You are on page 1of 9

Letter

http://pubs.acs.org/journal/aelccp

Direct H2S Decomposition by Plasmonic


Photocatalysis: Efficient Remediation plus
Sustainable Hydrogen Production
Minghe Lou, Junwei Lucas Bao, Linan Zhou, Gopal Narmada Naidu, Hossein Robatjazi,*
Aaron I. Bayles, Henry O. Everitt,* Peter Nordlander,* Emily A. Carter,* and Naomi J. Halas*
Cite This: ACS Energy Lett. 2022, 7, 3666−3674 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
Downloaded via RICE UNIV on October 5, 2022 at 15:24:24 (UTC).

ABSTRACT: Plasmonic metal nanostructures have garnered rapidly increasing


interest as heterogeneous photocatalysts, facilitating chemical bond activation and
overcoming the high energy demands of conventional thermal catalysis. Here we
report the highly efficient plasmonic photocatalysis of the direct decomposition of
hydrogen sulfide into hydrogen and sulfur, an alternative to the industrial Claus
process. Under visible light illumination and with no external heat source, up to a
20-fold reactivity enhancement compared to thermocatalysis can be observed. The
substantially enhanced reactivity can be attributed to plasmon-mediated hot carriers
(HCs) that modify the reaction energetics. With a shift in the rate-determining step
of the reaction, a new reaction pathway is made possible with a lower apparent
reaction barrier. Light-driven one-step decomposition of hydrogen sulfide
represents an exciting opportunity for simultaneous high-efficiency hydrogen production and low-temperature sulfur
recovery, important in many industrial processes.

H ydrogen sulfide (H2S) is a key byproduct of the


hydrodesulfurization process in oil and gas refineries
and the petrochemical industry (R-SH + H2 → RH +
H2S), contributing to the majority of the sulfur supply
worldwide.1 As a toxic and corrosive gas, H2S is listed as an
be fed back into the hydrodesulfurization reaction, forming a
closed loop and mitigating the need for an external H2 supply
(which, in current industrial practice, typically originates from
a CO2-emitting source). However, the direct decomposition of
H2S is a highly endothermic process (with a reaction enthalpy
extremely hazardous substance by the United States (U.S.) of +169.4 kJ/mol)8 with limited thermodynamic conversion at
Environment Protection Agency, causing irreversible harmful temperatures below 1000 K.6 Aside from thermocatalytic
effects to humans at concentrations as low as 20 ppm in efforts for direct H2S decomposition,9,10 there have been
ambient air.2 The U.S. National Emission Inventory reports attempts to achieve faster kinetic rates at milder temperatures
annual H2S emissions of ∼9800 tons in the U.S. alone.3 Hence, using electrocatalytic,11,12 photocatalytic,13−19 plasma-in-
H2S decomposition is critical for reducing sulfur emissions, duced,20 and microwave-assisted approaches.21,22 Chemical
while also potentially producing hydrogen (H2), a high-value looping23,24 has also been implemented to circumvent the
product. The traditionally established technology for H2S thermodynamic limitation of this reaction. While these studies
decomposition is the energy-intensive two-step Claus process: represent valuable progress toward the direct decomposition of
H2S first is partially oxidized in the air into SO2 in combustion H2S at milder temperatures, difficulties with scale-up due to
chambers at temperatures >1100 K (thermal stage), followed the energy source or the catalyst are still major challenges.
by a catalytic reaction of H2S with SO2 (Claus reaction) to Here we demonstrate highly efficient, visible-light-driven,
produce S2 and H2O.4 The produced S2 can further polymerize
heterogeneous photocatalysis for the direct decomposition of
or transform into other more stable forms of sulfur, such as S8.
In contrast, direct decomposition of H2S into H2 and sulfur
(2H2S(g) → 2H2(g) + S2(g)) is a potentially attractive way to Received: August 3, 2022
address the high overall energy cost of the Claus process5 and Accepted: September 14, 2022
simultaneously obtain clean H2 during sulfur recovery.6,7 This
approach can be implemented for H2 production from other,
nonpetrochemical H2S-containing resources, such as chemical
and animal wastes and sewer gases. The produced H2 can also

© XXXX American Chemical Society https://doi.org/10.1021/acsenergylett.2c01755


3666 ACS Energy Lett. 2022, 7, 3666−3674
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Figure 1. Optical characterization of the photocatalyst and its catalytic properties. (a) Transmission electron microscope (TEM) image of 5−
10 nm Au NPs dispersed in chloroform (scale bar: 5 nm). (b) TEM image of SiO2-supported Au NPs (scale bar: 10 nm). (c) Measured (solid
line) and Monte Carlo simulated (dashed line) diffuse reflectance spectra of SiO2-supported Au NPs. (d) H2S turnover frequency at different
surface temperatures in thermocatalytic (528−773 K) and photocatalytic (white light laser, 400−900 nm, 4.0−10.0 W/cm2 without external
heating, surface temperature monitored with IR camera) H2S decomposition. Error bars represent standard deviations of three parallel
measurements under the same conditions, using samples from the same batch. Solid lines are the linear fitting of the data points with R2 =
0.993 and 0.97 for thermocatalysis and photocatalysis, respectively.

H2S into H2 and sulfur using SiO2-supported gold (Au) use of only visible light, with no external heating, to drive a
nanoparticles (NPs) as the plasmonic photocatalyst. Plasmonic one-step H2S decomposition process allows for relatively
metal nanostructures constitute a transformative approach for straightforward future scale-up using renewable solar energy or
enhancing catalysis at substantially milder operating temper- highly efficient solid-state lighting systems (e.g., LEDs).37
atures than traditional processes by taking advantage of the The photocatalyst, consisting of Au NPs (5−10 nm
strong interactions between metal NPs and light.25 Illumina- diameter; Figure 1a and Figure S2), was synthesized colloidally
tion of plasmonic metal NPs can excite collective electronic using a previously reported method with slight modifications
oscillations, known as localized surface plasmons.26 Decay of (see the Supporting Information for details) and supported on
localized surface plasmons generates energetic HCs that powdery SiO2 at a 2% weight percentage using a wet
facilitate surface reactions through nonthermal bond activation impregnation method (Figure 1b) to increase its stability.38
mechanisms or through the excitation of phonons, producing The extinction spectrum of the NPs dispersed in chloroform
heat.26,27 shows a dipolar plasmon peak at ∼512 nm, consistent with
Compared to traditional thermocatalysis, hot-carrier-medi- spectra simulated using the finite difference time domain
ated plasmonic photocatalysis has been reported to be capable (FDTD) method (Figure S2D). The ensemble absorption
of enhancing both reactivity and selectivity,28,29 making it an spectrum of the NPs on the support measured by diffuse
attractive and more sustainable alternative to traditional reflectance spectrometry and reproduced using Monte Carlo
phonon-driven chemistry.30 Recent studies also show that simulations also shows a strong dipolar plasmon peak with a
nonthermal bond activation in plasmonic photocatalysis is slight red shift to ∼516 nm due to the increase in the
different from photocatalytic redox chemistry on semi- surrounding dielectric environment of the NPs (Figure 1c).
conductors that involves charge transfer to adsorbate(s), Polycrystalline X-ray diffraction (PXRD) of the supported
because the net result of HC generation in photoexcited catalyst Au/SiO2 shows characteristic peaks of SiO2 and Au
metal nanostructures is energy transfer.25 This feature of (Figure S3). The broadened peaks indicate that both the Au
plasmonic photocatalysts results in a positive dependence of and SiO2 are on the nanoscale. X-ray photoelectron spectros-
quantum efficiency on photon flux and operating temper- copy (XPS) analysis of the catalyst indicates a surface Au:Si
ature,27 allowing for higher reaction rates under illumination atomic ratio of 1:109 (Figure S4). A Brunauer−Emmett−
and a broader scope of chemical reactions that can be enabled Teller (BET) analysis of the adsorption isotherm of the
with light. The use of light for promoting heterogeneous catalyst gives a surface area of 196.4 m2/g (Figure S5).
catalysis on plasmonic NPs has been demonstrated recently for When the as-prepared catalyst was used for H 2 S
several high-impact chemical reactions, including, but not decomposition, a reduction of the apparent activation barrier
limited to, CO2 reduction,28,31,32 methane dry reforming,33,34 from 0.61 eV for thermocatalysis to 0.32 eV for photocatalysis
C−F bond activation,35 and olefin epoxidation.36 was observed (Figure 1d) (see the Supporting Information for
In this work we report the direct illumination of a SiO2- experimental details of photocatalysis and thermocatalysis
supported Au NP catalyst resulting in surface temperature measurements). Such a significant change in apparent
increases of up to 500 °C (at ∼13 W cm−2), producing up to a activation barrier indicates a nontrivial involvement of HCs
20-fold increase in reactivity compared to thermocatalysis at in the photocatalytic process, as has been reported in several
similar surface temperatures. H2 was produced at a rate of up previous studies.33,35,39 The photocatalytic process also
to 5 μmol H2 s−1 per gram of catalyst (turnover frequency exhibits 1−2 orders of magnitude higher reactivity than
TOF = 0.28 s−1 and external quantum yield EQY = 1.8%; thermocatalysis at the same temperature (Figure 1d), depend-
Figure S1) in photocatalysis, which is on par with or better ing on the optical intensity and power used. A set of
than most previously reported results, albeit without the need photocatalysis experiments was performed to understand these
to use a basic solution and additional reagents (Table S1). The properties further.
3667 https://doi.org/10.1021/acsenergylett.2c01755
ACS Energy Lett. 2022, 7, 3666−3674
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Figure 2. Photocatalytic behavior of the Au/SiO2 photocatalyst. Photocatalytic H2S turnover frequency from H2S decomposition (blue) and
the corresponding light-induced surface temperature of the catalyst (red) as a function of incident laser power density (a, 525 ± 25 nm
illumination) and illumination wavelength (b, 8.8 W/cm2 from each 50 nm filtered bandwidth of the light source). The estimated
photothermal contributions (black curves) were obtained by inserting the maximum surface temperature at each data point into the
Arrhenius fit of the thermocatalytic process in Figure 1d. The dashed line in (b) corresponds to the extinction spectrum of the Au NPs. All
measurements were performed under a 10 sccm H2S flow. Error bars represent the standard deviations of three parallel measurements under
the same conditions using samples from the same batch. Solid lines are included as visual guides.

Photocatalytic reaction rates as a function of incident power temperature increase to estimate photothermal reactivity, we
show a pseudolinear power dependence with a TOF as high as overestimate the actual photothermal heating of photocatalysis.
∼0.28 s−1 at an incident light intensity of ∼13 W/cm2 (Figure Even so, the photothermal contribution is found to account for
2a, blue). Due to the small thermal conductivity of H2S40 and less than 10% of the total photocatalytic TOF. This result
the low flow rates applied over an Au-NP-based catalyst known strongly supports the contribution of nonthermal plasmonic
for efficient photothermal heating, the light-induced maximum effects (HC injection and near-field enhancement)25,43 as the
surface temperature of the sample pellet reaches ∼773 K predominant bond activation mechanism for this process.
(Figure 2a, red). It should be mentioned that the assessments To understand how plasmonically generated HCs efficiently
of temperature homogeneity based on the previously proposed catalyze this highly endothermic reaction, quantum mechanical
empirical model41 and our calculations (see the Supporting simulations were performed. The optimized geometries and
Information for details) suggested the influence of collective energies of the key structures along minimum energy paths
heating, where a light-induced local temperature increase on an (MEPs) were calculated, assuming the steps
individual nanoparticle is assumed to be similar to that of its H 2S(g) + * H 2S* (1)
surroundings, as resolved by the camera. The wavelength-
dependent reactivity was also measured from 450 to 675 nm H 2S* + * HS* + H* (2)
(Figure 2b, blue) and followed the extinction spectrum of the
catalyst (Figure 2b, dashed), with a maximum reactivity near HS* + * S* + H* (3)
the dipolar plasmon wavelength of the Au NPs (525 nm).
Figure 2b also displays wavelength-dependent catalyst 2H* H 2(g) + 2* (4)
heating (red) that qualitatively follows the trend of the
reactivity response (blue). Despite the significant photo- 2S* S2* + * (5)
thermal heating produced, these data clearly show that there
is a difference in reactivity at excitation wavelengths that S2* S2(g) + * (6)
induce similar surface temperatures. For example, illumination where * denotes active sites on the Au surface. These involve
in the wavelength range between 575 and 625 nm results in a the adsorption of gaseous H2S (step 1), the sequential scission
substantially lower photocatalytic reactivity than illumination of two H−S bonds (steps 2 and 3), the associative desorption
in the 450−500 nm wavelength range, despite similar surface of gaseous H2 (step 4), and the formation (step 5) and
temperatures. This observation strongly suggests that non- desorption (step 6) of S2, as generally accepted for H2S
thermal effects are contributing to the reaction pathway, even decomposition on metal catalysts.7,8,44−47 The reaction
at these elevated temperatures, and that photothermal catalyst initiates via the adsorption of H2S molecules on the
heating is only a minor contributor to the overall reactivity.42 photocatalyst surface with an adsorption energy of 0.60 eV
To provide insight into the relative contributions of thermal (Figure 3a), matching well with experimental values of 0.4−0.5
and nonthermal effects to the photocatalytic reaction, we eV.48 Both previous work47,49 and our optimizations (Figure
estimated the photocatalytic reactivities associated with S6) show that the physisorbed H2S molecules attach to the
photothermal heating alone by inserting the maximum surface Au(111) surface with an S atom on top of an Au atom.
temperatures under illumination into the Arrhenius expression Subsequent scissions of the two H−S bonds then are facilitated
for thermocatalytic reactivity that would have been measured by a polarization of the Au surface through interaction with the
in the dark at elevated temperatures (Figure 2a,b, black). By adsorbed H2S, generating adsorbed H*, HS*, and S* species.
assuming a uniform light-induced temperature profile within The associative desorption of these species yields H2 and S2 as
the bulk of the catalyst and using the maximum surface the final products.
3668 https://doi.org/10.1021/acsenergylett.2c01755
ACS Energy Lett. 2022, 7, 3666−3674
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Figure 3. emb-CASPT2 energy curves for ground and excited states along the MEPs. (a) Ground-state potential energy diagram of H2S
decomposition on Au(111) constructed using the optimized MEPs. (b, c) emb-CASPT2 energy curves for ground and excited states along
the MEPs for HS* scission ((b) 0 to ∼3 Å along the reaction coordinate s; the 1.18 eV barrier at s = 4.8 Å is for (unnecessary) *S migration
(see text)), S2 formation ((c) 0−4.8 Å along s), and S2 desorption ((c) s = 4.8−10.5 Å). Blue and orange dashed curves are possible excited
state trajectories.

The potential energy diagram obtained from the optimized CASPT2).51 Previous work has shown that HCs generated
MEP geometries (Figures S6−S9) calculated using plane-wave during plasmonic photocatalysis can transfer energy to
density functional theory (PW-DFT; see the Supporting adsorbed species on the catalyst surface by exciting them
Information for details) indicates three key steps with high into electronic or vibronic excited states.35,36,39,57 In the Au/
activation barriers in the ground state (Figure 3a). These steps SiO2 catalysts, the insulating SiO2 support can confine the
are the scission of the second H−S bond (step 3) and the energetic HCs on the Au NPs, making it more likely for the
formation and desorption of S2 (steps 5 and 6). In steps 5 and plasmon-mediated decomposition of H2S to happen on the Au
6, the energies labeled are the barriers normalized to forming surface instead of the SiO2 surface.
or desorbing “half” of an S2 molecule, so that the actual barrier The emb-CASPT2 ground-state activation barrier of step 3
is twice the labeled value. Although step 3 has a lower barrier (reaction coordinate 0.70−1.05 Å in Figure 3b) was calculated
compared to the other two steps, previous studies have shown to be 1.36 eV without zero-point-energy corrections. However,
that S-covered Au surfaces may increase the activation barrier the HS* at reaction coordinate 0.70 Å can absorb, for instance,
of step 3.47 Based on these analyses, steps 3, 5, and 6 are likely ∼1.68 eV of energy from a plasmon decay and jump to the
to compete to be the rate-determining step (RDS) for the ninth excited state (S9). At reaction coordinate ∼0.96 Å, the
ground state. potential energy surface (PES) crosses to S2 or S3 non-
The accuracy of this theory has been proven in numerous adiabatically, and the reaction trajectory then follows the
previous studies on different systems,33,39,50−52 but we still see excited state MEPs adiabatically until it reaches the transition
a discrepancy between the measured (0.61 eV) and calculated state at 1.39 Å for S3 or 1.99 Å for S2. The PES then crosses
(>1.36 eV) reaction barriers, regardless of the RDS. We ascribe again to either S2 or S1 nonadiabatically at 2.58 Å or before,
this discrepancy mainly to the nature of small Au NPs and the before jumping back to the ground state near 4 Å. These
idealized Au(111) model we applied. For Au NPs between 2 reaction trajectories can lower the activation barrier of the H−
and 10 nm, the terraces, corners, or step edges dominate their S* bond-breaking step to 0.30−0.56 eV. Note that the PES
catalytic reactivity, giving rise to a lower barrier than that for an between 3.8 and 5.8 Å corresponds to S* migration from a
Au(111) surface.53,54 Small Au NPs have also been reported to hexagonal-close-packed (hcp) site to a face-centered-cubic site
be prone to surface reconstruction under various reaction on Au (111) (Figure S7), which is a possible pathway but does
conditions.55,56 Although our model Au(111) surface does not not appear to be necessary for the following steps of S2
provide an accurate estimation of the ground-state barrier for formation and desorption to occur.
thermocatalysis, it is still practical to use for comparison Similarly, the activation barriers of steps 5 and 6 also can be
between the ground and excited states. reduced to ∼0.40−0.45 eV (Figure 3c). In this case, the two
Potential energies along the MEPs of nine excited states S1− S* (occupying two hcp sites but slightly shifted toward the
S9 for steps 3 (Figure 3b), 5, and 6 (Figure 3c, with both bridge sites, as shown in Figure S9) could gain 1.4 eV of energy
singlet states S1−S9 and triplet states T1−T9 presented from the plasmon decay, jumping to S9 at 0 Å. Because S9 is
because S2(g) is a ground state triplet) were calculated with nearly degenerate with T4 (the notation Tn denotes a triplet
embedded correlated wave function (ECW) theory using state), an intersystem crossing facilitated by spin−orbit
complete active space second-order perturbation theory (emb- coupling may occur and the reaction trajectory then may go
3669 https://doi.org/10.1021/acsenergylett.2c01755
ACS Energy Lett. 2022, 7, 3666−3674
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Figure 4. Photoinduced changes in the kinetics of H2S decomposition. (a) Measurements of turnover frequencies as a function of H2S partial
pressure for photocatalysis under different incident power densities (triangles, solid lines) and dark thermocatalysis under external heating
(cubes, dashed lines). Photocatalysis and thermocatalysis measurements for a given maximum surface temperature share the same color.
Error bars represent the standard deviations of three parallel measurements under the same condition using samples from the same batch.
(b) Illustration of photo- and thermocatalytic H2S decomposition on the Au surface. We consider H−S bond scission in HS* to be the rate-
determining step (RDS) in the thermocatalytic process. Plasmon-induced hot electrons can facilitate the H−S bond scission in HS* while
shifting the RDS to the associative desorption of S2 in photocatalysis, thus resulting in a substantial enhancement in the reaction rates
relative to pure thermocatalysis. (c) Schematic showing the role of plasmon-induced hot electrons for activating the H−S bond in HS*
surface intermediates. HCs with sufficient energy can deposit energy into the adsorbate, resulting in electronic and/or vibrational excitation
of the adsorbate, which subsequently evolves along the reaction coordinate, thus effectively lowering the energy barrier for bond activation.

along the PES of T4 until crossing nonadiabatically to T2/T3 coverage of H2S* is low because the scission of the first H−S
at 1.60 Å. The reaction can go either along T2/T3 to give an bond is faster than scission of the second H−S bond. This
∼0.45 eV RDS barrier (more specifically, the dashed blue results in a lower probability of H2S* species being excited
curve shows a small barrier of ∼0.17 eV at 0.8 Å and two compared to other species with higher surface coverages, such
consecutive RDS barriers, both of which are ∼0.45 eV) or as S* and HS*.
spin-cross to T0 at ∼2.8 Å to experience an ∼0.20 eV barrier Both our photocatalytic reactivity measurements (see also
(dashed orange curve) until the final formation of adsorbed S2 Figures S10−S13) and our quantum mechanical calculations
at 4.80 Å. point to a substantial contribution from nonthermal effects in
In the desorption step, at 4.80 Å, T0 and S0 are nearly the observed light-activated direct H2S decomposition. To
degenerate with only a 0.13 eV energy gap, making ground- ascertain which elementary step is benefiting most from this
state spin crossover possible. The HC excitation therefore can mechanism, a kinetic study of photocatalysis and thermocatal-
populate the system directly from S0 to S9 by the transfer of ysis was performed. We measured the apparent reaction order
1.69 eV of energy to the adsorbed S2, which can travel along
for H2S under photocatalytic and thermocatalytic conditions
the S9 PES by acquiring 0.40 eV of thermal energy, then hop
by measuring the reaction rate as a function of H2S partial
back to the triplet manifold at 5.30 Å. At 7.30 Å, the trajectory
pressure (Figure 4a). The steady-state measurements of the
could cross back to T5 and finally deexcite to T0 at the
desorption limit of 10.40 Å (the dashed blue line shows this reaction order for thermocatalytic H2S decomposition exhibit a
possible path). Although this analysis suggests that it might be slight temperature-dependent behavior that decreases from 0.8
possible to lower the barriers of S2 formation and desorption to 0.6 with increasing temperature from 573 to 673 K. In
via additional excitations (see vertical arrows in Figure 3c), the contrast, we observe a near zeroth-order dependence of the
linear power dependence suggests that this photocatalysis only reaction rate on H2S in photocatalysis (measured under 4.4
involves a single excitation, namely that which lowers the W/cm2), that remained statistically unchanged (−0.10 to 0.10)
barrier for the second H−S bond scission. with increasing illumination intensities up to 7.2 W/cm2.
Whether these excited state pathways are feasible also Given that the light-induced maximum surface temperature
depends on the surface coverages of the adsorbed species within the intensities used is less than 673 K, the substantial
involved. For example, although the activation barrier can be photon-induced changes to the reaction order compared to
lowered to ∼0.41 eV for step 2 (Figure S6), the surface that of thermocatalysis provide additional definitive proof that
3670 https://doi.org/10.1021/acsenergylett.2c01755
ACS Energy Lett. 2022, 7, 3666−3674
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

nonthermal effects dominate this high-temperature photo- electronic and/or vibrational excitation of the adsorbate along
catalytic reaction. the reaction coordinate that subsequently lowers the energy
Light-induced changes in the reaction order directly barrier for the H−S* bond scission more than the other
correlate with changes in coverages of surface species elementary reaction steps, shifting the RDS toward S2
(intermediates) involved in the RDS. Previous studies formation and desorption (Figure 4b,c). This shift in the
suggested that nonthermal photoexcitation of adsorbate− RDS causes an increase in the surface coverage of S* and S2*,
surface states mediated by HCs can promote reactivity in which leaves only limited free sites for gaseous H2S, leading to
light-driven catalysis by modifying nanoparticle surface cover- the near-zero reaction order. According to our experiments
ages of the adsorbed species, shifting the reaction (Figure 4a), this occurs even at illumination intensities as low
RDS.35,39,58,59 as 4.4 W cm−2, for which the light-induced surface temperature
To understand the observed photoinduced changes to the (∼550 K) is lower than the lowest surface temperature for
reaction order, we performed a microkinetic analysis, deriving thermocatalysis (573 K).
rate expressions using a quasi-equilibrium approximation to In conclusion, we demonstrated the direct and efficient
calculate the reaction orders (nH2S) (see the Supporting decomposition of H2S using visible light and no external
Information for details). We considered three scenarios, heating, on SiO2-supported Au NP plasmonic photocatalysts.
assuming the RDS was the H−S bond scission in HS*, the Reaction rates up to 20 times higher than thermocatalysis at
formation of S2*, or the desorption of S2*. Equations 7−9 nominally the same surface temperatures were obtained. We
account for changes to the reaction orders as a function of thoroughly examined the plasmon-induced nonthermal effects
adsorbate surface coverage, where [i*] is the surface coverage responsible for this enhanced reactivity, showing that photo-
of adsorbate i. If the scission of the H−S bond in HS* is the generated HCs can accelerate the second H−S bond scission.
RDS, Our results clearly point to light-driven catalysis resulting in a
shift in the RDS to S2* formation and desorption, away from
n H 2S = 1 2[H 2S*] 2[HS*] (7) the competing RDSs observed for thermocatalysis, substan-
tially decreasing the apparent reaction barrier. This light-based
and HS* will be the dominant species. If the formation of direct decomposition reaction represents an exciting oppor-
adsorbed S2* is the RDS, tunity for improving desulfurization processes and developing
heterogeneous approaches for sustainable sulfur chemistry and
n H 2S = 2 2[H 2S*] 2[HS*] 2[S*] (8) hydrogen production. The continuous drop in the cost of
renewable electricity over the past decade and the increasingly
and S* will be the dominant species. If the desorption of improving efficiency of the LED technology has also made
gaseous S2 is the RDS, cheap renewable photons available, allowing for future
n H 2S = 2 [H 2S*] [HS*] [S*] 2[S*2 ] sustainable production of value-added chemicals and fuels
(9) through photocatalytic and photochemistry approaches.37
and S2* will be the dominant species.
For the thermocatalytic reaction (Figure 4a, dashed curves),
we observe temperature- and coverage-dependent reaction

*
ASSOCIATED CONTENT
sı Supporting Information
orders. Inserting the measured reaction orders into eqs 7−9 The Supporting Information is available free of charge at
gives a free site coverage [*] ranging from 0.9 to 0.4 if one https://pubs.acs.org/doi/10.1021/acsenergylett.2c01755.
assumes [*] = 1 − [D*], where D represents the dominant
species in each case. The measured thermocatalytic reaction Synthesis details, details of the optical and Monte Carlo
order of ∼0.8 for H2S at 573 K indicates that free active sites simulations, microkinetic analysis, and details of the
are relatively abundant. Increasing the temperature by 100 K quantum mechanical studies, including all atomic
slightly decreased the reaction order for H2S, indicating there structures used in the quantum mechanical simulations
are fewer available free sites for gaseous H2S. This observed (PDF)
decrease in the reaction order suggests that the kinetics of H−
S bond scission in HS* benefit slightly more from the
temperature increase than the other steps, leading to an
increase in [S*] and [S2*] coverage on the surface. The
■ AUTHOR INFORMATION
Corresponding Authors
comparable surface coverages of the intermediate species and Hossein Robatjazi − Department of Chemistry, Rice
the relatively close energy barriers for these three steps in the University, Houston, Texas 77005, United States; Laboratory
ground state suggest that, for the thermocatalytic reaction, the of Nanophotonics, Rice University, Houston, Texas 77005,
reaction rate on the Au surface is controlled by competing United States; Syzygy Plasmonics Inc., Houston, Texas
RDSs. 77054, United States; orcid.org/0000-0002-5101-263X;
For the photocatalytic reaction (Figure 4a, solid curves) we Email: hr10@rice.edu
observe near-zero reaction orders over the same temperature Henry O. Everitt − Department of Electrical and Computer
range, confirming that illumination substantially alters the Engineering and Laboratory of Nanophotonics, Rice
kinetics and energetics of the elementary reaction steps. University, Houston, Texas 77005, United States; U.S. Army
Directing light energy into the photoexcitation of adsorbate− DEVCOM Army Research Laboratory, Houston, Texas
metal states is known to induce nonthermal dissociation and 77005, United States; orcid.org/0000-0002-8141-3768;
desorption that may modify the coverage of surface species and Email: he5@rice.edu
thus the reaction pathway.25,29,33,39,59,60 For H2S photo- Peter Nordlander − Department of Electrical and Computer
decomposition, nonthermal energy transfer from plasmon- Engineering, Laboratory of Nanophotonics, and Department
mediated HCs into unoccupied orbitals of HS* could induce of Physics and Astronomy, Rice University, Houston, Texas
3671 https://doi.org/10.1021/acsenergylett.2c01755
ACS Energy Lett. 2022, 7, 3666−3674
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

77005, United States; orcid.org/0000-0002-1633-2937;


Email: nordlander@rice.edu
■ REFERENCES
(1) Apodaca, L. E. Sulfur. In Mineral Commodity Summaries 2020; U.
Emily A. Carter − Department of Mechanical and Aerospace S. Geological Survey: 2020; pp 160−161.
Engineering and the Andlinger Center for Energy and the (2) National Research Council. Hydrogen Sulfide Acute Exposure
Environment, Princeton University, Princeton, New Jersey Guideline Levels. In Acute Exposure Guideline Levels for Selected
08544, United States; orcid.org/0000-0001-7330-7554; Airborne Chemicals; The National Academies Press: 2010; Vol. 9, pp
Email: eac@princeton.edu 173−218.
Naomi J. Halas − Department of Chemistry, Rice University, (3) 2017 National Emissions Inventory Data. https://www.epa.gov/
Houston, Texas 77005, United States; Department of air-emissions-inventories/2017-national-emissions-inventory-nei-data
Electrical and Computer Engineering and Laboratory of (accessed 2022-02-14).
(4) Elsner, M. P.; Menge, M.; Müller, C.; Agar, D. W. The Claus
Nanophotonics, Rice University, Houston, Texas 77005, Process: Teaching an Old Dog New Tricks. Catal. Today 2003, 79−
United States; orcid.org/0000-0002-8461-8494; 80, 487−494.
Email: halas@rice.edu (5) Rahman, R. K.; Raj, A.; Ibrahim, S.; Khan, I. M.; Al Muhairi, N.
O. Reduction in Natural Gas Consumption in Sulfur Recovery Units
Authors through Kinetic Simulation Using a Detailed Reaction Mechanism.
Minghe Lou − Department of Chemistry, Rice University, Ind. Eng. Chem. Res. 2018, 57 (5), 1417−1428.
Houston, Texas 77005, United States; Laboratory of (6) De Crisci, A. G.; Moniri, A.; Xu, Y. Hydrogen from Hydrogen
Nanophotonics, Rice University, Houston, Texas 77005, Sulfide: Towards a More Sustainable Hydrogen Economy. Int. J.
United States Hydrogen Energy 2019, 44 (3), 1299−1327.
Junwei Lucas Bao − Department of Mechanical and (7) Startsev, A. N.; Bulgakov, N. N.; Ruzankin, S. P.; Kruglyakova,
Aerospace Engineering and the Andlinger Center for Energy O. V.; Paukshtis, E. A. The Reaction Thermodynamics of Hydrogen
Sulfide Decomposition into Hydrogen and Diatomic Sulfur. J. Sulfur
and the Environment, Princeton University, Princeton, New Chem. 2015, 36 (3), 234−239.
Jersey 08544, United States; Department of Chemistry, (8) Startsev, A. N.; Kruglyakova, O. V.; Kravtsov, E. A.; Larina, T.
Boston College, Chestnut Hill, Massachusetts 02467, United V.; Paukshtis, E. A. Low Temperature Catalytic Decomposition of
States; orcid.org/0000-0002-4967-663X Hydrogen Sulfide into Hydrogen and Diatomic Gaseous Sulfur. Top.
Linan Zhou − Department of Chemistry, Rice University, Catal. 2013, 56 (11), 969−980.
Houston, Texas 77005, United States; Laboratory of (9) Guldal, N. O.; Figen, H. E.; Baykara, S. Z. Production of
Nanophotonics, Rice University, Houston, Texas 77005, Hydrogen from Hydrogen Sulfide with Perovskite Type Catalysts:
United States LaMO3. Chem. Eng. J. 2017, 313, 1354−1363.
Gopal Narmada Naidu − Department of Electrical and (10) Guldal, N. O.; Figen, H. E.; Baykara, S. Z. Perovskite Catalysts
Computer Engineering and Laboratory of Nanophotonics, for Hydrogen Production from Hydrogen Sulfide. Int. J. Hydrogen
Energy 2018, 43 (2), 1038−1046.
Rice University, Houston, Texas 77005, United States (11) Zhang, M.; Guan, J.; Tu, Y.; Chen, S.; Wang, Y.; Wang, S.; Yu,
Aaron I. Bayles − Department of Chemistry, Rice University, L.; Ma, C.; Deng, D.; Bao, X. Highly Efficient H2 Production from
Houston, Texas 77005, United States; Laboratory of H2S via a Robust Graphene-Encapsulated Metal Catalyst. Energy
Nanophotonics, Rice University, Houston, Texas 77005, Environ. Sci. 2020, 13 (1), 119−126.
United States (12) Karaismailoglu, M.; Guldal, N. O.; Figen, H. E.; Baykara, S. Z.
Complete contact information is available at: Molybdenum- and Vanadium-Containing Perovskite Electrocatalysts
for Dissociation of H2S. Int. J. Energy Res. 2020, 44 (3), 2368−2374.
https://pubs.acs.org/10.1021/acsenergylett.2c01755 (13) Patil, S. S.; Patil, D. R.; Apte, S. K.; Kulkarni, M. V.; Ambekar, J.
D.; Park, C. J.; Gosavi, S. W.; Kolekar, S. S.; Kale, B. B. Confinement
Author Contributions of Ag3PO4 Nanoparticles Supported by Surface Plasmon Resonance
M.L., L.Z., and N.J.H. designed the experiments. M.L. of Ag in Glass: Efficient Nanoscale Photocatalyst for Solar H2
conducted the synthesis and experiments. G.N.N. and P.N. Production from Waste H2S. Appl. Catal. B Environ. 2016, 190,
75−84.
performed the optical simulations. J.L.B. and E.A.C. designed (14) Chaudhari, N. S.; Bhirud, A. P.; Sonawane, R. S.; Nikam, L. K.;
and performed the quantum mechanics calculations. H.R. Warule, S. S.; Rane, V. H.; Kale, B. B. Ecofriendly Hydrogen
assisted in analyzing and interpreting the data. H.O.E. assisted Production from Abundant Hydrogen Sulfide Using Solar Light-
in designing the experiments. A.I.B. performed the BET and Driven Hierarchical Nanostructured ZnIn2S4 Photocatalyst. Green
XRD measurements and part of the TEM imaging of the Chem. 2011, 13 (9), 2500−2506.
(15) Bai, X.-f.; Cao, Y.; Wu, W. Photocatalytic Decomposition of
samples. The manuscript was written with contributions from H2S to Produce H2 over CdS Nanoparticles Formed in HY-Zeolite
all authors. Pore. Renew. Energy 2011, 36 (10), 2589−2592.
Notes (16) Yu, S.; Wu, F.; Zou, P.; Fan, X. B.; Duan, C.; Dan, M.; Xie, Z.;
Zhang, Q.; Zhang, F.; Zheng, H.; Zhou, Y. Highly Value-Added
The authors declare no competing financial interest. Utilization of H2S in Na2SO3 Solution over Ca-CdS Nanocrystal

■ ACKNOWLEDGMENTS
We gratefully acknowledge support from the Robert A. Welch
Photocatalysts. Chem. Commun. 2020, 56 (91), 14227−14230.
(17) Uesugi, Y.; Nagakawa, H.; Nagata, M. Highly Efficient
Photocatalytic Degradation of Hydrogen Sulfide in the Gas Phase
Using Anatase/TiO2(B) Nanotubes. ACS Omega 2022, 7 (14),
Foundation under Grants C-1220 (N.J.H.) and C-1222 (P.N.) 11946−11955.
and support from the Air Force Office of Scientific Research (18) Mersel, M.-A.; Fodor, L.; Pekker, P.; Makó, É .; Horváth, O.
(FA9550-15-1-0022) and the Defense Threat Reduction Effects of Preparation Conditions on the Efficiency of Visible-Light-
Driven Hydrogen Generation Based on Ni(II)-Modified Cd0.25Zn0.75S
Agency (HDTRA 1-16-1-0042). We also thank the Rice Photocatalysts. Molecules 2022, 27 (13), 4296.
Shared Equipment Authority for use of electron microscopy, (19) Xie, Z.; Yu, S.; Fan, X. B.; Wei, S.; Yu, L.; Zhong, Y.; Gao, X.
Raman, XRD, XPS, and BET. W.; Wu, F.; Zhou, Y. Wavelength-Sensitive Photocatalytic H2

3672 https://doi.org/10.1021/acsenergylett.2c01755
ACS Energy Lett. 2022, 7, 3666−3674
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Evolution from H2S Splitting over g-C3N4 with S,N-Codoped Carbon (36) Christopher, P.; Xin, H.; Linic, S. Visible-Light-Enhanced
Dots as the Photosensitizer. J. Energy Chem. 2021, 52, 234−242. Catalytic Oxidation Reactions on Plasmonic Silver Nanostructures.
(20) Zhao, L.; Wang, Y.; Wang, A.; Li, X.; Song, C.; Hu, Y. Cr− Nat. Chem. 2011, 3 (6), 467−472.
Doped ZnS Semiconductor Catalyst with High Catalytic Activity for (37) Schroeder, E.; Christopher, P. Chemical Production Using
Hydrogen Production from Hydrogen Sulfide in Non-Thermal Light: Are Sustainable Photons Cheap Enough? ACS Energy Lett.
Plasma. Catal. Today 2019, 337, 83−89. 2022, 7 (2), 880−884.
(21) Zhu, J.; Xu, W.; Chen, J.; Gan, Z.; Wang, X.; Zhou, J. (38) Zheng, N.; Fan, J.; Stucky, G. D. One-Step One-Phase
Development of Core−Shell Structured Mo2C@BN as Novel Synthesis of Monodisperse Noble-Metallic Nanoparticles and Their
Microwave Catalysts for Highly Effective Direct Decomposition of Colloidal Crystals. J. Am. Chem. Soc. 2006, 128 (20), 6550−6551.
H2S into H2 and S at Low Temperature. Catal. Sci. Technol. 2020, 10 (39) Zhou, L.; Swearer, D. F.; Zhang, C.; Robatjazi, H.; Zhao, H.;
(20), 6769−6779. Henderson, L.; Dong, L.; Christopher, P.; Carter, E. A.; Nordlander,
(22) Luo, M.; Zhou, J.; Xu, W.; Chen, J.; Xiang, M.; Peng, K. P.; Halas, N. J. Quantifying Hot Carrier and Thermal Contributions
Development of Composite Microwave Catalysts (ABSx/CNTs, A = in Plasmonic Photocatalysis. Science 2018, 362, 69−72.
Co, B = Ni, Mo) for the Highly Effective Direct Decomposition of (40) Huber, M. L.; Harvey, A. H. Thermal Conductivity of Gases. In
H2S into H2 and S. Fuel 2020, 281, 118729. CRC Handbook of Chemistry and Physics; CRC Press: 2011.
(23) Gillis, R. J.; Al-Ali, K.; Green, W. H. Thermochemical (41) Baffou, G.; Berto, P.; Bermúdez Ureña, E.; Quidant, R.;
Production of Hydrogen from Hydrogen Sulfide with Iodine Monneret, S.; Polleux, J.; Rigneault, H. Photoinduced Heating of
Thermochemical Cycles. Int. J. Hydrogen Energy 2018, 43 (29), Nanoparticle Arrays. ACS Nano 2013, 7 (8), 6478−6488.
12939−12947. (42) Baffou, G.; Bordacchini, I.; Baldi, A.; Quidant, R. Simple
(24) Zagoruiko, A. Low-Temperature Chemisorption-Enhanced Experimental Procedures to Distinguish Photothermal from Hot-
Catalytic Decomposition of Hydrogen Sulfide: Thermodynamic Carrier Processes in Plasmonics. Light Sci. Appl. 2020, 9, 108.
Analysis and Process Concept. Catal. Today 2019, 329, 171−176. (43) Baffou, G.; Quidant, R. Nanoplasmonics for Chemistry. Chem.
(25) Robatjazi, H.; Yuan, L.; Yuan, Y.; Halas, N. J. Heterogeneous Soc. Rev. 2014, 43 (11), 3898−3907.
Plasmonic Photocatalysis: Light-Driven Chemical Reactions Intro- (44) Mohamadi, S.; Bashiri, H. Kinetic Study of Hydrogen Sulfide
duce a New Approach to Industrially-Relevant Chemistry. In Emerging Decomposition on Pt(111) Surface. Int. J. Chem. Kinet. 2020, 52 (1),
Trends in Chemical Applications of Lasers; Berman, M. R., Young, L., 16−22.
Dai, H., Eds.; ACS Publications: 2021; pp 363−387. (45) Jiang, D. E.; Carter, E. A. Adsorption, Diffusion, and
(26) Brongersma, M. L.; Halas, N. J.; Nordlander, P. Plasmon- Dissociation of H2S on Fe(100) from First Principles. J. Phys.
Induced Hot Carrier Science and Technology. Nat. Nanotechnol. Chem. B 2004, 108 (50), 19140−19145.
(46) Jiang, D. E.; Carter, E. A. First Principles Study of H2S
2015, 10 (1), 25−34.
(27) Christopher, P.; Xin, H.; Marimuthu, A.; Linic, S. Singular Adsorption and Dissociation on Fe(110). Surf. Sci. 2005, 583 (1),
Characteristics and Unique Chemical Bond Activation Mechanisms of 60−68.
(47) Jiang, Z.; Li, M.; Qin, P.; Fang, T. Insight into the Adsorption
Photocatalytic Reactions on Plasmonic Nanostructures. Nat. Mater.
and Decomposition Mechanism of H2S on Clean and S-Covered Au
2012, 11 (12), 1044−1050.
(100) Surface: A Theoretical Study. Appl. Surf. Sci. 2014, 311, 40−46.
(28) Zhang, X.; Li, X.; Zhang, D.; Su, N. Q.; Yang, W.; Everitt, H.
(48) Leavitt, A. J.; Beebe, T. P. Chemical Reactivity Studies of
O.; Liu, J. Product Selectivity in Plasmonic Photocatalysis for Carbon
Hydrogen Sulfide on Au(111). Surf. Sci. 1994, 314 (1), 23−33.
Dioxide Hydrogenation. Nat. Commun. 2017, 8, 1−9. (49) Alfonso, D. R. First-Principles Studies of H2S Adsorption and
(29) Swearer, D. F.; Zhao, H.; Zhou, L.; Zhang, C.; Robatjazi, H.;
Dissociation on Metal Surfaces. Surf. Sci. 2008, 602 (16), 2758−2768.
Martirez, J. M. P.; Krauter, C. M.; Yazdi, S.; McClain, M. J.; Ringe, E.; (50) Martirez, J. M. P.; Carter, E. A. Prediction of a Low-
Carter, E. A.; Nordlander, P.; Halas, N. J. Heterometallic Antenna- Temperature N2 Dissociation Catalyst Exploiting Near-IR-to-Visible
Reactor Complexes for Photocatalysis. Proc. Natl. Acad. Sci. U. S. A. Light Nanoplasmonics. Sci. Adv. 2017, 3 (12), 1−10.
2016, 113 (32), 8916−8920. (51) Martirez, J. M. P.; Bao, J. L.; Carter, E. A. First-Principles
(30) Linic, S.; Aslam, U.; Boerigter, C.; Morabito, M. Photochemical Insights into Plasmon-Induced Catalysis. Annu. Rev. Phys. Chem.
Transformations on Plasmonic Metal Nanoparticles. Nat. Mater. 2021, 72, 99−119.
2015, 14 (6), 567−576. (52) Bao, J. L.; Carter, E. A. Rationalizing the Hot-Carrier-Mediated
(31) Yu, S.; Wilson, A. J.; Heo, J.; Jain, P. K. Plasmonic Control of Reaction Mechanisms and Kinetics for Ammonia Decomposition on
Multi-Electron Transfer and C-C Coupling in Visible-Light-Driven Ruthenium-Doped Copper Nanoparticles. J. Am. Chem. Soc. 2019,
CO2 Reduction on Au Nanoparticles. Nano Lett. 2018, 18 (4), 2189− 141 (34), 13320−13323.
2194. (53) Haruta, M. Size- and Support-Dependency in the Catalysis of
(32) Robatjazi, H.; Zhao, H.; Swearer, D. F.; Hogan, N. J.; Zhou, L.; Gold. Catal. Today 1997, 36 (1), 153−166.
Alabastri, A.; McClain, M. J.; Nordlander, P.; Halas, N. J. Plasmon- (54) Haruta, M. Catalysis of Gold Nanoparticles Deposited on Metal
Induced Selective Carbon Dioxide Conversion on Earth-Abundant Oxides. Cattech 2002, 6 (3), 102−115.
Aluminum-Cuprous Oxide Antenna-Reactor Nanoparticles. Nat. (55) Piccolo, L. Restructuring Effects of the Chemical Environment
Commun. 2017, 8 (1), 1−9. in Metal Nanocatalysis and Single-Atom Catalysis. Catal. Today 2021,
(33) Zhou, L.; Martirez, J. M. P.; Finzel, J.; Zhang, C.; Swearer, D. 373, 80−97.
F.; Tian, S.; Robatjazi, H.; Lou, M.; Dong, L.; Henderson, L.; (56) Albrecht, W.; Arslan Irmak, E.; Altantzis, T.; Pedrazo-Tardajos,
Christopher, P.; Carter, E. A.; Nordlander, P.; Halas, N. J. Light- A.; Skorikov, A.; Deng, T. S.; van der Hoeven, J. E. S.; van Blaaderen,
Driven Methane Dry Reforming with Single Atomic Site Antenna- A.; Van Aert, S.; Bals, S. 3D Atomic-Scale Dynamics of Laser-Light-
Reactor Plasmonic Photocatalysts. Nat. Energy 2020, 5, 61−70. Induced Restructuring of Nanoparticles Unraveled by Electron
(34) Liu, H.; Dao, T. D.; Liu, L.; Meng, X.; Nagao, T.; Ye, J. Light Tomography. Adv. Mater. 2021, 33, 2100972.
Assisted CO2 Reduction with Methane over Group VIII Metals: (57) Kumar, P. V.; Norris, D. J. Tailoring Energy Transfer from Hot
Universality of Metal Localized Surface Plasmon Resonance in Electrons to Adsorbate Vibrations for Plasmon-Enhanced Catalysis.
Reactant Activation. Appl. Catal. B Environ. 2017, 209, 183−189. ACS Catal. 2017, 7 (12), 8343−8350.
(35) Robatjazi, H.; Bao, J. L.; Zhang, M.; Zhou, L.; Christopher, P.; (58) Kale, M. J.; Avanesian, T.; Xin, H.; Yan, J.; Christopher, P.
Carter, E. A.; Nordlander, P.; Halas, N. J. Plasmon-Driven Carbon− Controlling Catalytic Selectivity on Metal Nanoparticles by Direct
Fluorine (C(sp3)−F) Bond Activation with Mechanistic Insights into Photoexcitation of Adsorbate-Metal Bonds. Nano Lett. 2014, 14 (9),
Hot-Carrier-Mediated Pathways. Nat. Catal. 2020, 3 (7), 564−573. 5405−5412.

3673 https://doi.org/10.1021/acsenergylett.2c01755
ACS Energy Lett. 2022, 7, 3666−3674
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

(59) Qi, J.; Resasco, J.; Robatjazi, H.; Alvarez, I. B.; Abdelrahman,
O.; Dauenhauer, P.; Christopher, P. Dynamic Control of Elementary
Step Energetics via Pulsed Illumination Enhances Photocatalysis on
Metal Nanoparticles. ACS Energy Lett. 2020, 5 (11), 3518−3525.
(60) Spata, V. A.; Carter, E. A. Mechanistic Insights into
Photocatalyzed Hydrogen Desorption from Palladium Surfaces
Assisted by Localized Surface Plasmon Resonances. ACS Nano
2018, 12 (4), 3512−3522.

3674 https://doi.org/10.1021/acsenergylett.2c01755
ACS Energy Lett. 2022, 7, 3666−3674

You might also like