You are on page 1of 6

pubs.acs.

org/JPCC Article

Oxidizing Ethanol and 2‑Propanol by Hypochlorous Acid Generated


from Chloride Ions on HxWO3 Photoelectrodes
Andrew G. Breuhaus-Alvarez, Siqi Li, Nathaniel Z. Hardin, and Bart M. Bartlett*
Cite This: J. Phys. Chem. C 2021, 125, 26307−26312 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.

ABSTRACT: Photoelectrochemically oxidizing aqueous chloride


electrolytes (1 M NaCl, pH 3) to hypochlorous acid proceeds on
Downloaded via UNIV OF WISCONSIN-MADISON on June 23, 2022 at 15:10:35 (UTC).

HxWO3 semiconducting photoelectrodes using 405 nm illumination


at 100 mW/cm2 with 1.0 V versus Ag/AgCl constant potential
chronoamperometry. The hypochlorous acid serves as an oxidation
agent upon adding 100 mM ethanol or 2-propanol into the solution.
1
H NMR spectroscopy shows that 90% of the charge passed results in
oxidized products (acetaldehyde and acetic acid or acetone).
Chemically oxidizing ethanol by calcium hypochlorite shows that
both light and the HxWO3 surface enhance the reaction rate.

■ INTRODUCTION
Solar energy conversion to form fuels and commodity
vast abundance in terrestrial seawater. Despite its suitability to
carry out mediated reactions for photoelectrochemical syn-
thesis, using chloride is largely scarce.
chemicals remains a pressing, active area of research, with
Here, we report ethanol and 2-propanol oxidation by
photoelectrochemical water splitting as the most well-studied
photoelectrochemically generated hypochlorous acid from
example given the abundance of water on the Earth’s surface.1
aqueous chloride solution, mimicking seawater, on an n-type
During this reaction, hydrogen evolution is coupled with HxWO3 photoelectrode with ∼90% yield and stable, high
oxygen evolution, a four-proton/four-electron oxidation current density. The material also presents excellent stability in
reaction with a large overpotential.2,3 As a result, oxygen acid, which is beneficial to the coupled target cathodic
evolution is rate-limiting, the gas has little value, and it poses hydrogen evolution reaction.


an explosion risk with hydrogen if not properly separated.
As a result, replacing oxygen evolution with faster anodic
METHODS
reactions that provide valuable products has been inves-
tigated.4 Among them, oxidizing alcohols to aldehydes, Chemicals. The water used in all synthesis and experiments
ketones, and carboxylic acids is of high importance in was filtered by a Millipore filtration system (18 MΩ·cm).
producing pharmaceuticals and fine chemicals, especially for Ethanol (200 proof) was purchased from Decon Laboratories.
the likes of ethanol, which can be derived from biomass Ammonium metatungstate (NH4)6H2W12O40·xH2O (AMT)
feedstock.5−9 While alcohol oxidation is strategically favored, and poly(ethylene glycol) (Mw = 300 Da, PEG-300) were
direct alcohol oxidation on photoelectrodes usually requires purchased from Sigma-Aldrich. Hydrochloric acid (37 w/w %)
large applied bias, and the reactant and/or intermediates are and sodium chloride were purchased from Fisher Scientific.
prone to adsorb to the electrode.10 On the other hand, indirect Fluorine tin oxide (FTO, Pilkington Glass TEC-15) substrate
oxidation by introducing redox mediators circumvents these was cut into 1.5 × 2.54 cm2 strips for HxWO3 spin coat
problems. The mediator is rapidly oxidized at the anode, which synthesis. The FTO substrate was first coarsely cleaned by
subsequently oxidizes the alcohol in solution.11 Nitroxyl scrubbing with an acetone wetted Kimwipe. This cleaning
radicals12 and metal complexes13 are usually employed as procedure was followed by sonicating in the following solvents:
acetone, Sparkleen detergent, water, and acetone again for 10
mediators for alcohol oxidation.
min each, with each solvent wash, followed by drying under a
However, these reagents require using organic solvents, and
there are possible detrimental environmental impacts from the
heavy metal cations. From a green chemistry perspective, more Received: July 14, 2021
abundant and environmentally friendly mediators are sought. Revised: November 5, 2021
Inspired by industrial wastewater treatment, where bleach is Published: November 22, 2021
used to carry out oxidative destruction of organic waste in
aqueous solution,14,15 we set out to explore chloride as a redox
mediator for the targeted ethanol oxidation reaction due to its

© 2021 American Chemical Society https://doi.org/10.1021/acs.jpcc.1c06286


26307 J. Phys. Chem. C 2021, 125, 26307−26312
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

flowing N2 stream. Ag/AgCl reference electrodes were off and the cell was allowed to sit under stirring for 1 h before
purchased from CH Instruments and filled with saturated being placed in a −10 °C fridge for 10 min. The cell was then
KCl aqueous solution. Pt disk electrodes (2 mm diameter) removed and shaken to dissolve any alcohol oxidation product
were also purchased from CH Instruments. 99.95% pure Pt condensed on the cell walls. The 1H NMR sample was then
wire (24 gauge, P/N 1981) was purchased from Sure Pure made from the working compartment electrolyte.
1
Chemetals. H NMR Quantification of Alcohol Oxidation Prod-
Synthesis of HxWO3 Electrodes. HxWO3 electrodes were ucts. Quantification of products was carried out using
synthesized by spin coating AMT sol. The precursor was quantitative single pulse 1H NMR performed on 700 MHz
prepared by adding 2.51 g (833 μg) of AMT to a 100 mL Varian NMR using a relaxation delay of 25 s and a pulse angle
round-bottom flask and then 10 mL of water to dissolve of 90°. A typical NMR sample was composed of 200 μL of 100
completely with vigorous stirring. In a scintillation vial, 6.6 g mM sodium formate dissolved in D2O as an internal standard
(22 mmol) of PEG-300 was dissolved in 10 mL of ethanol. and 300 μL of the electrolyte from the working compartment.
Using a syringe pump, the ethanol solution was added at a rate The added sodium formate results in a 40 mM formic acid
of 120 mL·h−1. A faint, off-white sol precursor resulted, which standard. Because the acetaldehyde product can react with
was 0.5 M tungsten. The electrodes were prepared by dropping water to form hydrate in solution, the total yield of
70 μL of the precursor sol onto a clean, 1.5 × 1.5 cm2 square acetaldehyde reported was the sum of the observed
area (2.25 cm2) of FTO, which was masked off on a 1.5 × 2.54 acetaldehyde and acetaldehyde hydrate yields.16
cm2 strip. This was spun at 2500 rpm for 30 s using a Laurell UV−Vis Spectrophotometry. UV−vis solution absorp-
spin coater. The precursor was constantly and vigorously tion spectra were recorded using a Varian Cary 5000
stirred and then sealed tightly when not in use. Following spin spectrophotometer. A PEC CPC experiment was performed
coat deposition, the electrodes were transferred to a muffle using HxWO3 in both 1 M NaCl at pH 3 and 1 M HCl. No
furnace at 500 °C for 30 min. These deposition and annealing alcohol substrate was added. The HxWO3 electrode was poised
cycles were repeated 10 times to achieve films with at 1.0 V versus Ag/AgCl for 1 h. Following 1 h of electrolysis,
approximately 1 μm thickness. solution UV−vis absorption was performed on the undiluted
Material Characterization. XRD, SEM images, and electrolyte. Solution UV−vis absorption spectra were also
diffuse reflectance UV−vis spectroscopy were collected and obtained for Ca(ClO)2 solutions in order to observe the HClO
are presented in the Supporting Information. XRD was absorption feature. 71.5 mg of Ca(ClO)2 was added to a 25
performed on a Panalytical Empyrean diffractometer. The mL volumetric flask. The volumetric flask was then filled to the
operating power was 1.8 kW (45 kV, 40 mA) using Cu Kα (λ = line using either 1 M NaCl or 1 M HCl to make 20 mM
1.5418 Å) in θ−θ geometry. The SEM images were collected solution of Ca(ClO)2. Once all the solid had dissolved, the
with a JEOL JSM-7800FLV field emission scanning electron solution was quickly diluted by a factor of 10 using the
microscope. An operating voltage of 15.00 kV and an appropriate 1 M chloride electrolyte. The solution UV−vis
Everhart−Thornley secondary electron detector were used. spectrum shown in Figure S10 was recorded immediately after
Diffuse reflectance UV−vis spectroscopy was measured on a the solution was made. In the case of 1 M NaCl, the pH was
Varian Cary 5000 spectrophotometer with an external diffuse set to 3 after the dilution and immediately prior to the UV−vis
reflectance accessory and barium sulfate as the 100% measurement. When investigating the comproportionation of
reflectance standard. The spectra were collected in reflectance hypochlorous acid and chloride shown in Figure S16, 20 mM
mode and converted into absorbance by the Kubelka−Munk Ca(ClO)2 solution was made in 1 M NaCl and the pH was set
function. to 3. This solution was allowed to react for 45 min. At the end
Photoelectrochemistry. The photoelectrochemical of the 45 min period, the solution pH had increased and was
(PEC) experiments were performed using a CH Instruments set back to 3 before recording the UV−vis spectrum.
Series 760E electrochemical workstation. The light source used Chemical Oxidation of Ethanol by Ca(ClO)2. All
consisted of four 405 nm LEDs (Osram Sylvania, LZ1- samples were prepared in 5 mL vials. A 20 mM Ca(ClO)2
10UB00-01U7) attached to a CPU heat sink and powered by a solution was made by dissolving 286.0 mg of Ca(ClO)2 in 100
700 mA constant current LED driver (ERP Power, ESS015W- mL of 1 M NaCl solution. This concentration of Ca(ClO)2
0700-18). The 405 nm light was adjusted to an irradiance of produces a HClO concentration similar to the amount
100 mW·cm−2 with a thermopile detector (Newport 818P- generated during the PEC experiments presented in this
015-19) and an optical power meter (Newport 1918-R). The paper. The pH was then immediately set to 3, and this was
custom-made photoelectrochemical cell was glass with quartz combined with 460.7 mg of ethanol (100 mM). The
light windows and featured two compartments with a Nafion photochemical reactions were performed with the 405 nm
membrane separating the working electrode and counter light entering the bottom of the dram vial set to 100 mW·cm−2
electrode compartments. The Nafion membrane was used in irradiance. Experiments with tungsten oxide powder were
all PEC experiments reported in this work. The HxWO3 performed by adding 5 mg of the tungsten oxide powder to the
electrode was positioned such that it was backlit by the LED 4 mL reaction volume. At the end of the reaction, the WO3
light through the cell’s quartz window. Ag/AgCl reference powder was filtered out before analyzing the solution with
electrodes and Pt wire counter electrodes were used in all quantitative 1H NMR spectroscopy. Three trials were averaged
experiments. The electrolytes used in this work were 1 M NaCl in reporting the results.
set to pH 3 with HCl and 1 M Na2SO4 set to pH 3 with Starch−Iodine Titration. Quantification of free chlorine
sulfuric acid, both prepared in a volumetric flask before pH (HClO and Cl2) was carried out using starch−iodine titration.
adjustment. The alcohol substrate, either ethanol or 2- The starch−iodine test does not discriminate between HClO
propanol, was added to the electrolyte to make 100 mM and Cl2, both of which are two-electron oxidation products of
alcohol solution. Following the 3 h controlled potential Cl− oxidation. Immediately following a PEC CPC experiment,
chronoamperometry (CPC) experiments, the light was shut the working compartment solution was used to fill a 25.00 mL
26308 https://doi.org/10.1021/acs.jpcc.1c06286
J. Phys. Chem. C 2021, 125, 26307−26312
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

volumetric flask containing KI(s) (415.0 mg, 2.5 mM) to the


line to make a solution with an iodide concentration of 100
mM. Potassium iodide reacts with dissolved chlorine according
to the following equation
2I−(aq) + Cl 2(aq) → I 2(aq) + 2Cl−(aq)
or with hypochlorous acid
HClO(aq) + 2I−(aq) + H+(aq)
→ Cl−(aq) + I 2(aq) + H 2O(l)
Both reactions result in a 1:1 correspondence between the 2-
electron oxidation product of Cl2 or HClO and I2. 5 mL of the
25 mL iodide/iodine solution was then titrated with a stock
Na2S2O3 solution and a starch indicator. The thiosulfate
Figure 1. LSV traces of the chloride electrolyte with and without
reduces the generated iodine according to the equation added alcohols. The solid traces are recorded under illumination, and
I 2(aq) + 2S2 O32 −(aq) → 2I−(aq) + S4 O6 2 −(aq) the dashed traces are the corresponding dark measurements.

For the 300 s CPC experiment, a 10 mM sodium thiosulfate


stock solution was used. A 100 mM sodium thiosulfate was ethanol did not show significant changes, indicating that
used for the 10,800 s CPC experiment. similar reactions are occurring on the electrode regardless of

■ RESULTS AND DISCUSSION


PEC Chloride Oxidation with the Presence of
the presence of alcohol. The UV−vis spectrum of the solution
recorded immediately after performing the PEC CPC
experiment (vide infra) for 1 h (Figure S10) further supports
Alcohols. Voltammetry. Previous work by our group has chloride oxidation to hypochlorous acid at pH 3. The peak
shown that chloride ions can be oxidized in acidic aqueous absorbance at 233 nm corresponds to HClO with no shoulder
solution on HxWO3 photoelectrodes17 to form hypochlorous centered at 229 nm from dissolved Cl2.19
acid (pKa = 7.4)18 with near-unity Faradaic efficiency The Tafel plot (Figure S11) of both 1 M NaCl at pH 3 and
according to 1 M HCl electrolytes recorded on HxWO3 semiconducting
photoelectrodes exhibits a slope of ∼35 mV·dec−1, evidence
Cl−(aq) + H 2O(l) → HClO(aq) + H+(aq) + 2e−, that the rate-determining step of chloride oxidation is not
affected by large changes in proton activity and that HClO is
E° = 1.49 − 0.0296 × pH VNHE (1) ultimately the favored product of chloride oxidation on
In this work, we performed a series of PEC experiments on HxWO3 at these pH values. Our solution is unbuffered, and
HxWO3 photoelectrodes in the chloride electrolyte under 405 this result mitigates concern that the chloride oxidation
nm light at an irradiance of 100 mW·cm−2 to test the product is affected by swings in pH during CPC electrolysis.
possibility of mediating alcohol oxidation. Characterization of Controlled Potential Chronoamperometry. Oxidizing
the HxWO3 photoelectrode and quantification of the oxidized alcohols using hypochlorite anions has been previously
alcohol products are provided in the Supporting Information: demonstrated using TEMPO or nickel (II) salts as a catalyst,
1 suggesting that the HClO produced at the HxWO3 electrode
H NMR spectra (Figures S1−S5), XRD patterns (Figure S6),
SEM images (Figure S7), and a diffuse reflectance UV−vis may also carry out these oxidation reactions.20−24 To test the
spectrum with the associated Tauc plot (Figures S8 and S9). efficiency of the chloride oxidation coupled alcohol oxidation
Figure 1 shows the linear sweep voltammetry (LSV) of the 1 process, we conducted CPC experiments. Figure 2a shows the
M NaCl electrolyte with 100 mM ethanol or 2-propanol on reaction progress from quantitative 1H NMR spectroscopy
HxWO3. It can be observed that under illumination, the j−E over the 8 1/2 hour CPC experiment. Despite the large
curves overlap with or without the presence of alcohol, chloride concentration, we detect no peak splitting indicative
indicating that the same chemical reaction occurs. From this of chlorinated products by 1H NMR spectroscopy. Accord-
data, we surmise that chloride oxidation, and not direct alcohol ingly, the relevant reactions for generating acetaldehyde and
oxidation, occurs at the HxWO3 electrode. Moreover, in the acetic acid from ethanol and HClO are
LSV experiments using either 1 M Na2SO4 or 1 M NaClO4 CH3CH 2OH + HClO → CH3CHO + H+ + Cl− + H 2O
electrolytes, oxoanions with the central atoms in their highest
(2)
formal oxidation states show an increase in the j−E response
and an ∼200 mV shift to more negative potential upon adding CH3CH 2OH + 2HClO
the 100 mM ethanol substrate (Figure S15). This shift is the
same for both oxoanions, hinting that the anodic reaction is → CH3CH 2O + 2H+ + 2Cl− + H 2O (3)
likely direct ethanol oxidation because both the persulfate/ The yield of the oxidized ethanol product shown in Figure
sulfate and perchlorate anion/perchlorate radical couples are 2b is calculated according to the equation
>2 V versus NHE. The significant difference between the LSV
with ethanol in the chloride electrolyte and the two oxoanion F(2 × nacetaldehyde + 4 × nacetic acid)
electrolytes further indicates that ethanol is not directly yield = × 100%
q (4)
oxidized with the presence of chloride (i.e., chloride acts to
mediate the reaction). Additionally, the Nyquist plots obtained where F is Faraday’s constant (96,485 C/mol), n is the number
under irradiation (Figure S16) with and without 100 mM of moles of the product formed, and q is the charge passed in
26309 https://doi.org/10.1021/acs.jpcc.1c06286
J. Phys. Chem. C 2021, 125, 26307−26312
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

cm2, which is the average current density observed during the


PEC CPC experiment (Figure 2b).
Furthermore, the data in Figure 2 show that acetaldehyde
essentially behaves as an extremely long-lived intermediate
during ethanol oxidation to acetic acid. When we used the
same experimental conditions in a 16 h PEC CPC experiment
(j−t curve shown in Figure S13), 1H NMR spectroscopy
revealed a final solution composition of 3 mM acetaldehyde
(2% yield) and 75 mM acetic acid (86% yield).
Secondary alcohols can also be oxidized to ketones, and
Table 1 shows the product yields following a 3 h CPC

Table 1. Oxidation of 100 mM Ethanol and 2-Propanol in 1


M NaCl and Na2SO4 Electrolytes at pH 3
i
EtOH PrOH
electrolyte % MeCHO % MeCOOH % Me2CO
NaCl 50 ± 3 39 ± 2 92 ± 4
Na2SO4 16 ± 2 7±4 16 ± 2

experiment described previously with either 100 mM ethanol


or 2-propanol in 1 M NaCl and 1 M Na2SO4 electrolytes as a
control experiment. Unsurprisingly, a single product, acetone,
results from 2-propanol oxidation. Plots of the photocurrent
Figure 2. Ethanol oxidation using a HxWO3 photoelectrode at 1 V response during these PEC CPC experiments are shown in
versus Ag/AgCl under illumination. (a) Disappearance of reactants Figure S14. In 1 M NaCl, both ethanol and 2-propanol
and appearance of products as a function of time as measured by oxidation by HClO generated on HxWO3 proceeds with 90%
quantitative 1H NMR spectroscopy and (b) CPC trace and total of the charge passed contributing to an oxidized product.
product yield as a function of time. Previous work by our group has shown that the oxygen
evolution reaction is the dominant reaction on HxWO3 in the
Coulombs at the HxWO3 electrode at a given time point. At sulfate electrolyte,25 which should result in a lower yield for
early time points (low charge passed), a low yield is observed. alcohol oxidation in the sulfate electrolyte. Indeed, in a control
It increases until ∼200 C of charge pass at 2.5 h, at which point experiment containing 1 M Na2SO4, we observe significantly
the yield begins to stabilize ∼90%. A starch−iodine test shows lower conversion. This result further indicates the superiority
that HClO is present throughout the experiment, indicating of our chloride-oxidation-based indirect ethanol oxidation
that, not surprisingly, the rates of the solution-based chemical mechanism.
steps are slower than that of PEC chloride oxidation at the Decoupling Dark, Homogeneous Chemical Oxidation
HxWO3 electrode. The HClO concentration increases to 0.2 Steps from Light−Driven Reactions. The irradiation used
mM after passing 150 C of charge but increases by only about in the PEC CPC experiments is an intense energy input into
50 μM after another 150 C of charge was passed. As the HClO the reaction. Furthermore, metal oxide surfaces are well known
concentration increases, so does the rate of the subsequent to serve as heterogeneous oxidation catalysts.26,27 Therefore,
chemical oxidation steps until the rate is rapid enough to keep we carried out a series of experiments to decouple the PEC
pace with the rate of HClO generation to maintain a consistent chloride oxidation from the chemical steps of alcohol
yield of 90%. oxidation, as shown in Figure 3. First, as a control experiment,
Initially, we note that the rate of acetaldehyde formation is we combined the 100 mM ethanol solution in 1 M NaCl and
greater than the rate of acetic acid formation. However, as the
acetaldehyde concentration increases, the rate of its subsequent
oxidation to acetic acid also increases. Under these
experimental conditions, the plateau occurs when acetaldehyde
reaches 20 mM concentration. To test that acetaldehyde is at
steady state under these conditions, a linear fit (zero-order
kinetics) was performed on the concentration of both ethanol
and acetic acid versus time kinetic profile after 400 C of charge
was passed (Figure S12). The slopes d[EtOH]/dt and
d[CH3COOH]/dt are (−1.77 ± 0.04) × 10−6 M/s and
(1.53 ± 0.02) × 10−6 M/s, respectively. These rates are in
excellent agreement, and we note any discrepancy is likely due
to the volatility of acetaldehyde lost in the headspace. From the
two fits, the rate at which both ethanol and acetaldehyde are
oxidized is (3.30 ± 0.04) × 10−6 M/s. The current density Figure 3. Product yield of 100 mM ethanol oxidation after 3 h under
corresponding to this rate from Faraday’s law given the 2.25 varying conditions to distinguish the activity of WO3, HClO (at pH
cm2 geometric area of our HxWO3 electrodes is 9.9 ± 0.3 mA/ 3), and illumination on the yield and selectivity.

26310 https://doi.org/10.1021/acs.jpcc.1c06286
J. Phys. Chem. C 2021, 125, 26307−26312
The Journal of Physical Chemistry C


pubs.acs.org/JPCC Article

adjusted the pH to 3 with commercial WO3 powder (5 mg in 4 ASSOCIATED CONTENT


mL reaction volume) and observed no conversion after *
sı Supporting Information
reacting for 3 h in the dark. We conclude that ethanol The Supporting Information is available free of charge at
oxidation by dissolved O2 on WO3 does not occur. Next, we https://pubs.acs.org/doi/10.1021/acs.jpcc.1c06286.
carried out chemical oxidation of 100 mM ethanol by 20 mM 1
H NMR spectra, XRD patterns, SEM images, UV−vis
Ca(ClO)2 in 1 M NaCl solution at pH 3, still in the dark, to
spectra, Tafel plots, GC data, and CPC and LSV traces
measure the propensity for HClO to serve as the oxidant. The
(PDF)
pH was set after all the salts were dissolved, and the 4 mL
solution reacted for 3 h in a sealed vial. The yields of
acetaldehyde and acetic acid are 11.3 ± 0.6 and 6.8 ± 0.7%,
respectively.
■ AUTHOR INFORMATION
Corresponding Author
Then, adding WO3 powder to a solution containing 20 mM Bart M. Bartlett − Department of Chemistry, University of
Ca(ClO)2, still in the dark, results in the doubling of the Michigan, Ann Arbor, Michigan 48109-1055, United States;
acetaldehyde yield to 26.0 ± 0.2% and an increase in the yield orcid.org/0000-0001-8298-5963; Email: bartmb@
of acetic acid to 10.2 ± 0.2%. The WO3 powder enhances the umich.edu
rate of oxidation by HClO, hinting that the surface of our
Authors
HxWO3 semiconductor photoelectrodes may assist in the PEC
Andrew G. Breuhaus-Alvarez − Department of Chemistry,
CPC conversion of ethanol to acetaldehyde and acetic acid. University of Michigan, Ann Arbor, Michigan 48109-1055,
Finally, we illuminated the reaction mixture, now without WO3 United States
powder, with 405 nm light at 100 mW/cm2 entering the base of Siqi Li − Department of Chemistry, University of Michigan,
the vial reaction. In this experiment, the yields of acetaldehyde Ann Arbor, Michigan 48109-1055, United States
and acetic acid increased to 29.3 ± 0.4 and 10.7 ± 0.2%, Nathaniel Z. Hardin − Department of Chemistry, University
respectively. The solution temperature was 23.3 °C in the dark of Michigan, Ann Arbor, Michigan 48109-1055, United
and was measured at 24.2 °C during the irradiated experiments States
(nearly constant). We therefore attribute the increase in yield
Complete contact information is available at:
to the light rather than the <1 °C increase in temperature. https://pubs.acs.org/10.1021/acs.jpcc.1c06286
Moreover, hypochlorous acid undergoes homolytic bond
cleavage with near-UV light to generate the highly reactive Notes
Cl• and OH• radicals, both strong oxidants.28−30 While the The authors declare no competing financial interest.


major absorption maximum of the aqueous chlorine-containing
species (i.e., - Cl2, HClO, and ClO−) exists at λ < 300 nm, ACKNOWLEDGMENTS
using longer wavelength light to generate radicals has been
This work was supported by the U.S. Department of Energy,
previously demonstrated.28,29,31−34 The UV−vis spectrum of Office of Science, Basic Energy Sciences, under Award DE-
20 mM solution of Ca(ClO)2 is shown in Figure S17. The 405 SC0006587 (Catalysis Science). A.G.B.-A. thanks the Horace
nm light output by the LED overlaps with the tail in the H. Rackham Graduate School at the University of Michigan for
absorption band at λmax ≈ 325 nm. a Rackham Merit Fellowship. The authors acknowledge the

■ CONCLUSIONS
We demonstrate that chloride can be oxidized photo-
financial support from the University of Michigan College of
Engineering and NSF grant #DMR-9871177 and #DMR-
0723032 and technical support from the Michigan Center for
material characterization.


electrochemically to form hypochlorous acid in acidic aqueous
solution on HxWO3 semiconductor photoelectrodes with 98%
REFERENCES
Faradaic efficiency. The hypochlorous acid performs ethanol
and 2-propanol oxidation in high yield; 90% of the charge (1) Nielander, A. C.; Shaner, M. R.; Papadantonakis, K. M.; Francis,
S. A.; Lewis, N. S. A taxonomy for solar fuels generators. Energy
passed results in alcohol oxidation. In the ethanol case, the Environ. Sci. 2015, 8, 16−25.
yield for acetaldehyde is 50 ± 3% and the yield for acetic acid (2) Lhermitte, C. R.; Sivula, K. Alternative Oxidation Reactions for
is 39 ± 2% after 3 h. Then, 86% acetic acid results after 16 h. 2- Solar-Driven Fuel Production. ACS Catal. 2019, 9, 2007−2017.
Propanol oxidation produces acetone with 92 ± 4% yield. (3) McCrory, C. C. L.; Jung, S.; Ferrer, I. M.; Chatman, S. M.;
Using a sulfate electrolyte dramatically suppresses the alcohol Peters, J. C.; Jaramillo, T. F. Benchmarking Hydrogen Evolving
oxidation yield, showing that chloride oxidation must precede Reaction and Oxygen Evolving Reaction Electrocatalysts for Solar
Water Splitting Devices. J. Am. Chem. Soc. 2015, 137, 4347−4357.
alcohol oxidation. Furthermore, we used Ca(ClO)2 to
(4) Bender, M. T.; Yuan, X.; Choi, K. S. Alcohol oxidation as
demonstrate that the tungsten oxide surface catalyzes the alternative anode reactions paired with (photo)electrochemical fuel
conversion of alcohol to acetaldehyde and acetic acid. The 405 production reactions. Nat. Commun. 2020, 11, 4594.
nm, 100 mW/cm2 light used in the CPC PEC experiments was (5) Ley, S. V.; Norman, J.; Griffith, W. P.; Marsden, S. P.
also shown to increase the rate of ethanol conversion, likely by Tetrapropylammonium Perruthenate, Pr4N+RuO4−, TPAP : A
increasing the •OH and •Cl radical concentrations through Catalytic Oxidant for Organic Synthesis. Synthesis 1994, 1994,
photoinduced homolytic cleavage of hypochlorous acid. Our 639−666.
(6) Arends, I. W. C. E.; Sheldon, R. A. Modern Oxidation of
work reveals the capability to perform more attractive alcohol Alcohols Using Environmentally Benign Oxidants. Modern Oxidation
oxidation reactions coupled with the PEC chloride oxidation, Methods; John Wiley & Sons, Ltd, 2005; Chapter 4; pp 83−118.
which is an overall more valuable anodic reaction to couple (7) Hudlicky, M. Oxidations in Organic Chemistry; American
with the hydrogen evolution reaction in fuel-forming systems. Chemical Society; Washington DC, 1990.

26311 https://doi.org/10.1021/acs.jpcc.1c06286
J. Phys. Chem. C 2021, 125, 26307−26312
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

(8) Davis, S. E.; Ide, M. S.; Davis, R. J. Selective Oxidation of Selective Oxidation of Primary Alcohols. Nat. Commun. 2017, 8,
Alcohols and Aldehydes over Supported Metal Nanoparticles. Green 14039.
Chem. 2013, 15, 17−45. (28) Remucal, C. K.; Manley, D. Emerging investigators series: the
(9) Zerk, T. J.; Moore, P. W.; Harbort, J. S.; Chow, S.; Byrne, L.; efficacy of chlorine photolysis as an advanced oxidation process for
Koutsantonis, G. A.; Harmer, J. R.; Martínez, M.; Williams, C. M.; drinking water treatment. Environ. Sci.: Water Res. Technol. 2016, 2,
Bernhardt, P. V. Elucidating the Mechanism of the Ley-Griffith 565−579.
(TPAP) Alcohol Oxidation. Chem. Sci. 2017, 8, 8435−8442. (29) Bulman, D. M.; Mezyk, S. P.; Remucal, C. K. The Impact of pH
(10) DiMeglio, J. L.; Terry, B. D.; Breuhaus-Alvarez, A. G.; Whalen, and Irradiation Wavelength on the Production of Reactive Oxidants
M. J.; Bartlett, B. M. Base-Assisted Nitrate Mediation as the during Chlorine Photolysis. Environ. Sci. Technol. 2019, 53, 4450−
Mechanism of Electrochemical Benzyl Alcohol Oxidation. J. Phys. 4459.
Chem. C 2021, 125, 8148−8154. (30) Yan, X.; Cheng, H.; Zare, R. N. Two-Phase Reactions in
(11) Badalyan, A.; Stahl, S. S. Cooperative electrocatalytic alcohol Microdroplets without the Use of Phase-Transfer Catalysts. Angew.
oxidation with electron-proton-transfer mediators. Nature 2016, 535, Chem., Int. Ed. 2017, 56, 3562−3565.
406−410. (31) Izadifard, M.; Langford, C.; Achari, G. Alternative View of
(12) de Nooy, A. E. J.; Besemer, A. C.; van Bekkum, H. Selective Chlorine Oxidation Stimulated by Longer Wavelength Light. J.
oxidation of primary alcohols mediated by nitroxyl radical in aqueous Environ. Eng. 2016, 142, 04016048.
solution. Kinetics and mechanism. Tetrahedron 1995, 51, 8023−8032. (32) Forsyth, J. E.; Zhou, P.; Mao, Q.; Asato, S. S.; Meschke, J. S.;
(13) Shapley, P. A.; Zhang, N.; Allen, J. L.; Pool, D. H.; Liang, H.-C. Dodd, M. C. Enhanced Inactivation of Bacillus Subtilis Spores During
Selective alcohol oxidation with molecular oxygen catalyzed by Os− Solar Photolysis of Free Available Chlorine. Environ. Sci. 2013, 47,
Cr and Ru− Cr complexes. J. Am. Chem. Soc. 2000, 122, 1079−1091. 12976−12984.
(14) Chung, M.; Jin, K.; Zeng, J. S.; Manthiram, K. Mechanism of (33) Shu, Z.; Li, C.; Belosevic, M.; Bolton, J. R.; El-Din, M. G.
Chlorine-Mediated Electrochemical Ethylene Oxidation in Saline Application of a Solar UV/Chlorine Advanced Oxidation Process to
Water. ACS Catal. 2020, 10, 14015−14023. Oil Sands Process-Affected Water Remediation. Environ. Sci. 2014, 48,
(15) Behin, J.; Akbari, A.; Mahmoudi, M.; Khajeh, M. Sodium 9692−9701.
hypochlorite as an alternative to hydrogen peroxide in Fenton process (34) Horton, J. A.; Laura, M. A.; Kalbag, S. M.; Petterson, R. C.
for industrial scale. Water Res. 2017, 121, 120−128. Reaction of Hypochlorous Acid with Ketones. Novel Baeyer-Villiger
(16) Scheithauer, A.; Brächer, A.; Grützner, T.; Zollinger, D.; Thiel, Oxidation of Cyclobutanone with Hypochlorous Acid. J. Org. Chem.
W. R.; Von Harbou, E.; Hasse, H. Online 1H NMR Spectroscopic 1969, 34, 3366−3368.
Study of the Reaction Kinetics in Mixtures of Acetaldehyde and Water
using a New Microreactor Probe Head. Ind. Eng. Chem. Res. 2014, 53,
17589−17596.
(17) Breuhaus-Alvarez, A. G.; Cheek, Q.; Cooper, J. J.; Maldonado,
S.; Bartlett, B. M. Chloride Oxidation as an Alternative to the Oxygen
Evolution Reaction on HxWO3 Photoelectrodes. J. Phys. Chem. C
2021, 125, 8543−8550.
(18) Perrin, D. D. Ionisation Constants of Inorganic Acids and Bases in
Aqueous Solution; Elsevier Science: Burlington, 2013.
(19) Belz, M.; Boyle, W. J. O.; Klein, K.-F.; Grattan, K. T. V. Smart-
sensor approach for a fibre-optic-based residual chlorine monitor.
Sens. Actuators, B 1997, 39, 380−385.
(20) Lucio Anelli, P.; Biffi, C.; Montanari, F.; Quici, S. Fast and
selective oxidation of primary alcohols to aldehydes or to carboxylic
acids and of secondary alcohols to ketones mediated by
oxoammonium salts under two-phase conditions. J. Org. Chem.
1987, 52, 2559−2562.
(21) Grill, J. M.; Ogle, J. W.; Miller, S. A. An Efficient and Practical
System for the Catalytic Oxidation of Alcohols, Aldehydes, and α,β-
Unsaturated Carboxylic Acids. J. Org. Chem. 2006, 71, 9291−9296.
(22) Cheng, L.; Wang, J.; Wang, M.; Wu, Z. Mechanistic Insight into
the Alcohol Oxidation Mediated by an Efficient Green [CuBr2(2,2′-
bipy)]-TEMPO Catalyst by Density Functional Method. Inorg. Chem.
2010, 49, 9392−9399.
(23) Liu, A. T.; Kunai, Y.; Cottrill, A. L.; Kaplan, A.; Zhang, G.; Kim,
H.; Mollah, R. S.; Eatmon, Y. L.; Strano, M. S. Solvent-Induced
Electrochemistry at an Electrically Asymmetric Carbon Janus Particle.
Nat. Commun. 2021, 12, 3415.
(24) Das, A.; Stahl, S. S. Noncovalent Immobilization of Molecular
Electrocatalysts for Chemical Synthesis: Efficient Electrochemical
Alcohol Oxidation with a Pyrene-TEMPO Conjugate. Angew. Chem.,
Int. Ed. 2017, 56, 8892−8897.
(25) Breuhaus-Alvarez, A. G.; DiMeglio, J. L.; Cooper, J. J.;
Lhermitte, C. R.; Bartlett, B. M. Kinetics and Faradaic Efficiency of
Oxygen Evolution on Reduced HxWO3 Photoelectrodes. J. Phys.
Chem. C 2019, 123, 1142−1150.
(26) Védrine, J. C. Metal Oxides in Heterogeneous Oxidation
Catalysis: State of the Art and Challenges for a More Sustainable
World. ChemSusChem 2019, 12, 577−588.
(27) Zhao, G.; Yang, F.; Chen, Z.; Liu, Q.; Ji, Y.; Zhang, Y.; Niu, Z.;
Mao, J.; Bao, X.; Hu, P.; et al. Metal/Oxide Interfacial Effects on the

26312 https://doi.org/10.1021/acs.jpcc.1c06286
J. Phys. Chem. C 2021, 125, 26307−26312

You might also like