The following article is Open access

ATR-SEIRAS Method to Measure Interfacial pH during Electrocatalytic Nitrate Reduction on Cu

, and

Published 15 April 2024 © 2024 The Author(s). Published on behalf of The Electrochemical Society by IOP Publishing Limited
, , Citation Elizabeth R. Corson et al 2024 J. Electrochem. Soc. 171 046503 DOI 10.1149/1945-7111/ad3a22

1945-7111/171/4/046503

Abstract

This study reports the accuracy and applications of an attenuated total reflectance–surface-enhanced infrared absorption spectroscopy (ATR–SEIRAS) technique to indirectly measure the interfacial pH of the electrolyte within 10 nm of the electrocatalyst surface. This technique can be used in situ to study aqueous electrochemical reactions with a calibration range from pH 1–13, time resolution down to 4 s, and an average 95% confidence interval of 14% that varies depending on the pH region (acidic, neutral, or basic). The method is applied here to electrochemical nitrate reduction at a copper cathode to demonstrate its capabilities, but is broadly applicable to any aqueous electrochemical reaction (such as hydrogen evolution, carbon dioxide reduction, or oxygen evolution) and the electrocatalyst may be any SEIRAS-active thin film (e.g., silver, gold, or copper). The time-resolved results show a dramatic increase in the interfacial pH from pH 2–7 in the first minute of operation during both constant current and pulsed current experiments where the bulk pH is unchanged. Attempts to control the pH polarization at the surface by altering the electrochemical operating conditions—lowering the current or increasing the pulse frequency—showed no significant change, demonstrating the challenge of controlling the interfacial pH.

Highlights

  • Infrared spectroscopy method measures interfacial pH within 10 nm of the electrode.

  • This method can be used to study aqueous electrochemical reactions from pH 1–13.

  • Interfacial pH rises from pH 2–7 during nitrate reduction; the bulk pH is constant.

Export citation and abstract BibTeX RIS

This is an open access article distributed under the terms of the Creative Commons Attribution 4.0 License (CC BY, http://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse of the work in any medium, provided the original work is properly cited.

In the face of global climate change, electrochemical conversion for the production of chemicals is a renewable alternative to traditional thermochemical processes. Rather than burning fossil fuels for energy, the driving force can come from renewable electricity, and petroleum-based reactants are replaced with abundant precursors such as water, carbon dioxide (CO2), and other small molecules. These electrochemical reactions often consume or produce protons (H+) and hydroxide (OH). Because heterogeneous electrocatalysis occurs at the electrode–electrolyte interface, these reactions often generate a pH gradient between the electrode surface and the bulk electrolyte. This results in an interfacial pH that can greatly differ from the bulk pH. 13 The pH can strongly influence the selectivity and activity of electrochemical reactions. For example, acidic conditions during nitrate (${\mathrm{NO}}_{3}^{-}$) reduction have been shown to favor ammonia (NH3) formation on copper (Cu) and titanium (Ti) electrodes, 47 and basic conditions during CO2 reduction have been shown to favor the formation of C2+ products at Cu electrodes. 8,9 Therefore, it is important to understand the pH of the reaction environment and develop strategies for controlling the interfacial pH.

Numerous techniques to measure the interfacial pH have been reported, and were recently reviewed by Monteiro et al. 3 Key differences between these techniques are how close the measurement occurs to the electrode–electrolyte interface, whether or not the results are spatially resolved, measurement time, mass transport control, and compatibility of the probe molecule with a given electrochemical system. Various types of scanning probe microscopy (SPM) can spatially resolve the local pH under reaction conditions, either as the probe is moved at a fixed distance across the surface of the working electrode or by moving the probe away from the surface. However, the proximity to the electrode surface is limited by the size of the probe tip. SPM resolution is typically on the order of μm, but the use of nanoelectrodes can bring the tip to within 100 nm of the working electrode surface. 10 A key challenge with SPM techniques is difficulty in synthesizing the probes. Potentiometric pH sensors have a response time on the order of a few seconds while voltammetric pH sensors can go faster, depending on the time it takes for a cyclic voltammetry (CV) scan. A rotating ring-disc electrode (RRDE) technique may be used to measure local pH, which has the benefit of having well-defined mass transport conditions but provides no spatial resolution. The response time is a function of the rotation rate as well as the material of the pH sensor. For example, Zimer et al. used a RRDE with iridium oxide (IrOx) as the potentiometric pH sensor and found response times of 4–12 s (1800 rpm and 300 rpm, respectively). 11 Fluorescence microscopy does provide spatial resolution and, unlike SPM techniques, a spatially resolved map of the local pH can be generated at every time point. For electrochemical systems, the measurements start around 1 μm from the electrode surface and have a resolution limit of 200 nm. 3,12 Xin et al. reported a time resolution of 3.8 frame s−1. 12 This technique is limited by the stability of the fluorescent probe molecule and the possibility that it may affect the electrochemical reaction.

In this study we explore the accuracy and applications of an attenuated total reflectance–surface-enhanced infrared absorption spectroscopy (ATR–SEIRAS) technique to measure the interfacial pH. The surface enhancement enables quantitative measurement of the average pH in the first 5–10 nm of electrolyte from the working electrode surface, 13 much closer to the electrode–electrolyte interface than the techniques listed above. However, unlike SPM and fluorescence methods, ATR–SEIRAS provides no spatial resolution. Although not employed in this study, it is possible to perform ATR–SEIRAS with a flow cell, achieving tunable mass transport conditions, 14 albeit not as well-defined as the mass transport of a RRDE. The measurement time can be tuned by changing the resolution (i.e., number of data points collected), number of averaged spectra, and wavenumber range; in this study the fastest measurement time was 4 s. The surface enhancement can be achieved through a number of SEIRAS-active thin films: silver (Ag), gold (Au), and Cu are commonly used for SEIRAS and as electrocatalysts. 13 In infrared (IR) spectroscopy, the peak areas change as a function of concentration. Despite this, individual peaks cannot be used to quantify concentration because their size is so dependent on the film morphology, which can change over the course of an experiment. While the concentration of H+ and OH cannot be directly measured by IR spectroscopy, the interfacial pH can be indirectly measured through the ratio of peak areas of two species that compose a buffer system. Different buffers may be used; the selected buffer must be IR-active, cover the relevant pH region, and ideally have peaks that do not overlap with the reactant and product peaks of the reaction of interest. This technique has been used to study the interfacial pH during CO2 reduction: a bicarbonate (HCO3 ) buffer with a Au electrode from pH 7–13, 15,16 and a phosphate buffer with a Cu electrode from pH 5–13. 17 The same concept of measuring the pH indirectly through buffers can also be performed using surface-enhanced Raman spectroscopy (SERS). 18,19 Choosing between SEIRAS and SERS depends on the buffer and reactant identity because peaks that overlap in one technique may appear distinct in the other due to the different selection rules. 18 The ATR–SEIRAS interfacial pH measurement presented here may be applied to any electrochemical reaction that can occur in a buffered aqueous solution at an electrode made from a SEIRAS-active material. In addition to CO2 reduction, other possible reactions include ${\mathrm{NO}}_{3}^{-}$ reduction, the hydrogen evolution reaction (HER), and the oxygen evolution reaction (OER).

Here we describe the interfacial pH calibration procedure for a phosphate buffer using ATR–SEIRAS, expanding the calibration range to pH 1–13 and quantifying the accuracy of the method, which has not been previously reported. We then demonstrate the capabilities of this technique by comparing time-resolved studies of ${\mathrm{NO}}_{3}^{-}$ reduction using constant current and pulsed current to evaluate the possible use of pulsed electrolysis to control the interfacial pH.

Experimental

Electrode preparation

An ATR crystal served as the support for the thin film Cu working electrode (Fig. S1). The 60° germanium (Ge) ATR crystal (013-3132, Pike Technologies) was polished with diamond suspensions of 15, 3, and 1 μm (MF-2051, MF-2059, and MF-2054, BASi) using TexMet C pads (401102, Buehler) for the 15 μm polish and MicroCloth pads (407212, Buehler) for the 3 and 1 μm polishes. The crystal was sonicated in water in between each polishing step. Nanopure water (resistivity: 18.2 MΩ · cm, Millipore Sigma) was used for all experiments and measurements. The crystal surface was cleaned with water and isopropyl alcohol using lint-free wipes and blow-dried with nitrogen (N2). The crystal was cleaned with oxygen (O2) plasma for 3 min on high power (Pico, Diener Electronic). A thin film (<100 nm) of Cu was deposited on the crystal surface by contacting the surface with a plating solution (40 mM cupric sulfate (CuSO4·5H2O) + 0.2 M ethylenediaminetetraacetic acid (EDTA) + 0.2 M glyoxylic acid + 1 mM 2,2'-bipyridine, adjusted to pH 13 with sodium hydroxide (NaOH)) at 70 °C for 4 min. 20 The deposition was performed in a custom-designed cell 21 (Fig. S1) in a water bath to maintain a constant temperature for both the crystal and the plating solution. To confirm a successful deposition, the resistance across the surface of the thin film was measured by a multimeter and verified to be consistent (1–2 Ω for all synthesized films in this study).

ATR–SEIRAS

ATR–SEIRAS experiments were performed on a Thermo Scientific Nicolet iS50 FTIR Spectrometer with a Pike Technologies VeeMax III attachment in a custom, single-chamber electrochemical cell (Fig. S1). 21 The counter electrode was a graphite rod (40766, Alfa Aesar), and the reference electrode was a leakless Ag/AgCl electrode (3.4 M potassium chloride (KCl), ET072-1, eDAQ). All currents were converted to current density by dividing by the geometric surface area of the working electrode (2 cm2). All potentials were converted to the reversible hydrogen electrode (RHE) scale using the bulk pH unless otherwise noted. Each thin film Cu electrode was electrochemically conditioned by three CV scans from −0.40 to 0.50 VRHE at 50 mv s−1 for three cycles in 0.1 M sodium perchlorate (NaClO4) (Fig. S2).

SEIRAS spectra for calibration and constant current experiments were an average of 32 spectra taken at a resolution of 4 cm−1 for a total duration of 22 s. SEIRAS spectra for pulsed current experiments were an average of 4 spectra taken at a resolution of 16 cm−1 for a total duration of 4 s. The pulsed current method on the potentiostat and the ATR–SEIRAS time series were started at the same time and the timestamp of each spectrum was correlated with the time recorded by the potentiostat to ensure the methods were synchronized. A background spectrum of water was taken at the start of each set of experiments. The negative logarithm of the ratio between the single-beam sample spectrum (R) and the single-beam background spectrum (R0) gives the absorbance (A) spectrum of the sample in absorbance units (a.u.): $A=-\mathrm{log}(\tfrac{R}{{R}_{0}})$.

The bulk pH was measured with a pH meter (FiveEasy, Mettler Toledo) before and after each experiment. At the conclusion of each experiment, chronoamperometry (CA) was performed at 0.1 VRHE and a SEIRAS spectrum was collected after 1 min to check the stability of the film.

Interfacial pH calibration data were gathered in 0.5 M phosphate buffer solutions during CA at 0.1 VRHE after 1 min. The pH was adjusted in increments of approximately 0.2 pH units by adding 0.1 M NaOH or 0.1 M phosphoric acid (H3PO4). All experiments, including calibration, were performed at room temperature, because Keq changes as a function of temperature.

Results and Discussion

Phosphate peak identification

The four species that comprise the phosphate buffer system are H3PO4, dihydrogen phosphate (H2PO4 ), hydrogen phosphate (HPO${}_{4}^{2-}$), and the phosphate ion (PO${}_{4}^{3-}$). The phosphate buffer equilibrium reactions are

Equation (1)

Equation (2)

Equation (3)

The relationship between pH and the calibration ratios is derived from the equilibrium Eqs. (Table S1). For example, rearranging the equation for K1, we find that the pH is proportional to the log of the calibration ratio (derivation shown in the Supplemental Material).

Equation (4)

Three calibration regions were defined: acidic (pH 1–3), neutral (pH 5–9), and basic (pH 9–13). Representative ATR–SEIRAS spectra in each of the three pH regions are shown in Fig. 1. In each region, the calibration curve is based on the ratio of the peak areas of the two phosphate species with the highest concentration in that region (Fig. 2). The relationship between the pH and the calibration ratio is shown in Eq. 4. The calibration ratios, average peak positions, and standard deviation (SD) of peak positions are shown in Table I.

Figure 1.

Figure 1. Representative ATR–SEIRAS spectra at pH 1.5, pH 6.4, and pH 11.8 demonstrating the three calibration regions: acidic (pH 1-3), neutral (pH 5-9), and basic (pH 9-13). Black curves represent the baseline-corrected spectra and colored curves show the Gaussian fit for each peak. Shaded peaks are used for the calibration ratios (Table I) and are labeled with the vibrational mode and phosphate species. Measurements taken in 0.5 M sodium phosphate buffer electrolyte at 0.1 VRHE at a Cu thin film electrode.

Standard image High-resolution image
Figure 2.

Figure 2. Phosphate speciation from pH 0–14. Curves represent the relative fraction of each phosphate species at a given pH, as calculated from the Keq values (Table S1). Symbols show the normalized peak areas from one data replicate for the peaks used in the calibration ratios. Data for all replicates are shown in Fig. S6. Measurements taken in 0.5 M sodium phosphate buffer electrolyte at 0.1 VRHE at a Cu thin film electrode.

Standard image High-resolution image

Table I. Calibration ratios for the three calibration regions: acidic (pH 1–3), neutral (pH 5–9), and basic (pH 9–13). Reported peak positions are the average values from the three calibration data replicates. The relationship between the pH and the calibration ratio is shown in Eq. 4.

Calibration Ratios
RegionpH rangeCalibration ratioPeak position/cm−1 Standard deviation/cm−1
AcidicpH 1–3 $\displaystyle \frac{{\nu }_{s}(\mathrm{PO}){{\rm{H}}}_{3}{\mathrm{PO}}_{4}}{{\nu }_{a}(\mathrm{POH}){{\rm{H}}}_{2}{\mathrm{PO}}_{4}^{-}}$ 100818.4
   9472.7
NeutralpH 5–9 $\displaystyle \frac{{\nu }_{a}(\mathrm{PO}){{\rm{H}}}_{2}{\mathrm{PO}}_{4}^{-}}{{\nu }_{s}(\mathrm{PO}){\mathrm{HPO}}_{4}^{2-}}$ 11359.6
   9901.2
BasicpH 9–13 $\displaystyle \frac{{\nu }_{a}(\mathrm{PO}){\mathrm{HPO}}_{4}^{2-}}{{\nu }_{a}(\mathrm{PO}){\mathrm{PO}}_{4}^{3-}}$ 10825.2
   10186.0

Peaks were identified from literature and confirmed through peak area trends as the pH was adjusted. For example, in the acidic region, the asymmetrical νa(POH) vibrations of ${{\rm{H}}}_{2}{\mathrm{PO}}_{4}^{-}$ peak at 947 cm−117 increased in area as the pH increased, and the symmetrical νs(PO) vibrations of H3PO4 peak at 1008 cm−1 22 decreased in area with increasing pH (Fig. S3), as expected for phosphate speciation in this pH region. The peak evolution in the acidic, neutral, and basic calibration regions can be seen in Figs. S3, S4, and S5, respectively.

In the acidic pH range (pH 1–3), the calibration ratio was the peak area from symmetrical νs(PO) vibrations of H3PO4 (1008 cm−1, SD 18.4) 22 to the peak area from asymmetrical νa(POH) vibrations of ${{\rm{H}}}_{2}{\mathrm{PO}}_{4}^{-}$ (947 cm−1, SD 2.7). 17 These two peaks were selected for the calibration ratio because the area trends were consistent with the concentration trends of the individual species (Fig. S3). In contrast, the area trends of the peaks at 1070 and 1155 cm−1 revealed contributions from both H3PO4 and ${{\rm{H}}}_{2}{\mathrm{PO}}_{4}^{-}$ that could not be deconvoluted.

In the neutral pH range (pH 5–9), the calibration ratio was the peak area from asymmetrical νa(PO) vibrations of ${{\rm{H}}}_{2}{\mathrm{PO}}_{4}^{-}$ (1135 cm−1, SD 9.6) to the peak area from symmetrical νs(PO) vibrations of HPO${}_{4}^{2-}$ (990 cm−1, SD 1.2). 17 The peak at 940 cm−1 was not selected due to low sensitivity (Fig. S4) and the peak at 1080 cm−1 was not used because the area trends revealed contributions from both ${{\rm{H}}}_{2}{\mathrm{PO}}_{4}^{-}$ and HPO${}_{4}^{2-}$ that could not be deconvoluted.

In the basic pH range (pH 9–13), the calibration ratio was the peak area from asymmetrical νa(PO) vibrations of HPO${}_{4}^{2-}$ (1082 cm−1, SD 5.2) to the peak area from asymmetrical νa(PO) vibrations of PO${}_{4}^{3-}$ (1018 cm−1, SD 6.0). 17 The peak at 990 cm−1 was not selected due to low sensitivity (Fig. S4).

Due to the similar appearance of the four peaks in the acidic and neutral regions (Fig. 1), a quantitative method using the position of the second peak was established for peak assignment. If the second peak position was ≥998 cm−1 then the peak was assigned to νs(PO) H3PO4 and the acidic calibration curve was used. If the second peak position was <998 cm−1 then the peak was assigned to νs(PO) HPO${}_{4}^{2-}$ and the neutral calibration curve was used. The cutoff value was set to be 2 SD above the average position for νs(PO) HPO${}_{4}^{2-}$ from the constant current data (986 cm−1, SD 6.1). The νs(PO) HPO${}_{4}^{2-}$ peak was selected because there were more data points than the νs(PO) H3PO4 peak (n${}_{{\nu }_{{\rm{s}}}(\mathrm{PO}){\mathrm{HPO}}_{4}^{2-}}$ = 319, ${{\rm{n}}}_{{\nu }_{{\rm{s}}}(\mathrm{PO}){{\rm{H}}}_{3}{\mathrm{PO}}_{4}}=26$), and the constant current data had higher variance than the pulsed current data.

Excellent agreement is seen between the normalized peak areas and the relative fraction of each phosphate species (Fig. 2) calculated from the equilibrium constants (Keq) (Table S1). 23 The normalized peak areas for all data replicates are shown in Fig. S6, demonstrating the same trend. There is a seamless transition at pH 9 from the neutral to basic region. Due to the low sensitivity of the νs(PO) H3PO4 peak (Fig. S3), there is a gap from pH 3–5 at the transition from the acidic to neutral region, and the normalized peak area does not follow the calculated trend line quite as well as the other peaks.

Interfacial pH calibration

Figure 3 shows all calibration data plotted as pH against the log of the calibration ratio, as derived from the equilibrium equation for phosphate speciation (Table S1, Eq. 4). Three complete sets of calibration data were collected for each calibration region (Table S2). The average calibration curves—used to calculate the interfacial pH during electrochemical experiments—are the average of the best fit linear calibration curves for each of the three data replicates in a given calibration region. The slope, intercept, and R2 value of each calibration curve are shown in Table S3.

Figure 3.

Figure 3. All calibration data plotted as interfacial pH against log of the calibration ratio in the (a) acidic, (b) neutral, and (c) basic regions. Calibration ratios are defined in Table I. The average calibration curves (black) are the average of the best fit linear calibration curves for each of the three data replicates (rep) in a given calibration region. The slope, intercept, and R2 value of each calibration curve is shown in Table S3. The shaded area represents the 95% CI (Eq. S1). The margin of error expressed as a percentage (Eq. S2) and as a function of pH is shown in Fig. S7. Measurements taken in 0.5 M sodium phosphate buffer electrolyte at 0.1 VRHE at a Cu thin film electrode.

Standard image High-resolution image

The accuracy for the calibration varies with the pH, as shown by the shaded 95% confidence interval (CI) in Fig. 3 (Eq. S1). The margin of error (MOE) expressed as a percentage (Eq. S2) and as a function of pH is shown in Fig. S7. The region with the lowest accuracy is around pH 1, where the 95% CI is the largest. For example, at pH 1.25 the 95% CI is pH 1.01–1.48, a variability of 38%. The 95% CI for the acidic region improves as the pH increases, with a 17% variability at pH 1.75 and a 1% variability at pH 2.5. In the neutral region, the 95% CI is more evenly distributed, with a 30% variability at pH 5.5 decreasing to a 12% variability at pH 9. The basic region has the most consistent 95% CI, with 6%–8% variability across pH 9–13.

This pH-dependent variability demonstrates the importance of using repeat data sets to establish an average calibration curve and report the margin of error for the technique. This statistical analysis defines the accuracy of calculations based on the interfacial pH, such as the overpotential at the electrode interface. It also sets the limits when comparing with other studies and computational models, because we cannot distinguish between results with values within the 95% CI. The variation is due to subtle changes in the thin film morphology, which can occur during the deposition process, electrochemical conditioning, and electrochemical experiments. Because the thin film morphology and material can strongly influence the sensitivity of the SEIRAS spectra, there will also be some variability in the accuracy of this method between researchers. Thus, it is important to adhere to a well-defined experimental procedure for thin film deposition and consider the individual reported accuracy when comparing results between research groups. In the acidic region, the higher margin of error and the gap from pH 3–5 may be improved by changing the material from Cu to Ag or Au, 13 decreasing the thickness of the film, and increasing the surface roughness. However, these methods of enhancing the SEIRAS signal must also be balanced with film stability, which is increasingly important at higher current densities. Raising the temperature of the electrolyte would alter the phosphate buffer Keq values which could also improve the acidic calibration.

Time evolution of interfacial pH with constant current

${\mathrm{NO}}_{3}^{-}$ reduction was used as an example to demonstrate the capability of this technique to measure the interfacial pH during an aqueous electrochemical reaction. The expected reactions on Cu are ${\mathrm{NO}}_{3}^{-}$ reduction to nitrite (${\mathrm{NO}}_{2}^{-}$) (Eq. 5), ${\mathrm{NO}}_{3}^{-}$ reduction to NH3 (Eq. 6), and hydrogen (H2) evolution (Eq. 7). The corresponding reactions in acidic media are shown in Eqs. S3–S5.

Equation (5)

Equation (6)

Equation (7)

${\mathrm{NO}}_{3}^{-}$ reduction to ${\mathrm{NO}}_{2}^{-}$ and HER each have a generation rate of 1 OH per electron, and ${\mathrm{NO}}_{3}^{-}$ reduction to NH3 has a generation rate of 1.13 OH per electron. Cu has been experimentally 4,2432 and theoretically 33 shown to be selective toward NH3 production, although the selectivity may flip to favor ${\mathrm{NO}}_{2}^{-}$ depending on the applied potential, surface structure, and electrolyte. At the same current density, an electrocatalyst with a higher selectivity toward NH3 (1.13 OH per electron) will experience a greater pH polarization than an electrocatalyst with higher selectivity toward ${\mathrm{NO}}_{2}^{-}$ and H2 (1 OH per electron).

The time evolution of the interfacial pH during ${\mathrm{NO}}_{3}^{-}$ reduction under constant current conditions was investigated in 0.5 M sodium phosphate buffer electrolyte with a bulk pH of 1.84 (0.25 M H3PO4 + 0.25 M sodium dihydrogen phosphate (NaH2PO4)) with 10 mM sodium nitrate (NaNO3) (Fig. 4). ATR–SEIRAS measurements were collected every 22 s for 10 min at four different constant current densities (−3, −5, −6, and −8 mA cm−2) and each condition was repeated three times (Table S2). A subset of corresponding spectra are shown in Fig. S10. Corresponding voltages are shown in Table S4. In all constant current experiments the bulk pH was unchanged.

Figure 4.

Figure 4. Interfacial pH vs. time at −3, −5, −6, and −8 mA cm−2 for data replicate 3. ATR–SEIRAS measurements were collected every 22 s for 10 min while applying a constant current; a subset of corresponding spectra are shown in Fig. S10. Measurements taken in 0.5 M sodium phosphate buffer electrolyte (bulk pH 1.84) with 10 mM NaNO3 at a Cu thin film electrode. The bulk pH remained constant. The final point is taken without applied current and shows the interfacial pH returning to the bulk pH value. The interfacial pH plotted against charge passed can be seen in Fig. S11, average values are shown in Fig. S12, and the voltage is listed in Table S4.

Standard image High-resolution image

As shown in Fig. 4, a gradual increase in the interfacial pH was observed at −3 mA cm−2, the smallest current density investigated. The interfacial pH increased for approximately 3 min before reaching a plateau of pH 5.7. At −5 mA cm−2, the interfacial pH increased more quickly over the span of 2 min, reaching a plateau of pH 6.5. Similar behavior was observed for both −6 and −8 mA cm−2. During the time when the interfacial pH is initially increasing, the potential sharply increases in magnitude before reaching a region of slow growth (Fig. S9). This dynamic period is due to charging of the double layer and changes in the concentration profile in the diffusion layer. 34

The plateau behavior of the interfacial pH demonstrates that the rate of OH produced by the cathodic reactions is matched by phosphate buffering, maintaining the interfacial pH near pKa2 = 7.20 (Eq. 2, Table S1). This plateau phenomenon at pKa2 was also observed by Yang et al. during CO2 reduction in 0.2, 0.5, and 1.0 M phosphate buffer at a Cu electrode. 17 While the buffer is necessary for this type of interfacial pH measurement, differences between electrochemical operating conditions will be suppressed near the buffer pKas. Researchers employing this technique may strategically select the buffer bulk pH and concentration based on the pH range of interest with respect to the pKas of the buffer species.

In the data set shown in Fig. 4 (rep 3), there was a statistically significant difference in the final pH (average of the last three data points, ca. 1 min) between −3 mA cm−2 and the larger current densities. When this interfacial pH data is plotted against charge passed (Fig. S11), the interfacial pH at −3 mA cm−2 during the initial ramp period is still lower than the interfacial pH at the higher current densities at the same charge passed, indicating that the lower current density condition may slightly suppress the pH polarization at the interface. However, when all three data sets are averaged (Fig. S12), there is no statistically significant difference in the final pH between these four current densities. This result indicates that, if there is a slight difference in the interfacial pH between −3 to −8 mA cm−2, it is within the margin of error for this technique. At all four current densities, the plateau pH represents a significant increase in the interfacial pH of 5 pH points on average from the bulk pH (pH 1.84–7.15), even in a heavily buffered electrolyte.

These results demonstrate the importance of a technique that can quantify the pH so close to the electrode–electrolyte interface, often revealing a greater pH polarization than is measured by other methods. This pH polarization perturbs the reaction thermodynamics and kinetics, and under-estimating the extent of the pH gradient can lead to incorrect mechanistic interpretations and an overestimation of the catalytic overpotential due to polarization loss. 2,8,17 A more precise calculation of the overpotential at the electrode surface enables a more accurate comparison between different studies. For example, scanning electrochemical microscopy (SECM, a type of SPM) was used to measure the local pH during ${\mathrm{NO}}_{3}^{-}$ reduction at a Cu electrode by Santos et al. 35 The IrOx-coated Au microelectrode tip had a radius of 10 μm and the closest achievable measurement to the surface was 10 μm. In a 0.1 M sodium sulfate (Na2SO4) solution with 0.5 mM ${\mathrm{NO}}_{3}^{-}$ and a bulk pH of 2 at −1.0 VAg/AgCl (O(-1 mA cm−2)), the local pH at a distance of 10 μm was found to be pH 5.8. Considering this value is lower than the results presented here in a 0.5 M buffered solution at a similar current density (average pH 6.6 at −3 mA cm−2, Fig. S12), we can predict that the interfacial pH in 0.1 M Na2SO4 was likely higher than pH 5.8 at the electrode–electrolyte interface. The SECM measurement at 10 μm from the electrode surface underestimates the actual interfacial pH, which we can more accurately quantify through ATR–SEIRAS because the measurement distance is three orders of magnitude closer to the electrode–electrolyte interface: just 10 nm rather than 10 μm.

A more accurate measurement of the interfacial pH under reaction conditions is important because the proton concentration can impact both the selectivity and kinetics of electrochemical reactions. In another report of ${\mathrm{NO}}_{3}^{-}$ reduction on Cu, Pérez-Gallent et al. found that acidic media favored the formation of NH3 and nitric oxide (NO), while alkaline media favored ${\mathrm{NO}}_{2}^{-}$ and hydroxylamine (NH2OH). 4 Wang et al. performed a computational study of ${\mathrm{NO}}_{3}^{-}$ reduction on Cu and confirmed higher selectivity toward NH3 in acidic conditions. 6 If NH3 is the desired product of ${\mathrm{NO}}_{3}^{-}$ reduction, then an acidic reaction environment is preferred.

Potential strategies to maintain such acidic conditions include buffered electrolytes, mass transport control, and pulsed electrolysis. Clearly, even highly concentrated buffers cannot maintain the desired acidic conditions at the electrode–electrolyte interface, as demonstrated in this study using a 0.5 M buffer and the report by Yang et al. measuring the interfacial pH during CO2 reduction in 0.1–1.0 M buffers. 17 Guo et al. studied the effect of mass transport on ${\mathrm{NO}}_{3}^{-}$ reduction at a Ti electrode by varying the flow rate. 7 As the flow rate increased, the diffusion layer thickness decreased, enhancing the nitrate diffusion driving force. These changes increased nitrate reduction activity, producing more ${\mathrm{NO}}_{2}^{-}$, NH3, and OH (Eqs. 5 and 6). While the smaller diffusion layer thickness would also enhance the OH diffusion driving force away from the electrode surface, a continuum model simulation found that the interfacial pH actually increased with increasing flow rate. This pH trend demonstrates that improving the mass transport increases the production of OH more than it enhances the dissipation of OH under the experimental conditions. In the next Section we explore the potential application of pulsed electrolysis to control the interfacial pH.

Time evolution of interfacial pH with pulsed current

The time evolution of the interfacial pH during ${\mathrm{NO}}_{3}^{-}$ reduction under pulsed current conditions was investigated in 0.5 M sodium phosphate buffer electrolyte with a bulk pH of 1.84 (0.25 M H3PO4 + 0.25 M NaH2PO4) with 10 mM NaNO3 (Fig. 5). Experiments were conducted for 10 min at four different pulsed current conditions (-5 mA cm−2 for 8 s and −5, −6, and −8 mA cm−2 for 4 s), and each condition was repeated three times (Table S2). A subset of corresponding spectra are shown in Fig. S13. Corresponding voltages are listed in Table S5. During pulsed current experiments, the applied current was a symmetric square wave alternating between current on and current off (zero current, open circuit). For example, at the −5 mA cm−2, 4 s pulse condition, −5 mA cm−2 was applied for 4 s followed by open circuit for 4 s, and repeated over the course of 10 min. An example of the pulsed current and corresponding voltage is shown in Fig. S14. ATR–SEIRAS spectra were collected every 4 s, one for each pulse under the 4 s conditions and two for each pulse under the 8 s conditions. In all pulsed current experiments the bulk pH was unchanged.

Figure 5.

Figure 5. Interfacial pH vs. time from data replicate 2 while the current was pulsed on and off (zero current, open circuit) in a symmetric square wave for 10 min at (a) −5 mA cm−2 for 8 s, (b) −5 mA cm−2 for 4 s, (c) −6 mA cm−2 for 4 s, and (d) −8 mA cm−2 for 4 s. ATR–SEIRAS measurements were collected every 4 s; a subset of corresponding spectra are shown in Fig. S13. The final point is taken without applied current and shows the interfacial pH returning to the bulk pH value. The ramp time observed in (a)–(d) to switch from the acidic to neutral region is (e) shown for the average of all three data replicates (values for each data set shown in Fig. S15). The average pH of the final minute (f) is shown for the average of all three data replicates for both constant and pulsed current experiments; the values for each data replicate are plotted separately for constant current (Fig. S12) and pulsed current (Fig. S16). Corresponding voltages are listed in Table S5. An example of the pulsed current and corresponding voltage is shown in Fig. S14. Measurements taken in 0.5 M sodium phosphate buffer electrolyte with 10 mM NaNO3 at a Cu thin film electrode. The bulk pH remained constant. Error bars represent one SD in each direction.

Standard image High-resolution image

As shown in Figs. 5a–5d, when the current was off (open circuit) the interfacial pH consistently returned to a value close to that of the bulk pH for all conditions, regardless of the cathodic current magnitude. When the current was on, each condition showed an initial ramp period where the pH steadily increased before reaching a plateau pH for the remainder of the 10 min experiment. The duration of these ramp periods (0.3–1.5 min) were very similar to those observed in the constant current experiments (Fig. 4). Across all pulsed current data replicates, there was a statistically significant trend of the pH ramp time decreasing with increasing pulsed current density (Fig. 5e and Fig. S15). The pH ramp times from the constant current experiments were not statistically analyzed because the longer ATR–SEIRAS measurement time (22 s for constant current versus 4 s for pulsed current) did not have sufficient time resolution to capture the switch from the acidic to neutral region.

The average interfacial pH of the final minute is shown for each data replicate and the average of all three data replicates in Fig. S16, demonstrating no statistically significant difference in final pH between the different pulsed conditions. We selected the final minute as a standard of comparison between data sets and methods because it was the furthest from the dynamic pH ramp period observed at the start of experiments. In Fig. 5f, the average final pH for the constant current and pulsed current conditions is compared. While the average value of the final pH for the 4 s pulsed current at −5, −6, and −8 mA cm−2 was lower than the average value of the final pH for the corresponding constant currents, it was not statistically significant. Thus, while the pulsed current method may slightly suppress the increase in interfacial pH, it is within the margin of error for this technique. There is a transient period when the current is switched from off to on when the interfacial pH must be increasing from the near-bulk value to an average pH of 6.1, but this change occurs faster than the 4 s measurement time.

Pulsed electrolysis has been explored as a method to suppress the concentration polarization that occurs in the diffusion layer during electrocatalysis by promoting ion diffusion that periodically renews the electrode–electrolyte interface, 9,12,36,37 including during ${\mathrm{NO}}_{3}^{-}$ reduction. 7,3840 Guo et al. found that pulsed electrolysis (10 s symmetric pulse) at a Ti cathode during ${\mathrm{NO}}_{3}^{-}$ reduction in a non-buffered electrolyte showed an improvement in NH3 selectivity but also an increase in H2 production when compared to constant potential experiments under the same conditions, and had similar product selectivity to a 0.5 M phosphate buffer with a bulk pH of 10.5. 7 The authors concluded that the pulsed electrolysis in the non-buffered electrolyte resulted in an interfacial pH similar to that of the 0.5 M phosphate buffer and lower than the interfacial pH under a constant applied potential. While this study finds no significant difference in the interfacial pH between the constant current and pulsed current conditions in 0.5 M phosphate buffer, it is possible that the pH polarization in a non-buffered solution is so much greater than in a buffered solution that the pulsed electrolysis is effective at maintaining a lower interfacial pH. Alternatively, the transient lower interfacial pH conditions achieved each time the current is turned back on may be enough to impact the product selectivity. Finally, it may be that resetting the interfacial concentration of ${\mathrm{NO}}_{3}^{-}$ and other electrolyte species accounts for the selectivity changes. Huang et al. similarly found an enhancement in nitrate-to-ammonia selectivity at a carbon-supported ruthenium (Ru) and indium (In) intermetallic compound (RuIn3) during pulsed electrolysis (4 s cathodic, 0.5 s anodic) in a non-buffered electrolyte and used Raman spectroscopy to demonstrate the restoration of the ${\mathrm{NO}}_{3}^{-}$ concentration at the interface during the anodic pulse. 40 The open question of how the pulsed electrolysis is impacting product selectivity in these two studies highlights a challenge with the ATR–SEIRAS technique in that it must be performed in a buffered electrolyte; we cannot confirm the interfacial pH under the non-buffered condition.

Despite this limitation, with the ATR–SEIRAS technique we are able to confirm that the electrolyte renewal does indeed reach the first 10 nm of the electrode–electrolyte interface (Fig. 5). Xin et al. used a fluorescent probe technique to show a restoration of the initial local pH conditions at 1 μm from the electrode–electrolyte interface during chromate (${\mathrm{HCrO}}_{4}^{-}$) reduction to the chromic ion (${\mathrm{Cr}}_{3}^{+}$) (2.3 OH per e). 12 The fluorescent probe method enabled spatially resolved visualization of the diffusion layer (40–80 μm) with a resolution of 200 nm. In the unbuffered electrolyte at 1 μm of the surface, the local pH increased from pH 3 to pH 5 when the voltage was pulsed on and returned to pH 3 when the voltage was pulsed off (5 s and 2.5 s symmetric pulse patterns, no current density reported). Considering our buffered system that produces 1–1.13 OH per electron shows an even higher pH increase (average pH 6.1, Fig. 5f), it is likely that the fluorescent probe technique is not capturing the full pH polarization at the interface. This finding demonstrates the value of the ATR–SEIRAS method to better understand the environment at the electrode–electrolyte interface.

While pulsed electrolysis appears to be effective at restoring the interfacial electrolyte conditions while the current is off, the concentration polarization quickly returns when the current is applied. Thus, pulsed electrolysis may beneficially impact product selectivity by periodically increasing reactant concentrations and decreasing product concentrations at the interface, but it is not effective at maintaining a higher H+ concentration, at least at a 4 s symmetric pulse rate in 0.5 M phosphate buffer. Xin et al. saw a decrease in the pH polarization (pH 4.5 instead of 5) within 1–2 μm of the surface when the pulse time was decreased to 1 s. 12 However, the local pH under these conditions increased over the course of 30 s, similar to the pH ramp observed in this study (Fig. 5e). Had the authors continued the measurement beyond 30 s, it is possible that the local pH at the 1 s pulse time may have reached the same level as the 2.5 and 5 s pulse times. While decreasing the pulse time may be a viable strategy to controlling the interfacial pH, there are limits. The pulse duration must be longer than the charge and discharge times of the double layer, otherwise the current will not reach the desired Faradaic current when pulsed on and it will not reach zero current when pulsed off. 9,37 Future work to decrease the measurement time for the ATR–SEIRAS technique will enable further exploration into the possibility of using pulsed electrolysis to maintain a desired interfacial pH.

Conclusions

Using ATR–SEIRAS to quantify the interfacial pH during electrocatalysis is a powerful tool to develop a better understanding of the reaction environment. While other methods can provide spatial resolution of the diffusion layer on the order of 0.1–100 μm, 3 the ATR–SEIRAS technique probes the average pH within 10 nm of the electrode–electrolyte interface. There is flexibility in the selection of the working electrode material as well as the identity and concentration of the buffer, permitting application to a wide range of conditions and aqueous electrochemical reactions. However, despite the flexibility in buffer selection, the requirement of having a buffered electrolyte can be limiting. While we can generally conclude that a non-buffered system would have a greater pH polarization at the interface than what we measure in a buffered system, we cannot quantify the interfacial pH under the non-buffered conditions. Additionally, while the phosphate buffer provides an excellent way to indirectly measure the interfacial pH, it may alter the product distribution by acting as a proton donor and increasing selectivity toward H2 through electrochemical deprotonation. 41,42

In this study we demonstrated that ATR–SEIRAS calibration curves for a phosphate buffer can be created from pH 1–13 with a variable margin of error: average of 16% for pH 1–3, 20% for pH 5–9, and 7% for pH 9–13. We showed that the time-resolved interfacial pH can be quantified for both constant current and pulsed current experiments with a resolution of 4 s. We found a marked increase in the interfacial pH from pH 2–7 under both constant current and pulsed current conditions with no change to the bulk pH. The pH polarization at the interface is not significantly altered by lowering the current density or increasing the pulse frequency. Future work will look at further reducing the measurement time to enable exploration of the highly dynamic pulsed electrolysis environment.

Acknowledgments

This work was supported by the National Science Foundation EFRI program (Award 2132007) and the Chemical Engineering Department at Stanford University. The authors acknowledge support from the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, Chemical Sciences, Geosciences, and Biosciences Division, Catalysis Science Program to the SUNCAT Center for Interface Science and Catalysis. E.R.C. and J.G. acknowledge support from the TomKat Center for Sustainable Energy. Part of this work was performed at the Stanford Nano Shared Facilities (SNSF), supported by the National Science Foundation under award ECCS-2026822.

Please wait… references are loading.

Supplementary data (6.5 MB PDF)