A.; Lasia, A. J. Appl. Electrochem. 1992, 22 (4), 376−382. (49) Hibble, S. J.; Rice, D. A.; Pickup, D. M.; Beer, M. P. Inorg. Chem.
(15) Machado, S. A. S.; Avaca, L. A. Electrochim. Acta 1994, 39 (10), 1995, 34 (21), 5109−5113.
1385−1391. (50) Hibble, S. J.; Walton, R. I.; Pickup, D. M.; Hamon, A. C. J. Non-
(16) Petrov, Y.; Schosger, J.-P.; Stoynov, Z.; de Bruijn, F. Int. J. Cryst. Solids 1998, 232, 434−439.
Hydrogen Energy 2011, 36 (20), 12715−12724. (51) Hibble, S. J.; Feaviour, M. R.; Almond, M. J. J. Chem. Soc., Dalton
(17) Endoh, E.; Otouma, H.; Morimoto, T.; Oda, Y. Int. J. Hydrogen Trans. 2001, 6, 935−940.
Energy 1987, 12 (7), 473−479. (52) Jiao, H. J.; Li, Y. W.; Delmon, B.; Halet, J. F. J. Am. Chem. Soc.
(18) Huq, A. K. M. S.; Rosenberg, A. J.; Makrides, A. C. J. 2001, 123 (30), 7334−7339.
Electrochem. Soc. 1964, 111 (3), 278−286. (53) Hibble, S. J.; Wood, G. B. J. Am. Chem. Soc. 2004, 126 (3), 959−
(19) Birry, L.; Lasia, A. J. Appl. Electrochem. 2004, 34 (7), 735−749. 965.
(20) Dominey, R. N.; Lewis, N. S.; Bruce, J. A.; Bookbinder, D. C.; (54) Weber, M. F.; Dignam, M. J. J. Electrochem. Soc. 1984, 131 (6),
Wrighton, M. S. J. Am. Chem. Soc. 1982, 104 (2), 467−482. 1258−1265.
(21) Li, Y.; Somorjai, G. A. Nano Lett. 2010, 10 (7), 2289−2295. (55) Markovic, N. M.; Grgur, B. N.; Ross, P. N. J. Phys. Chem. B
(22) Boettcher, S. W.; Spurgeon, J. M.; Putnam, M. C.; Warren, E. L.; 1997, 101 (27), 5405−5413.
Turner-Evans, D. B.; Kelzenberg, M. D.; Maiolo, J. R.; Atwater, H. A.; (56) Sheng, W.; Gasteiger, H. A.; Shao-Horn, Y. J. Electrochem. Soc.
Lewis, N. S. Science 2010, 327 (5962), 185−187. 2010, 157 (11), B1529−B1536.
(23) Hou, Y. D.; Abrams, B. L.; Vesborg, P. C. K.; Bjorketun, M. E.; (57) Pourbaix, M., Atlas of Electrochemical Equilibria in Aqueous
Herbst, K.; Bech, L.; Setti, A. M.; Damsgaard, C. D.; Pedersen, T.; Solutions, 2nd ed.; NACE International: Houston, TX, 1974.
Hansen, O.; Rossmeisl, J.; Dahl, S.; Norskov, J. K.; Chorkendorff, I. (58) Karunadasa, H. I.; Montalvo, E.; Sun, Y. J.; Majda, M.; Long, J.
Nat. Mater. 2011, 10 (6), 434−438. R.; Chang, C. J. Science 2012, 335 (6069), 698−702.
(24) Jaramillo, T. F.; Jorgensen, K. P.; Bonde, J.; Nielsen, J. H.; (59) Kotz, R.; Carlen, M. Electrochim. Acta 2000, 45 (15−16), 2483−
2498.
Horch, S.; Chorkendorff, I. Science 2007, 317 (5834), 100−102.
(60) Trasatti, S.; Petrii, O. A. Pure Appl. Chem. 1991, 63 (5), 711−
(25) Jaramillo, T. F.; Bonde, J.; Zhang, J. D.; Ooi, B. L.; Andersson,
734.
K.; Ulstrup, J.; Chorkendorff, I. J. Phys. Chem. C 2008, 112 (45),
(61) Bard, A. J.; Faulkner, L. R. Electrochemical Methods:
17492−17498.
Fundamentals and Applications; Wiley: New York, 2001.
(26) Bonde, J.; Moses, P. G.; Jaramillo, T. F.; Norskov, J. K.;
(62) Conway, B. E.; Birss, V.; Wojtowicz, J. J. Power Sources 1997, 66
Chorkendorff, I. Faraday Discuss. 2008, 140, 219−231.
(1−2), 1−14.
(27) Chen, Z. B.; Cummins, D.; Reinecke, B. N.; Clark, E.; Sunkara,
M. K.; Jaramillo, T. F. Nano Lett. 2011, 11 (10), 4168−4175.
(28) Li, Y. G.; Wang, H. L.; Xie, L. M.; Liang, Y. Y.; Hong, G. S.; Dai,
H. J. J. Am. Chem. Soc. 2011, 133 (19), 7296−7299.
(29) Merki, D.; Fierro, S.; Vrubel, H.; Hu, X. L. Chem. Sci. 2011, 2
(7), 1262−1267.
(30) Merki, D.; Hu, X. Energy Environ. Sci. 2011, 4 (10), 3878−3888.
(31) Laursen, A. B.; Kegnaes, S.; Dahl, S.; Chorkendorff, I. Energy
Environ. Sci. 2012, 5, 2.
(32) Jaramillo, T. F.; Jorgensen, K. P.; Bonde, J.; Nielsen, J. H.;
Horch, S.; Chorkendorff, I. Science 2007, 317 (5834), 100−102.
(33) Merki, D.; Vrubel, H.; Rovelli, L.; Fierro, S.; Hu, X. Chem. Sci.
2012, 3 (8), 2515−2525.

1923 dx.doi.org/10.1021/cs300451q | ACS Catal. 2012, 2, 1916−1923

You might also like