You are on page 1of 7

pubs.acs.

org/JPCC Article

Base-Assisted Nitrate Mediation as the Mechanism of


Electrochemical Benzyl Alcohol Oxidation
John L. DiMeglio, Bradley D. Terry, Andrew G. Breuhaus-Alvarez, Matthew J. Whalen,
and Bart M. Bartlett*
Cite This: J. Phys. Chem. C 2021, 125, 8148−8154 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: Nitrate anion (NO3−) oxidation to nitrate radical


Downloaded via UNIV OF WISCONSIN-MADISON on June 23, 2022 at 15:09:59 (UTC).

(NO3•) is chemically irreversible in acetonitrile (MeCN) solvent


due to solvent-based hydrogen-atom transfer (HAT). Introducing
benzyl alcohol (PhCH2OH) leads to competition with MeCN for
electrochemically generated NO3• and affords benzaldehyde
(PhCHO) product with ∼80% faradaic efficiency (FE) in 250
mM PhCH2OH. Stoichiometric HNO3 forms during HAT
reactions (observed by UV−vis spectroscopy) and exists as an
electrochemically inert and weak electrolyte; this off-cycle form of
nitrate can be reintroduced to the catalytic cycle upon
deprotonation by 2,6-lutidine while maintaining the base-free FE.
Oxygen reduction complements nitrate oxidation during catalysis
and reduced oxygen species (ROS) generated during proton-
coupled oxygen reduction are identified through rotating ring-disk electrochemistry; proton-coupled oxygen reduction indicates
ROS are capable of rendering NO3− catalytic when collocal. Directly observing ROS as the stoichiometric base generated during
nitrate anion oxidation resolves differences in photocatalytic vs photoelectrochemical reactivity of NO3− in base-free conditions and
points toward HAT as the general mode of reactivity for nitrate radical in acetonitrile solutions.

■ INTRODUCTION
Redox mediators are valuable tools used to facilitate electron-
received considerable attention, other classes remain woefully
underexplored.
transfer reactions, thus improving reaction rates and product Recently, indirect alcohol oxidation through nitrate radical
selectivity in a variety of catalytic transformations. This indirect (NO3•) formed from nitrate anion (NO3−) has been revisited
process has been implemented under electrochemical control with photocatalytic and photoelectrochemical demonstrations
to drive aliphatic C−H functionalization,1,2 aziridination of on CdS nanowires12 and BiVO4 photoanodes.13 On CdS
alkenes,3 alcohol oxidation, and more.4 Selective alcohol nanowires, NO3− mediation is observed; on BiVO4 photo-
oxidation is a challenging endeavor commonly achieved electrodes, NO3− reacts stoichiometrically. NO3• chemistry is
using toxic chromium-based oxidantssuch as chromium ubiquitously observed in atmospheric gas-phase oxidation
trioxide, dichromate, and pyridinium chlorochromatein the reactions involving remediation of volatile organic species
pursuit of selective aldehyde and carboxylic acid generation, of through hydrogen-atom abstraction or alkene addition.14
which serve as starting materials for polymer production and Electrochemical oxidation of NO3− was first reported in
fine-chemical synthesis.5 Implementing nontoxic redox medi- 1980, where NO3• generated at an electrode reacts under a
ators offers a safer alternative and can be designed to adhere to proposed hydrogen-atom transfer (HAT) mechanism.15 With
metal-free syntheses. N-oxyl mediators dominate the field of this revival of NO3• deployment in catalytic alcohol oxidation,
indirect alcohol oxidation due to their reversible redox the need for a thorough mechanistic investigation emerged.
properties, relatively low redox potentials, and high selectivity This work aims to describe the reactivity differences observed
for desired aldehyde and carboxylic acid products.6−9 These
homogenous catalysts (mediators) have been used to achieve Received: November 20, 2020
lignin oxidation, a nonedible biomass feedstock, yielding Revised: April 2, 2021
promise for sustainable petrochemical-free routes to fuel and Published: April 14, 2021
commodity chemical production.10,11 Given the bevy of
alcohol-derived targets, preparing suitable mediators to meet
this demand is imperative. Although N-oxyl mediators have

© 2021 American Chemical Society https://doi.org/10.1021/acs.jpcc.0c10476


8148 J. Phys. Chem. C 2021, 125, 8148−8154
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

and to explain the general behavior of NO3−/NO3• in enabled by the integration ratio of the analyte and
acetonitrile solutions using platinum (Pt) electrodes. chlorobenzene signals in relation to an obtained calibration

■ EXPERIMENTAL SECTION
Materials and Methods. Lithium nitrate (LiNO3), lithium
curve. For the FE vs time plot, 1H NMR spectroscopy (d3-
acetonitrile) was used to track the formation of benzaldehyde
product at 20 min time points using a known concentration of
nitromethane internal standard.


hexafluorophosphate (LiPF6), benzyl alcohol-α,α-d2, and nitric
acid (HNO3) were purchased from Sigma Aldrich; hydro-
chloric acid (HCl) and acetonitrile (MeCN) were purchased RESULTS
from Fisher Scientific; benzyl alcohol (PhCH2OH), chlor- To demonstrate the ability of electrochemically generated
obenzene, and 2,6-lutidine were purchased from Acros NO3• to facilitate indirect PhCH2OH oxidation, cyclic
Organics. All were used without further purification. voltammograms were collected on platinum electrodes in
Tetrabutylammonium hexafluorophosphate (Bu4NPF6) was MeCN solutions containing 50 mM lithium nitrate (LiNO3)
purchased from TCI and purified through hot ethanol and 100 mM tetrabutylammonium hexafluorophosphate
recrystallization. To achieve 50 mM LiNO3, mixtures of (Bu4NPF6) electrolyte. The cyclic voltammetry (CV) traces
LiNO3, MeCN, and PhCH2OH were prepared and then in Figure 1a highlight the anodic current achieved only when
sonicated under heat (∼50 °C) for 20 min to give a clear,
colorless solution.
Electrochemistry. Cyclic voltammetry (CV) was con-
ducted using either a CH Instruments Electrochemical
Workstation 660C, 760E, or 1000A with a CHI Pt (area =
0.0314 cm2) working electrode. Working electrodes were
polished using a slurry of 0.05 μm alumina powder in
deionized water on a MicroCloth pad. A silver wire was used as
the reference electrode and inserted into a sealed glass tube
with a Vycor tip and filled with the electrolyte solution. The
counter electrode was either a graphite rod or a Pt wire (∼0.33
cm2), of which the latter was cleaned between experiments by
dipping into aqua regia for 15 s. Unless otherwise stated,
voltammetry experiments were conducted at a scan rate of 25
mV s−1.
Constant potential chronoamperometry (CPC) experiments
were conducted using a custom-built two-compartment glass
cell separated by either an ultrafine glass frit or a Sterlitech
polypropylene filter paper with a pore size of 0.1 μm (Figure
S10). The working compartment contained 12 mL of solution,
and the counter compartment had 8 mL of solution. CPC
utilized a Pt wire working electrode (area: ∼0.35 cm2), with a
silver wire reference electrode and graphite counter as used in
the voltammetry experiments. Over the course of the CPC
experiment, the working compartment solution was agitated
using a stir bar.
Rotating ring-disk electrode (RRDE) experiments were
conducted at 30 Hz with a Pt ring and disk on a RRDE-3A
version 1.2 apparatus from CH Instruments. The ring and disk
were polished using a 0.05 μm Alpha Alumina slurry on a
MicroCloth pad. After polishing, the ring and disk were
sonicated in deionized water for 10 min. The ring and disk
were then inserted into the fixture and dipped in aqua regia for Figure 1. (a) CV on Pt electrodes in MeCN solutions of 100 mM
15 s, then rinsed with Millipore water. Finally, the RRDE lithium hexafluorophosphate (LiPF6), 100 mM Bu4NPF6 electrolyte,
assembly was spun at 30 Hz in acetonitrile for 5 min. The now or 50 mM LiNO3 at 25 mV s−1. (b) Scan rate-dependent CV traces
cleaned RRDE assembly was placed into 35 mL of the analyte on Pt in MeCN solutions containing 100 mM Bu4NPF6 supporting
solution using a graphite rod counter and silver wire reference electrolyte and 50 mM LiNO3 at 25, 50, 100, 250, 500, and 1000 mV
electrodes. s−1 from bottom (black) to top (red).
Product Analysis. Substrate and product concentrations
were measured using gas chromatography-flame ionization nitrate is in solution, corresponding to the one-electron
detection (GC-FID) by periodically removing an aliquot (10 oxidation of NO3− to NO3•. This nitrate redox couple is
μL) from the reaction mixture and combining with a irreversible, as evidenced by the lack of detectable cathodic
chlorobenzene standard (50 μL of a 21.6 mM standard) and current on reverse scans even at scan rates up to 1000 mV s−1
acetonitrile (940 μL). The samples were analyzed on either a (Figure 1b). The peak current for NO3− oxidation shifts to
Shimadzu QP-2010 GC/MS or Thermo Fisher Trace 13-10 more positive potentials with increasing scan rate due to
gas chromatograph equipped with a capillary column and FID quasireversibility and remains diffusion-controlled, noted by
detector. During analysis, the column oven was warmed from the linear dependence on the square root of scan rate (inset,
40 to 290 °C at a 10 °C min−1 ramp rate. Quantification was Figure 1b). The inability to detect any oxidized intermediates
8149 https://doi.org/10.1021/acs.jpcc.0c10476
J. Phys. Chem. C 2021, 125, 8148−8154
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

during the reverse scan is likely due to a solvent-based raphy-flame ionization detection (GC-FID) shows benzalde-
chemical reaction given the known high reactivity of NO3• in hyde (PhCHO) production with 80 ± 16% FE. This result
atmospheric chemistry and the reported ability of oxidized supports that the electrochemically generated NO3• can
electrolyte anions to react similarly with MeCN.14,16 facilitate PhCH2OH oxidation.
Linear sweep voltammetry (LSV) traces were collected in The irreversible NO3− oxidation involved in indirect
the presence of PhCH2OH to explore the reactivity of PhCH2OH oxidation mirrors the photoelectrochemical exam-
electrochemically generated NO3• in acetonitrile solutions. ple on BiVO4 films13 but differs from the mediation behavior
Surprisingly, the presence of alcohol substrate does not observed photochemically on CdS nanowires.12 The presence
influence the NO3−/NO3• voltammetric response (Figure 2a, of lithium cations was shown not to influence the voltammetry
(Figure 2a, red), confirming that the improvements afforded by
LiNO3 are due to nitrate anion.
To glean mechanistic insight regarding NO3− oxidation
coupled to a solution-based chemical step with PhCH2OH, we
collected concentration-dependent LSV traces. Figure 3a

Figure 2. (a) LSV recorded at 25 mV s−1 scan rate on Pt in MeCN


solutions containing 100 mM Bu4NPF6 under the listed conditions, Figure 3. (a) LSV on a Pt electrode in MeCN solutions (100 mM
where [PhCH2OH] = 250 mM and [LiNO3] = [LiPF6] = 50 mM; (b) Bu4NPF6) measuring current dependence on [LiNO3] and (b)
CPC performed at 1.68 V vs Fc+/0 under the listed conditions using a measuring current dependence on [PhCH2OH]. The scan rate is 25
Pt wire electrode. mV s−1.

green trace vs blue trace) suggesting that either NO3• is shows the current dependence on [LiNO3] while maintaining
incapable of oxidizing PhCH2OH or the reaction between a constant 250 mM PhCH2OH concentration; as the [NO3−]
NO3• and PhCH2OH does not regenerate NO3− (noted by the increases, the peak current increases in a linear fashion,
preservation in peak current). To identify NO3• reactivity revealing a first-order dependence on NO3− (Figure S2). In a
toward PhCH2OH oxidation, constant potential chronoamper- parallel experiment, increasing the [PhCH2OH] in solution
ometry (CPC) was performed at the most positive potential while keeping the [NO3−] constant at 50 mM (Figure 3b)
before PhCH2OH-only solutions show anodic current on Pt reveals a peak current that does not change with the
electrodes (1.68 V vs Fc+/0). Figure 2b displays the current concentration (or presence) of alcohol, i.e., the electrochemical
density that results from CPC in PhCH2OH solutions step is zero-order in PhCH2OH (Figure S3). Electrochemical
containing varying electrolytes while stirring to exclude mass NO3− oxidation is the rate-determining step and remains
transport limitations in the measured current. Without a nitrate distinct from the ensuing chemical steps in which PhCH2OH is
salt, current densities <0.3 mA cm−2 are observed, as predicted oxidized to PhCHO.
by LSV traces in Figure 2a. With a nitrate salt (LiNO3), Given the insensitivity to [PhCH2OH] during the electro-
current densities of ∼1.5 mA cm−2 result (additional trials are chemical step, the consumption of benzyl alcohol (and
provided in Figure S1). Product analysis by gas chromatog- formation of benzaldehyde) was tracked over time during a
8150 https://doi.org/10.1021/acs.jpcc.0c10476
J. Phys. Chem. C 2021, 125, 8148−8154
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

CPC experiment targeting high conversion. To effect a large Since NO3− oxidation remains irreversible in all solution
conversion, we performed a CPC reaction at 1.68 V vs Fc+/0 in conditions, another mechanism must be at play. The
acetonitrile containing 20 mM PhCH2OH, 100 mM Bu4NPF6, electrochemical system detailed here closely matches our
and 50 mM LiNO3 using a small two-compartment cell. Figure recent findings of nitrate-assisted (but not mediated) reactivity
4 displays the result of CPC where the concentrations of on BiVO4 photoelectrodes.13 In that report, the irreversibility
benzyl alcohol and benzaldehyde were tracked at 20 min time of the NO3−/NO3• couple is thought to result from
points by 1H NMR spectroscopy. stoichiometric nitric acid formation; HNO3 is a weak
electrolyte in acetonitrile (pKa, MeCN ∼ 9).18 To identify the
presence of HNO3 as an off-cycle species, we carried out CPC
in alcohol-free acetonitrile solutions with 50 mM LiNO3 at
1.68 V vs Fc+/0 for 16 h before investigating the nitrate
speciation using UV−vis spectroscopy (Figure 5). Before CPC,

Figure 5. UV−vis spectra of acetonitrile solutions before CPC (red)


and after CPC (blue) containing 50 mM LiNO3. The black trace is
from an acetonitrile solution with HNO3(aq) added externally.

only the NO3− n → π* transition at 300 nm in acetonitrile is


observed; after CPC, this peak is still present in addition to a
new peak at ∼270 nm. Externally added HNO3(aq) displays a
single peak in acetonitrile at 268 nm, supporting its formation
Figure 4. (a). Concentration of benzyl alcohol reactant (black) and
as a direct result of electrochemical NO3− oxidation in
benzaldehyde product (green) measured at 20 min intervals during acetonitrile. A similar system detailing UV-light-driven
CPC at 1.68 V vs Fc+/0 in acetonitrile solutions starting from 20 mM cyanation reports the n → π* transition for AgNO3 at 311
PhCH2OH, 50 mM LiNO3, and 100 mM Bu4NPF6. (b). First-order nm and 273 nm for CH3ONO2, closely aligning with neutral
kinetics plot of ln[PhCH2OH] vs time as measured from (a). (HONO2) and anionic nitrate (Bu4NNO3) assignments in
Figure 5.19
Following the identification of nitric acid in this system, its
At high conversion (> 50%), we note a static concentration electrochemical activity was probed by cyclic voltammetry.
of benzaldehyde product (Figure 4a), indicating a competitive HNO3 displays an oxidation potential > 1.75 V vs Fc+/0more
pathway for nitrate radical that occurs when the alcohol positive than the CPC operating potential of 1.68 V vs Fc+/0
concentration in solution is on the order of the aldehyde (Figure S4)representing a dead end for nitrate, rendering
concentration. The further oxidation of benzaldehyde has been the system stoichiometric. The formation of nitric acid from
reported, with volatile CO and benzene as products.17 The alcohol-free nitrate oxidation demonstrates the HAT mode of
experiment starting with 20 mM benzyl alcohol concludes with reactivity for nitrate radical toward acetonitrile. GC-MS
a FE ∼ 45%, much lower than the previous CPC starting with analysis reveals solvent-based oligomer formation occurred
250 mM yielding ∼80% FE; both loadings produced around both with and without PhCH2OH in solution as a result of
10 mM benzaldehyde, demonstrating that a high loading of nitrate radical’s reactivity (Figures S5 and S6), mirroring a
benzyl alcohol acts to suppress reactions with aldehyde report detailing acetonitrile polymerization coupled to
product. From Figure 4b, we observe first-order kinetics for perchlorate anion oxidation.16 This result explains the benzyl
the consumption of benzyl alcohol in the chemical step, alcohol loading-dependent FE due to competition for electro-
distinct from the zero-order dependence revealed by cyclic chemically generated nitrate radicals.
voltammetry (Figure 3). To liberate trapped off-cycle nitric acid in solution, a suitable
So far, the system-dependent stoichiometric behavior of stoichiometric Brønsted base is required. Previous catalysis
NO3• remains unexplained. If NO3• reacts with PhCH2OH by involving N-oxyl mediators and primary alcohols has shown
single-electron transfer to yield an oxygen-centered radical, that 2,6-lutidine does not interfere with Pt-based electro-
then one would expect to see NO3− regenerated, i.e., mediator catalysis,8 and its conjugate acid2,6-lutidiniumhas a
behavior in the voltammetry where the peak current increases. pKa,MeCN of 14.1, which should be sufficiently basic to
8151 https://doi.org/10.1021/acs.jpcc.0c10476
J. Phys. Chem. C 2021, 125, 8148−8154
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

deprotonate nitric acid in acetonitrile.20 Therefore, we predict With this depiction of base-assisted nitrate mediation, we
the balanced half-reaction for nitrate-catalyzed benzyl alcohol then questioned how seemingly base-free mediation on CdS
oxidation at the anode to be: nanowires was observed photochemically. We applied rotating
ring-disk electrochemistry (RRDE) on Pt to track the species
present (or formed) near the electrode surface as a result of
electrochemical NO3− oxidation. In the RRDE experiments, we
The red traces of Figure 6a demonstrate that adding 2,6- held the disk electrode either at the open-circuit potential or at
lutidine removes the previously observed NO3− oxidation peak the same potential we conducted the CPC experiments at
(1.68 V vs Fc+/0). We then recorded LSV traces at the ring
during rotation, affording the data shown in Figure 7. Without
NO3−, we observed no ring current, demonstrating the high
electrochemical stability of Bu4NPF6 electrolyte in MeCN.

Figure 7. LSV traces displaying ring current from rotating ring-disk


experiments using a 30 Hz rotation rate in the listed MeCN solutions,
where [Bu4NPF6] = 100 mM and [LiNO3] = 50 mM. The disk
potential is either at the open-circuit potential (OCP) or 1.68 V vs
Fc+/0 as indicated.

However, when NO3− is introduced to the solution (solid


lines), a cathodic ring current is measured that increases
sharply when the disk is poised at 1.68 V vs Fc+/0 (solid red
line). This cathodic current cannot be the one-electron
reduction of NO3• back to NO3− since the potential onset
(−0.7 V vs Fc+/0) is far too negative from the nitrate oxidation
Figure 6. (a) LSV on Pt electrodes recorded at 25 mV s−1 scan rate in potential (+1.5 V vs Fc+/0), shown in Figure 1a. Instead, we
acetonitrile solutions containing [LiNO3] = 50 mM, [PhCH2OH] = considered the oxygen reduction reaction (ORR) since it is
250 mM, and [2,6-Lutidine] = 25 mM as designated. (b) CPC at 1.68 known to occur acetonitrile solutions on negatively polarized
V vs Fc+/0 on Pt in acetonitrile solutions with 0.5 mM LiNO3, 250 electrodes.21 Subsequent LSVs in MeCN with added HNO3 on
mM PhCH2OH, and 25 mM 2,6-lutidine as noted.
Pt show a reduction that is sensitive to the presence of protons
and cathodic current that is lost upon degassing and purging
with argon gas (Figure S8). These results suggest that the ring
(black trace), indicating that the surface concentration of current observed only during NO3− oxidation is not re-
NO3− is no longer diffusion-limited due to rapid deprotonation reduction of oxidized nitrate species but instead the proton-
of HNO3. Voltammograms of dilute nitrate solutions also coupled oxygen reduction reaction (ORR) in acetonitrile.
display current enhancements with added 2,6-lutidine (Figure Oxygen reduction facilitated by the acidification provided
S7) like the concentrated solutions shown in Figure 6a. during, and collocal to, HNO3 formation generates reduced
Corresponding dilute nitrate CPC experiments (0.5 mM) oxygen species (ROS) that provide the requisite base needed
confirm that the FE for PhCHO, 83 ± 5%, is retained with 2,6- to render nitrate catalytic.
lutidine in solution (Figure 6b, red trace); importantly, the
measured [PhCHO] is ∼5 mM, nearly an order of magnitude
higher than the [NO3−], signifying catalytic nitrate behavior.
■ DISCUSSION
NO3− oxidation in MeCN produces an oxidized intermediate
Control experiments in Figure 6 (dashed traces) confirm that that is capable of subsequently oxidizing PhCH2OH to
increases in rate observed with 2,6-lutidine require NO3−. The PhCHO with ∼80% FE under stoichiometric (base-free) or
results here confirm that the crucial factor for high FE is the catalytic (2,6-lutidine present) conditions. On Pt, electro-
substrate loading or ratio of substrate to acetonitrile/ chemical NO3− oxidation is natively irreversible, where nitrate
benzaldehyde. terminates as HNO3 in the absence of an external base; BiVO4
8152 https://doi.org/10.1021/acs.jpcc.0c10476
J. Phys. Chem. C 2021, 125, 8148−8154
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

photoanodes showed the same irreversible nitrate consump- Scheme 1. Electrochemical Nitrate-Mediated PhCH2OH
tion.13 Interestingly, under photochemical control, NO3− Oxidation on Pt Electrodes (Top) Juxtaposed with the
behaves catalytically on CdS nanowires toward PhCH2OH Photocatalytic Scheme Operative on CdS Nanowires to
oxidation without an added base.12 Revisiting this result Describe General Nitrate Behavior in Acetonitrile Solutions
sparked new questions regarding the system-dependent
behavior of nitrate. Here, we confirmed that off-cycle nitric
acid could be reintroduced by deprotonation with a suitable
external base, 2,6-lutidine (Figure 6).
Nitrate mediation on Pt electrodes enabled by an external
base suggests that a latent proton acceptor present during
photocatalysis on CdS is responsible for the unique base-free
nitrate mediation observed in that reaction. Revisiting the
fundamental differences between these three modes of
PhCH2OH oxidation (photochemical, photoelectrochemical,
and electrochemical), we consider the role of the collocal
reduction reaction on CdS, where an oxygen atmosphere is
required for catalysis to proceed through ORR as the redox
counterpart to NO3− oxidation. By RRDE on Pt, we identified
that acidification of acetonitrile solutions, due to HNO3
formation, facilitates proton-coupled ORR. By LSV, we
measure the reduction potential of the ORR at −1.6 V vs
Fc+/0 (Figure S8) in the absence of acid, compared to −0.7 V
vs Fc+/0 at the ring in RRDE when NO3− oxidation occurs at
the disk electrode. Our data support that the ORR collocal to
NO3− oxidation on CdS particles is responsible for more than
the completion of the light-driven redox cycle, namely, that in
situ generated reduced oxygen species (ROS) serve as the
requisite base that renders nitrate catalytic. In photo-
electrochemical and electrochemical reactions, the reduction
reaction is removed from the working electrode and occurs at a


separate counter electrode ∼5 cm away. This distance
precludes any ROS existing at the working electrode and CONCLUSIONS
thus requires an added Brønsted base to render nitrate catalytic
Nitrate anion is electrochemically oxidized in acetonitrile
(Scheme 1).
solutions to form a highly reactive nitrate radical intermediate.
Previous computational work identified bond dissociation This reactive intermediate is active toward indirect benzyl
energy (BDE) of the resultant nitroxyl mediator-based O−H alcohol oxidation when present. The FE for benzaldehyde
bond as a descriptor for the ability of N-oxyl mediators to product is ∼80% with and without a base in 250 mM benzyl
proceed through the HAT mechanism.22 The authors report alcohol, depending on the loading of alcohol substrate and not
that activity for indirect PhCH2OH oxidation by HAT occurs the amount of nitrate. Acetonitrile-derived oligomers detected
when BDEs were >78 kcal mol−1 for N-oxyl NO−H bonds, through GC-MS analysis are observed as a side product due to
and reactivity increases with larger BDE. We surmise that the a solvent-based hydrogen-atom transfer (HAT) reaction with
resultant BDE of nitric acid (O2NO−H) governs the reactivity nitrate radical, demonstrating the need for high substrate
of NO3• here. HNO3 forms as a byproduct of CPC loading to suppress this competitive pathway. Electrochemical
experiments both with and without alcohol substrate present nitrate anion oxidation is rate-limiting (first order) zero-order
due to a high BDE for O2NO−H of 101 kcal mol−1,23 making in alcohol substrate; however, the subsequent chemical step
it at least 15 kcal mol−1 stronger than the highest BDE shows first-order kinetics in benzyl alcohol consumption.
reported for the N-oxyl class of mediators (THICA).22 Nitric acid formation occurs under all solution conditions due
Competitive H-atom abstraction with solvent represents a to a large O2NO−H bond dissociation energy, making HAT
fundamental limitation when housing the electrochemical the general mode of reactivity for NO3• in MeCN. Improving
process in acetonitrile. To slow down solvent-based side the system design, namely removing ancillary C−H bonds,
reactions, electrochemical NO3− oxidation in d3-acetonitrile remains an ongoing effort to promote selective nitrate-
was attempted, but we still do not observe a reverse wave mediated alcohol oxidation.
feature at 1 V s−1, signifying nitrate radical consumption
(Figure S9). Using aprotic solvents, hexafluorobenzene and
trichloroacetonitrile proved unsuccessful due to the limited

*
ASSOCIATED CONTENT
sı Supporting Information
solubility of the nitrate salts. Future work involves exploring The Supporting Information is available free of charge at
reactions in neat alcohol (solvent-free conditions), where there https://pubs.acs.org/doi/10.1021/acs.jpcc.0c10476.
are no ancillary C−H bonds and the product concentration
remains significantly less than the alcohol concentration to Additional electrochemistry data, CG-MS analysis, and a
ensure that the desired aldehyde product is not consumed. photograph of the electrochemical cell (PDF)

8153 https://doi.org/10.1021/acs.jpcc.0c10476
J. Phys. Chem. C 2021, 125, 8148−8154
The Journal of Physical Chemistry C


pubs.acs.org/JPCC Article

AUTHOR INFORMATION (9) De Nooy, A. E. J.; Besemer, A. C.; van Bekkum, H. Selective
oxidation of primary alcohols mediated by nitroxyl radical in aqueous
Corresponding Author solution. Kinetics and mechanism. Tetrahedron 1995, 51, 8023−8032.
Bart M. Bartlett − Department of Chemistry, University of (10) Zakzeski, J.; Bruijnincx, P. C. A.; Jongerius, A. L.; Weckhuysen,
Michigan, Ann Arbor, Michigan 48109, United States; B. M. The Catalytic Valorization of Lignin for the Production of
orcid.org/0000-0001-8298-5963; Email: bartmb@ Renewable Chemicals. Chem. Rev. 2010, 110, 3552−3599.
umich.edu (11) Rahimi, A.; Azarpira, A.; Kim, H.; Ralph, J.; Stahl, S. S.
Chemoselective Metal-Free Aerobic Alcohol Oxidation in Lignin. J.
Authors Am. Chem. Soc. 2013, 135, 6415−6418.
John L. DiMeglio − Department of Chemistry, University of (12) DiMeglio, J. L.; Breuhaus-Alvarez, A. G.; Li, S.; Bartlett, B. M.
Michigan, Ann Arbor, Michigan 48109, United States; Nitrate Mediated Alcohol Oxidation on Cadmium Sulfide Photo-
orcid.org/0000-0001-5975-5967 catalysts. ACS Catal. 2019, 9, 5732−5741.
Bradley D. Terry − Department of Chemistry, University of (13) Terry, B. D.; DiMeglio, J. L.; Cousineau, J. P.; Bartlett, B. M.
Nitrate Radical Facilitates Indirect Benzyl Alcohol Oxidation on
Michigan, Ann Arbor, Michigan 48109, United States Bismuth(III) Vanadate Photoelectrodes. ChemElectroChem 2020, 7,
Andrew G. Breuhaus-Alvarez − Department of Chemistry, 3776−3782.
University of Michigan, Ann Arbor, Michigan 48109, United (14) Hering, T.; Slanina, T.; Hancock, A.; Wille, U.; König, B.
States Visible light photooxidation of nitrate: the dawn of a nocturnal
Matthew J. Whalen − Department of Chemistry, University of radical. Chem. Commun. 2015, 51, 6568−6571.
Michigan, Ann Arbor, Michigan 48109, United States (15) Leonard, J. E.; Scholl, P. C.; Steckel, T. P.; Lentsch, S. E.; Van
De Mark, M. R. Electrochemical Oxidation of Alcohols: Part II
Complete contact information is available at: Preparative Anodic Oxidation of Secondary Alkanols Employing
https://pubs.acs.org/10.1021/acs.jpcc.0c10476 Lithium Nitrate. Tetrahedron Lett. 1980, 21, 4695−4698.
(16) Tourillon, G.; Lacaze, P-C.; Dubois, J-E. Electrochemical
Author Contributions formation of thin polyacetonitrile films on a Pt surface: P.m.t., i.r.,
J.L.D. and B.D.T.contributed equally. The manuscript was x.p.s. and s.i.m.s. analyses and study of formation mechanism. J.
written through the contributions of all authors. All authors Electroanal. Chem. Interfacial Electrochem. 1979, 100, 247−262.
have given approval to the final version of the manuscript. (17) Iyo, O.; Akiho, S.; Iino, M. Kinetics for Reactions of the Nitrate
Radical (NO3•) with Aldehydes in Acetonitrile. J. Phys. Chem. C
Notes 1989, 93, 4079−4083.
The authors declare no competing financial interest.


(18) Eckert, F.; Leito, I.; Kaljurand, I.; Kütt, A.; Klamt, A.;
Diedenhofen, M. Prediction of Acidity in Acetonitrile Solution with
ACKNOWLEDGMENTS COSMO-RS. J. Comput. Chem. 2009, 30, 799−810.
(19) Zou, S.; Li, R.; Kobayashi, H.; Liu, J.; Fan, J. Photo-assisted
This work was supported by the U.S. Department of Energy,
cyanation of transition metal nitrates coupled with room temperature
Office of Science, Basic Energy Sciences, Catalysis Science C−C bond cleavage of acetonitrile. Chem. Commun. 2013, 49, 1906−
Program, under Award DE-SC0006587. The authors thank 1908.
U−M Department of Chemistry glassblower, Roy Wentz for (20) Kaljurand, I.; Kütt, A.; Sooväli, L.; Rodima, T.; Mäemets, V.;
designing and hand crafting the electrochemical cell used in Leito, I.; Koppel, I. A. Extension of the self-consistent spectrophoto-
the experiments. metric basicity scale in acetonitrile to a full span of 28 pKa units:


unification of different basicity scales. J. Org. Chem. 2005, 70, 1019−
REFERENCES 1028.
(21) Maricle, D. L.; Hodgeson, W. G. Reduction of Oxygen to
(1) Rafiee, M.; Wang, F.; Hruszkewycz, D. P.; Stahl, S. S. N- Superoxide Anion in Aprotic Solvents. Anal. Chem. 1965, 37, 1562−
Hydroxyphthalimide-Mediated Electrochemical Iodination of Meth- 1565.
ylarenes and Comparison of Electron-Transfer-Initiated C−H (22) Dao, R.; Zhao, C.; Yao, J.; Li, H. Distinguishing Ionic and
Functionalization. J. Am. Chem. Soc. 2018, 140, 22−25. Radical Mechanisms of the Hydroxylamines Mediated Electrocatalytic
(2) Li, W.-C.; Zeng, C.-C.; Hu, L.-M.; Tian, H.-Y.; Little, R. D. Alcohol Oxidation using the NO-H Bond Dissociation Energy. Phys.
Efficient Indirect Electrochemical Synthesis of 2-Substituted Benzox- Chem. Chem. Phys. 2018, 20, 28249−28256.
azoles Using Sodium Iodide as Mediator. Adv. Synth. Catal. 2013, (23) Benson, S. W. III − Bond Energies. J. Chem. Educ. 1965, 42,
355, 2884−2890. No. 502.
(3) Chen, J.; Yan, W.-Q.; Lam, C. M.; Zeng, C.-C.; Hu, L.-M.; Little,
D. R. Electrocatalytic Aziridination of Alkenes Mediated by n-Bu4NI:
A Radical Pathway. Org. Lett. 2015, 17, 986−989.
(4) Yan, M.; Kawamata, Y.; Baran, P. S. Synthetic Organic
Electrochemical Methods Since 2000: On the Verge of a Renaissance.
Chem. Rev. 2017, 117, 13230−13319.
(5) Ciriminna, R.; Pagliaro, M. Industrial Oxidations with Organo-
catalyst TEMPO and Its Derivatives. Org. Process Res. Dev. 2010, 14,
245−251.
(6) Badalyan, A.; Stahl, S. S. Cooperative electrocatalytic alcohol
oxidation with electron-proton-transfer mediators. Nature. 2016, 535,
406−410.
(7) Lauber, M. B.; Stahl, S. S. Efficient Aerobic Oxidation of
Secondary alcohols at Ambient Temperature with an ABNO/NOx
Catalyst System. ACS Catal. 2013, 3, 2612−2616.
(8) Taitt, B. J.; Bender, M. T.; Choi, K.-S. Impacts of the
Regeneration Pathways of the Oxoammonium Cation on Electro-
chemical Nitroxyl Radical-Mediated Alcohol Oxidation. ACS Catal.
2020, 10, 265−275.

8154 https://doi.org/10.1021/acs.jpcc.0c10476
J. Phys. Chem. C 2021, 125, 8148−8154

You might also like