Operando Electron Paramagnetic Resonance for Elucidating the

May 29, 2019 - One of the most important challenges in chemistry with direct implication in biochemistry is probing the mechanism of electron transfer...
1 downloads 0 Views 3MB Size
Article Cite This: J. Phys. Chem. C 2019, 123, 16058−16064

pubs.acs.org/JPCC

Operando Electron Paramagnetic Resonance for Elucidating the Electron Transfer Mechanism of Coenzymes Mian A. Ali,† Ayaz Hassan,† Graziela C. Sedenho,† Renato V. Gonçalves,‡ Daniel R. Cardoso,† and Frank N. Crespilho*,† †

São Carlos Institute of Chemistry, University of São Paulo, 13560-970 São Carlos, SP, Brazil São Carlos Institute of Physics, University of São Paulo, 13560-970 São Carlos, SP, Brazil



Downloaded via KEAN UNIV on July 19, 2019 at 12:33:56 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: One of the most important challenges in chemistry with direct implication in biochemistry is probing the mechanism of electron transfer originating from biological molecules. On the basis of protein film voltammetry, mediated electron transfer and molecular adsorption followed by heterogeneous catalysis result in similar responses for steadystate currents; both processes increase the Faradaic current at a low overpotential. This is typical of NAD-dependent alcohol dehydrogenase (ADH), an oxidoreductase enzyme that uses the interconversion of NAD+/NADH coenzyme to catalyze the oxidation of alcohol to aldehyde. We propose a setup based on operando electron paramagnetic resonance (EPR) spectroscopy to investigate the NADH/NAD+ redox reaction and introduce how to probe free electrons on a carbon electrode surface and correlate them with the electrocatalytic mechanism. Since knowledge of the g-factor may provide information about the electronic structure of the paramagnetic center at the carbon surface, it was found that the concentration of unpaired free electrons responds to both applied overpotential and NADH oxidation, enabling measurement of the in situ dynamics of the electron transfer reaction. A new correlation for the spin concentration reveals an increasing number of free unpaired electrons with increasing applied overpotential and NADH oxidation, which corroborates the controversial hypothesis that quinone groups act as electrocatalysts and not as redox mediators toward the oxidation of NADH to NAD+. Furthermore, operando EPR provides useful information in probing the electron transfer dynamics on a carbon surface and may be extended to other chemical systems involving electron transfer reactions.



INTRODUCTION Electron transfer involving biological molecules is a significant phenomenon in both biochemistry and technology. However, important questions remain to be answered, such as how to determine the dominant reaction pathways for efficient proton and electron motion from their initial positions to their final positions in biomolecules. In particular, bioelectrocatalysis using coenzymes usually follows complex reaction pathways and is therefore difficult to explain. For instance, dehydrogenase enzymes have pivotal roles in metabolism and oxidize substrates by transferring hydrogen to electron acceptors. NAD+ (an oxidized form of nicotinamide adenine dinucleotide) is one of the most important coenzymes essential to the growth and survival of organisms and for technological applications in organic molecule oxidation for producing fuel and energy. One typical example is NAD-dependent alcohol dehydrogenase (ADH), an oxidoreductase enzyme that uses the interconversion of the NAD+/NADH redox couple1−3 to catalyze the oxidation of alcohol to aldehyde. The coenzyme binds to the enzyme by oxidizing the alcoholic moieties in it. The dissociation of the enzyme−NADH complex is the ratedetermining step according to a recently reported mechanism.4 © 2019 American Chemical Society

Currently, voltammetry is used to measure the kinetics and dynamics of NAD+/NADH interconversion, where carbon electrodes functionalized with quinone groups are usually used to study this reaction. The terms “electrocatalysis” and “catalytic electro-oxidation” have been frequently used to explain the NADH oxidation5−9 mechanism on quinonemodified electrodes, mainly owing to the shape of the cyclic voltammogram. The “electrocatalytic” oxidation of NADH has been typically reported for ortho-quinones, anthraquinones (see the scheme in Figure 1), and their derivatives. In contrast, several authors have reported that quinones act as redox mediators.10−15 Even though the quinone−electrolyte interface plays an important role as an electron donor/acceptor, there is no consensus about the mechanism of electron transfer. As far as the experimental approach is concerned, there is an intrinsic problem in solving the reaction mechanism: both the mediated electron transfer mechanism and electrocatalysis can result in similar responses in steady-state (SS) electrochemistry because Received: February 5, 2019 Revised: April 29, 2019 Published: May 29, 2019 16058

DOI: 10.1021/acs.jpcc.9b01160 J. Phys. Chem. C 2019, 123, 16058−16064

Article

The Journal of Physical Chemistry C

Figure 1. (a) Illustration of immobilized ADH (PDB 4W6Z) on a carbon electrode highlighting the electro-oxidation of the coenzyme NADH to NAD+ involving different quinone groups on the carbon electrode surface. (b) Chemical structure of the coenzyme NAD+. (c) General reaction pathway of NADH electro-oxidation involving quinones, as frequently reported in the literature.5,8,25

Figure 2. (a) Possible reaction pathway of FCF oxidation by KMnO4 under acidic conditions. (b) Electrochemical grafting of AQ groups to FCF in a HCl solution (0.5 mol L−1) containing 1-aminoanthraquinone (1.0 mmol L−1) and NaNO2 (4.0 mmol L−1) in an Ar atmosphere.

both increase the Faradaic current at a low overpotential. However, if there is a catalytic effect on a reaction taking place at the surface of a quinone-modified electrode due to the electrode itself, an increase in the electrochemical reaction rate is expected and the electronic properties of the electrode surface should change. However, to the best of our knowledge, there are no studies in the literature so far. In addition, the NADH electrocatalysis mechanism involving direct electron transfer has never been reported in terms of the spectroscopic data of the electrode. Electron paramagnetic resonance (EPR) spectroscopy is a powerful technique that has been used to investigate in situ electrochemical phenomena16−22 or electrochemical systems under operando conditions.23,24 This is because EPR is a highly specific technique capable of detecting the formation and disappearance of unpaired electrons or radicals involved in redox processes. In this context, we propose a new experimental approach that may contribute to the understanding of the NADH oxidation mechanism (and eventually to other redox coenzymes) on a quinone-modified carbon electrode. This approach is based on operando EPR spectroscopy of modified carbon electrodes under NADH electro-oxidation, wherein the quantity of unpaired electron spins can be measured with an applied overpotential, thus revealing the role of the quinone functional groups on the electro-oxidation of NADH. The

versatility of operando EPR is discussed to clarify mechanistic ambiguities and in terms of broader application to other electrocatalytic systems. In this context, we used operando EPR measurements, because EPR is a powerful technique used to investigate in situ electrochemical reactions or electrochemical systems under operando conditions. This is because EPR is a highly specific technique capable of detecting the formation and disappearance of unpaired electrons or radicals involved in redox processes.



RESULTS AND DISCUSSION Carbon Electrodes Functionalized with Quinones. There are several ways to promote the functionalization of a carbon electrode surface. We have selected two different methodologies to modify the surfaces of carbon electrodes with quinone groups toward the electro-oxidation of NADH to NAD+. The first one consists of covalent functionalization through electrochemical grafting to introduce anthraquinone functionalities to flexible carbon fiber (FCF) electrodes, named FCF-AQ here. This methodology has been consolidated from several studies described elsewhere.26−28 The second methodology consists of the modification of FCF with quinone groups by using a one-pot reaction with KMnO4 under acidic conditions.3,29,30 This methodology has been proven to be very useful for engineering more robust and stable electrodes 16059

DOI: 10.1021/acs.jpcc.9b01160 J. Phys. Chem. C 2019, 123, 16058−16064

Article

The Journal of Physical Chemistry C

Figure 3. SEM images of (a) FCF, (b) FCF-O, and (c) FCF-AQ. Raman spectra of (d) FCF, FCF-O, and FCF-AQ. All spectra were fitted using a Lorentzian curve with a linear baseline and normalized according to the G-band intensities. XPS spectra of the (e) C 1s and (f) O 1s regions for FCF, FCF-O, and FCF-AQ. Plots of the concentration versus the functional groups of FCF, FCF-O, and FCF-AQ in the (g) C 1s region and (h) O 1s regions of the XPS spectra in (e) and (f).

(PAN) filaments through a wet spinning procedure.21 Figure 3b shows an SEM image of an FCF-O electrode, which shows an increase in the depths of the grooves and the formation of defects on the surface. An SEM image of an FCF-AQ electrode (Figure 3c) shows a rougher surface with wider and deeper stretches. On the basis of the SEM images, it can be inferred that the surfaces of both of the FCF-O and FCF-AQ electrodes are strongly affected by the functionalization processes. As observed from the SEM results, the presence of defects on the electrode surfaces after the functionalization processes is clear. In this context, micro-Raman spectroscopy was conducted to characterize these defects. Figure 3d shows the first-order region of the Raman spectra of pristine FCF, FCF-O, and FCF-AQ electrodes. Each spectrum consists of two main characteristic bands: D and G bands centered at about 1365 and 1603 cm−1, respectively. The G band is associated with the presence of sp2 carbon networks in the electrodes, whereas the D-band is a defect-induced Raman feature related to the presence of amorphous or disordered carbon forms.31 The deconvolution of the raw spectra (see Figure S3) shows the presence of a weak shoulder, the A band at 1555 cm−1. This band is assigned to amorphous forms of carbon32 and is commonly observed for carbon fibers.33,34 The ratio of the intensities of the D and G bands (ID/IG)31 was used to characterize the number of defects present in carbon fibers promoted by surface modification methodologies. The calculated values of ID/IG were 0.98, 1.27, and 1.03 for FCF, FCF-O, and FCF-AQ, respectively (see Table S1). Since the values of ID/IG increase with increasing disorder,31 the results indicate that the introduction of quinone functionalities by both methodologies increases the number of defects in the carbon fiber structure. In addition, it can be concluded that the chemical treatment with KMnO4 promotes a larger number of defects on the FCF-O surface than functionalization through electrochemical grafting. The defects are intrinsically related to the presence of oxygenated groups, as we shall show in the following discussion.

for applications in enzyme bioelectrochemistry, particularly for the high electrochemical performance of dehydrogenase-based electrodes. In this case, one-pot synthesis improves the electrode surface by subjecting it to oxidation through a direct chemical reaction, avoiding the complicated separation and purification of the intermediates involved, which sufficiently reduces the time and the use of chemicals. The FCF electrode modified by this procedure is named FCF-O. The proposed reaction pathways26−30 of both of the functionalized FCF-O and FCF-AQ electrodes are shown in Figure 2. To obtain FCF-AQ electrodes, the protocol initially involves the in situ generation of an anthraquinone diazonium cation in an aqueous solution containing 1-aminoanthraquinone, NaNO2, and HCl.26−28 Then, cyclic voltammograms obtained during this process are shown in Figure S1a in the Supporting Information, where the formation of a reduction wave at 0.10 V is indicative of the attachment of anthraquinonyl groups to the surface of FCF. The decrease in the reduction current density and the formation of well-defined redox peaks in the subsequent cycles are consistent with the surface functionalization. This electrode was further characterized by cyclic voltammetry, and the results are compared to those of pristine FCF, as shown in Figure S1b,c. The FCF-AQ anode showed a very well-defined electroactivity for both acid and alkaline electrolytes. The surface density of anthraquinone groups obtained from the charge of the anodic peak was found to be 2.8 × 10−9 mol cm−2, which is in accordance with the literature.26,28 In a similar way, the surface density of oquinones present on FCF-O was found to be 1.3 × 10−9 mol cm−2. The electrodes were further electrochemically characterized using the K4[Fe(CN)6]/K3[Fe(CN)6] redox probe (see Figure S2). Pristine FCF, FCF-O, and FCF-AQ anodes were further characterized by scanning electron microscopy (SEM), micro-Raman spectroscopy, and X-ray photoelectron spectroscopy (XPS). Figure 3a displays an SEM image of pristine FCF, which exhibits a uniform surface with parallel lines and shallow grooves. This morphology is characteristic of FCF obtained by the carbonization process of polyacrylonitrile 16060

DOI: 10.1021/acs.jpcc.9b01160 J. Phys. Chem. C 2019, 123, 16058−16064

Article

The Journal of Physical Chemistry C To obtain chemical information on the surface of pristine FCF, FCF-O, and FCF-AQ electrodes, XPS spectra were recorded. Figure 3e shows the deconvoluted C 1s core-level XPS spectra for pristine FCF, FCF-O, and FCF-AQ electrodes. The C 1s spectra for FCF-O electrode was fitted with five peaks using a pseudo-Voigt function, whereas four peaks were observed for FCF-AQ. For all electrodes, peak I is attributed to sp2 carbon (C−C aromatic) observed at 284.7 eV for pristine FCF, 283.9 eV for FCF-O, and 284.4 eV for FCF-AQ. The presence of sp3-bonded carbon atoms (C−H, C−N) is indicated by peak II located at 284.5 eV for FCF-O and 285.1 eV for FCF-AQ. Peak III belongs to phenolic, alcohol, or ether groups and is observed at 286.1 eV for pristine FCF, 285.1 eV for FCF-O, and 286.1 for FCF-AQ. Carbonyl or quinone groups are associated with peak IV and are observed at 286.0 eV for FCF-O and 288.2 eV for FCF-AQ.3,35,36 The surface concentrations of different functional groups, as calculated on the basis of the peaks observed in the XPS spectra, are shown in Figure 3g, wherein a decrease in the concentration of sp 2 carbon and an increase in the concentration of sp3 carbon (C−H, C−N) for the FCF-O and FCF-AQ electrodes, respectively, clearly confirm the surface functionalization of these electrodes. On the other hand, the increases in the carbonyl peak at 286.0 eV for FCF-O and 288.2 eV for FCF-AQ clearly confirm the formation of quinone molecules on their surfaces. Figure 3f shows the deconvoluted XPS spectra in the O 1s region for pristine FCF, FCF-O, and FCF-AQ electrodes, wherein three peaks were observed for pristine FCF and the FCF-O and FCF-AQ electrodes. Here, peak I is related to carbonyl or quinone and is observed at 531 eV for pristine FCF, 531.6 eV for FCF-O, and 531.4 eV for FCF-AQ.3,35 Peak II corresponds to carbon in phenolic, alcoholic, or ether groups, which is observed at 532.0, 532.6, and 532.7 eV for pristine FCF, FCF-O, and FCF-AQ electrodes, respectively.3,35 Additional peaks were observed at 533.0, 533.6, and 533.7 eV for pristine FCF, FCF-O, and FCF-AQ electrodes, respectively, which correspond to carboxylic acid groups. The deconvoluted XPS spectra in the O 1s region corroborate the spectra in the C 1s region, exhibiting an increase in surface oxidation for both electrodes. Figure 3h shows a decrease in the concentration of C−O groups and an increase in the concentration of CO groups on the FCF-O and FCF-AQ electrodes compared with pristine FCF. Electro-oxidation of NADH. The electro-oxidation of NADH was evaluated for both FCF-O and FCF-AQ electrodes. First, cyclic voltammetry was performed in a phosphate buffer solution (0.1 mol L−1, pH 7.5) in the absence of NADH, as shown in Figure 4a−d (showing magnified regions of Figure 4a,b, respectively). A well-resolved redox behavior is observed for both electrodes, which is consistent with surface modification with quinones. No redox process was observed for pristine FCF (see Figure S4). The values of Epa and Epc are, respectively, −0.05 and −0.13 V for FCF-O and, respectively, −0.33 and −0.84 V for FCF-AQ. To obtain the heterogeneous electron transfer (HET) constant for both electrodes, the Butler−Volmer equation was used since the quinone reaction is controlled by charge transfer at the electrode and not by mass transfer. By using the Butler− Volmer model (see Figure S5), the HET constants at the zero overpotential condition (k0) were estimated to be 6.4 and 11.94 s−1 for FCF-O and FCF-AQ, respectively, and these values are consistent with the literature.29 The electrochemical

Figure 4. Cyclic voltammograms obtained in a phosphate buffer solution (0.1 mol L−1, pH 7.5) for (a) FCF-O and (b) FCF-AQ. (c,d) Magnified regions of FCF-O (a) and FCF-AQ (b) at different scan rates of 5−1000 mV s−1 (FCF-O) and 5−500 mV s−1 (FCF-AQ). Cyclic voltammograms obtained in a phosphate buffer solution (0.1 mol L−1, pH 7.5) containing NADH (1.0 mmol L−1) for (e) pristine FCF (black line) and FCF-O (red line) and (f) pristine FCF (black line) and FCF-AQ (red line) at a scan rate of 5 mV s−1. All experiments were performed in an Ar atmosphere at T = 25 °C.

behavior toward NADH oxidation was then evaluated in the presence of 1.0 mmol L−1 NADH, as shown in Figure 4e,f. Unlike pristine FCF, the presence of NADH in solution causes in drastic changes in the cyclic voltammograms, wherein the anodic peak current increases and the cathode peak disappears. The wave shape indicates the typical electrooxidation of NADH on the surfaces of both electrodes. The value of Epa is 0.62 V for FCF-O and 0.72 V for FCF-AQ in the presence of NADH in solution. The FCF-AQ electrode shows a higher activity than FCF-O, as confirmed by the peak current density of 0.24 mA cm−2, which is 3 times higher than 0.08 mA cm−2 observed for the FCF-O electrode. As expected, the presence of quinones on the electrode surface enhanced the Faradaic current. In the following text, the nature of this behavior is described in terms of operando EPR. Operando EPR and NADH Electrocatalysis. Operando EPR experiments were carried out using an X-band EPR spectrometer coupled with a potentiostat/galvanostat (see Figure S6) for the electrochemical regeneration of NAD+ from NADH catalyzed by quinone-modified carbon-based electrode. The electro-oxidation of NADH is extremely necessary to regenerate an NAD-dependent enzyme. Previous studies showed in situ electrochemical EPR for investigation of charge transfer across the liquid/liquid interface16 and to probe electrical double-layer capacitance,17 as well as to study the electrochemistry of polyaniline films,18,21 carbon nanotubes and fullerene,20 enzyme,22 the formation of mossy lithium and dendrite on lithium anodes, and redox mechanisms in lithiumion batteries.23,24 Here, we propose observing how a redox process taking place on the surface of carbon electrodes affects the concentration of unpaired electrons. An operando EPR electrochemical cell was designed for use with aqueous electrolyte. This cell consists of a capillary system with three electrodes comprising pristine FCF, FCF-O, or FCF-AQ as the working electrode (WE), platinum as the counter electrode, and saturated Ag/AgCl as the reference electrode, as schematically shown in Figure 5a. Using this 16061

DOI: 10.1021/acs.jpcc.9b01160 J. Phys. Chem. C 2019, 123, 16058−16064

Article

The Journal of Physical Chemistry C

Figure 5. (a) Operando EPR cell design for applications in electrochemistry. X-band EPR spectra of pristine FCF (black line), FCF-O (red line), and FCF-AQ (blue line) in a phosphate buffer solution (0.1 mol L−1, pH 7.5) in (b) the absence of NADH and (c) the presence of NADH (1.0 mmol L−1) at 0.60 V. (d) Spin concentration obtained with FCF-O (orange bars) and FCF-AQ (green bars) in a phosphate buffer solution (0.1 mol L−1, pH 7.5) in the absence of NADH and in the presence of NADH (1.0 mmol L−1). (e) Plot of the relative spin concentration versus the applied potential for FCF-O in the absence of NADH (red dots) and in the presence of NADH (1.0 mmol L−1) (black dots). All measurements were carried out at 273 K.

concentration with the applied potential suggests that the kinetics of NADH electro-oxidation has a Butler−Volmer behavior, as in the polarization curve. This result also corroborates that the limiting step in the reaction is NADH oxidation since the observed values of k0 were 6.4 and 11.94 s−1 for FCF-O and FCF-AQ, respectively. Since an unpaired electron is associated with the semiquinone radical, the EPR results suggest a new mechanism for electron transfer, as shown in Figure 6. The proposed mechanism is based on clear surface. This is a typical characteristic of electrocatalytic behavior. A two-step mechanism could be assumed to explain the electro-oxidation of NADH by the quinones on the surface of the electrode. It is possible that there is proton-coupled

experimental setup, the X-band EPR spectra are recorded simultaneously with chronoamperometric measurements after 50 s, when the SS current is obtained. Figure 5b shows the Xband EPR spectra of pristine FCF (black), FCF-O (red), and FCF-AQ (blue) electrodes in a phosphate buffer (0.1 mol L−1, pH 7.5) in the absence of NADH at 0.60 V. Increase in the EPR signal intensities is observed for the FCF-O and FCF-AQ electrodes compared with pristine FCF, suggesting that the presence of quinone functionalities causes an increase in the number of unpaired electrons on the electrode surface. Interestingly, in the presence of NADH (Figure 5c), the concentration of unpaired electrons increases for all electrodes. However, for FCF-O and FCF-AQ, the EPR signal is significantly higher compared to that for pristine FCF. For FCF-O, the concentration of unpaired electrons (spin concentration, SC) increased to 44% (ΔSC) in the presence of NADH whereas ΔSC was 80% for FCF-AQ (Figure 5d). This value indicates that FCF-AQ presents a superior activity for the oxidation of NADH. The spin ratio between FCF-O and FCF-AQ (SRFCF‑O/FCF‑AQ) provides useful quantitative information about the electro-oxidation process. SRFCF‑O/FCF‑AQ can be determined as follows SR FCF ‐ O/FCF ‐ AQ =

ΔSCFCF ‐ O ΔSCFCF ‐ AQ

(1)

SRFCF‑O/FCF‑AQ is 0.75, which indicates a higher unpaired electron concentration for FCF-AQ under the operando condition (in the presence of NADH under 0.60 V). This value agrees with the ratio of the surface concentration of quinone groups for FCF-O (1.3 × 10−9 mol cm−2) and FCFAQ (2.8 × 10−9 mol cm−2), which is 0.46. This concordance corroborates the hypothesis that quinone interacts with NADH molecules during the electro-oxidation reaction. Figure 5e shows a plot of the relative spin concentration versus the potential for FCF-O in a phosphate buffer in the absence of NADH (shown by the red line) and in the presence of NADH (shown by the black line). The increase in the spin

Figure 6. Mechanism of the formation of unpaired electrons on the surface of a quinone-modified carbon electrode. Semiquinone is a free radical. This scheme shows an anion radical as an intermediate between the fully reduced and fully oxidized states catalyzed by NADH. The NADH intramolecular reactions are omitted in this model. 16062

DOI: 10.1021/acs.jpcc.9b01160 J. Phys. Chem. C 2019, 123, 16058−16064

The Journal of Physical Chemistry C electron transfer taking place at the surface. However, quinones are also known to undergo single electron reduction to a semiquinone. EPR measurements were carried out with an applied potential (SS current), and two steps can be considered for electron transfer during NADH oxidation: (1) the protoncoupled electron transfer from NADH to quinone and (2) the reduction of hydroquinone. Semiquinone is a free radical that can be assumed as a possible intermediate electron donor to the electrode surface, promoting an increase of free unpaired electrons. The scheme in Figure 6 shows an anion radical as an intermediate between the fully reduced and fully oxidized states catalyzed by NADH. In the past, this reaction was proposed in terms of abortive side reactions caused by the initial reaction between catechol (from NADH) and o-quinone, “presumably caused by an intermediate semiquinone”.8 For the first time, we show a new correlation for the spin content that clearly reveals an increasing number of free unpaired electrons with increasing applied overpotential and NADH oxidation, which corroborates the fact that quinone groups on carbon surfaces act as electrocatalysts toward the oxidation of NADH to NAD+.

ACKNOWLEDGMENTS



REFERENCES

(1) May, S. W.; Padgette, S. R. Oxidoreductase Enzymes in Biotechnology: Current Status and Future Potential. Nat. Biotechnol. 1983, 1, 677−686. (2) Godbole, S. S.; D’Souza, S. F.; Nadkarni, G. B. Regeneration of NAD(H) by Alcohol Dehydrogenase in Gel-Entrapped Yeast Cells. Enzyme Microb. Technol. 1983, 5, 125−128. (3) Pereira, A.; de Souza, J.; Gonçalves, A.; Pagnoncelli, K.; Crespilho, F. Bioelectrooxidation of Ethanol Using NAD-Dependent Alcohol Dehydrogenase on Oxidized Flexible Carbon Fiber Arrays. J. Braz. Chem. Soc. 2017, 28, 1698−1707. (4) de Souza, J. C. P.; Silva, W. O.; Lima, F. H. B.; Crespilho, F. N. Enzyme Activity Evaluation by Differential Electrochemical Mass Spectrometry. Chem. Commun. 2017, 53, 8400−8402. (5) Carlson, B. W.; Miller, L. L. Mechanism of the Oxidation of NADH by Quinones. Energetics of One-Electron and Hydride Routes. J. Am. Chem. Soc. 1985, 107, 479−485. (6) Jaegfeldt, H.; Torstensson, A. B. C.; Gorton, L. G. O.; Johansson, G. Catalytic Oxidation of Reduced Nicotinamide Adenine Dinucleotide by Graphite Electrodes Modified with Adsorbed Aromatics Containing Catechol Functionalities. Anal. Chem. 1981, 53, 1979− 1982. (7) c̆eá s, N. K.; Kanapieniené, J. J.; Kulys, J. J. NADH Oxidation by Quinone Electron Acceptors. Biochim. Biophys. Acta, Bioenerg. 1984, 767, 108−112. (8) Gorton, L.; Domínguez, E. Electrocatalytic Oxidation of NAD(P) H at Mediator-Modified Electrodes. Rev. Mol. Biotechnol. 2002, 82, 371−392. (9) Murata, K.; Nakamura, N.; Ohno, H. Electrocatalytic Oxidation of NADH on Electrochemically Oxidized Carbon Fiber Paper Electrodes for a Simple Biofuel Cell Anode. Electrochemistry 2008, 76, 563−565. (10) Haas, J.; Schätzle, M. A.; Husain, S. M.; Schulz-Fincke, J.; Jung, M.; Hummel, W.; Müller, M.; Lüdeke, S. A Quinone Mediator Drives Oxidations Catalysed by Alcohol Dehydrogenase-Containing Cell Lysates. Chem. Commun. 2016, 52, 5198−5201. (11) Liu, T.; Frith, J. T.; Kim, G.; Kerber, R. N.; Dubouis, N.; Shao, Y.; Liu, Z.; Magusin, P. C. M. M.; Casford, M. T. L.; Garcia-Araez, N.; Grey, C. P. The Effect of Water on Quinone Redox Mediators in Nonaqueous Li-O 2 Batteries. J. Am. Chem. Soc. 2018, 140, 1428− 1437. (12) Preger, Y.; Gerken, J. B.; Biswas, S.; Anson, C. W.; Johnson, M. R.; Root, T. W.; Stahl, S. S. Quinone-Mediated Electrochemical O2 Reduction Accessing High Power Density with an Off-Electrode CoN/C Catalyst. Joule 2018, 2, 2722−2731. (13) Zhu, W.; Yu, D.; Shi, M.; Zhang, Y.; Huang, T. QuinoneMediated Microbial Goethite Reduction and Transformation of Redox Mediator, Anthraquinone-2,6-Disulfonate (AQDS). Geomicrobiol. J. 2017, 34, 27−36. (14) van der Zee, F. P.; Bouwman, R. H. M.; Strik, D. P. B. T. B.; Lettinga, G.; Field, J. A. Application of Redox Mediators to Accelerate the Transformation of Reactive Azo Dyes in Anaerobic Bioreactors. Biotechnol. Bioeng. 2001, 75, 691−701.

CONCLUSIONS Operando EPR spectroscopy proved to be very useful for investigating the NADH/NAD+ redox reaction. We introduce how to probe the number of free electrons on a carbon electrode surface and correlate it with the electrocatalytic mechanism. The correlation with the spin concentration reveals an increasing number of free unpaired electrons with increasing applied overpotential and NADH oxidation, which corroborates the controversial hypothesis that quinone groups act as electrocatalysts toward the oxidation of NADH to NAD+. Furthermore, operando EPR provides useful information by probing the electron transfer dynamics on a carbon surface and we anticipate that this approach can be extended to other chemical dynamics systems involving electron transfer. ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.9b01160. Experimental procedure; studies of the effect of pH and the scan rate and calculation of the heterogeneous electron transfer rate, Raman spectroscopy, and electrochemical behavior of pristine FCF (Figures S1−S7); and positions of D, G, and A bands and the calculated ID/IG ratios (Table S1) (PDF)





The authors would like to acknowledge the Third World Academy of Sciences (TWAS) and gratefully acknowledge the financial support provided by the Sao Paulo Research Foundation (FAPESP) for the ongoing projects of F.N.C. (Grant Nos. 2015/16672-3 and 2013/14262-7), A.H. (Grant No. 2016/25806-6), D.R.C. (Grant No. 2017/01189-0) and G.C.S. (Grant No. 2015/22973-6). F.N.C. and D.R.C. would also like to acknowledge Conselho Nacional de Desenvolvimento Cientifico e Tecnologico (CNPq) for financial support (Project No. 478525/2013-3 and 306491/2015-0).





Article

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Graziela C. Sedenho: 0000-0001-8696-5978 Renato V. Gonçalves: 0000-0002-3372-6647 Frank N. Crespilho: 0000-0003-4830-652X Notes

The authors declare no competing financial interest. 16063

DOI: 10.1021/acs.jpcc.9b01160 J. Phys. Chem. C 2019, 123, 16058−16064

Article

The Journal of Physical Chemistry C

Tensile Strength of PAN-Based Carbon Fibers. Mater. Chem. Phys. 2010, 122, 548−555. (34) Tao, L.; Wang, Q.; Dou, S.; Ma, Z.; Huo, J.; Wang, S.; Dai, L. Edge-Rich and Dopant-Free Graphene as a Highly Efficient MetalFree Electrocatalyst for the Oxygen Reduction Reaction. Chem. Commun. 2016, 52, 2764−2767. (35) de Souza, J. C. P.; Iost, R. M.; Crespilho, F. N. Nitrated Carbon Nanoblisters for High-Performance Glucose Dehydrogenase Bioanodes. Biosens. Bioelectron. 2016, 77, 860−865. (36) Fujimoto, A.; Yamada, Y.; Koinuma, M.; Sato, S. Origins of Sp3 C Peaks in C 1s X-Ray Photoelectron Spectra of Carbon Materials. Anal. Chem. 2016, 88, 6110−6114.

(15) Yeh, S. Y.; Wang, C. M. Anthraquinone-Modified Electrodes, Preparations and Characterizations. J. Electroanal. Chem. 2006, 592, 131−138. (16) Webster, R. D.; Dryfe, R. A. W.; Coles, B. A.; Compton, R. G. In Situ Electrochemical EPR Studies of Charge Transfer across the Liquid/Liquid Interface. Anal. Chem. 1998, 70, 792−800. (17) Wang, B.; Fielding, A. J.; Dryfe, R. A. W. In situ electrochemical electron paramagnetic resonance spectroscopy as a tool to probe electrical double layer capacitance. Chem. Commun. 2018, 54, 3827− 3830. (18) Tang, J.; Allendoerfer, R. D.; Osteryoung, R. A. Simultaneous EPR and Electrochemical Measurements on Polyaniline in Ambient Temperature Molten Salts. J. Phys. Chem. 1992, 96, 3531−3536. (19) Piette, L. H.; Ludwig, P.; Adams, R. N. Electron Paramagnetic· Resonance and Electrochemistry Studies of Electrochemically Generated Radical Ions in Aqueous Solution. Anal. Chem. 1962, 34, 916−921. (20) Tarábek, J.; Kavan, L.; Kalbác,̌ M.; Rapta, P.; Zukalová, M.; Dunsch, L. In situ EPR spectroelectrochemistry of single-walled carbon nanotubes and C60 fullerene peapods. Carbon 2006, 44, 2147−2154. (21) Lapkowski, M.; Geniés, E. M. Evidence of two kinds of spin in polyaniline from in situ EPR and electrochemistry. J. Electroanal. Chem. Interfacial Electrochem. 1990, 279, 157−168. (22) Zumft, W. G.; Mortenson, L. E.; Palmer, G. ElectronParamagnetic-Resonance Studies on Nitrogenase Investigation of the Oxidation-Reduction Behaviour of Azoferredoxin and Molybdoferredoxin with Potentiometric and Rapid-Freeze Techniques. Eur. J. Biochem. 1974, 46, 525−535. (23) Sathiya, M.; Leriche, J.-B.; Salager, E.; Gourier, D.; Tarascon, J.M.; Vezin, H. Electron paramagnetic resonance imaging for real-time monitoring of Li-ion batteries. Nat. Commun. 2015, 6, No. 6276. (24) Wandt, J.; Marino, C.; Gasteiger, H. A.; Jakes, P.; Eichel, R.-A.; Granwehr, J. Operando electron paramagnetic resonance spectroscopy − formation of mossy lithium on lithium anodes during charge− discharge cycling. Energy Environ. Sci. 2015, 8, 1358−1367. (25) Tse, D. C.-S.; Kuwana, T. Electrocatalysis of Dihydronicotinamide Adenosine Diphosphate with Quinones and Modified Quinone Electrodes. Anal. Chem. 1978, 50, 1315−1318. (26) Kullapere, M.; Seinberg, J.-M.; Mäeorg, U.; Maia, G.; Schiffrin, D. J.; Tammeveski, K. Electroreduction of Oxygen on Glassy Carbon Electrodes Modified with in Situ Generated Anthraquinone Diazonium Cations. Electrochim. Acta 2009, 54, 1961−1969. (27) Kullapere, M.; Marandi, M.; Sammelselg, V.; Menezes, H. A.; Maia, G.; Tammeveski, K. Surface Modification of Gold Electrodes with Anthraquinone Diazonium Cations. Electrochem. Commun. 2009, 11, 405−408. (28) Weissmann, M.; Crosnier, O.; Brousse, T.; Bélanger, D. Electrochemical Study of Anthraquinone Groups, Grafted by the Diazonium Chemistry, in Different Aqueous Media-Relevance for the Development of Aqueous Hybrid Electrochemical Capacitor. Electrochim. Acta 2012, 82, 250−256. (29) Martins, M. V. A.; Pereira, A. R.; Luz, R. A. S.; Iost, R. M.; Crespilho, F. N. Evidence of Short-Range Electron Transfer of a Redox Enzyme on Graphene Oxide Electrodes. Phys. Chem. Chem. Phys. 2014, 16, 17426−17436. (30) Pereira, A. R.; de Souza, J. C. P.; Iost, R. M.; Sales, F. C. P. F.; Crespilho, F. N. Application of Carbon Fibers to Flexible Enzyme Electrodes. J. Electroanal. Chem. 2016, 780, 396−406. (31) Pimenta, M. A.; Dresselhaus, G.; Dresselhaus, M. S.; Cançado, L. G.; Jorio, A.; Saito, R. Studying Disorder in Graphite-Based Systems by Raman Spectroscopy. Phys. Chem. Chem. Phys. 2007, 9, 1276−1290. (32) Jawhari, T.; Roid, A.; Casado, J. Raman Spectroscopic Characterization of Some Commercially Available Carbon Black Materials. Carbon 1995, 33, 1561−1565. (33) Liu, J.; Tian, Y.; Chen, Y.; Liang, J.; Zhang, L.; Fong, H. A Surface Treatment Technique of Electrochemical Oxidation to Simultaneously Improve the Interfacial Bonding Strength and the 16064

DOI: 10.1021/acs.jpcc.9b01160 J. Phys. Chem. C 2019, 123, 16058−16064