10 research outputs found
Simplifying the Evaluation of Graphene Modified Electrode Performance Using Rotating Disk Electrode Voltammetry
Graphene modified electrodes have been fabricated by
electrodeposition
from an aqueous graphene oxide solution onto conducting Pt, Au, glassy
carbon, and indium tin dioxide substrates. Detailed investigations
of the electrochemistry of the [RuÂ(NH<sub>3</sub>)<sub>6</sub>]<sup>3+/2+</sup> and [FeÂ(CN)<sub>6</sub>]<sup>3‑/4‑</sup> and
hydroquinone and uric acid oxidation processes have been undertaken
at glassy carbon and graphene modified glassy carbon electrodes using
transient cyclic voltammetry at a stationary electrode and near steady-state
voltammetry at a rotating disk electrode. Comparisons of the data
with simulation suggest that the transient voltammetric characteristics
at graphene modified electrodes contain a significant contribution
from thin layer and surface confined processes. Consequently, interpretations
based solely on mass transport by semi-infinite linear diffusion may
result in incorrect conclusions on the activity of the graphene modified
electrode. In contrast, steady-state voltammetry at a rotating disk
electrode affords a much simpler method for the evaluation of the
performance of graphene modified electrode since the relative importance
of the thin layer and surface confined processes are substantially
diminished and mass transport is dominated by convection. Application
of the rotated electrode approach with carbon nanotube modified electrodes
also should lead to simplification of data analysis in this environment
Polyoxometalate-Promoted Electrocatalytic CO<sub>2</sub> Reduction at Nanostructured Silver in Dimethylformamide
Electrochemical reduction
of CO<sub>2</sub> is a promising method to convert CO<sub>2</sub> into
fuels or useful chemicals, such as carbon monoxide (CO), hydrocarbons,
and alcohols. In this study, nanostructured Ag was obtained by electrodeposition
of Ag in the presence of a Keggin type polyoxometalate, [PMo<sub>12</sub>O<sub>40</sub>]<sup>3–</sup> (PMo). Metallic Ag is formed
upon reduction of Ag<sup>+</sup>. Adsorption of PMo on the surface
of the newly formed Ag lowers its surface energy thus stabilizes the
nanostructure. The electrocatalytic performance of this Ag–PMo
nanocomposite for CO<sub>2</sub> reduction was evaluated in a CO<sub>2</sub> saturated dimethylformamide medium containing 0.1 M [<i>n</i>-Bu<sub>4</sub>N]ÂPF<sub>6</sub> and 0.5% (v/v) added H<sub>2</sub>O. The results show that this Ag–PMo nanocomposite
can catalyze the reduction of CO<sub>2</sub> to CO with an onset potential
of −1.70 V versus Fc<sup>0/+</sup>, which is only 0.29 V more
negative than the estimated reversible potential (−1.41 V)
for this process and 0.70 V more positive than that on bulk Ag metal.
High faradaic efficiencies of about 90% were obtained over a wide
range of applied potentials. A Tafel slope of 60 mV dec<sup>–1</sup> suggests that rapid formation of *CO<sub>2</sub><sup>•–</sup> is followed by the rate-determining protonation step. This is consistent
with the voltammetric data which suggest that the reduced PMo interacts
strongly with CO<sub>2</sub> (and presumably CO<sub>2</sub><sup>•–</sup>) and hence promotes the formation of CO<sub>2</sub><sup>•–</sup>
PdCu@Pd Nanocube with Pt-like Activity for Hydrogen Evolution Reaction
The
electronic properties of metal surfaces can be modulated to weaken
the binding energy of adsorbed H-intermediates on the catalyst surface,
thus enhancing catalytic activity for the hydrogen evolution reaction
(HER). Here we first prepare PdCu alloy nanocubes (NCs) by coreduction
of CuÂ(acac)<sub>2</sub> (acac = acetylacetonate) and Na<sub>2</sub>PdCl<sub>4</sub> in the presence of oleylamine (OAm) and trioctylphosphine
(TOP). The PdCu NC coated glassy carbon electrode is then anodized
at a constant potential of 0.51 V vs Ag/AgCl at room temperature in
0.5 M H<sub>2</sub>SO<sub>4</sub> solution for 10 s, which converts
PdCu NCs into core@shell PdCu@Pd NCs that show much enhanced Pt-like
activity for the HER and much more robust durability. The improvements
in surface property and HER activity are rationalized based on strain
and ligand effects that enhance the activity of the edge-exposed Pd
atoms on core@shell PdCu@Pd structure. This work opens up a new perspective
for simultaneously reducing metal Pd cost and achieving excellent
performance toward the HER
Voltammetric Determination of the Reversible Potentials for [{Ru<sub>4</sub>O<sub>4</sub>(OH)<sub>2</sub>(H<sub>2</sub>O)<sub>4</sub>}(γ-SiW<sub>10</sub>O<sub>36</sub>)<sub>2</sub>]<sup>10–</sup> over the pH Range of 2–12: Electrolyte Dependence and Implications for Water Oxidation Catalysis
Voltammetric studies of the Ru-containing
polyoxometalate water oxidation molecular catalyst [{Ru<sub>4</sub>O<sub>4</sub>(OH)<sub>2</sub>(H<sub>2</sub>O)<sub>4</sub>}Â(γ-SiW<sub>10</sub>O<sub>36</sub>)<sub>2</sub>]<sup>10–</sup> ([<b>1</b>(γ-SiW<sub>10</sub>O<sub>36</sub>)<sub>2</sub>]<sup>10–</sup> where <b>1</b> represents the {Ru<sub>4</sub>O<sub>4</sub>(OH)<sub>2</sub>(H<sub>2</sub>O)<sub>4</sub>} core and <b>1</b>(0) stands for its initial form with all ruthenium centers
in the oxidation state IV) have been carried out in aqueous media
over a wide range of pH (2–12 using Britton–Robinson
buffer) and ionic strength. Well-defined voltammograms in buffered
media are only obtained when Frumkin double layer effects are suppressed
by the presence of a sufficient concentration of additional supporting
electrolyte (LiNO<sub>3</sub>, NaNO<sub>3</sub>, KNO<sub>3</sub>,
CaÂ(NO<sub>3</sub>)<sub>2</sub>, MgÂ(NO<sub>3</sub>)<sub>2</sub>, MgSO<sub>4</sub>, or Na<sub>2</sub>SO<sub>4</sub>). A combination of data
derived from dc cyclic, rotating disk electrode, and Fourier transformed
large amplitude ac voltammetry allow the assignment of two processes
related to reduction of the framework and the complete series of Ru<sup>III/IV</sup> and Ru<sup>IV/V</sup> redox processes and also provide
their reversible potentials. Analysis of these data reveals that K<sup>+</sup> has a significantly stronger interaction with <b>1</b>(1) (the number inside bracket stands for the number of electrons
removed from <b>1</b>(0)) than found for the other cations investigated,
and hence its presence significantly alters the pH dependence of the <b>1</b>(0)/<b>1</b>(1) reversible potential. Comparison of
experimental data with theory developed in terms of equilibrium constants
for process <b>1</b>(0)/<b>1</b>(1) reveals that both
H<sup>+</sup> and K<sup>+</sup> interact competitively with both <b>1</b>(0) and <b>1</b>(1). Importantly, reversible potential
data reveal that (i) proton transfer does not necessarily need to
be coupled to all electron transfer steps to achieve catalytic oxidation
of water, (ii) the four-electron oxidized form, <b>1</b>(4),
is capable of oxidizing water under all conditions studied, and (iii)
under some conditions, the three-electron oxidized form, <b>1</b>(3), also exhibits considerable catalytic activity
Voltammetric and Spectroscopic Studies of α- and β‑[PW<sub>12</sub>O<sub>40</sub>]<sup>3–</sup> Polyoxometalates in Neutral and Acidic Media: Structural Characterization as Their [(<i>n</i>‑Bu<sub>4</sub>N)<sub>3</sub>][PW<sub>12</sub>O<sub>40</sub>] Salts
The structure of the Keggin-type
β-[PW<sub>12</sub>O<sub>40</sub>]<sup>3–</sup> (PW<sub>12</sub>) polyoxometalate,
with <i>n</i>-Bu<sub>4</sub>N<sup>+</sup> as the countercation,
has been determined for the first time by single-crystal X-ray analysis
and compared to data obtained from a new determination of the structure
of the α-PW<sub>12</sub> isomer, having the same countercation.
Analysis of cyclic voltammograms obtained in CH<sub>3</sub>CN (0.1
M [<i>n</i>-Bu<sub>4</sub>N]Â[PF<sub>6</sub>]) reveals that
the reversible potential for the β-PW<sub>12</sub> isomer always
remains ca. 100 mV more positive than that of the α-PW<sub>12</sub> isomer on addition of the acid CF<sub>3</sub>SO<sub>3</sub>H. Simulations
of the cyclic voltammetry as a function of acid concentration over
the range 0–5 mM mimic experimental data exceptionally well.
These simulation–experiment comparisons provide access to reversible
potentials and acidity constants associated with α and β
fully oxidized and one- and two-electron reduced systems and also
explain how the two well-resolved one-electron WÂ(VI)/WÂ(V) processes
converge into a single two-electron process if sufficient acid is
present. <sup>183</sup>W NMR spectra of the oxidized forms of the
PW<sub>12</sub> isomers are acid dependent and in the case of β-PW<sub>12</sub> imply that the bridging oxygens between the W<sub>I</sub> and W<sub>II</sub> units are preferentially protonated in acidic
media. EPR data on frozen solutions of one-electron reduced β-[PW<sup>V</sup>W<sup>VI</sup><sub>11</sub>O<sub>40</sub>]<sup>4–</sup> indicate that either the W<sub>I</sub> or the W<sub>III</sub> unit
in β-PW<sub>12</sub> is reduced in the β-[PW<sup>VI</sup><sub>12</sub>O<sub>40</sub>]<sup>3–</sup>/β-[PW<sup>V</sup>W<sup>VI</sup><sub>11</sub>O<sub>40</sub>]<sup>4–</sup> process. In the absence of acid, reversible potentials obtained
from the α- and β-isomers of PW<sub>12</sub> and [SiW<sub>12</sub>O<sub>40</sub>]<sup>4–</sup> exhibit a linear relationship
with solvent properties such as Lewis acidity, acceptor number, and
polarity index
Voltammetric and Spectroscopic Studies of α- and β‑[PW<sub>12</sub>O<sub>40</sub>]<sup>3–</sup> Polyoxometalates in Neutral and Acidic Media: Structural Characterization as Their [(<i>n</i>‑Bu<sub>4</sub>N)<sub>3</sub>][PW<sub>12</sub>O<sub>40</sub>] Salts
The structure of the Keggin-type
β-[PW<sub>12</sub>O<sub>40</sub>]<sup>3–</sup> (PW<sub>12</sub>) polyoxometalate,
with <i>n</i>-Bu<sub>4</sub>N<sup>+</sup> as the countercation,
has been determined for the first time by single-crystal X-ray analysis
and compared to data obtained from a new determination of the structure
of the α-PW<sub>12</sub> isomer, having the same countercation.
Analysis of cyclic voltammograms obtained in CH<sub>3</sub>CN (0.1
M [<i>n</i>-Bu<sub>4</sub>N]Â[PF<sub>6</sub>]) reveals that
the reversible potential for the β-PW<sub>12</sub> isomer always
remains ca. 100 mV more positive than that of the α-PW<sub>12</sub> isomer on addition of the acid CF<sub>3</sub>SO<sub>3</sub>H. Simulations
of the cyclic voltammetry as a function of acid concentration over
the range 0–5 mM mimic experimental data exceptionally well.
These simulation–experiment comparisons provide access to reversible
potentials and acidity constants associated with α and β
fully oxidized and one- and two-electron reduced systems and also
explain how the two well-resolved one-electron WÂ(VI)/WÂ(V) processes
converge into a single two-electron process if sufficient acid is
present. <sup>183</sup>W NMR spectra of the oxidized forms of the
PW<sub>12</sub> isomers are acid dependent and in the case of β-PW<sub>12</sub> imply that the bridging oxygens between the W<sub>I</sub> and W<sub>II</sub> units are preferentially protonated in acidic
media. EPR data on frozen solutions of one-electron reduced β-[PW<sup>V</sup>W<sup>VI</sup><sub>11</sub>O<sub>40</sub>]<sup>4–</sup> indicate that either the W<sub>I</sub> or the W<sub>III</sub> unit
in β-PW<sub>12</sub> is reduced in the β-[PW<sup>VI</sup><sub>12</sub>O<sub>40</sub>]<sup>3–</sup>/β-[PW<sup>V</sup>W<sup>VI</sup><sub>11</sub>O<sub>40</sub>]<sup>4–</sup> process. In the absence of acid, reversible potentials obtained
from the α- and β-isomers of PW<sub>12</sub> and [SiW<sub>12</sub>O<sub>40</sub>]<sup>4–</sup> exhibit a linear relationship
with solvent properties such as Lewis acidity, acceptor number, and
polarity index
Voltammetric and Spectroscopic Studies of α- and β‑[PW<sub>12</sub>O<sub>40</sub>]<sup>3–</sup> Polyoxometalates in Neutral and Acidic Media: Structural Characterization as Their [(<i>n</i>‑Bu<sub>4</sub>N)<sub>3</sub>][PW<sub>12</sub>O<sub>40</sub>] Salts
The structure of the Keggin-type
β-[PW<sub>12</sub>O<sub>40</sub>]<sup>3–</sup> (PW<sub>12</sub>) polyoxometalate,
with <i>n</i>-Bu<sub>4</sub>N<sup>+</sup> as the countercation,
has been determined for the first time by single-crystal X-ray analysis
and compared to data obtained from a new determination of the structure
of the α-PW<sub>12</sub> isomer, having the same countercation.
Analysis of cyclic voltammograms obtained in CH<sub>3</sub>CN (0.1
M [<i>n</i>-Bu<sub>4</sub>N]Â[PF<sub>6</sub>]) reveals that
the reversible potential for the β-PW<sub>12</sub> isomer always
remains ca. 100 mV more positive than that of the α-PW<sub>12</sub> isomer on addition of the acid CF<sub>3</sub>SO<sub>3</sub>H. Simulations
of the cyclic voltammetry as a function of acid concentration over
the range 0–5 mM mimic experimental data exceptionally well.
These simulation–experiment comparisons provide access to reversible
potentials and acidity constants associated with α and β
fully oxidized and one- and two-electron reduced systems and also
explain how the two well-resolved one-electron WÂ(VI)/WÂ(V) processes
converge into a single two-electron process if sufficient acid is
present. <sup>183</sup>W NMR spectra of the oxidized forms of the
PW<sub>12</sub> isomers are acid dependent and in the case of β-PW<sub>12</sub> imply that the bridging oxygens between the W<sub>I</sub> and W<sub>II</sub> units are preferentially protonated in acidic
media. EPR data on frozen solutions of one-electron reduced β-[PW<sup>V</sup>W<sup>VI</sup><sub>11</sub>O<sub>40</sub>]<sup>4–</sup> indicate that either the W<sub>I</sub> or the W<sub>III</sub> unit
in β-PW<sub>12</sub> is reduced in the β-[PW<sup>VI</sup><sub>12</sub>O<sub>40</sub>]<sup>3–</sup>/β-[PW<sup>V</sup>W<sup>VI</sup><sub>11</sub>O<sub>40</sub>]<sup>4–</sup> process. In the absence of acid, reversible potentials obtained
from the α- and β-isomers of PW<sub>12</sub> and [SiW<sub>12</sub>O<sub>40</sub>]<sup>4–</sup> exhibit a linear relationship
with solvent properties such as Lewis acidity, acceptor number, and
polarity index
Electrooxidation of Ethanol and Methanol Using the Molecular Catalyst [{Ru<sub>4</sub>O<sub>4</sub>(OH)<sub>2</sub>(H<sub>2</sub>O)<sub>4</sub>}(γ-SiW<sub>10</sub>O<sub>36</sub>)<sub>2</sub>]<sup>10–</sup>
Highly efficient electrocatalytic
oxidation of ethanol and methanol
has been achieved using the ruthenium-containing polyoxometalate molecular
catalyst, [{Ru<sub>4</sub>ÂO<sub>4</sub>Â(OH)<sub>2</sub>Â(H<sub>2</sub>O)<sub>4</sub>}Â(γ-SiW<sub>10</sub>O<sub>36</sub>)<sub>2</sub>]<sup>10–</sup> ([<b>1</b>(γ-SiW<sub>10</sub>ÂO<sub>36</sub>)<sub>2</sub>]<sup>10–</sup>).
Voltammetric studies with dissolved and surface-confined forms of
[<b>1</b>(γ-SiW<sub>10</sub>O<sub>36</sub>)<sub>2</sub>]<sup>10–</sup> suggest that the oxidized forms of <b>1</b> can act as active catalysts for alcohol oxidation in both aqueous
(over a wide pH range covering acidic, neutral, and alkaline) and
alcohol media. Under these conditions, the initial form of <b>1</b> also exhibits considerable reactivity, especially in neutral solution
containing 1.0 M NaNO<sub>3</sub>. To identify the oxidation products,
preparative scale bulk electrolysis experiments were undertaken. The
products detected by NMR, gas chromatography (GC), and GC-mass spectrometry
from oxidation of ethanol are 1,1-diethoxyethane and ethyl acetate
formed from condensation of acetaldehyde or acetic acid with excess
ethanol. Similarly, the oxidation of methanol generates formaldehyde
and formic acid which then condense with methanol to form dimethoxymethane
and methyl formate, respectively. These results demonstrate that electrocatalytic
oxidation of ethanol and methanol occurs via two- and four-electron
oxidation processes to yield aldehydes and acids. The total faradaic
efficiencies of electrocatalytic oxidation of both alcohols exceed
94%. The numbers of aldehyde and acid products per catalyst were also
calculated and compared with the literature reported values. The results
suggest that <b>1</b> is one of the most active molecular electrocatalysts
for methanol and ethanol oxidation
Observation of Ferromagnetic Exchange, Spin Crossover, Reductively Induced Oxidation, and Field-Induced Slow Magnetic Relaxation in Monomeric Cobalt Nitroxides
The
reaction of [Co<sup>II</sup>(NO<sub>3</sub>)<sub>2</sub>]·6H<sub>2</sub>O with the nitroxide radical, 4-dimethyl-2,2-diÂ(2-pyridyl)
oxazolidine-<i>N</i>-oxide (L<sup>•</sup>), produces
the mononuclear transition-metal complex [Co<sup>II</sup>(L<sup>•</sup>)<sub>2</sub>]Â(NO<sub>3</sub>)<sub>2</sub> (<b>1</b>), which
has been investigated using temperature-dependent magnetic susceptibility,
electron paramagnetic resonance (EPR) spectroscopy, electrochemistry,
density functional theory (DFT) calculations, and variable-temperature
X-ray structure analysis. Magnetic susceptibility measurements and
X-ray diffraction (XRD) analysis reveal a central low-spin octahedral
Co<sup>2+</sup> ion with both ligands in the neutral radical form
(L<sup>•</sup>) forming a linear L<sup>•</sup>···CoÂ(II)···L<sup>•</sup> arrangement. This shows a host of interesting magnetic
properties including strong cobalt-radical and radical–radical
intramolecular ferromagnetic interactions stabilizing a <i>S</i> = <sup>3</sup>/<sub>2</sub> ground state, a thermally induced spin
crossover transition above 200 K and field-induced slow magnetic relaxation.
This is supported by variable-temperature EPR spectra, which suggest
that <b>1</b> has a positive <i>D</i> value and nonzero <i>E</i> values, suggesting the possibility of a field-induced
transverse anisotropy barrier. DFT calculations support the parallel
alignment of the two radical π*<sub>NO</sub> orbitals with a
small orbital overlap leading to radical–radical ferromagnetic
interactions while the cobalt-radical interaction is computed to be
strong and ferromagnetic. In the high-spin (HS) case, the DFT calculations
predict a weak antiferromagnetic cobalt-radical interaction, whereas
the radical–radical interaction is computed to be large and
ferromagnetic. The monocationic complex [Co<sup>III</sup>(L<sup>–</sup>)<sub>2</sub>]Â(BPh<sub>4</sub>) (<b>2</b>) is formed by a rare,
reductively induced oxidation of the Co center and has been fully
characterized by X-ray structure analysis and magnetic measurements
revealing a diamagnetic ground state. Electrochemical studies on <b>1</b> and <b>2</b> revealed common Co-redox intermediates
and the proposed mechanism is compared and contrasted with that of
the Fe analogues
Observation of Ferromagnetic Exchange, Spin Crossover, Reductively Induced Oxidation, and Field-Induced Slow Magnetic Relaxation in Monomeric Cobalt Nitroxides
The
reaction of [Co<sup>II</sup>(NO<sub>3</sub>)<sub>2</sub>]·6H<sub>2</sub>O with the nitroxide radical, 4-dimethyl-2,2-diÂ(2-pyridyl)
oxazolidine-<i>N</i>-oxide (L<sup>•</sup>), produces
the mononuclear transition-metal complex [Co<sup>II</sup>(L<sup>•</sup>)<sub>2</sub>]Â(NO<sub>3</sub>)<sub>2</sub> (<b>1</b>), which
has been investigated using temperature-dependent magnetic susceptibility,
electron paramagnetic resonance (EPR) spectroscopy, electrochemistry,
density functional theory (DFT) calculations, and variable-temperature
X-ray structure analysis. Magnetic susceptibility measurements and
X-ray diffraction (XRD) analysis reveal a central low-spin octahedral
Co<sup>2+</sup> ion with both ligands in the neutral radical form
(L<sup>•</sup>) forming a linear L<sup>•</sup>···CoÂ(II)···L<sup>•</sup> arrangement. This shows a host of interesting magnetic
properties including strong cobalt-radical and radical–radical
intramolecular ferromagnetic interactions stabilizing a <i>S</i> = <sup>3</sup>/<sub>2</sub> ground state, a thermally induced spin
crossover transition above 200 K and field-induced slow magnetic relaxation.
This is supported by variable-temperature EPR spectra, which suggest
that <b>1</b> has a positive <i>D</i> value and nonzero <i>E</i> values, suggesting the possibility of a field-induced
transverse anisotropy barrier. DFT calculations support the parallel
alignment of the two radical π*<sub>NO</sub> orbitals with a
small orbital overlap leading to radical–radical ferromagnetic
interactions while the cobalt-radical interaction is computed to be
strong and ferromagnetic. In the high-spin (HS) case, the DFT calculations
predict a weak antiferromagnetic cobalt-radical interaction, whereas
the radical–radical interaction is computed to be large and
ferromagnetic. The monocationic complex [Co<sup>III</sup>(L<sup>–</sup>)<sub>2</sub>]Â(BPh<sub>4</sub>) (<b>2</b>) is formed by a rare,
reductively induced oxidation of the Co center and has been fully
characterized by X-ray structure analysis and magnetic measurements
revealing a diamagnetic ground state. Electrochemical studies on <b>1</b> and <b>2</b> revealed common Co-redox intermediates
and the proposed mechanism is compared and contrasted with that of
the Fe analogues