ABSTRACT
Aqueous organic electrosynthesis such as nucleophile oxidation reaction (NOR) is an economical and green approach. However, its development has been hindered by the inadequate understanding of the synergy between the electrochemical and non-electrochemical steps. In this study, we unravel the NOR mechanism for the primary alcohol/vicinal diol electrooxidation on NiO. Thereinto, the electrochemical step is the generation of Ni3+-(OH)ads, and the spontaneous reaction between Ni3+-(OH)ads and nucleophiles is an electrocatalyst-induced non-electrochemical step. We identify that two electrophilic oxygen-mediated mechanisms (EOMs), EOM involving hydrogen atom transfer (HAT) and EOM involving C–C bond cleavage, play pivotal roles in the electrooxidation of primary alcohol to carboxylic acid and the electrooxidation of vicinal diol to carboxylic acid and formic acid, respectively. Based on these findings, we establish a unified NOR mechanism for alcohol electrooxidation and deepen the understanding of the synergy between the electrochemical and non-electrochemical steps during NOR, which can guide the sustainable electrochemical synthesis of organic chemicals.
Keywords: nucleophile oxidation reaction, alcohol electrooxidation, organic electrosynthesis, nickel-based electrocatalysts, C–C bond cleavage
This work unravels the electrophilic oxygen-mediated mechanism for alcohol electrooxidation on NiO, and elaborates the synergy between the electrochemical and non-electrochemical steps for the alcohol electrooxidation on NiO.
INTRODUCTION
Under the escalating severity of environmental and energy-related concerns, organic electrosynthesis in aqueous electrolytes, such as electrooxidation/electroreduction of organic compounds, electrochemical CO2 reduction reaction (eCO2RR), etc., has started to garner increasing attention as a sustainable and green approach [1–3]. Aqueous organic electrosynthesis differs from the traditional aqueous electrochemical reactions, e.g. hydrogen/oxygen evolution reaction (HER/OER), because it may involve both the electrochemical step and non-electrochemical process involving organic compounds [4,5]. However, electrochemists sometimes excessively emphasize the roles of the electrocatalyst and electrochemical step, ignoring the non-electrochemical process. This severely hinders the development of aqueous organic electrosynthesis. Recently, numerous aqueous organic electrosynthesis reactions, which sounded almost impossible to achieve previously, have been realized by regulating spontaneous non-electrochemical processes (such as coupling and rearrangement) [6–9]. Researchers have gradually realized the importance of the synergy between the electrochemical and non-electrochemical steps so as to develop new organic electrosynthesis reactions. Nucleophile (such as alcohols, aldehydes, amines, etc.) oxidation reaction (NOR) based on nickel-based electrocatalysts is an essential part of aqueous organic electrosynthesis [2,10–14]. Nevertheless, almost all existing NOR systems were proposed decades ago, suggesting a lack of breakthroughs in this area [10–14]. Establishing a unified NOR mechanism is critical for developing new NOR systems, and the key lies in comprehensively understanding the synergy between electrochemical and non-electrochemical steps.
Notably, the nucleophile oxidation pathway is almost the same across different nickel-based electrocatalysts [2,10–15]. However, the oxidation pathways of different nucleophiles may differ for the same nickel-based electrocatalyst [16,17]. To research the NOR mechanism, consider alcohol electrooxidation for instance. On nickel-based electrocatalysts, primary alcohol (R-CH2OH) can be electrochemically oxidized to carboxylic acid (R-COOH) [2,10]. Vicinal diol (R-CHOH-CH2OH) can be electrochemically oxidized to R-COOH and formic acid (HCOOH) accompanied by the C–C bond cleavage [16–18]. It remains unknown as to why the R-CH2OH and R-CHOH-CH2OH electrooxidation pathways differ significantly. As such, to gain an insight into the NOR mechanism, it is necessary to address the following three key issues: (1) What are the electrochemical and non-electrochemical steps in NOR? (2) How do the electrochemical/non-electrochemical steps collaborate during NOR? (3) How does the synergy between electrochemical and non-electrochemical steps determine the nucleophile electrooxidation pathway?
Here, we explore alcohol electrooxidation on nickel oxide (NiO). The electrochemical step is directly related to the electrocatalyst. We identify two types of non-electrochemical steps, i.e. (1) electrocatalyst-induced non-electrochemical step and (2) electrocatalyst-irrelevant non-electrochemical step. Alcohol electrooxidation on NiO is an indirect electrooxidation reaction with Ni3+-(OH)ads containing electrophilic oxygen as the redox mediator, and it follows the electrophilic oxygen-mediated mechanism (EOM), which comprises the electrochemical step (i.e. electrochemical generation of Ni3+-(OH)ads) and electrocatalyst-induced non-electrochemical step (e.g. Ni3+-(OH)ads–induced hydrogen atom transfer and Ni3+-(OH)ads–induced C–C bond cleavage). Consequently, there are two electrocatalyst functions for alcohol electrooxidation on NiO, i.e. (1) EOM involving hydrogen atom transfer (HAT) and (2) EOM involving C–C bond cleavage. Owing to the synergy of the EOM involving HAT and hydration reaction (which is an electrocatalyst-irrelevant non-electrochemical step), R-CH2OH can be electrochemically oxidized to R-COOH on NiO. Because of the synergy between the EOM involving HAT and EOM involving C–C bond cleavage, R-CHOH-CH2OH can be electrochemically oxidized to R-COOH and HCOOH on NiO. As such, we establish a unified NOR mechanism for alcohol electrooxidation, which can pave the way for developing new NOR systems and other organic electrosynthesis reactions.
RESULTS AND DISCUSSION
Electrophilic oxygen-mediated mechanism involving hydrogen atom transfer
The unique relevance between Ni2+/Ni3+ redox couple and NOR results in excellent NOR performance of nickel-based electrocatalysts, which have been extensively used in NOR, e.g. alcohol electrooxidation [2,10–18]. In this study, NiO was adopted as the model NOR electrocatalyst to study the electrocatalyst function for alcohol electrooxidation. NiO hexagonal nanosheets with a lateral size range of 40–60 nm were synthesized successfully; the detailed characterizations are shown in Figs S1–S5. A 1 M KOH solution and nucleophile-containing 1 M KOH were employed as the electrolytes for the OER and NOR systems, respectively. Under an oxidation potential, the anode will lose electrons to generate electron-deficient species, which may be able to attack the nucleophile as the active NOR intermediate [19,20]. The OER and NOR systems based on NiO were analyzed and compared to unravel the electrocatalyst function in the NOR system.
The onset potentials of OER and NOR (the model NOR system: the ethanol electrooxidation) for NiO are ∼1.55 and ∼1.35 V, respectively (Fig. 1a). The Raman spectrum of fresh NiO electrode shows a broad peak at ∼485 cm−1, which is fitted by employing two components including two-phonon (2P) transverse optical (TO(Δ); peak at ∼460 cm−1) and 2P longitudinal optical (LO(Δ); peak at ∼505 cm−1) modes [21,22]. For the in situ Raman spectra of NiO in the OER system, a new peak at ∼559 cm−1, attributed to the stretching vibration of OH adsorption on Ni3+ site (Ni3+-(OH)ads), appears at a potential above 1.35 V, indicating the electrochemical generation of Ni3+-(OH)ads (Figs 1b and S6) [23–26]. However, during the NOR on NiO at a potential above 1.35 V, the Ni3+-(OH)ads intermediate cannot be detected by in situ Raman spectroscopy (Fig. 1c). Operando electrochemical impedance spectroscopy (EIS) was performed to analyze the reaction interface for the NiO electrode in the OER/NOR system (Figs S7–S10, Tables S1 and S2) [27–29]. According to analysis of the reaction interface, both OER and NOR occur at the low-frequency interface between the surface hydrous layer and electrical double layer (EDL) (Fig. 1d–f) [30–32]. Hence, the OER intermediates i.e. OHads, Oads, and OOHads, are generated at the low-frequency interface, and the NOR active intermediate may be one of the OER intermediates [19].
Figure 1.
OER and NOR systems based on NiO. (a) Anodic polarization curves of NiO in the OER system (1 M KOH) and NOR system (1 M KOH with 50 mM ethanol). (b and c) In situ Raman spectra of NiO electrode in the OER system (b) and NOR system (1 M KOH with 0.5 M ethanol) (c). (d and e) Bode plots for the NiO electrode at different potentials in the OER system (d) and NOR system (1 M KOH with 0.5 M ethanol) (e). (f) Equivalent resistances (R1 and R2) versus potential for the NiO electrode in the OER and NOR systems. (g and h) Schematic diagrams illustrating the OER (g) and NOR (h) systems based on NiO.
There are two processes for NiO in the OER system (Fig. 1g) [31]: (1) Between ∼1.35 to ∼1.55 V, the hydroxyl ion is electrochemically adsorbed to form Ni3+-(OH)ads, i.e. OHads, but it cannot be further oxidized to oxygen, thereby resulting in the accumulation of Ni3+-(OH)ads [19,27]; (2) At a potential above ∼1.55 V, Ni3+-(OH)ads can be electrochemically oxidized to oxygen [32]. There are two possibilities for the undetectable Ni3+-(OH)ads during NOR on NiO (Fig. 1h) [33]: (1) Nucleophiles can be directly electrochemically oxidized, without generating Ni3+-(OH)ads; (2) The NOR on NiO is an indirect electrooxidation with Ni3+-(OH)ads as the redox mediator, and the rate of the reaction between Ni3+-(OH)ads and nucleophiles is far larger than that of the electrochemical generation of Ni3+-(OH)ads, thereby yielding the undetectable Ni3+-(OH)ads.
The Ni3+-(OH)ads accumulates at a high potential (e.g. 1.50 V) in the OER system, and the electroreduction of Ni3+-(OH)ads occurs at a low potential (e.g. 0.94 V) (Fig. S11) [34]. According to the multi-potential steps for the NiO electrode in the OER system and synchronous in situ Raman spectra, after accumulating Ni3+-(OH)ads at 1.50 V, the generated Ni3+-(OH)ads still existed in 1 M KOH under an open-circuit condition, and the accumulated Ni3+-(OH)ads could be electrochemically reduced at 0.94 V (Fig. S12). To identify the role of Ni3+-(OH)ads in NOR, the electrochemical generation of Ni3+-(OH)ads should be separated from NOR. At 1.50 V in 1 M KOH (0 to 100 s), Ni3+-(OH)ads was formed and accumulated; the generated Ni3+-(OH)ads still existed in 1 M KOH under an open-circuit condition (100 to 150 s); after adding 0.5 M ethanol into 1 M KOH at 150 s, the accumulated Ni3+-(OH)ads intermediates were consumed completely by ethanol under the open-circuit condition (Fig. 2a and b). Furthermore, the oxidation product for the reaction between Ni3+-(OH)ads and ethanol is acetic acid (Fig. S13). These results prove that Ni3+-(OH)ads can spontaneously catalyze the oxidative dehydrogenation of alcohol (e.g. ethanol) to carboxylic acid without an applied voltage.
Figure 2.
Electrocatalyst function for NOR on NiO. (a) Current density versus time for the NiO electrode (0 to 100 s: 1.5 V; 100 to 200 s: open-circuit condition; 200 to 300 s: 0.94 V) in different electrolytes (0 to 150 s: 1 M KOH; 150 to 300 s: 1 M KOH with 0.5 M ethanol). (b) Synchronous in situ Raman spectra for the NiO electrode in the electrochemical testing. (c) EOM involving HAT including the electrochemical generation of Ni3+-(OH)ads- and Ni3+-(OH)ads–induced HAT.
The hydrogen atom transfer (HAT) reaction between Ni3+-(OH)ads and nucleophiles (e.g. R-CH2OH) plays an important role in the nucleophile dehydrogenation oxidation pathway for NOR [10,35]. Although both the Ni3+ site and electron-deficient OHads can work as the electron acceptor, the proton acceptor can only be the electrophilic oxygen in Ni3+-(OH)ads, instead of the Ni3+ site [10,19]. The electrophilic oxygen in Ni3+-(OH)ads can spontaneously grab the hydrogen atom from nucleophiles, and this electrocatalyst-induced non-electrochemical step is defined as Ni3+-(OH)ads–induced HAT (Fig. 2c). The NOR on NiO is a unique indirect electrooxidation reaction with the Ni3+-(OH)ads containing electrophilic oxygen as the redox mediator [3,10]. Hence, in the NOR system, NiO follows the electrophilic oxygen-mediated mechanism (EOM) including the electrochemical and electrocatalyst-induced non-electrochemical steps [3,36,37]. The EOM involving HAT includes the electrochemical generation of Ni3+-(OH)ads (Ni2+ + OH− = Ni3+-(OH)ads + e−circuit) and Ni3+-(OH)ads–induced HAT (Ni3+-(OH)ads + X-H = Ni2+ + H2O + X·); thereinto, the former is the rate-limiting step in the electrocatalyst function of NOR. During the EOM involving HAT, the proton and electron of the X−H bond in nucleophiles are spontaneously captured by the electrophilic oxygen in Ni3+-(OH)ads, generated in the electrochemical step, to form X· free radical and H2O (Fig. 2c) [35].
Electrooxidation of primary alcohol on NiO
The NOR system presents a tremendous potential in alcohol electrooxidation [2,13,16–18]. On the NiO electrode, R-CH2OH could be electrochemically oxidized to R-COOH (Figs S14 and S15). During the R-CH2OH electrooxidation, two processes, i.e. (1) oxidation of R-CH2OH to aldehyde (R-CHO) and (2) oxidation of R-CHO to R-COOH, are involved. The electrooxidation of R-CH2OH to R-CHO can be explained by the EOM involving HAT. During two Ni3+-(OH)ads–induced HAT steps, both the hydroxyl group and alpha-methylene (−CH2−) in R-CH2OH lose a hydrogen atom each to produce R-CHO (Fig. S16). Outwardly, only one oxygen atom is transferred for the electrooxidation of R-CHO to R-COOH, but two Ni3+-(OH)ads–induced HAT steps should occur in this two-electron transfer process. Hence, there must be a reaction process involving the transfer of two hydrogen atoms and one oxygen atom (maybe H2O is involved) during the electrooxidation of R-CHO to R-COOH.
While R-CHO dissolves in water, H2O can attack the electrophilic carbonyl carbon in R-CHO to form aldehyde hydrate (geminal diol molecule; R-CH(OH)2), which is a hydration reaction [38,39]. By dissolving R-CHO in 18O labeled water (H218O), almost all oxygen-16 atoms (16O) in the aldehyde group were replaced with oxygen-18 atoms (18O) from H218O within ∼5 minutes (Fig. S17). Hence, the hydration of R-CHO is a spontaneous and reversible reaction, i.e. R-CHO + H2O ⇌ R-C(OH)2, and it is an electrocatalyst-irrelevant non-electrochemical step in NOR [39]. The oxidation of R-CH2OH to R-COOH includes three processes: (1) dehydrogenation of R-CH2OH to R-CHO, (2) hydration of R-CHO to R-C(OH)2, and (3) dehydrogenation of R-C(OH)2 to R-COOH (Fig. S18).
The NOR mechanism comprises the electrochemical step, electrocatalyst-induced non-electrochemical step, and electrocatalyst-irrelevant non-electrochemical step (Fig. 3a). There are three basic steps for the R-CH2OH electrooxidation on NiO (Fig. 3b): (1) electrochemical generation of Ni3+-(OH)ads, (2) Ni3+-(OH)ads–induced HAT, and (3) reversible hydration of R-CHO. Thereinto, the EOM involving HAT comprises the electrochemical generation of Ni3+-(OH)ads and Ni3+-(OH)ads–induced HAT. On the NiO electrode, R-CH2OH can be electrochemically oxidized to R-COOH due to the synergy between the EOM involving HAT and hydration of R-CHO (Fig. 3c). Although the electrochemical performance of NOR depends on the electrochemical step, the nucleophile oxidation pathway is directly determined by the non-electrochemical steps involving nucleophiles, instead of the electrochemical step.
Figure 3.
Synergy between the electrochemical and non-electrochemical steps during the electrooxidation of primary alcohol on NiO. (a) Three basic types of elementary steps in NOR. (b) Three basic electrochemical/non-electrochemical steps in the electrooxidation of R-CH2OH on NiO. (c) Schematic diagram showing the electrooxidation of R-CH2OH on NiO. (d) BDFEs of C–H and O–H bonds in R-CH2OH and R-CH(OH)2 (R: CH3−, CH3-CH2−, and C6H5−). (e) Reaction pathway for the electrooxidation of R-CH2OH to R-COOH based on non-electrochemical steps.
The free energy of the X–H bond cleavage can be described by the bond dissociation free energy (BDFE) [35,40]. Theoretically, the X–H bond with a lower BDFE is more likely to be broken. According to the DFT calculations of R-CH2OH and R-CH(OH)2 (R: CH3−, CH3-CH2−, and C6H5−), the BDFE of the alpha-C-H bond is lower than that of the O–H bond (Figs 3d and S19–S21). Hence, in R-CH2OH or R-CH(OH)2, the dehydrogenation of the alpha-C–H bond takes precedence over that of the O–H bond. Consequently, we proposed the R-CH2OH oxidation pathway based on these five non-electrochemical steps (Figs 3e and S19–S21): (1) dehydrogenation of R-CH2OH to R-·CHOH (HAT1), (2) dehydrogenation of R-·CHOH to R-CHO (HAT2), (3) hydration of R-CHO to R-CH(OH)2, (4) dehydrogenation of R-CH(OH)2 to R-·C(OH)2 (HAT3), and (5) dehydrogenation of R-·C(OH)2 to R-COOH (HAT4).
Electrophilic oxygen-mediated mechanism involving C–C bond cleavage
The NiO electrode cannot catalyze the C–C bond cleavage during the R-CH2OH electrooxidation due to two main reasons, i.e. (1) the limited electrophilicity (oxidizability) of Ni3+-(OH)ads, and (2) the inert C–C bond in R-CH2OH. For vicinal diols (R-CHOH-CHOH), two hydroxyl groups (electron-withdrawing group) significantly reduce the bond dissociation energy of the C–C bond so that the C–C bond cleavage can be catalyzed during the R-CHOH-CHOH electrooxidation on NiO [10,16–18]. On the NiO electrode, the ethylene glycol (CH2OH-CH2OH; the simplest vicinal diol) electrooxidation occurs at a potential above ∼1.35 V (Fig. 4a). During the CH2OH-CH2OH electrooxidation on NiO, Ni3+-(OH)ads can be generated, but cannot be accumulated due to the EOM, thereby yielding the undetectable Ni3+-(OH)ads (Fig. 4b). According to analysis of these products, one CH2OH-CH2OH molecule can be electrochemically oxidized to two HCOOH molecules on the NiO electrode, accompanied by the C–C bond cleavage (Figs 4c and S22). If the reaction mechanism of the CH2OH-CH2OH electrooxidation is the same as that of the R-CH2OH electrooxidation, the electrooxidation product of CH2OH-CH2OH should be oxalic acid (COOH-COOH), instead of HCOOH (Fig. 4d). There must be a special process involving the C–C bond cleavage for the CH2OH-CH2OH electrooxidation on NiO (Fig. S23).
Figure 4.

Electrooxidation of CH2OH-CH2OH on NiO. (a) Anodic polarization curve of NiO in 1 M KOH with 50 mM CH2OH-CH2OH. (b) In situ Raman spectra of NiO electrode in 1 M KOH with 0.5 M CH2OH-CH2OH. (c) Concentration changes of CH2OH-CH2OH and its electrooxidation products versus charge passed during the electrooxidation of CH2OH-CH2OH on NiO. (d) Schematic diagram illustrating the theoretical/actual pathway for the CH2OH-CH2OH electrooxidation on NiO.
Among three C–C bond cleavage types, the oxidative C–C bond cleavage is the only possible means during alcohol electrooxidation since non-spontaneous reduction reactions cannot occur in the NOR system (Fig. S24) [41–43]. For the NOR on NiO, it is challenging to distinguish the oxidative C–C bond cleavage process from other electrochemical/non-electrochemical steps [44]. To avoid the irrelevant non-electrochemical steps (i.e. Ni3+-(OH)ads–induced HAT and hydration of R-CHO), the alpha ketonic acid (R-CO-COOH) without hydroxyl/aldehyde group was used in the NOR system for studying the oxidative C–C bond cleavage (Fig. S25) [45].
The electrooxidation of pyruvic acid (CH3-CO-COOH; the model compound for R-CO-COOH) on NiO occurs at a potential above ∼1.35 V (which is similar to other NOR systems), and CH3-CO-COOH can be electrochemically oxidized to CH3-COOH and carbonic acid (H2CO3) (Figs 5a and S26). Hence, the CH3-CO-COOH electrooxidation on NiO involves two basic steps: (1) the electrochemical generation of Ni3+-(OH)ads and (2) oxidative C–C bond cleavage of CH3-CO-COOH (Fig. 5b). On the NiO electrode, the generated Ni3+-(OH)ads can react with CH3-CO-COOH to generate CH3-COOH and H2CO3 so that the Ni3+-(OH)ads cannot be detected by in situ Raman spectroscopy (Figs 5c and S27). For example, the accumulated Ni3+-(OH)ads cannot be oxidized to oxygen on the NiO electrode at 1.40 V in the OER system; however, for the NOR system containing CH3-CO-COOH at 1.40 V, the Ni3+-(OH)ads can catalyze oxidative C–C bond cleavage at the low-frequency interface (Figs 5c, d, S28 and S29).
Figure 5.
Electrochemical oxidation-induced C–C bond cleavage on NiO. (a) Anodic polarization curves of NiO in 1 M KOH with/without nucleophiles (CH3-CH2OH, CH2OH-CH2OH, and CH3-CO-COOH). (b) Model NOR system for researching oxidative C–C bond cleavage on NiO: electrooxidation of CH3-CO-COOH. (c and d) In situ Raman spectra (c) and Bode plots (d) of NiO in 1 M KOH with/without 0.5 M CH3-CO-COOH. (e) Current density versus time for the NiO electrode (0 to 100 s: 1.5 V; 100 to 200 s: open-circuit condition; 200 to 300 s: 0.94 V; 300 to 400 s: 1.5 V) in different electrolytes (0 to 150 s: 1 M KOH; 150 to 400 s: 1 M KOH with 0.5 M CH3-CO-COOH), inset: Synchronous in situ Raman spectra. (f and g) Schematic diagrams illustrating the CH3-CO-COOH electrooxidation on NiO (f) and the EOM involving C–C bond cleavage (g).
To further understand the reaction mechanism between the C–C bond and Ni3+-(OH)ads, the electrochemical generation of Ni3+-(OH)ads and the oxidative C–C bond cleavage of CH3-CO-COOH should be separated (Figs 5e and S30). Ni3+-(OH)ads was formed and accumulated at 1.50 V in 1 M KOH (0 to 100 s), and the generated Ni3+-(OH)ads still existed in 1 M KOH under an open-circuit condition (100 to 150 s); however, after adding 0.5 M CH3-CO-COOH into 1 M KOH at 150 s, the Ni3+-(OH)ads was consumed completely under the open-circuit condition (150 to 200 s), accompanied by the oxidative C–C bond cleavage (Fig. 5e). Hence, CH3-CO-COOH can spontaneously react with Ni3+-(OH)ads to generate CH3-COOH and H2CO3, and this process is defined as the Ni3+-(OH)ads–induced C–C bond cleavage, which is an electrocatalyst-induced non-electrochemical step.
We consider that two Ni3+-(OH)ads intermediates attack the C–C bond in CH3-CO-COOH to form two C–OH bonds, thereby yielding the spontaneous C–C bond cleavage (Fig. 5f) [46]. Hence, in addition to the EOM involving HAT, there is another electrocatalyst function for the NOR on NiO, i.e. the EOM involving C–C bond cleavage including the electrochemical generation of Ni3+-(OH)ads and Ni3+-(OH)ads–induced C–C bond cleavage (Figs 5g and S31). The EOM involving C–C bond cleavage is accompanied by the formation of the O-H bond, and the C–C bond cleavage products of a hydroxymethyl group (–CH2OH), aldehyde group (–CHO), carboxyl group (–COOH), secondary hydroxyl group (R-CHOH–), and ketone group (R-CO–) are formaldehyde hydrate (H2C(OH)2), HCOOH, H2CO3, aldehyde hydrate (R-CH(OH)2), and R-COOH, respectively (Fig. S32). Generally, there are two electrocatalyst functions, i.e. (1) the EOM involving HAT and (2) EOM involving C–C bond cleavage. According to the density functional theory (DFT) calculations, the electrochemical generation of Ni3+-(OH)ads is the rate-limiting step in these two EOMs, because Ni3+-(OH)ads can spontaneously catalyze HAT or C–C bond cleavage (Fig. S33).
Electrooxidation of vicinal diol on NiO
Compared with the R-CH2OH electrooxidation, the NiO electrode exhibits better electrochemical performances for the R-CHOH-CH2OH electrooxidation because of more reaction sites in R-CHOH-CH2OH (Figs 6a and S34). According to analysis of the products, R-CHOH-CH2OH can be electrochemically oxidized to R-COOH and HCOOH accompanied by the C–C bond cleavage (Fig. S35). As for polyhydric alcohols, multiple C–C bonds in polyhydric alcohols can be broken during NOR; e.g. one glycerol (CH2OH-CHOH-CH2OH) molecule can be electrochemically oxidized to three HCOOH molecules during the electrooxidation over NiO [16–18]. We further prove that the R-CHOH-CH2OH oxidation accompanied by the C–C bond cleavage (e.g. the oxidation of CH2OH-CH2OH to HCOOH) can be spontaneously catalyzed by Ni3+-(OH)ads without an applied voltage (Fig. S36). There may be four basic steps in the electrooxidation of R-CHOH-CH2OH to R-COOH and HCOOH on NiO, i.e. (1) the electrochemical generation of Ni3+-(OH)ads, (2) Ni3+-(OH)ads–induced HAT, (3) Ni3+-(OH)ads–induced C–C bond cleavage, and (4) hydration of R-CHO (Fig. 6b). Considering the CH2OH-CH2OH (the simplest vicinal diol) electrooxidation as an example, the possible electrooxidation products are COOH-COOH, HCOOH, and H2CO3. The sluggish deep electrooxidation of HCOOH to H2CO3 on NiO can influence the concentrations of HCOOH and H2CO3 during the CH2OH-CH2OH electrooxidation (Fig. S37). Hence, the COOH-COOH selectivity is an important piece of evidence to identify the CH2OH-CH2OH electrooxidation pathway.
Figure 6.
Synergy between the electrochemical and non-electrochemical steps during the electrooxidation of vicinal diol on NiO. (a) Current densities (at 1.5 V) for the electrooxidations of different primary alcohols and vicinal diols on NiO. (b) Basic steps for the electrooxidation of R-CHOH-CH2OH to R-COOH and HCOOH on NiO. (c) Selectivities of HCOOH and COOH-COOH for the electrooxidations of HCHO, CHO-COOH, CH2OH-COOH, CHO-CHO, CH2OH-CHO, and CH2OH-CH2OH on NiO. (d and e) Reaction pathways of the electrooxidations of CH2OH-CH2OH (d) and R-CHOH-CH2OH (e) based on non-electrochemical steps. (f) Schematic diagram illustrating the electrooxidation of CH2OH-CH2OH or R-CHOH-CH2OH on NiO.
According to analyses performed on the electrooxidations of all possible reaction intermediates (i.e. HCHO, CHO-COOH, CH2OH-COOH, CHO-CHO, and CH2OH-CHO) on NiO, we estimated the probabilities of the basic reaction processes for the CH2OH-CH2OH electrooxidation on NiO through a backward induction strategy (Figs S38–S44) [47]. Thereinto, only the HCHO electrooxidation exhibits a higher HCOOH selectivity (∼98%) than that attained from the CH2OH-CH2OH electrooxidation (∼95%), and the COOH-COOH selectivities for the electrooxidations of other possible reaction intermediates (i.e. CHO-COOH, CH2OH-COOH, CHO-CHO, and CH2OH-CHO) far outweigh that for the CH2OH-CH2OH electrooxidation (∼1.4%) (Fig. 6c). These results demonstrate that the EOM involving C–C bond cleavage takes priority during the CH2OH-CH2OH electrooxidation on NiO, and the dominant reaction intermediate is CH2(OH)2, rather than other C2 intermediates (Fig. S44).
According to the BDFEs of CH2(OH)2 or R-CH(OH)2, the dehydrogenation of the alpha–C–H bond takes precedence over that of the O-H bond (Fig. S45). The CH2OH-CH2OH electrooxidation pathway on NiO is described as follows (Fig. 6d): (1) CH2OH-CH2OH is oxidized to two CH2(OH)2 molecules during the Ni3+-(OH)ads–induced C–C bond cleavage; (2) The C–H and O–H bonds in CH2(OH)2 undergo dehydrogenation in turn to form HCOOH during two Ni3+-(OH)ads–induced HAT steps. Both the EOM involving C–C bond cleavage and EOM involving HAT work during the R-CHOH-CH2OH electrooxidation on NiO (Fig. 6e and f). In the first stage, R-CHOH-CH2OH is electrochemically oxidized to R-CH2(OH)2 and CH2(OH)2via the EOM involving C–C bond cleavage. In the second stage, due to the EOM involving HAT, the alpha-C–H and O–H bonds in CH2(OH)2 or R-CH2(OH)2 undergo dehydrogenation in turn to form HCOOH or RCOOH, respectively.
CONCLUSION
In this study, we investigated the alcohol (i.e. R-CH2OH and R-CHOH-CH2OH) electrooxidation on NiO. The alcohol electrooxidation on NiO is an indirect electrooxidation reaction with the Ni3+-(OH)ads containing electrophilic oxygen as the redox mediator. The electrochemical step is the electrochemical generation of Ni3+-(OH)ads. Ni3+-(OH)ads–induced HAT and Ni3+-(OH)ads–induced C–C bond cleavage are electrocatalyst-induced non-electrochemical steps. Hence, there are two EOMs including the electrochemical step and electrocatalyst-induced non-electrochemical step, i.e. (1) the EOM involving HAT and (2) EOM involving C–C bond cleavage. The hydration of R-CHO is an electrocatalyst-irrelevant non-electrochemical step. The NOR system depends on the synergy between the electrochemical and non-electrochemical steps. The synergy between EOM involving HAT and hydration of R-CHO results in the electrooxidation of R-CH2OH to R-COOH on NiO. The synergy between EOM involving HAT and EOM involving C–C bond cleavage causes the electrooxidation of R-CHOH-CH2OH to R-COOH and HCOOH on NiO. As such, this study establishes a unified NOR mechanism for alcohol electrooxidation, which will open up new possibilities for developing NOR. Our work highlights the synergy between the electrochemical and non-electrochemical steps in the NOR system. We firmly believe that, by tuning electrochemical and non-electrochemical steps, the NOR system can achieve various aqueous organic electrooxidation reactions toward the green synthesis of high-value organic compounds.
MATERIALS AND METHODS
Ni(NO3)2·6H2O, NaOH, and KOH were purchased from Sinopharm Chemical Reagent Co., Ltd. 5 wt% Nafion solution was purchased from DuPont™. All organic compounds were purchased from Sigma-Aldrich.
The morphologies and crystalline structures of electrocatalysts were identified by scanning electron microscope (SEM; Hitachi S-4800, Hitachi Corporation, Japan), transmission electron microscope (TEM; FEI Tecnai G20, FEI Company, USA), X-ray diffraction (XRD; Bruker D8 Advance diffractometer, Bruker, Germany). The Ni K-edge X-ray absorption was measured using Taiwan Photon Source (TPS) Quick-scanning X-ray absorption spectroscopy beamline 44A1 under ambient pressure at the National Synchrotron Radiation Research Center (NSRRC), Hsinchu. XAS measurement was made in transmission mode using ion chamber detectors. And pure cobalt metal foil was used for energy calibration. All spectra were carried out with 1 Hz oscillating frequency for 2 min. We have 240 spectra to average and increase the S/N ratio and normalize unit step height in the absorption coefficient from well below to well above the edges. All X-ray absorption spectroscopy (XAS) spectra were aligned, merged, deglitched, and normalized using the Athena (version number 0.9.26) module implemented in the IFEFFIT software packages. The electrode surface species was identified by X-ray photoelectron spectroscopy (XPS; Axis Supra, Kratos Company, England) and Raman spectrum (Alpha300R, WETEC, Germany). The XPS spectra were calibrated by adventitious carbon (C1s at 284.8 eV).
Preparation of NiO
The NiO hexagonal nanosheets were synthesized with the β-Ni(OH)2 hexagonal nanosheets as precursor. Then 5 mmol Ni(NO3)2·6H2O was dissolved in 30 mL deionized water. The Ni(NO3)2 solution was mixed with 2 M NaOH solution, thus obtaining the β-Ni(OH)2 suspension (pH: ∼13.5). The β-Ni(OH)2 suspension was transferred into a 100-mL Teflon-lined stainless-steel autoclave, and the autoclave was placed in an oven at 160°C for 6 h. After cooling, the synthesized β-Ni(OH)2 nanosheets were repeatedly washed by deionized water and anhydrous ethanol, then dried in a vacuum drying chamber for 10 h at 60°C. Finally, the β-Ni(OH)2 hexagonal nanosheets were calcined under an air atmosphere at 400°C for 2 h (the heating/cooling rate: 5°C min−1), thus synthesizing the NiO hexagonal nanosheets.
Supplementary Material
ACKNOWLEDGEMENTS
The authors would like to thank the Shiyanjia Lab (www.shiyanjia.com) for the language editing service.
Contributor Information
Wei Chen, State Key Laboratory of Chemo/Bio-Sensing and Chemometrics, College of Chemistry and Chemical Engineering, Advanced Catalytic Engineering Research Center of the Ministry of Education, Hunan University, Changsha 410082.
Jianqiao Shi, State Key Laboratory of Chemo/Bio-Sensing and Chemometrics, College of Chemistry and Chemical Engineering, Advanced Catalytic Engineering Research Center of the Ministry of Education, Hunan University, Changsha 410082.
Chao Xie, State Key Laboratory of Chemo/Bio-Sensing and Chemometrics, College of Chemistry and Chemical Engineering, Advanced Catalytic Engineering Research Center of the Ministry of Education, Hunan University, Changsha 410082; College of Chemistry and Chemical Engineering, Hunan Normal University, Changsha 410081.
Wang Zhou, College of Materials Science and Engineering, Hunan University, Changsha 410082.
Leitao Xu, State Key Laboratory of Chemo/Bio-Sensing and Chemometrics, College of Chemistry and Chemical Engineering, Advanced Catalytic Engineering Research Center of the Ministry of Education, Hunan University, Changsha 410082.
Yingying Li, State Key Laboratory of Chemo/Bio-Sensing and Chemometrics, College of Chemistry and Chemical Engineering, Advanced Catalytic Engineering Research Center of the Ministry of Education, Hunan University, Changsha 410082.
Yandong Wu, State Key Laboratory of Chemo/Bio-Sensing and Chemometrics, College of Chemistry and Chemical Engineering, Advanced Catalytic Engineering Research Center of the Ministry of Education, Hunan University, Changsha 410082.
Binbin Wu, State Key Laboratory of Chemo/Bio-Sensing and Chemometrics, College of Chemistry and Chemical Engineering, Advanced Catalytic Engineering Research Center of the Ministry of Education, Hunan University, Changsha 410082.
Yu-Cheng Huang, Research Center for X-ray Science & Department of Physics, Tamkang University, New Taipei City 25137.
Bo Zhou, State Key Laboratory of Chemo/Bio-Sensing and Chemometrics, College of Chemistry and Chemical Engineering, Advanced Catalytic Engineering Research Center of the Ministry of Education, Hunan University, Changsha 410082.
Ming Yang, State Key Laboratory of Chemo/Bio-Sensing and Chemometrics, College of Chemistry and Chemical Engineering, Advanced Catalytic Engineering Research Center of the Ministry of Education, Hunan University, Changsha 410082.
Jilei Liu, College of Materials Science and Engineering, Hunan University, Changsha 410082.
Chung-Li Dong, Research Center for X-ray Science & Department of Physics, Tamkang University, New Taipei City 25137.
Tehua Wang, State Key Laboratory of Chemo/Bio-Sensing and Chemometrics, College of Chemistry and Chemical Engineering, Advanced Catalytic Engineering Research Center of the Ministry of Education, Hunan University, Changsha 410082.
Yuqin Zou, State Key Laboratory of Chemo/Bio-Sensing and Chemometrics, College of Chemistry and Chemical Engineering, Advanced Catalytic Engineering Research Center of the Ministry of Education, Hunan University, Changsha 410082; Shenzhen Institute of Hunan University, Shenzhen 518057.
Shuangyin Wang, State Key Laboratory of Chemo/Bio-Sensing and Chemometrics, College of Chemistry and Chemical Engineering, Advanced Catalytic Engineering Research Center of the Ministry of Education, Hunan University, Changsha 410082; Shenzhen Institute of Hunan University, Shenzhen 518057.
FUNDING
This work was supported by the National Key R&D Program of China (2020YFA0710000), the National Natural Science Foundation of China (22122901, 21902047, 21825201 and U19A2017), the Provincial Natural Science Foundation of Hunan (2020JJ5045, 2021JJ20024 and 2021RC3054), and the Shenzhen Science and Technology Program (JCYJ20210324140610028).
AUTHOR CONTRIBUTIONS
S.W. conceived the project. W.C. carried out most of the experiments and wrote the manuscript. J.S., C.X., Y.W. and B.W. performed the product characterization. W.Z., B.Z., M.Y. and J.L. performed in situ Raman spectra. L.X. and Y.L. performed the theoretical calculations. Y.H. and C.D. performed XAS measures. T.W. performed EIS and analysis. Y.Z. conducted part of the synthesis of catalysts and characterizations. All authors discussed the results and commented on the manuscript.
Conflict of interest statement
None declared.
REFERENCES
- 1. Li F, Thevenon A, Rosas-Hernández Aet al. Molecular tuning of CO2-to-ethylene conversion. Nature 2020; 577: 509–13. 10.1038/s41586-019-1782-2 [DOI] [PubMed] [Google Scholar]
- 2. You B, Sun Y. Innovative strategies for electrocatalytic water splitting. Acc Chem Res 2018; 51: 1571–80. 10.1021/acs.accounts.8b00002 [DOI] [PubMed] [Google Scholar]
- 3. Francke R, Little RD. Redox catalysis in organic electrosynthesis: basic principles and recent developments. Chem Soc Rev 2014; 43: 2492–521. 10.1039/c3cs60464k [DOI] [PubMed] [Google Scholar]
- 4. Martínez-Huitle CA, Ferro S. Electrochemical oxidation of organic pollutants for the wastewater treatment: direct and indirect processes. Chem Soc Rev 2006; 35: 1324–40. 10.1039/B517632H [DOI] [PubMed] [Google Scholar]
- 5. Sui Y, Ji X. Anticatalytic strategies to suppress water electrolysis in aqueous batteries. Chem Rev 2021; 121: 6654–95. 10.1021/acs.chemrev.1c00191 [DOI] [PubMed] [Google Scholar]
- 6. Xia C, Xia Y, Zhu Pet al. Direct electrosynthesis of pure aqueous H2O2 solutions up to 20% by weight using a solid electrolyte. Science 2019; 366: 226–31. 10.1126/science.aay1844 [DOI] [PubMed] [Google Scholar]
- 7. Lee KM, Jang JH, Balamurugan Met al. Redox-neutral electrochemical conversion of CO2 to dimethyl carbonate. Nat Energy 2021; 6: 733–41. 10.1038/s41560-021-00862-1 [DOI] [Google Scholar]
- 8. Chen C, Zhu X, Wen Xet al. Coupling N2 and CO2 in H2O to synthesize urea under ambient conditions. Nat Chem 2020; 12: 717–24. 10.1038/s41557-020-0481-9 [DOI] [PubMed] [Google Scholar]
- 9. Leow WR, Lum Y, Ozden Aet al. Chloride-mediated selective electrosynthesis of ethylene and propylene oxides at high current density. Science 2020; 368: 1228–33. 10.1126/science.aaz8459 [DOI] [PubMed] [Google Scholar]
- 10. Chen W, Xie C, Wang Yet al. Activity origins and design principles of nickel-based catalysts for nucleophile electrooxidation. Chem 2020; 6: 2974–93. 10.1016/j.chempr.2020.07.022 [DOI] [Google Scholar]
- 11. Chen W, Xu L, Zhu Xet al. Unveiling the electrooxidation of urea: intramolecular coupling of the N–N bond. Angew Chem Int Ed 2021; 60: 7297–307. 10.1002/anie.202015773 [DOI] [PubMed] [Google Scholar]
- 12. Vértes G, Horányi G. Some problems of the kinetics of the oxidation of organic compounds at oxide-covered nickel electrodes. J Electroanal Chem 1974; 52: 47–53. 10.1016/S0022-0728(74)80100-2 [DOI] [Google Scholar]
- 13. Robertson PM. On the oxidation of alcohols and amines at nickel oxide electrodes: mechanistic aspects. J Electroanal Chem 1980; 111: 97–104. 10.1016/S0022-0728(80)80079-9 [DOI] [Google Scholar]
- 14. Hui BS, Huber CO. Amperometric detection of amines and amino acids in flow injection systems with a nickel oxide electrode. Anal Chim Acta 1982; 134: 211–8. 10.1016/S0003-2670(01)84191-X [DOI] [Google Scholar]
- 15. Hu X, Zhu J, Li Jet al. Urea electrooxidation: current development and understanding of Ni-based catalysts. ChemElectroChem 2020; 7: 3211–28. 10.1002/celc.202000404 [DOI] [Google Scholar]
- 16. Li Y, Wei X, Chen Let al. Nickel-molybdenum nitride nanoplate electrocatalysts for concurrent electrolytic hydrogen and formate productions. Nat Commun 2019; 10: 5335. 10.1038/s41467-019-13375-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Bott-Neto JL, Martins TS, Machado Set al. Electrocatalytic oxidation of methanol, ethanol, and glycerol on Ni(OH)2 nanoparticles encapsulated with poly[Ni(salen)] film. ACS Appl Mater Inter 2019; 11: 30810–8. 10.1021/acsami.9b08441 [DOI] [PubMed] [Google Scholar]
- 18. Lin QH, Wei YM, Liu WTet al. Electrocatalytic oxidation of ethylene glycol and glycerol on nickel ion implanted-modified indium tin oxide electrode. Int J Hydrogen Energy 2017; 42: 1403–11. 10.1016/j.ijhydene.2016.10.011 [DOI] [Google Scholar]
- 19. Tao HB, Xu Y, Huang Xet al. A general method to probe oxygen evolution intermediates at operating conditions. Joule 2019; 3: 1498–509. 10.1016/j.joule.2019.03.012 [DOI] [Google Scholar]
- 20. Pfeifer V, Jones TE, Vélez JJVet al. In situ observation of reactive oxygen species forming on oxygen-evolving iridium surfaces. Chem Sci 2017; 8: 2143–9. 10.1039/C6SC04622C [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21. Wang D, Xu S, Wu Let al. Spin-phonon coupling in NiO nanoparticle. J Appl Phys 2020; 128: 133905. 10.1063/5.0022668 [DOI] [Google Scholar]
- 22. Wang WZ, Liu YK, Xu CKet al. Synthesis of NiO nanorods by a novel simple precursor thermal decomposition approach. Chem Phys Lett 2002; 362: 119–22. 10.1016/S0009-2614(02)00996-X [DOI] [Google Scholar]
- 23. Yeo BS, Bell AT. In situ Raman study of nickel oxide and gold-supported nickel oxide catalysts for the electrochemical evolution of oxygen. J Phys Chem C 2012; 116: 8394–400. 10.1021/jp3007415 [DOI] [Google Scholar]
- 24. Qiu Z, Ma Y, Edvinsson T. In operando Raman investigation of Fe doping influence on catalytic NiO intermediates for enhanced overall water splitting. Nano Energy 2019; 66: 104118. 10.1016/j.nanoen.2019.104118 [DOI] [Google Scholar]
- 25. Trześniewski BJ, Diaz-Morales O, Vermaas DAet al. In situ observation of active oxygen species in Fe-containing Ni-based oxygen evolution catalysts: the effect of pH on electrochemical activity. J Am Chem Soc 2015; 137: 15112–21. 10.1021/jacs.5b06814 [DOI] [PubMed] [Google Scholar]
- 26. Wang YH, Wang XT, Ze Het al. Spectroscopic verification of adsorbed hydroxy intermediates in the bifunctional mechanism of the hydrogen oxidation reaction. Angew Chem Int Ed 2021; 60: 5708–11. 10.1002/anie.202015571 [DOI] [PubMed] [Google Scholar]
- 27. Bredar ARC, Chown AL, Burton ARet al. Electrochemical impedance spectroscopy of metal oxide electrodes for energy applications. ACS Appl Energy Mater 2020; 3: 66–98. 10.1021/acsaem.9b01965 [DOI] [Google Scholar]
- 28. Li W, Zhao L, Jiang Xet al. Confinement engineering of electrocatalyst surfaces and interfaces. Adv Funct Mater 2022; 32: 2207727. 10.1002/adfm.202207727 [DOI] [Google Scholar]
- 29. Schichlein H, Müller AC, Voigts Met al. Deconvolution of electrochemical impedance spectra for the identification of electrode reaction mechanisms in solid oxide fuel cells. J Appl Electrochem 2002; 32: 875–82. 10.1023/A:1020599525160 [DOI] [Google Scholar]
- 30. Xie C, Chen W, Du Set al. In-situ phase transition of WO3 boosting electron and hydrogen transfer for enhancing hydrogen evolution on Pt. Nano Energy 2020; 71: 104653. 10.1016/j.nanoen.2020.104653 [DOI] [Google Scholar]
- 31. Smith RDL, Berlinguette CP. Accounting for the dynamic oxidative behavior of nickel anodes. J Am Chem Soc 2016; 138: 1561–7. 10.1021/jacs.5b10728 [DOI] [PubMed] [Google Scholar]
- 32. Song J, Wei C, Huang ZFet al. A review on fundamentals for designing oxygen evolution electrocatalysts. Chem Soc Rev 2020; 49: 2196–214. 10.1039/C9CS00607A [DOI] [PubMed] [Google Scholar]
- 33. Martinez-Huitle CA, Ferro S. Electrochemical oxidation of organic pollutants for the wastewater treatment: direct and indirect processes. Chem Soc Rev 2006; 35: 1324–40. 10.1039/B517632H [DOI] [PubMed] [Google Scholar]
- 34. Chen A, Lipkowski J. Electrochemical and spectroscopic studies of hydroxide adsorption at the Au(111) electrode. J Phys Chem B 1999; 103: 682–91. 10.1021/jp9836372 [DOI] [Google Scholar]
- 35. Wise CF, Mayer JM. Electrochemically determined O–H bond dissociation free energies of NiO electrodes predict proton-coupled electron transfer reactivity. J Am Chem Soc 2019; 141: 14971–5. 10.1021/jacs.9b07923 [DOI] [PubMed] [Google Scholar]
- 36. Zhang N, Chai Y. Lattice oxygen redox chemistry in solid-state electrocatalysts for water oxidation. Energy Environ Sci 2021; 14: 4647–71. 10.1039/D1EE01277K [DOI] [Google Scholar]
- 37. Grimaud A, Diaz-Morales O, Han Bet al. Activating lattice oxygen redox reactions in metal oxides to catalyse oxygen evolution. Nat Chem 2017; 9: 457–65. 10.1038/nchem.2695 [DOI] [PubMed] [Google Scholar]
- 38. Hilal SH, Bornander LL, Carreira LA. Hydration equilibrium constants of aldehydes, ketones and quinazolines. QSAR Comb Sci 2005; 24: 631–8. 10.1002/qsar.200430913 [DOI] [Google Scholar]
- 39. Guthrie JP. Hydration of carbonyl compounds, an analysis in terms of multidimensional Marcus theory. J Am Chem Soc 2000; 122: 5529–38. 10.1021/ja992992i [DOI] [Google Scholar]
- 40. Darcy JW, Koronkiewicz B, Parada GAet al. A continuum of proton-coupled electron transfer reactivity. Acc Chem Res 2018; 51: 2391–9. 10.1021/acs.accounts.8b00319 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41. Subbotina E, Rukkijakan T, Marquez-Medina MDet al. Oxidative cleavage of C–C bonds in lignin. Nat Chem 2021; 13: 1118–25. 10.1038/s41557-021-00783-2 [DOI] [PubMed] [Google Scholar]
- 42. Murphy SK, Park JW, Cruz FAet al. Rh-catalyzed C–C bond cleavage by transfer hydroformylation. Science 2015; 347: 56–60. 10.1126/science.1261232 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Shi SH, Liang Y, Jiao N. Electrochemical oxidation induced selective C–C bond cleavage. Chem Rev 2021; 121: 485–505. 10.1021/acs.chemrev.0c00335 [DOI] [PubMed] [Google Scholar]
- 44. Parpot P, Nunes N, Bettencourt AP. Electrocatalytic oxidation of monosaccharides on gold electrode in alkaline medium: structure–reactivity relationship. J Electroanal Chem 2006; 596: 65–73. 10.1016/j.jelechem.2006.07.006 [DOI] [Google Scholar]
- 45. Sen Gupta KK, Sarkar T. Kinetics of the chromic acid oxidation of glyoxylic and pyruvic acids. Tetrahedron 1975; 31: 123–7. [Google Scholar]
- 46. Parpot P, Santos PRB, Bettencourt AP. Electro-oxidation of d-mannose on platinum, gold and nickel electrodes in aqueous medium. J Electroanal Chem 2007; 610: 154–62. 10.1016/j.jelechem.2007.07.011 [DOI] [Google Scholar]
- 47. Bambagioni V, Bevilacqua M, Bianchini Cet al. Ethylene glycol electrooxidation on smooth and nanostructured Pd electrodes in alkaline media. Fuel Cells 2010; 10: 582–90. 10.1002/fuce.200900120 [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.





