You are on page 1of 7

pubs.acs.

org/JPCL Letter

Perspectives on the Competition between the Electrochemical


Water and N2 Oxidation on a TiO2(110) Electrode
Ebrahim Tayyebi, Á rni Björn Höskuldsson,⊥ André Wark,⊥ Narges Atrak, Benjamin M. Comer,
Andrew James Medford, and Egill Skúlason*
Cite This: J. Phys. Chem. Lett. 2022, 13, 6123−6129 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via INDIAN INST OF TECH (ISM) DHANBAD on June 28, 2022 at 11:14:37 (UTC).

ABSTRACT: The electrochemical nitrogen oxidation reaction (NOR) has recently drawn attention due to
promising experimental and theoretical results. It provides an alternative, environmentally friendly route to
directly synthesize nitrate from N2(g). There is to date a limited number of investigations focused on the
electrochemical NOR. Herein, we present a detailed computational study on the kinetics of both the NOR
and the competing oxygen evolution reaction (OER) on the TiO2(110) electrode under ambient
conditions. The use of grand canonical density functional theory in combination with the linearized
Poisson−Boltzmann equation allows a continuous tuning of the explicitly applied electrical potential. We
find that the OER may either promote or suppress the NOR on TiO2(110) depending on reaction
conditions. The detailed atomistic insights provided on the mechanisms of these competing processes make possible further
developments toward a direct electrochemical NOR process.

S ince the early 20th century, ammonia has primarily been


synthesized via the Haber-Bosch process, in which
nitrogen and hydrogen gases are heated to 430 °C, pressurized
NH3 from solvated protons, electrons, and atmospheric
nitrogen:8,9

to 150 atm, and passed over a metal catalyst:1,2 N2(g) + 8(H+ + e−) → 2NH3(g) + H 2(g) (5)

Fe/Ru The ammonia produced in the soil can subsequently be


N2(g) + 3H 2(g) ⎯⎯⎯⎯⎯⎯⎯⎯→ 2NH3(g) (1) oxidized to nitrate by nitrifying bacteria:
The ammonia can further be converted to nitrate via the NH3(g) + O2 (g) + H 2O(l) → NO3− + 5(H+ + e−)
Ostwald process,3 where ammonia is oxidized to nitric oxide at (6)
relatively low pressures (4−10 atm) and high temperatures
While nitrate is produced from nitrogen fixation via
(600−800 °C) in the presence of a catalyst, such as platinum
ammonia, both naturally and artificially, realizing the direct
and rhodium in a 9:1 ratio, platinum metal on fused silica
electro-/photochemical oxidation of N2(g) to nitrate, via the
wool, copper, or nickel:
nitrogen oxidation reaction (NOR), would be of great interest.
cat Recently, the NOR has been the focus of several experimental
4NH3(g) + 5O2 (g) ⎯⎯⎯→ 4NO(g) + 6H 2O(g) (2)
and theoretical works, wherein a few potential (photo)-
Then, NO(g) is further oxidized, and nitric acid is produced: electrocatalysts were identified.10−14 However, some of the
experiments do not include state-of-the-art quantitative
2NO(g) + O2 (g) → 2NO2 (g) (3) isotopic NMR to quantify the fixed nitrogen in the measured
products (e.g., nitrate or ammonia) to rule out false-positive
3NO2 (g) + H 2O(l) → 2HNO3(aq) + NO(g) (4) results. In the theoretical parts, activation energies and the
presence of solvent are mostly neglected, and no consideration
Recent experimental studies show that nitric acid can be is given to the competing oxygen evolution reaction
converted into nitrate via the reaction of NO(g) with a (OER).10−12
hydroperoxyl radical at milder conditions (in the pressure and The role of TiO2 in (photo)electrocatalytic water splitting
temperature ranges of 0.10−0.78 atm and −50−50 °C, for hydrogen production has been extensively studied.15−35
respectively).4,5 Together, the industrial conversion of N2 to
ammonia via the Haber-Bosch process and the subsequent
production of NO(g) via the first half of the Ostwald process Received: March 15, 2022
account for most of the total energy demand of the nitrogen- Accepted: June 22, 2022
fertilizer industry.6,7
The reaction environments in the Haber-Bosch and Ostwald
processes are far from ambient conditions, under which the
enzyme nitrogenase in bacteria catalyzes the production of

© XXXX American Chemical Society https://doi.org/10.1021/acs.jpclett.2c00769


6123 J. Phys. Chem. Lett. 2022, 13, 6123−6129
The Journal of Physical Chemistry Letters pubs.acs.org/JPCL Letter

The widely accepted OER mechanism in acidic conditions where Ueq = 1.68 V vs RHE is the experimentally determined
consists of four proton−electron transfer steps,24,25 as shown equilibrium potential.
in black in Figure 1. The experimentally determined Two different NOR mechanisms are investigated in this
equilibrium potential for the overall reaction (7) is 1.23 V vs article. In the first mechanism, shown in red in Figure 1, water
the reversible hydrogen electrode (RHE). is oxidized to OH*, which then reacts electrochemically with
either N2 or N2O to produce N2O or N2O2, respectively. In the
2H 2O → O2 (g) + 4(H+ + e−) (7) second, depicted in blue in Figure 1, the O* generated from
OH* oxidation reacts nonelectrochemically with either N2 or
N2O to form N2O or N2O2, respectively. In total, four proton−
electron transfers are required to produce NO, but the further
oxidation of NO toward HNO3 should be facile4,5 compared to
the formation of N2O/N2O2 and is therefore not included
here.
Figure 2 shows free energy diagrams for the OER and the
NOR at 2.43 V vs RHE, which is the predicted onset potential

Figure 1. The oxygen evolution reaction pathway is shown in black,


while the electrochemical and nonelectrochemical nitrogen oxidation
reactions pathways are shown in red and blue, respectively. Each
catalytic cycle starts from the pristine TiO2 surface.

Anand et al.36 recently looked at the NOR and OER in the


absence of solvent on TiO2(110), IrO2(110), PtO2(110), and
PbO2(110), using the so-called thermochemical model (TCM)
and computational hydrogen electrode (CHE) method, which
relates thermochemical solid−gas phase chemistry with an
electrochemical system. The TCM-CHE method has been
shown to predict fairly accurate onset potentials for a number
of redox reactions.37−46 However, as also demonstrated here,
predicting the selectivity of a catalyst toward different products
in a multistep reaction as a function of potential requires more
complex models.47−52 Figure 2. Free energy diagrams for the competing OER and NOR on
In this study, grand canonical density functional theory the TiO2(110) surface at 2.43 V. The oxidation of N2 to NO is
(GC-DFT) calculations are used to explore the competing divided into two parts: (a) N2 oxidation toward N2O and (b) N2O
NOR and OER under ambient conditions as a function of oxidation toward NO. In both (a) and (b), the black curve shows the
explicitly applied potential on a semiconducting TiO2(110) predicted OER pathway, but for clarity the values of the free energy
photoelectrocatalyst. GC-DFT has recently been developed for barriers are only included in (a). The number of completed proton−
the purpose of simulating electrocatalytic and electrochemical electron transfer steps in the OER is used as a reaction coordinate.
The error bars for the OER barriers represent the maximum (a
interfaces with fixed electron and ion chemical potentials in the symmetry factor of 0.9) and the minimum (a symmetry factor of 0.5)
grand canonical ensemble.53−55 While GC-DFT was originally estimates of the activation free energies for O* and O2(g) formations,
used to compute activation energies for a series of reduction ±0.03 eV and ±0.2 eV, respectively. A symmetry factor of 0.7 was
reactions on transition metal surfaces,56−58 the theoretical used to estimate the barriers of O* and OOH* formation.
framework has recently been expanded to model the reaction
kinetics on semiconductor materials under applied poten-
tial.59−61 for the OER (Figure S3), in good agreement with the value
The effects of solvation and applied potential on the previously obtained by Man et al.62 For the remainder of the
predicted reaction pathways and activation energies are text, all reported potentials are referred to the RHE. The
accounted for with a hybrid implicit/explicit solvent model. oxidation of N2 to N2O is shown in Figure 2(a), and the
We find that TiO2 may, under certain conditions, catalyze the oxidation of N2O to NO is shown in Figure 2(b). The
electrochemical formation of NO(g) from solvated N2(g). The competing OER is shown in both Figures 2(a) and 2(b). The
overall electrochemical NOR toward NO(g) at acidic calculations were performed using the implicit solvation model
conditions is implemented in VASP, VASPsol,63,64 which enables the
application of a tunable explicit potential in the system. Four
N2(g) + 2H 2O(l) → 2NO(g) + 4(H+ + e−) (8) water molecules (in addition to the two water molecules
6124 https://doi.org/10.1021/acs.jpclett.2c00769
J. Phys. Chem. Lett. 2022, 13, 6123−6129
The Journal of Physical Chemistry Letters pubs.acs.org/JPCL Letter

Figure 3. The blue dots give the calculated activation free energy for the nonelectrochemical reaction of N2 + O* to N2O with only the implicit
solvent model (no explicit water molecules). The cyan colored dots show the calculated activation free energy for the same reaction using the
hybrid implicit/explicit solvent model at 1.2 and 2.43 V. The red dots give the calculated activation free energy for the electrochemical reaction of
N2 + OH* to N2O using the hybrid implicit/explicit solvent model. The black dashed and dotted lines give the estimated activation free energy for
the reaction O* + H2O to O2(g) as a function of the applied potential using the hybrid implicit/explicit solvent model. The dotted and dash-dotted
lines are obtained with a symmetry factor of β = 0.9 (steeper) and β = 0.5 (flatter), respectively, while the dashed line is obtained with a symmetry
factor of β = 0.7. The figure is divided into three regions based on which species are expected to occupy the two CUS sites in the unit cell at a given
applied potential. In the darkest region (U < 2 V), water molecules are expected to be adsorbed on both CUS sites in the unit cell. In the medium
dark region (2 V < U < 2.43 V), one of the adsorbed water molecules has oxidized to OH*, with the other water molecule remaining intact. Finally,
in the light dark region (U > 2.43 V), the OH* species has been further oxidized to O*. The blue and red lines are included for visual guidance.
Filled dots and solid lines denote that the expected surface coverage matches the given reaction step, whereas empty dots and dashed lines signify
that the expected surface coverage at that potential is not the one required for the given reaction step to take place. The insets above the figure
show the relaxed coadsorbed water molecule and the expected adsorbate on the other CUS (H2O*, OH*, and O*) for the three different potential
regions.

adsorbed on the active sites) were added to the simulation cell ion (Figure S9-a). A potential of 2.6 V is required for the
to account for explicit solvation effects (such as hydrogen subsequent oxidation of OH* to O* to occur spontaneously
bonding interactions) and to enable the simulation of proton− (Figure S9-b). The activation free energy for applied potentials
electron transfer reactions. The model system is shown in U < 2.6 V is estimated as the potential difference multiplied by
Figure S1(c,d) in the Supporting Information (SI), and the a symmetry factor (β) of 0.7, i.e., 0.7*(2.6 V − U). Using this,
computational settings used in the simulations are explained in an estimate of 0.12 eV is obtained for the activation free energy
section S1 of the SI. All OER proton−electron transfer steps at 2.43 V, as shown in Figure 2. For comparison, similar values
occur spontaneously when a sufficiently high electric potential of the activation free energy of OH* oxidation are obtained
is explicitly included in the simulation. Moreover, we introduce using the CI-NEB method at constant charge, with the details
a method to estimate OER activation free energies and given in section S4 of the SI.
compare the results with those obtained using the climbing- Next, O* is further oxidized to OOH*, with a predicted
image nudged elastic band method (CI-NEB)65,66 (section S4, reaction free energy of 0.38 eV at 0 V (Figure S3-b), and thus
Figures S6 and S7). Desorbed or physisorbed species are not −2.05 eV at 2.43 V according to the TCM-CHE. However, a
specifically marked in Figure 2 or the text. An isolated * kinetic barrier of around 0.7 eV is predicted when a potential
denotes a vacant active site, which for rutile TiO2(110) is the of 2.43 V is explicitly included in the simulation (Figures 2 and
coordinately unsaturated Ti site (CUS). A chemisorbed S11, section S5). The oxidation of O* is thus the predicted
reaction intermediate on the CUS is denoted by a * connected rate-limiting step (RLS) of the OER, while the oxidation of
to a chemical moiety (for instance, *OH). OH* is potential-limiting according to the TCM. The same
The OER is found to proceed for only one of the two water RLS was identified in a recent study of the OER on IrO2(110),
molecules adsorbed on the CUSs in the unit cell. An applied RuO2(110), PtO2(110), and RhO2(110), where explicit
potential of around 2.2 V is needed to directly oxidize solvent water molecules were included.67 A potential of 3.44
adsorbed water to OH*, with the proton subsequently V is required to spontaneously oxidize O* to O2(g) (section
relocating to the bulk of the solution forming a hydronium S5 and Figure S11), which is significantly higher than the 2.43
6125 https://doi.org/10.1021/acs.jpclett.2c00769
J. Phys. Chem. Lett. 2022, 13, 6123−6129
The Journal of Physical Chemistry Letters pubs.acs.org/JPCL Letter

V predicted by the TCM-CHE. The predicted activation free water molecule diffusing away from the O*. While this suggests
energy of 0.71 eV at 2.43 V is again obtained by multiplying that the O* would rather react with N2 than continuing the
the potential difference by a symmetry factor of 0.7 (Figure 2). OER, this finding could also originate from the limited number
These results reveal that the OOH* is directly oxidized to of explicit water molecules in the simulation. If that is the case,
O2(g) at such positive potentials (Figure S10), so the second the presence of the O*−OH2 bond might lead to an increase
half of the OER is combined into a single step for the in the activation free energy for N2O formation for applied
remainder of this study. potentials larger than 2.1 V. Excluding this possibility requires
Our results indicate that the N2, N2O, and N2O2 molecules calculations with, for instance, a full bilayer of water, which is
are all physisorbed on the surface for the duration of the beyond the scope of this study.
electrochemical NOR process, with the calculated binding free Anand et al.36 calculated the kinetic barriers for the
energies of the N2, N2O, N2O2, and H2O molecules listed in nonelectrochemical steps of N2 + O* to N2O and N2O +
Table S1. Hence, the NOR always begins with the adsorption O* to N2O2 in the absence of a solvent. They concluded the
of water, followed by the oxidation of water to OH*. Then, nonelectrochemical mechanism to be the most likely reaction
OH* is most likely further oxidized to O*, with a predicted pathway of the NOR on TiO2(110) in agreement with our
activation free energy of 0.12 eV at 2.43 V. Alternatively, the detailed study here. They obtained an energy barrier of 0.84 eV
OH* can react with N2 to form N2O in a combined when no potential was applied in the simulation for the
electrochemical process, where the proton is simultaneously reaction of gaseous N2 with O*. A posteriori corrections were
transferred to the solution. However, an activation free energy used to account for field effects and entropy loss in the
of 0.81 eV is obtained for that reaction at 2.43 V (Figures 2(a) transition state. From the field effects, they estimate the barrier
and S12). Likewise, OH* can react with N2O to form N2O2, to decrease by around 0.5 eV at a potential of 2.13 V, but from
with a similarly high activation free energy of 0.87 eV obtained the entropy loss, the free energy barrier is increased to around
at 2.43 V (Figures 2(b), S13, and S14). When O* has formed 1 eV. Our results agree qualitatively with those reported by
on the surface, a barrier of around 0.7 eV must be overcome to Anand et al., as we observe a substantial decrease of the barrier
form OOH* (using β = 0.7). Alternatively, it can react with of N2O formation at more positive potentials (Figure 3).
N2/N2O in a nonelectrochemical process to form N2O/N2O2, Contrary to their conclusion that the OER should be favored
with activation free energies of 0.57 and 0.49 eV at 2.43 V, over the NOR at all relevant potentials, our detailed analysis of
respectively (Figures 2(a) and 2(b)). The calculated activation the competing OER shows that the TiO2(110) might be
free energies of both N2O and N2O2 formations via the selective toward the NOR at potentials of around 2.5 V.
nonelectrochemical process are slightly smaller than that Furthermore, the inclusion of implicit solvent enables us to
predicted for O2 formation, suggesting these reactions will be predict the changes in surface coverage and local reaction
competitive. The cleavage of the N−N bond to N2O2 at 2.43 V environment as the electrode potential is continuously tuned.
is exothermic and has an activation free energy of 0.25 eV Such knowledge is essential for those who seek to optimize the
(Figures 2(b) and S15), which indicates rapid NO formation design and performance of the electrochemical cell.
upon the formation of N2O2. Recently, Kuang et al.10 investigated the NOR on various
Having calculated all relevant NOR activation free energies Ru-doped TiO2 electrocatalysts in a neutral aqueous solution,
at 2.43 V, the next step is to obtain them at different potentials. obtaining a Faradaic efficiency of above 25% at 1.8 V and a
Figure 3 shows the calculated activation free energies of N2O nitrate yield exceeding 150 μmol h−1 g−1 at 2.2 V. They also
formation in the absence/presence of water and the estimated studied the NOR on the pristine TiO2 surface and observed
activation free energies of O2 formation, for a range of positive only trace amounts of nitrate at 2.2 V, below the limit of
potentials. A corresponding plot for the formation of N2O2 detection. According to our results, OH* is the dominant OER
from N2O is shown in Figure S14. intermediate on the surface at 2.2 V (Figure 3), though some
The potential dependence of the activation free energy of O2(g) formation is expected at all potentials above 2 V, as can
the nonelectrochemical N2O formation (blue dots) is be observed from the linear sweep voltammetry curves in ref
explained by the fact that O* is destabilized at more positive 10. With the limited availability of O* at 2.2 V, nitrate
potentials, whereas the transition state remains unchanged (see formation via the nonelectrochemical NOR pathway could
Figure S16). While the nonelectrochemical N2O formation can occur, albeit with a very low efficiency. Thus, the trace
only occur at potentials where an O* coverage is expected, i.e. amounts of nitrate reported by Kuang et al. at 2.2 V are
for U > 2.43 V, the activation free energy is also given for U < consistent with our predictions of very limited NOR efficiency
2.43 V to show the predicted potential dependency of the in this potential range. Below, we propose some strategies to
reaction. increase the NOR efficiency of pristine TiO2 and similar
An O*−OH2 bond is formed between the adsorbed oxygen materials to achieve appreciable yields.
and the oxygen of the nearest solvent water molecule at all Based on the results, we conclude that a close competition
potentials larger than 2.1 V, here defined as the minimum exists between the NOR and OER on TiO2. The conclusion of
potential required for O*−OH2 bond formation. The bond which reaction prevails will depend on the details of the
strength increases, and the O*−OH2 distance decreases, as the reaction environment, the specific facet or active site
potential is increased. The estimated activation free energies investigated, and the level of theory and simulation techniques
for O2 formation for U < 2.6 V, using β = 0.7, are higher than employed. The findings suggest that further theoretical
the ones obtained for N2O formation. The latter ones are advances are needed to elucidate the precise potential-
obtained without accounting for the presence of water (only dependent barriers for the OER on TiO2 due to the instability
implicit solvent). However, the inclusion of explicit water of the final OOH* transition state and that it may be necessary
molecules has insignificant effects on the calculated barriers to use hybrid functionals or even higher levels of theory to
(cyan dots), as the O*−OH2 bond breaking is observed upon conclusively determine the prevailing reaction on a given active
the introduction of N2 in the vicinity of the surface, with the site under typical reaction conditions. The NOR is always at a
6126 https://doi.org/10.1021/acs.jpclett.2c00769
J. Phys. Chem. Lett. 2022, 13, 6123−6129
The Journal of Physical Chemistry Letters pubs.acs.org/JPCL Letter

disadvantage compared to the OER due to the limited Benjamin M. Comer − School of Chemical and Biomolecular
solubility of N2 in the solution. However, we suggest here Engineering, Georgia Institute of Technology, Atlanta,
several strategies which may be fruitful for promoting NOR Georgia 30318, United States; SUNCAT Center for Interface
selectivity. One possibility is tuning the relative chemical Science and Catalysis, Department of Chemical Engineering,
potentials of N2 and H2O to limit the availability of reactants Stanford University, Stanford, California 94305, United
for the OER. Alternatively, increasing the concentration of O2 States
may cause the OER to become equilibrium limited and Andrew James Medford − School of Chemical and
simultaneously promote O*-covered surfaces, leading to Biomolecular Engineering, Georgia Institute of Technology,
enhanced NOR activity. Finally, utilizing potential pulsing, Atlanta, Georgia 30318, United States; orcid.org/0000-
where the potenital is varied between higher values, which 0001-8311-9581
promote O* coverage, and lower values, which promote NOR Complete contact information is available at:
selectivity, may offer a viable strategy.68−70 Furthermore, our https://pubs.acs.org/10.1021/acs.jpclett.2c00769
detailed study indicates that N2O(g) may be formed during the
NOR process. Thus, quantifying N2O(g) yield will be a critical Author Contributions
experimental contribution that will provide additional insight ⊥
A.B.H. and A.W. contributed equally.
into the reaction mechanism.
To summarize, we explore whether N2 can be directly Notes
oxidized to nitric oxide through an electrochemical process on The authors declare no competing financial interest.
rutile TiO2(110). We use GC-DFT calculations to elucidate
the NOR mechanism toward several NOR products, N2O,
N2O2, and NO, and propose ways in which the selectivity of
■ ACKNOWLEDGMENTS
This work was supported by the Icelandic Research Fund
the NOR in an electrochemical reaction cell can be enhanced. (Grant no. 196437-051) and the Doctoral Fund of the
We predict that the NOR can occur on TiO2(110) under University of Iceland.


certain conditions at relatively high potentials (around 2.5 V),
albeit likely with a relatively low efficiency as reported in a
recent experiment. REFERENCES


(1) Jennings, J. R. Catalytic Ammonia Synthesis: Fundamentals and
ASSOCIATED CONTENT Practice; Plenum Press: New York, 1991; DOI: 10.1007/978-1-4757-
9592-9.
* Supporting Information
sı (2) Ammonia - Catalysis and Manufacture; Nielsen, A., Ed.; Springer:
The Supporting Information is available free of charge at 1995; DOI: 10.1007/978-3-642-79197-0.
https://pubs.acs.org/doi/10.1021/acs.jpclett.2c00769. (3) Higton, A.; Clemmet, M.; Golding, E.; Jones, A. V. Access to
Chemistry; Royal Society of Chemistry: 1999; DOI: 10.1039/
Section S1, computational and methodological details 9781847550040.
and model system; section S2, potential referencing (4) Butkovskaya, N. I.; Kukui, A.; Pouvesle, N.; Le Bras, G.
scheme; section S3, oxygen evolution reaction pathway Formation of Nitric Acid in the Gas-Phase HO2 + NO Reaction:
for different xc-functionals; section S4, comparison of Effects of Temperature and Water Vapor. J. Phys. Chem. A 2005, 109
explicit and explicit/implicit solvation models; section (29), 6509−6520.
S5, effects of explicitly applied potential on OER; section (5) Butkovskaya, N.; Kukui, A.; Le Bras, G. HNO3 Forming
Channel of the HO2 + NO Reaction as a Function of Pressure and
S6, binding free energies of relevant molecules; section Temperature in the Ranges of 72−600 Torr and 223−323 K. J. Phys.
S7, effects of explicitly applied potential on N2O and Chem. A 2007, 111 (37), 9047−9053.
N2O2; section S8, stabilizing/destabilizing effects of (6) Smith, C.; Hill, A. K.; Torrente-Murciano, L. Current and Future
potential on transition state of N2O formation; and Role of Haber−Bosch Ammonia in a Carbon-Free Energy Landscape.
section S9, alternative version of Figure 1 showing NOR Energy Environ. Sci. 2020, 13 (2), 331−344.
pathway toward NO (PDF) (7) Rouwenhorst, K. H. R.; Krzywda, P. M.; Benes, N. E.; Mul, G.;


Lefferts, L. Ammonia Production Technologies. Techno-Economic
Challenges Green Ammon. as an Energy Vector 2021, 41−83.
AUTHOR INFORMATION (8) Stryer, L. Stryer Biochemistry; 1995.
(9) Burgess, B. K.; Lowe, D. J. Mechanism of Molybdenum
Corresponding Author Nitrogenase. Chem. Rev. 1996, 96 (7), 2983−3011.
Egill Skúlason − Science Institute, University of Iceland, 107 (10) Kuang, M.; Wang, Y.; Fang, W.; Tan, H.; Chen, M.; Yao, J.; Liu,
Reykjavík, Iceland; Faculty of Industrial Engineering, C.; Xu, J.; Zhou, K.; Yan, Q. Efficient Nitrate Synthesis via Ambient
Mechanical Engineering and Computer Science, University of Nitrogen Oxidation with Ru-Doped TiO2/RuO2 Electrocatalysts.
Iceland, 107 Reykjavík, Iceland; orcid.org/0000-0002- Adv. Mater. 2020, 32 (26), 2002189.
0724-680X; Email: egillsk@hi.is (11) Fang, W.; Du, C.; Kuang, M.; Chen, M.; Huang, W.; Ren, H.;
Xu, J.; Feldhoff, A.; Yan, Q. Boosting Efficient Ambient Nitrogen
Authors Oxidation by a Well-Dispersed Pd on MXene Electrocatalyst. Chem.
Ebrahim Tayyebi − Science Institute, University of Iceland, Commun. 2020, 56 (43), 5779−5782.
107 Reykjavík, Iceland; orcid.org/0000-0002-9461-0410 (12) Dai, C.; Sun, Y.; Chen, G.; Fisher, A. C.; Xu, Z. J.
Á rni Björn Höskuldsson − Science Institute, University of
Electrochemical Oxidation of Nitrogen towards Direct Nitrate
Production on Spinel Oxides. Angew. Chem. 2020, 132 (24), 9504−
Iceland, 107 Reykjavík, Iceland 9508.
André Wark − Science Institute, University of Iceland, 107 (13) Yuan, S. J.; Chen, J. J.; Lin, Z. Q.; Li, W. W.; Sheng, G. P.; Yu,
Reykjavík, Iceland H. Q. Nitrate Formation from Atmospheric Nitrogen and Oxygen
Narges Atrak − Science Institute, University of Iceland, 107 Photocatalysed by Nano-Sized Titanium Dioxide. Nat. Commun.
Reykjavík, Iceland 2013, 4 (1), 2249.

6127 https://doi.org/10.1021/acs.jpclett.2c00769
J. Phys. Chem. Lett. 2022, 13, 6123−6129
The Journal of Physical Chemistry Letters pubs.acs.org/JPCL Letter

(14) Comer, B. M.; Medford, A. J. Analysis of Photocatalytic Level Understanding of the Challenges. Chem. Sci. 2021, 12 (18),
Nitrogen Fixation on Rutile TiO2(110). ACS Sustain. Chem. Eng. 6442−6448.
2018, 6 (4), 4648−4660. (37) Greeley, J.; Jaramillo, T. F.; Bonde, J.; Chorkendorff, I.;
(15) Fujishima, A.; Honda, K. Electrochemical Photolysis of Water Nørskov, J. K. Computational High-Throughput Screening of
at a Semiconductor Electrode. Nature 1972, 238 (5358), 37−38. Electrocatalytic Materials for Hydrogen Evolution. Nat. Mater.
(16) Hashimoto, K.; Irie, H.; Fujishima, A. TiO 2 Photocatalysis: A 2006, 5 (11), 909−913.
Historical Overview and Future Prospects. Japanese J. Appl. Physics, (38) Peterson, A. A.; Abild-Pedersen, F.; Studt, F.; Rossmeisl, J.;
Part 1 Regul. Pap. Short Notes Rev. Pap. 2005, 44 (12), 8269−8285. Nørskov, J. K. How Copper Catalyzes the Electroreduction of Carbon
(17) Deskins, N. A.; Rousseau, R.; Dupuis, M. Defining the Role of Dioxide into Hydrocarbon Fuels. Energy Environ. Sci. 2010, 3 (9),
Excess Electrons in the Surface Chemistry of TiO 2. J. Phys. Chem. C 1311−1315.
2010, 114 (13), 5891−5897. (39) Peterson, A. A.; Nørskov, J. K. Activity Descriptors for CO 2
(18) Deskins, N. A.; Rousseau, R.; Dupuis, M. Distribution of Ti3+ Electroreduction to Methane on Transition-Metal Catalysts. J. Phys.
Surface Sites in Reduced TiO2. J. Phys. Chem. C 2011, 115 (15), Chem. Lett. 2012, 3 (2), 251−258.
7562−7572. (40) Tayyebi, E.; Hussain, J.; Abghoui, Y.; Skulason, E. Trends of
(19) Nakamura, R.; Nakato, Y. Primary Intermediates of Oxygen Electrochemical CO2 Reduction Reaction on Transition Metal Oxide
Photoevolution Reaction on TiO2 (Rutile) Particles, Revealed by in Catalysts. J. Phys. Chem. C 2018, 122 (18), 10078−10087.
Situ FTIR Absorption and Photoluminescence Measurements. J. Am. (41) Atrak, N.; Tayyebi, E.; Skúlason, E. Effect of Co-Adsorbed
Chem. Soc. 2004, 126 (4), 1290−1298. Water on Electrochemical CO2 Reduction Reaction on Transition
(20) Di Valentin, C.; Pacchioni, G.; Selloni, A. Electronic Structure Metal Oxide Catalysts. Appl. Surf. Sci. 2021, 570, 151031.
of Defect States in Hydroxylated and Reduced Rutile TiO2(110) (42) Skúlason, E.; Bligaard, T.; Gudmundsdóttir, S.; Studt, F.;
Surfaces. Phys. Rev. Lett. 2006, 97 (16), 166803. Rossmeisl, J.; Abild-Pedersen, F.; Vegge, T.; Jónsson, H.; Nørskov, J.
(21) Shibuya, T.; Yasuoka, K.; Mirbt, S.; Sanyal, B. A Systematic K. A Theoretical Evaluation of Possible Transition Metal Electro-
Study of Polarons Due to Oxygen Vacancy Formation at the Rutile Catalysts for N2 Reduction. Phys. Chem. Chem. Phys. 2012, 14 (3),
TiO 2(110) Surface by GGA+U and HSE06 Methods. J. Phys.: 1235−1245.
Condens. Matter 2012, 24 (43), 435504. (43) Höskuldsson, Á . B.; Abghoui, Y.; Gunnarsdóttir, A. B.;
(22) Chrétien, S.; Metiu, H. Electronic Structure of Partially Skúlason, E. Computational Screening of Rutile Oxides for Electro-
Reduced Rutile TiO2(110) Surface: Where Are the Unpaired chemical Ammonia Formation. ACS Sustain. Chem. Eng. 2017, 5 (11),
Electrons Located? J. Phys. Chem. C 2011, 115 (11), 4696−4705. 10327−10333.
(23) Cheng, J.; Sprik, M. Aligning Electronic Energy Levels at the (44) Greeley, J.; Stephens, I. E. L.; Bondarenko, A. S.; Johansson, T.
TiO2/H2O Interface. Phys. Rev. B - Condens. Matter Mater. Phys. P.; Hansen, H. A.; Jaramillo, T. F.; Rossmeisl, J.; Chorkendorff, I.;
2010, 82 (8), 081406. Nørskov, J. K. Alloys of Platinum and Early Transition Metals as
(24) Siahrostami, S.; Vojvodic, A. Influence of Adsorbed Water on Oxygen Reduction Electrocatalysts. Nat. Chem. 2009, 1 (7), 552−
the Oxygen Evolution Reaction on Oxides. J. Phys. Chem. C 2015, 119 556.
(2), 1032−1037. (45) Gíslason, P. M.; Skúlason, E. Catalytic Trends of Nitrogen
(25) Rossmeisl, J.; Qu, Z. W.; Zhu, H.; Kroes, G. J.; Nørskov, J. K. Doped Carbon Nanotubes for Oxygen Reduction Reaction. Nanoscale
Electrolysis of Water on Oxide Surfaces. J. Electroanal. Chem. 2007, 2019, 11 (40), 18683−18690.
607 (1−2), 83−89. (46) Rossmeisl, J.; Qu, Z. W.; Zhu, H.; Kroes, G. J.; Nørskov, J. K.
(26) Valdés, Á .; Qu, Z. W.; Kroes, G. J.; Rossmeisl, J.; Nørskov, J. K. Electrolysis of Water on Oxide Surfaces. J. Electroanal. Chem. 2007,
Oxidation and Photo-Oxidation of Water on TiO2 Surface. J. Phys. 607 (1−2), 83−89.
Chem. C 2008, 112 (26), 9872−9879. (47) Hussain, J.; Jónsson, H.; Skúlason, E. Calculations of Product
(27) Chen, J.; Li, Y. F.; Sit, P.; Selloni, A. Chemical Dynamics of the Selectivity in Electrochemical CO2 Reduction. ACS Catal. 2018, 8
First Proton-Coupled Electron Transfer of Water Oxidation on TiO2 (6), 5240−5249.
Anatase. J. Am. Chem. Soc. 2013, 135 (50), 18774−18777. (48) Tayyebi, E.; Hussain, J.; Skúlason, E. Why Do RuO2electrodes
(28) Gono, P.; Ambrosio, F.; Pasquarello, A. Effect of the Solvent on Catalyze Electrochemical CO2reduction to Methanol Rather than
the Oxygen Evolution Reaction at the TiO2-Water Interface. J. Phys. Methane or Perhaps Neither of Those? Chem. Sci. 2020, 11 (35),
Chem. C 2019, 123 (30), 18467−18474. 9542−9553.
(29) Selcuk, S.; Selloni, A. Facet-Dependent Trapping and Dynamics (49) Xiao, H.; Cheng, T.; Goddard, W. A.; Sundararaman, R.
of Excess Electrons at Anatase TiO2 Surfaces and Aqueous Interfaces. Mechanistic Explanation of the PH Dependence and Onset Potentials
Nat. Mater. 2016, 15 (10), 1107−1112. for Hydrocarbon Products from Electrochemical Reduction of CO on
(30) Savory, D. M.; McQuillan, A. J. IR Spectroscopic Behavior of Cu (111). J. Am. Chem. Soc. 2016, 138 (2), 483−486.
Polaronic Trapped Electrons in TiO2 under Aqueous Photocatalytic (50) Cheng, T.; Xiao, H.; Goddard, W. A. Free-Energy Barriers and
Conditions. J. Phys. Chem. C 2014, 118 (25), 13680−13692. Reaction Mechanisms for the Electrochemical Reduction of CO on
(31) Di Valentin, C.; Fittipaldi, D. Hole Scavenging by Organic the Cu(100) Surface, Including Multiple Layers of Explicit Solvent at
Adsorbates on the TiO2 Surface: A DFT Model Study. J. Phys. Chem. PH 0. J. Phys. Chem. Lett. 2015, 6 (23), 4767−4773.
Lett. 2013, 4 (11), 1901−1906. (51) Tayyebi, E.; Abghoui, Y.; Skulason, E. Elucidating the
(32) Li, Y. F.; Liu, Z. P.; Liu, L.; Gao, W. Mechanism and Activity of Mechanism of Electrochemical N2 Reduction at the Ru(0001)
Photocatalytic Oxygen Evolution on Titania Anatase in Aqueous Electrode. ACS Catal. 2019, 9 (12), 11137−11145.
Surroundings. J. Am. Chem. Soc. 2010, 132 (37), 13008−13015. (52) Höskuldsson, Á . B.; Tayyebi, E.; Skúlason, E. Computational
(33) Stecher, T.; Reuter, K.; Oberhofer, H. First-Principles Free- Examination of the Kinetics of Electrochemical Nitrogen Reduction
Energy Barriers for Photoelectrochemical Surface Reactions: Proton and Hydrogen Evolution on a Tungsten Electrode. J. Catal. 2021,
Abstraction at TiO2 (110). Phys. Rev. Lett. 2016, 117 (27), 276001. 404, 362−370.
(34) Zhang, D.; Yang, M.; Dong, S. Hydroxylation of the Rutile (53) Gauthier, J. A.; Dickens, C. F.; Ringe, S.; Chan, K. Practical
TiO2(110) Surface Enhancing Its Reducing Power for Photocatalysis. Considerations for Continuum Models Applied to Surface Electro-
J. Phys. Chem. C 2015, 119 (3), 1451−1456. chemistry. ChemPhysChem 2019, 20 (22), 3074−3080.
(35) Deskins, N. A.; Rousseau, R.; Dupuis, M. Localized Electronic (54) Melander, M. M.; Kuisma, M. J.; Christensen, T. E. K.;
States from Surface Hydroxyls and Polarons in TiO 2(110). J. Phys. Honkala, K. Grand-Canonical Approach to Density Functional
Chem. C 2009, 113 (33), 14583−14586. Theory of Electrocatalytic Systems: Thermodynamics of Solid-Liquid
(36) Anand, M.; Abraham, C. S.; Nørskov, J. K. Electrochemical Interfaces at Constant Ion and Electrode Potentials. J. Chem. Phys.
Oxidation of Molecular Nitrogen to Nitric Acid-towards a Molecular 2019, 150 (4), 041706.

6128 https://doi.org/10.1021/acs.jpclett.2c00769
J. Phys. Chem. Lett. 2022, 13, 6123−6129
The Journal of Physical Chemistry Letters pubs.acs.org/JPCL Letter

(55) Melander, M. M. Grand Canonical Ensemble Approach to


Electrochemical Thermodynamics, Kinetics, and Model Hamilto-
nians. Curr. Opin. Electrochem. 2021, 29, 100749.
(56) Goodpaster, J. D.; Bell, A. T.; Head-Gordon, M. Identification
of Possible Pathways for C-C Bond Formation during Electrochemical
Reduction of CO2: New Theoretical Insights from an Improved
Electrochemical Model. J. Phys. Chem. Lett. 2016, 7 (8), 1471−1477.
(57) Van Den Bossche, M.; Skúlason, E.; Rose-Petruck, C.; Jónsson,
H. Assessment of Constant-Potential Implicit Solvation Calculations
of Electrochemical Energy Barriers for H2 Evolution on Pt. J. Phys.
Chem. C 2019, 123 (7), 4116−4124.
(58) Hossain, M. D.; Huang, Y.; Yu, T. H.; Goddard, W. A.; Luo, Z.
Reaction Mechanism and Kinetics for CO2 Reduction on Nickel
Single Atom Catalysts from Quantum Mechanics. Nat. Commun.
2020, 11 (1), 2256.
(59) Abidi, N.; Bonduelle-Skrzypczak, A.; Steinmann, S. N.
Revisiting the Active Sites at the MoS2/H2O Interface via Grand-
Canonical DFT: The Role of Water Dissociation. ACS Appl. Mater.
Interfaces 2020, 12 (28), 31401−31410.
(60) Liu, C.; Qian, J.; Ye, Y.; Zhou, H.; Sun, C. J.; Sheehan, C.;
Zhang, Z.; Wan, G.; Liu, Y. S.; Guo, J.; et al. Oxygen Evolution
Reaction over Catalytic Single-Site Co in a Well-Defined Brookite
TiO2 Nanorod Surface. Nat. Catal. 2021, 4 (1), 36−45.
(61) Bhati, M.; Chen, Y.; Senftle, T. P. Density Functional Theory
Modeling of Photo-Electrochemical Reactions on Semiconductors:
H2 Evolution on 3C-SiC. J. Phys. Chem. C 2020, 124 (49), 26625−
26639.
(62) Man, I. C.; Su, H.-Y.; Calle-Vallejo, F.; Hansen, H. A.;
Martínez, J. I.; Inoglu, N. G.; Kitchin, J.; Jaramillo, T. F.; Nørskov, J.
K.; Rossmeisl, J. Universality in Oxygen Evolution Electrocatalysis on
Oxide Surfaces. ChemCatChem. 2011, 3 (7), 1159−1165.
(63) Mathew, K.; Sundararaman, R.; Letchworth-Weaver, K.; Arias,
T. A.; Hennig, R. G. Implicit Solvation Model for Density-Functional
Study of Nanocrystal Surfaces and Reaction Pathways. J. Chem. Phys.
2014, 140 (8), 084106.
(64) Mathew, K.; Kolluru, V. S. C.; Mula, S.; Steinmann, S. N.;
Hennig, R. G. Implicit Self-Consistent Electrolyte Model in Plane-
Wave Density-Functional Theory. J. Chem. Phys. 2019, 151 (23),
234101.
(65) Henkelman, G.; Jónsson, H. Improved Tangent Estimate in the
Nudged Elastic Band Method for Finding Minimum Energy Paths
and Saddle Points. J. Chem. Phys. 2000, 113 (22), 9978−9985.
(66) Henkelman, G.; Uberuaga, B. P.; Jónsson, H. Climbing Image
Nudged Elastic Band Method for Finding Saddle Points and
Minimum Energy Paths. J. Chem. Phys. 2000, 113 (22), 9901−9904.
(67) Dickens, C. F.; Kirk, C.; Nørskov, J. K. Insights into the
Electrochemical Oxygen Evolution Reaction with Ab Initio
Calculations and Microkinetic Modeling: Beyond the Limiting
Potential Volcano. J. Phys. Chem. C 2019, 123 (31), 18960−18977.
(68) Dauenhauer, P. J.; Shetty, M.; Ardagh, M. A.; Pang, Y.;
Abdelrahman, O. A. Electric-Field-Assisted Modulation of Surface
Thermochemistry. ACS Catal. 2020, 10 (21), 12867−12880.
(69) Shetty, M.; Walton, A.; Gathmann, S. R.; Ardagh, M. A.;
Gopeesingh, J.; Resasco, J.; Birol, T.; Zhang, Q.; Tsapatsis, M.;
Vlachos, D. G.; et al. The Catalytic Mechanics of Dynamic Surfaces:
Stimulating Methods for Promoting Catalytic Resonance. ACS Catal.
2020, 10 (21), 12666−12695.
(70) Hanifpour, F.; Canales, C. P.; Fridriksson, E. G.;
Sveinbjörnsson, A.; Tryggvason, T. K.; Lewin, E.; Magnus, F.;
Ingason, Á . S.; Skúlason, E.; Flosadóttir, H. D. Investigation into the
Mechanism of Electrochemical Nitrogen Reduction Reaction to
Ammonia Using Niobium Oxynitride Thin-Film Catalysts. Electro-
chim. Acta 2022, 403, 139551.

6129 https://doi.org/10.1021/acs.jpclett.2c00769
J. Phys. Chem. Lett. 2022, 13, 6123−6129

You might also like