Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect
ACS Publications. Most Trusted. Most Cited. Most Read
Accessing In Situ Photocorrosion under Realistic Light Conditions: Photoelectrochemical Scanning Flow Cell Coupled to Online ICP-MS
My Activity
  • Open Access
Article

Accessing In Situ Photocorrosion under Realistic Light Conditions: Photoelectrochemical Scanning Flow Cell Coupled to Online ICP-MS
Click to copy article linkArticle link copied!

  • Ken J. Jenewein*
    Ken J. Jenewein
    Forschungszentrum Jülich GmbH, Helmholtz Institute Erlangen-Nürnberg for Renewable Energy (IEK-11), Egerlandstrasse 3, 91058 Erlangen, Germany
    Department of Chemical and Biological Engineering, Friedrich-Alexander-Universität Erlangen-Nürnberg, Egerlandstrasse 3, 91058 Erlangen, Germany
    *Email: [email protected]
  • Attila Kormányos
    Attila Kormányos
    Forschungszentrum Jülich GmbH, Helmholtz Institute Erlangen-Nürnberg for Renewable Energy (IEK-11), Egerlandstrasse 3, 91058 Erlangen, Germany
  • Julius Knöppel
    Julius Knöppel
    Forschungszentrum Jülich GmbH, Helmholtz Institute Erlangen-Nürnberg for Renewable Energy (IEK-11), Egerlandstrasse 3, 91058 Erlangen, Germany
    Department of Chemical and Biological Engineering, Friedrich-Alexander-Universität Erlangen-Nürnberg, Egerlandstrasse 3, 91058 Erlangen, Germany
  • Karl J. J. Mayrhofer
    Karl J. J. Mayrhofer
    Forschungszentrum Jülich GmbH, Helmholtz Institute Erlangen-Nürnberg for Renewable Energy (IEK-11), Egerlandstrasse 3, 91058 Erlangen, Germany
    Department of Chemical and Biological Engineering, Friedrich-Alexander-Universität Erlangen-Nürnberg, Egerlandstrasse 3, 91058 Erlangen, Germany
  • Serhiy Cherevko*
    Serhiy Cherevko
    Forschungszentrum Jülich GmbH, Helmholtz Institute Erlangen-Nürnberg for Renewable Energy (IEK-11), Egerlandstrasse 3, 91058 Erlangen, Germany
    *Email: [email protected]
Open PDFSupporting Information (1)

ACS Measurement Science Au

Cite this: ACS Meas. Sci. Au 2021, 1, 2, 74–81
Click to copy citationCitation copied!
https://doi.org/10.1021/acsmeasuresciau.1c00016
Published August 19, 2021

Copyright © 2021 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

High-impact photoelectrode materials for photoelectrochemical (PEC) water splitting are distinguished by synergistically attaining high photoactivity and stability at the same time. With numerous efforts toward optimizing the activity, the bigger challenge of tailoring the durability of photoelectrodes to meet industrially relevant levels remains. In situ photostability measurements hold great promise in understanding stability-related properties. Although different flow systems coupled to light-emitting diodes were introduced recently to measure time-resolved photocorrosion, none of the measurements were performed under realistic light conditions. In this paper, a photoelectrochemical scanning flow cell connected to an inductively coupled plasma mass spectrometer (PEC-ICP-MS) and equipped with a solar simulator, Air Mass 1.5 G filter, and monochromator was developed. The established system is capable of independently assessing basic PEC metrics, such as photopotential, photocurrent, incident photon to current efficiency (IPCE), and band gap in a high-throughput manner as well as the in situ photocorrosion behavior of photoelectrodes under standardized and realistic light conditions by coupling it to an ICP-MS. Polycrystalline platinum and tungsten trioxide (WO3) were used as model systems to demonstrate the operation under dark and light conditions, respectively. Photocorrosion measurements conducted with the present PEC-ICP-MS setup revealed that WO3 starts dissolving at 0.8 VRHE with the dissolution rate rapidly increasing past 1.2 VRHE, coinciding with the onset of the saturation photocurrent. The most detrimental damage to the photoelectrode is caused when subjecting it to a prolonged high potential hold, e.g., at 1.5 VRHE. By using standardized illumination conditions such as Air Mass 1.5 Global under 1 Sun, the obtained dissolution characteristics are translatable to actual devices under realistic light conditions. The gained insights can then be utilized to advance synthesis and design approaches of novel PEC materials with improved photostability.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2021 The Authors. Published by American Chemical Society

1. Introduction

Click to copy section linkSection link copied!

Photoelectrochemical (PEC) water splitting is a promising way of harnessing solar energy and to directly create energy-dense, storable fuels, e.g., hydrogen. (1) The generation of such fuels circumvents the intermittency of renewables and enables coupling of different energy sectors and thus contributing to their decarbonization. Unfortunately, PEC devices have yet to see widespread commercial application in part due to low efficiencies and poor durability of the employed photoactive components. (2) Most research efforts in the past decades have been devoted to improving the intrinsic activity of photoelectrocatalysts, but comparatively few investigations focused on the photostability. Hence, the instability of semiconductors in PEC environments is less understood than the catalytic processes. (3) From a material science perspective, only photoelectrodes accommodating high catalytic activity with long-term stability can be considered as a high-performance material.
Photostability is traditionally assessed by measuring the photocurrent retention over an extended chronoamperometic hold or potential cycling under illumination. Alternatively, morphological and structural changes in the sample can be investigated by post-mortem analysis using scanning electron microscopy (SEM), transmission electron microscopy (TEM), X-ray diffraction (XRD), and X-ray photoelectron spectroscopy (XPS). None of these analyses can provide insights into the degradation processes happening in situ. (3,4) Past reports on the revelation of in situ potential-dependent degradation processes in dark electrocatalysis involved a highly sensitive inductively coupled mass spectrometer (ICP-MS) coupled downstream to a scanning flow cell (SFC). (5,6) On the one hand, ICP-MS allows for the detection of multiple elements simultaneously with a very low detection limit. On the other hand, this system lacks the ability to provide information about the chemical state of dissolved or undissolved species. In more recent studies, flow cells have been equipped with light-emitting diodes (LEDs), channeled perpendicular to the photoelectrode surface, to investigate photocorrosion processes of WO3, (7) ZnO, (8) or BiVO4 (9) in a time-resolved fashion. While these studies provide great detail into degradation processes during PEC operation, the first two works irradiate the sample using one LED at a fixed wavelength. Alternatively, Zhang et al. (9) used a continuous light source filtered to 400–700 nm that recreates white light. In their work, the light is not channeled in into the cell but the sample is irradiated through the cell body, which could make exact adjustments of the irradiation condition at the sample area challenging. Photoelectrodes are commonly characterized under standardized 1 Sun Air Mass 1.5 Global (AM 1.5 G) conditions using a solar simulator. It is, therefore, inconclusive whether observed degradation behaviors using LEDs are also encountered under realistic light conditions. Notably, Sliozberg et al. have developed a scanning droplet cell system coupled to a solar simulator, yet, with clear focus on activity determination. (10) There remains a need for an experimental setup that combines simulated solar irradiation with a flow system, making in situ photocorrosion under standardized light conditions yet to be demonstrated for any phototelectode. Attempts to conduct in situ analysis on the photostability include a quartz crystal microbalance to detect mass losses during PEC operation. (11) Given the complexity of photodecomposition processes, a gravimetric approach is less conclusive of the underlying degradation mechanism, especially in multielement systems with diverse dissolution characteristics.
In this work, we developed a photoelectrochemical scanning flow cell connected to an inductively coupled plasma mass spectrometer (PEC-ICP-MS) equipped with a solar simulator, AM 1.5 G filter, and a monochromator. This system enables the assessment of basic PEC properties of a photoelectrode, such as photopotential, photocurrent, incident photon to current efficiency (IPCE), and band gap, while at the same time allows measuring real-time dissolution behavior under realistic solar irradiation. First, details on the hardware development are outlined together with the design concepts of the photoelectrochemical scanning flow cell (PEC-SFC) that allows for the coupling of a solar simulator to a flow cell. Then, initial system evaluations are conducted under dark electrocatalytic conditions to benchmark the setup against past literature reports for SFC systems. Finally, tungsten trioxide (WO3) is used as a model system to study its PEC performance and in situ photostability. WO3 is acknowledged to be a promising photoanode for solar water splitting due to its band gap being in the visible spectrum, electron transport properties, hole diffusion length, and chemical stability in acidic media. (12) Since maximum theoretical solar-to-hydrogen efficiencies are around 6%, PEC devices made solely of WO3 are not commercially viable. Nonetheless, modified WO3-based PEC devices have great potential to attain efficiencies beyond 10%, which would make such systems considerable for widespread use. (13)

2. Experimental Section

Click to copy section linkSection link copied!

2.1. PEC-ICP-MS Setup

The developed PEC-ICP-MS setup is schematically illustrated in Scheme 1. A photograph of the light components can be found in Figure S1. A 300 W ozone-free Xe lamp (I, Newport, 66984–300XF-R1) with a lens transmission range of 200–2500 nm was used as an illumination source. The emitted light was then passed through a focusing lens (II, Newport, 6199) and an AM 1.5 G filter (III, Newport, 81094) before channeling in the simulated sunlight into a monochromator with an integrated shutter (IV, Newport, Cornerstone, CS130B-1-MC). To ensure maximum light throughput, the distance between the focusing lens and the monochromator was adjusted in a way to channel as much light into the monochromator as possible. The monochromator can be operated in either a “bypass mode” in which the light passes through the device without any alteration or a “monochromatic mode” in which a selected wavelength can be extracted from the full light spectrum. The AM 1.5 G filter was removed when operating the monochromator in the latter mode. Having two modes available makes the setup flexible and modular, rendering the need to manually exchange the monochromator depending on the desired PEC experiment obsolete. The light exiting the monochromator was then fed into a filter wheel (V, Newport, USFW-100) equipped with long-pass filters of different cutoff wavelengths to exclude any higher-order harmonics produced on the diffraction grating in the monochromator. Finally, the collimated light source was converted into a fiber optic source by using a focusing assembly (VI, Newport, 77799) before being fed into a liquid light guide (VII, Newport, 77566) with a core diameter of 3 mm and a transmission range of 340–800 nm. Employing a liquid light guide is crucial especially when operating the setup in “monochromatic mode.” Here, light intensities drop significantly when filtering for just a single wavelength with few nanometers of bandwidth. Low light power can potentially manifest in very small photocurrents in the range of few microamperes, complicating data analysis. Hence, choosing a liquid light guide with a sufficient core diameter ensures high light throughput and superior longevity compared to glass fiber optics due to no risk of fiber breakage. White light was calibrated to 1 Sun (100 mW cm–2) at the PEC-SFC opening at the beginning of each measurement day using a reference solar cell (Newport, 91150V). Additionally, the light intensity of the solar simulator was measured at the end of a measurement day to verify no significant intensity losses over the course of a day. Power densities of the monochromatic light were measured using a photodiode sensor (Thorlabs, S120VC) coupled to a USB power and energy meter interface (Thorlabs, PM100USB).

Scheme 1

Scheme 1. Schematic Illustration of PEC-ICP-MS Setup Employed in This Studya

aThe optical components consist of: (I) 300 W ozone-free Xe lamp, (II) focusing lens, (III) AM 1.5 G filter, (IV) monochromator, (V) filter wheel with different cutoff filters, (VI) focusing assembly, and (VII) liquid light guide.

Basic (photo)electrochemical experiments were controlled with a potentiostat (BioLogic, SP 150) using a Ag/AgCl electrode in 3 M KCl (Metrohm) as the reference and a glassy carbon rod (HTW, SIGRADUR G) as the counter electrode. Investigated samples were contacted through a metal wire to establish good electrical contact. All experiments were conducted in 0.1 M HClO4 prepared by diluting appropriate amounts of concentrated HClO4 (VWR, Suprapur, 70%) with ultrapure water (Merck Millipore, Milli-Q). The electrolyte was constantly purged with Ar in a separate reservoir and was transported through the system within Tygon tubings via a peristaltic pump (Ismatec, Reglo ICC). Measured potentials, EAg/AgCl, were all corrected to the reversible hydrogen electrode (RHE) scale by determining the pH of the electrolyte using a pH meter (Mettler Toledo, FiveEasy Plus) and applying the following equation:
Samples were mounted onto an XYZ stage (Physik Instrumente, M-403), allowing rapid and accurate positioning under the PEC-SFC. The cell was fixed on a force sensor (ME-Messsysteme GmbH, KD45) to control the contact pressure to the sample using a proportional–integral–derivative controller. The contacting force was set to 2 N. All components within the setup were controlled and synchronized through an in-house made LabVIEW software.
In situ photostability measurements using an online ICP-MS were performed with an Nexion 350X instrument (PerkinElmer). Calibration of the ICP-MS was carried out daily for all investigated metals by a four-point calibration curve (0, 0.5, 1, 5 μg L–1) from freshly prepared standard solutions (Merck Centripur). 187Re at a concentration of 10 μg L–1 served as an internal standard to ensure good system performance. The flow rate of the electrolyte into the ICP-MS was 3.55 μL s–1.

2.2. Cell Design

Several cell designs used for SFC applications have been previously reported. (6,14,15) However, due to the necessity to accommodate a liquid light guide with an 11 mm ferrule into an SFC, a new PEC-SFC design was conceptualized for this setup. The V-shaped channels in an SFC allow the electrolyte to continuously flow above a sample. Therefore, it is crucial to optimize the cell design for individual applications to ensure that dissolved species are transported downstream toward the ICP-MS for trace element detection. (16) Flow profiles simulated with COMSOL Multiphysics show that incorporating a third light channel perpendicular to the sample plane shifts the electrolyte velocity maximum further away from the SFC opening compared to a conventional arrangement used for dark electrocatalysis (Figure S2a, b). An insufficient flow velocity at the sample plane may cause issues due to dissolved ions not being flushed away to the outlet. To remedy this problem, the channel angles can be widened up to 90° to decrease the headspace above the cell opening, thus bringing the maximum flow velocity closer to the measurement area (Figure S2c).
Polytetrafluoroethylene (PTFE) was chosen as the cell material for its nontransparency, chemical stability, resistance to UV light, and plasticity. The latter is especially advantageous for creating a hermetic seal between the cell opening and the sample due to the ability of PTFE to deform when pressing the cell onto the sample with a set force. As long as the contacting force is within reasonable ranges (e.g., 1–2 N), the sealing by deformation can be repeated multiple times. Therefore, there is no need to equip the cell opening with an additional silicone O-ring to ensure complete closure, as was practiced with previous designs, (14) which reduces manual labor. Additional grinding of the cell opening was performed to ensure alignment with the sample to account for manufacturing inaccuracies. To further prevent any oxygen from diffusing into the measurement area, a gas pocket around the cell opening was designed, which can be purged with an inert gas. The space between the pocket walls and the cell opening plane was 0.5 mm to prevent any contact of the pocket walls with the sample. This feature could become useful when studying, for instance, hydrogen evolution or CO2 reduction reactions, where oxygen reduction could interfere with the measurement. (17)
As evident from the established PEC-SFC design (Figure 1), a small volume of air will be entrapped between the ferrule tip and the electrolyte flow. This raises the question of whether the flowing electrolyte moves up toward the ferrule rather than being transported out of the PEC-SFC. To quickly evaluate this potential issue, a simplified PEC-SFC was 3D-printed using a transparent polymer and its flow profile was checked visually. Figure S3a shows that the electrolyte only flows from the inlet to the outlet side without migrating toward the liquid light guide. This proves that the newly designed PEC-SFC is capable of operating as a flow cell.

Figure 1

Figure 1. Cross-sectional 3D design of a PEC-SFC. (I) Electrolyte outlet channel, (II) reference electrode channel, (III) inert gas pocket, (IV) liquid light guide channel, (V) electrolyte inlet channel, and (VI) inert gas inlet.

The reference electrode was positioned on the outlet channel side to prevent potential contamination of chlorine due to diffusion from the reference electrode compartments, which would affect the measured stability. (18) The counter electrode was connected to the sample through the inlet side. A three-electrode configuration was established by contacting the sample through the cell opening. To determine the actual illuminated area of the PEC-SFC, a UV-sensitive paper (Astromedia, Solar-Fotopapier) was employed (Figure S3b). The illuminated spot shows homogeneous coverage of the total cell opening, verifying that the entire contacted sample area is also irradiated.

3. Results and Discussion

Click to copy section linkSection link copied!

3.1. Setup Validation under Dark Conditions

With a new cell design concept being established herein, reproducing results from older SFC designs is crucial to verify the viability of the current setup as a flow system. Hence, before benchmarking the setup for PEC applications, it was necessary to check that the system can reliably operate under dark electrocatalysis conditions. Initial system validation was performed by recording cyclic voltammograms (CVs) of a polycrystalline platinum foil (MaTeck, 99.99%) in 0.1 M HClO4. The surface of the foil was ground and polished with a 1 μm diamond suspension. Prior to the measurement, the surface was electrochemically cleaned by sweeping the working electrode potential between 0.05 and 1.4 VRHE with a scan rate of 200 mV s–1 for 30 cycles. Figure S4a shows a CV recorded at 10 mV s–1, indicating all expected oxidation and reduction peaks for Pt in an acidic medium. (19) Model samples such as polycrystalline platinum are susceptible to oxygen traces in the electrolyte, which manifest as a significant cathodic drop in the current below 1 VRHE due to concomitantly occurring oxygen reduction reactions. (20) Notably, the double-layer region between 0.4 and 0.6 VRHE lies close to 0 mA cm–2, indicating minimal diffusion of atmospheric oxygen into the measurement area. The employed air-shielding walls, together with the introduction of Ar around the cell opening, prove to be suitable strategies against gas contaminations entering the measurement area. Successive online dissolution measurements on the same sample (Figure S4b, c) revealed similar Pt dissolution traces to previous reports. (5) Dissolution peaks are slightly wider, originating from the larger PEC-SFC opening area employed here. Reproducing the reported Pt dissolution behavior encourages the suitability of the established PEC-SFC design for the application in online dissolution measurements.

3.2. Output Spectra

Figure 2 shows the corresponding light spectra for both operating modes (bypass and monochromatic) recorded with a concave grating spectrometer (StellarNet, BLACK-Comet) after the light has traversed through a 0.1 M HClO4 electrolyte. The relatively large bandwidth of the monochromatic spectra is a result of setting the slit size to its maximum width and height to achieve a compromise between light throughput and resolution. Using an AM 1.5 G filter reduces light intensities in the UV region below 350 nm and dampens the characteristic Xe peaks between 800 and 1000 nm in order to shape the emission spectrum close to the solar energy distribution.

Figure 2

Figure 2. Output spectrum of PEC-SFC setup operated in (a) monochromatic mode and (b) bypass mode. All spectra were recorded at the PEC-SFC opening after the light has traveled through a 0.1 M HClO4 solution.

3.3. Setup Validation under Illuminated Conditions

Performing general PEC measurements on a photoactive material can aid in estimating the performance and suitability for further focused assessments. (21−23) By using a PEC-SFC, such general PEC measurements can be performed rapidly in a serial manner, presenting a high-throughput platform capable of discovering new materials for solar water splitting applications. The validation was conducted on spray-coated WO3 samples synthesized using a peroxotungstic acid-based route. (24−26) Detailed descriptions of the sample preparation can be found in the Supporting Information.
Measuring the photopotential is the very first step to check whether a semiconductor is light-active. The absorption of photons leads to electron–hole pairs that induce an unbending of the valence and the conduction band of the semiconductor, which manifests in different open circuit potentials (OCP) under dark and light conditions. (23) The semiconducting type can be estimated depending on whether the potential changes in the anodic or cathodic direction. (27) The magnitude of the photopotential also reveals the driving force for the photoelectrochemical reaction (28) and can help in assessing the flat band potential used to estimate the band edge positions in photoelectrodes. (29) The spray-coated WO3 exhibits a drop in OCP when illuminating the sample after a short period in dark conditions (Figure 3a) and equilibrates around 0.48 VRHE. A shift toward cathodic potentials confirms WO3 being a n-type semiconductor.

Figure 3

Figure 3. (a) Photopotential under dark and illuminated state. (b) Photovoltammogram recorded with a scan rate of 10 mV s–1 and a chopping frequency of 0.2 Hz. Inset showing magnified potential region around onset potential. (c) IPCE measurement at 1.23 VRHE. (d) (IPCE hν)0.5 vs hν (gray) with linear fit (red). All measurements were conducted on spray-coated WO3 in Ar-saturated 0.1 M HClO4. (a, b) were measured with the monochromator in bypass mode to access the polychromatic spectrum of the solar simulator calibrated to 1 Sun and (c) was measured using the monochromatic mode.

The actual activity of the sample is most commonly assessed by recording photovoltammograms under chopped irradiation (Figure 3b). Here, it is crucial to utilize standardized conditions, such as 1 Sun AM 1.5 G, to ensure comparability between different studies. WO3 exhibits a broad oxidation peak around 0.4–0.6 VRHE representing the electrochromic redox response of HxWO3/WO3. (30,31) The onset potential for the photocurrent is around 0.5 VRHE, standing in agreement with previous reports (30,32) and the observed photopotential in Figure 3a. The photocurrent slowly increases with an inclining bias as more charge carriers get extracted from the semiconductor into the solution. The photocurrent starts to plateau past 1.2 VRHE indicating the maximum extractable charge carriers is reached. The measured photocurrent, which is the current density difference between a light-on and light-off state, is about 0.7 mA cm–2 and comparable with literature values. (33)
Another crucial diagnostic metric regarding material efficiency is the so-called incident photon to current efficiency (IPCE). This metric represents the semiconductor’s capability of converting the impinging photons into photoelectrochemical reaction products. Catalytic conversions are accounted for by the overall measured photocurrent (Jph(λ)) at a given wavelength, λ. The photon energy is given by the measured power density (Pmono(λ)).
Here, the faradaic efficiency is assumed to be 100% toward the oxygen evolution reaction (OER). The IPCE curve recorded at 1.23 VRHE (Figure 3c) shows a maximum efficiency of 36% at 340 nm and falls to 0% around 480 nm. Similar performances have been observed in other reports investigating WO3 in 1 M HClO4. (34) The decrease in IPCE is a sign of decreased photon absorption by the semiconductor as higher wavelengths cannot excite WO3 to form electron–hole pairs. The corresponding Tauc plot based on an indirect semiconductor ((IPCE hν)0.5) (35) is shown in Figure 3d. Extrapolating the linear region around the inflection point reveals the bandgap value of the semiconductor at the intercept with the x-axis (red fit in Figure 3d). Here, the bandgap energy is around 2.56 eV and is consistent with the indirect bandgap for crystalline WO3 known from literature. (35)
The validation of the developed PEC-SFC setup under PEC conditions demonstrated closely matching performance values for WO3 to those reported in literature where the measurements are conducted using bulk PEC cells. Thus, the current setup can give valid insights into PEC properties of a photoelectrode in an accurate and high-throughput manner.

3.4. In Situ Photocorrosion under Simulated Solar Irradiation

Besides measuring the intrinsic activity, it is equally important to assess the stability of photoelectrodes to better understand material properties and ultimately improve on the durability constrain. Achieving high activity while retaining good stability is one of the greatest challenges when designing novel materials for PEC applications. For this, the developed PEC-ICP-MS setup using a solar simulator serves as an excellent tool to assess these two important metrics simultaneously. By measuring photocorrosion under standardized illumination rather than monochromatic light, as it was previously performed, the determined photostability will be closer to what would be expected in a real-life application scenario.
To study the photocorrosion behavior under realistic operating conditions, a potential protocol consisting of two linear sweep voltammograms (LSV) and chronoamperometric (CA) holds was established. The first of the two LSVs and CA holds was performed in dark and the second in 1 Sun AM 1.5 G conditions, as depicted in Figure 4a. Measured currents (Figure 4b) in the dark correspond to the redox behavior of WO3. Greater performance is recorded under illumination, evident by visibly higher photocurrents. WO3 is well-known to be a stable photoanode in acidic conditions demonstrated by negligible dissolution rates during a dark LSV up to 1.5 VRHE shown in Figure 4c. A small dissolution peak is observed probably due to undercoordinated W sites dissolving at high potentials. Subjecting WO3 to harsher conditions such as a two minute CA hold at 1.5 VRHE in dark results in a noticeable W dissolution peak in dark. Assessing the real-time photocorrosion under illuminated LSVs reveals a W dissolution onset around 0.8 VRHE, which coincides with the steep increase in the photocurrent in this potential region. Previous PEC-ICP-MS setups utilizing a 385 nm LED detected an onset potential of dissolution around 1.0 VRHE for WO3. (7,36) With the emitted polychromatic spectrum from the solar simulator including wavelengths below 385 nm, it is suggested that the additional high energy irradiation leads to an altered dissolution characteristic compared to results using monochromatic light. Once the potential passes 1.2 VRHE (gray dashed line in Figure 4) the dissolution rate experiences a substantial incline indicated by the two red dashed lines in Figure 4c representing linear fits of the dissolution rate before and after 1.2 VRHE. Interestingly, 1.2 VRHE also marks the potential at which the saturation photocurrent starts to onset. This behavior highlights that when nearing the maximum extractable charge carriers for the photooxidation, the additional bias applied will contribute more to W dissolution than to the actual PEC reaction. Such photocorrosion behavior in conjunction with the current response was not observed in previous PEC-ICP-MS studies with WO3. (7,36) Encountered discrepancies in the photocorrosion behavior between 385 nm and AM 1.5 G implies the importance of utilizing realistic irradiation conditions to understand photostability properties for photoelectrodes in real devices. Applying light to a CA hold at 1.5 VRHE has a more detrimental effect on the photoanode, causing a steep increase in dissolution until peaking around 0.2 ng cm–2 s–1. Notably, WO3 exposed to an illuminated, high potential hold displays an elevated W baseline dissolution after returning to a cathodic hold of 0.4 VRHE in the dark. This observation might hint to permanent damage (e.g., pinhole) to the sample after harsh operating conditions. Newly generated imperfections in the photolectrode structure or lattice could promote irreversible material instability. (37,38) The degraded sample could also lose mechanical integrity, causing material detachment by the flowing electrolyte. The measurements were successfully repeated with Figure S5 summarizing obtained W dissolutions for the individual, potentiostatic techniques under dark and light with their respective reproducibility.

Figure 4

Figure 4. Current density and dissolution profile of spray-coated WO3 under dark and 1 Sun AM 1.5 G illuminated condition in 0.1 M HClO4. (a) Potential protocol with a scan rate of 10 mV s–1 for the LSVs. CA holds were performed for two minutes. Upper potential limit for all LSVs and CA holds was 1.5 VRHE. (b) Measured current density. (c) Dissolution rate of W (gray, measured signal; dashed red, linear fit of dissolution rate before and after dissolution rate increase; dashed gray, transition in dissolution rate). Yellow bars indicate the period of illumination.

To translate the amount of dissolved ions into a comparable metric across different studies, the S-number concept has been introduced first for dark OER catalysis. (39) The S-number originally relates the produced oxygen, assuming 100% faradaic efficiency, to the total amount of dissolved species (n(M)). However, it remains disputed whether OER is the sole reaction occurring on the WO3 surface with HClO4 as the electrolyte, as ClO4 can undergo decomposition reactions to form radical species. (40) Therefore, the S-number concept was adapted to account for any oxidized species using the total number of transferred electrons (n(e)). (36)
The S-number(e) based on the second CA hold under illumination was calculated to be 8850 and is in the ballpark of previously reported values for WO3 in HClO4 conducted in our group. (36) The calculated S-number(e) for the dark CA hold equals to about 21 700. The smaller value under light indicates that even after consideration of the total measured charge, WO3 is rendered more unstable by photogenerated charge carriers.
The developed PEC-ICP-MS system proves its viability as a tool to study in situ photocorrosion behaviors under AM 1.5 G conditions. This experimental setup is expected to play a vital role in unfolding photodecomposition phenomena guiding the synthesis and design of more stable photoactive materials. Moreover, flow cell systems are well suited for operation in a high-throughput manner. In fact, scanning droplet cells have been successfully employed for rapid (photo)electrochemical screening of (photo)electrocatalyst libraries. (10,41) We envision a full automation of the presented system to establish a platform that is capable of not only performing full sets of PEC measurements but at the same time assess the in situ photostability of a photoelectrode material library. Such a platform would enable material discovery, which is tailored to search not only for the most active but also for the most stable PEC material.

4. Conclusion

Click to copy section linkSection link copied!

In situ photodecomposition measurements of photoelectrodes are still a rarity in the field of photoelectrochemistry, yet they are crucial to unravel stability properties of promising photoelectrodes. Here, a PEC-ICP-MS setup was developed that is operated under standardized illumination conditions using a solar simulator and an AM 1.5 G filter. The established system was benchmarked for its suitability to function as a scanning flow design with expected observations under dark electrocatalytic conditions being met. Further, it was demonstrated that the PEC-SFC setup is capable of conducting general PEC measurements to assess photopotential, photocurrent, and IPCE of semiconducting materials, e.g., WO3. The obtained performance for spray-coated WO3 meets literature values. Finally, time-resolved PEC-ICP-MS measurements were successfully performed under simulated solar irradiation on WO3, confirming the pronounced dissolution of W under light compared to dark conditions. The dissolution behavior is closely tied to the photocurrent response as the dissolution onset coincides with the steep incline in the photocurrent. Applying additional bias after the saturation photocurrent is reached triggers an increase in W dissolution. Such behavior was not observed in the past when studying the in situ photocorrosion of WO3 at 385 nm, which demonstrates the importance of studying the time-resolved photostability under realistic light conditions.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsmeasuresciau.1c00016.

  • Electrolyte flow profile simulation, WO3 synthesis procedure, photograph of light components, photograph of transparent PEC-SFC with flowing electrolyte, photograph of UV sensitive paper to determine illuminated area, CV and dissolution profile of polycrystalline Pt, reproducibility of WO3 dissolution experiments (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
    • Ken J. Jenewein - Forschungszentrum Jülich GmbH, Helmholtz Institute Erlangen-Nürnberg for Renewable Energy (IEK-11), Egerlandstrasse 3, 91058 Erlangen, GermanyDepartment of Chemical and Biological Engineering, Friedrich-Alexander-Universität Erlangen-Nürnberg, Egerlandstrasse 3, 91058 Erlangen, GermanyOrcidhttps://orcid.org/0000-0001-8979-7252 Email: [email protected]
    • Serhiy Cherevko - Forschungszentrum Jülich GmbH, Helmholtz Institute Erlangen-Nürnberg for Renewable Energy (IEK-11), Egerlandstrasse 3, 91058 Erlangen, GermanyOrcidhttps://orcid.org/0000-0002-7188-4857 Email: [email protected]
  • Authors
    • Attila Kormányos - Forschungszentrum Jülich GmbH, Helmholtz Institute Erlangen-Nürnberg for Renewable Energy (IEK-11), Egerlandstrasse 3, 91058 Erlangen, GermanyOrcidhttps://orcid.org/0000-0002-2145-7419
    • Julius Knöppel - Forschungszentrum Jülich GmbH, Helmholtz Institute Erlangen-Nürnberg for Renewable Energy (IEK-11), Egerlandstrasse 3, 91058 Erlangen, GermanyDepartment of Chemical and Biological Engineering, Friedrich-Alexander-Universität Erlangen-Nürnberg, Egerlandstrasse 3, 91058 Erlangen, GermanyOrcidhttps://orcid.org/0000-0001-8355-3062
    • Karl J. J. Mayrhofer - Forschungszentrum Jülich GmbH, Helmholtz Institute Erlangen-Nürnberg for Renewable Energy (IEK-11), Egerlandstrasse 3, 91058 Erlangen, GermanyDepartment of Chemical and Biological Engineering, Friedrich-Alexander-Universität Erlangen-Nürnberg, Egerlandstrasse 3, 91058 Erlangen, GermanyOrcidhttps://orcid.org/0000-0002-4248-0431
  • Author Contributions

    K.J.J. developed the setup, conducted all measurements, simulated flow profiles, and wrote the manuscript, J.K. provided spray-coated WO3. A.K. and S.C. supervised the work. K.J.J.M., A.K., J.K., and S.C. participated in discussions and provided valuable comments on the manuscript. All authors have given approval to the final version of the manuscript.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

The authors would like to thank Jonas Möller for developing the custom-made LabVIEW software employed for the setup. Furthermore, the authors would like to acknowledge Achim Mannke, Stefan Borlein, and Stefan Fiegl for technical support. We appreciate the help from Renee Timmins with editing the manuscript.

Abbreviations

Click to copy section linkSection link copied!

AM 1.5 G

air mass 1.5 global

CA

chronoamperometric

CV

cyclic voltammogram

ICP-MS

inductively coupled plasma mass spectrometer

IPCE

incident photon to current efficiency

LED

light-emitting diode

LSV

linear sweep voltammogram

OCP

open circuit potential

OER

oxygen evolution reaction

PEC

photoelectrochemical

PEC-ICP-MS

photoelectrochemical scanning flow cell connected to an inductively coupled plasma mass spectrometer

PEC-SFC

photoelectrochemical scanning flow cell

PTFE

polytetrafluoroethylene

RHE

reversible hydrogen electrode

SEM

scanning electron microscopy

SFC

scanning flow cell

TEM

transmission electron microscopy

XPS

X-ray photoelectron spectroscopy

XRD

X-ray diffraction

References

Click to copy section linkSection link copied!

This article references 41 other publications.

  1. 1
    Gratzel, M. Photoelectrochemical cells. Nature 2001, 414 (6861), 33844,  DOI: 10.1038/35104607
  2. 2
    Tilley, S. D. Recent Advances and Emerging Trends in Photo-Electrochemical Solar Energy Conversion. Adv. Energy Mater. 2019, 9 (2), 1802877,  DOI: 10.1002/aenm.201802877
  3. 3
    Eichhorn, J.; Liu, G.; Toma, F. M. Degradation of Semiconductor Electrodes in Photoelectrochemical Devices: Principles and Case Studies. In Integrated Solar Fuel Generators; Royal Society of Chemistry, 2018; Chapter 8, pp 281303. DOI: 10.1039/9781788010313-00281
  4. 4
    Chen, S.; Huang, D.; Xu, P.; Xue, W.; Lei, L.; Cheng, M.; Wang, R.; Liu, X.; Deng, R. Semiconductor-based photocatalysts for photocatalytic and photoelectrochemical water splitting: will we stop with photocorrosion?. J. Mater. Chem. A 2020, 8 (5), 22862322,  DOI: 10.1039/C9TA12799B
  5. 5
    Topalov, A. A.; Cherevko, S.; Zeradjanin, A. R.; Meier, J. C.; Katsounaros, I.; Mayrhofer, K. J. J. Towards a comprehensive understanding of platinum dissolution in acidic media. Chem. Sci. 2014, 5 (2), 631638,  DOI: 10.1039/C3SC52411F
  6. 6
    Klemm, S. O.; Topalov, A. A.; Laska, C. A.; Mayrhofer, K. J. J. Coupling of a high throughput microelectrochemical cell with online multielemental trace analysis by ICP-MS. Electrochem. Commun. 2011, 13 (12), 15331535,  DOI: 10.1016/j.elecom.2011.10.017
  7. 7
    Knöppel, J.; Zhang, S.; Speck, F. D.; Mayrhofer, K. J. J.; Scheu, C.; Cherevko, S. Time-resolved analysis of dissolution phenomena in photoelectrochemistry - A case study of WO3 photocorrosion. Electrochem. Commun. 2018, 96, 5356,  DOI: 10.1016/j.elecom.2018.09.008
  8. 8
    Dworschak, D.; Brunnhofer, C.; Valtiner, M. Photocorrosion of ZnO Single Crystals during Electrochemical Water Splitting. ACS Appl. Mater. Interfaces 2020, 12 (46), 5153051536,  DOI: 10.1021/acsami.0c15508
  9. 9
    Zhang, S.; Rohloff, M.; Kasian, O.; Mingers, A. M.; Mayrhofer, K. J. J.; Fischer, A.; Scheu, C.; Cherevko, S. Dissolution of BiVO4 Photoanodes Revealed by Time-Resolved Measurements under Photoelectrochemical Conditions. J. Phys. Chem. C 2019, 123 (38), 2341023418,  DOI: 10.1021/acs.jpcc.9b07220
  10. 10
    Sliozberg, K.; Schafer, D.; Erichsen, T.; Meyer, R.; Khare, C.; Ludwig, A.; Schuhmann, W. High-throughput screening of thin-film semiconductor material libraries I: system development and case study for Ti-W-O. ChemSusChem 2015, 8 (7), 12708,  DOI: 10.1002/cssc.201402917
  11. 11
    Choi, Y.; Yim, C.; Baek, S.; Choi, M.; Jeon, S.; Yong, K. In situ measurement of photostability of CdSe/CdS/ZnO nanowires photoelectrode for photoelectrochemical water splitting. Sens. Actuators, B 2015, 221, 113119,  DOI: 10.1016/j.snb.2015.06.092
  12. 12
    Wang, Y.; Tian, W.; Chen, C.; Xu, W.; Li, L. Tungsten Trioxide Nanostructures for Photoelectrochemical Water Splitting: Material Engineering and Charge Carrier Dynamic Manipulation. Adv. Funct. Mater. 2019, 29 (23), 1809036,  DOI: 10.1002/adfm.201809036
  13. 13
    Zheng, G.; Wang, J.; Liu, H.; Murugadoss, V.; Zu, G.; Che, H.; Lai, C.; Li, H.; Ding, T.; Gao, Q.; Guo, Z. Tungsten oxide nanostructures and nanocomposites for photoelectrochemical water splitting. Nanoscale 2019, 11 (41), 1896818994,  DOI: 10.1039/C9NR03474A
  14. 14
    Schuppert, A. K.; Topalov, A. A.; Katsounaros, I.; Klemm, S. O.; Mayrhofer, K. J. J. A Scanning Flow Cell System for Fully Automated Screening of Electrocatalyst Materials. J. Electrochem. Soc. 2012, 159 (11), F670F675,  DOI: 10.1149/2.009211jes
  15. 15
    Grote, J. P.; Zeradjanin, A. R.; Cherevko, S.; Mayrhofer, K. J. Coupling of a scanning flow cell with online electrochemical mass spectrometry for screening of reaction selectivity. Rev. Sci. Instrum. 2014, 85 (10), 104101,  DOI: 10.1063/1.4896755
  16. 16
    Shkirskiy, V.; Speck, F. D.; Kulyk, N.; Cherevko, S. On the Time Resolution of Electrochemical Scanning Flow Cell Coupled to Downstream Analysis. J. Electrochem. Soc. 2019, 166 (16), H866H870,  DOI: 10.1149/2.1401915jes
  17. 17
    Benn, E. E.; Gaskey, B.; Erlebacher, J. D. Suppression of Hydrogen Evolution by Oxygen Reduction in Nanoporous Electrocatalysts. J. Am. Chem. Soc. 2017, 139 (10), 36633668,  DOI: 10.1021/jacs.6b10855
  18. 18
    Geiger, S.; Cherevko, S.; Mayrhofer, K. J. J. Dissolution of Platinum in Presence of Chloride Traces. Electrochim. Acta 2015, 179, 2431,  DOI: 10.1016/j.electacta.2015.03.059
  19. 19
    Shinozaki, K.; Zack, J. W.; Richards, R. M.; Pivovar, B. S.; Kocha, S. S. Oxygen Reduction Reaction Measurements on Platinum Electrocatalysts Utilizing Rotating Disk Electrode Technique. J. Electrochem. Soc. 2015, 162 (10), F1144F1158,  DOI: 10.1149/2.1071509jes
  20. 20
    Klemm, S. O.; Karschin, A.; Schuppert, A. K.; Topalov, A. A.; Mingers, A. M.; Katsounaros, I.; Mayrhofer, K. J. J. Time and potential resolved dissolution analysis of rhodium using a microelectrochemical flow cell coupled to an ICP-MS. J. Electroanal. Chem. 2012, 677–680, 5055,  DOI: 10.1016/j.jelechem.2012.05.006
  21. 21
    Chen, Z.; Jaramillo, T. F.; Deutsch, T. G.; Kleiman-Shwarsctein, A.; Forman, A. J.; Gaillard, N.; Garland, R.; Takanabe, K.; Heske, C.; Sunkara, M.; McFarland, E. W.; Domen, K.; Miller, E. L.; Turner, J. A.; Dinh, H. N. Accelerating materials development for photoelectrochemical hydrogen production: Standards for methods, definitions, and reporting protocols. J. Mater. Res. 2010, 25 (1), 316,  DOI: 10.1557/JMR.2010.0020
  22. 22
    Shi, X.; Cai, L.; Ma, M.; Zheng, X.; Park, J. H. General Characterization Methods for Photoelectrochemical Cells for Solar Water Splitting. ChemSusChem 2015, 8 (19), 3192203,  DOI: 10.1002/cssc.201500075
  23. 23
    Bedoya-Lora, F. E.; Holmes-Gentle, I.; Hankin, A. Electrochemical techniques for photoelectrode characterisation. Current Opinion in Green and Sustainable Chemistry 2021, 29, 100463,  DOI: 10.1016/j.cogsc.2021.100463
  24. 24
    Nanba, T.; Takano, S.; Yasui, I.; Kudo, T. Structural study of peroxopolytungstic acid prepared from metallic tungsten and hydrogen peroxide. J. Solid State Chem. 1991, 90 (1), 4753,  DOI: 10.1016/0022-4596(91)90170-M
  25. 25
    Kwong, W. L.; Savvides, N.; Sorrell, C. C. Electrodeposited nanostructured WO3 thin films for photoelectrochemical applications. Electrochim. Acta 2012, 75, 371380,  DOI: 10.1016/j.electacta.2012.05.019
  26. 26
    Vidmar, T.; Topič, M.; Dzik, P.; Opara Krašovec, U. Inkjet printing of sol-gel derived tungsten oxide inks. Sol. Energy Mater. Sol. Cells 2014, 125, 8795,  DOI: 10.1016/j.solmat.2014.02.023
  27. 27
    Krishnan, R. Fundamentals of Semiconductor Electrochemistry and Photoelectrochemistry. In Encyclopedia of Electrochemistry; Bard, A.J., Ed.; Wiley, 2007.
  28. 28
    White, J. R.; Fan, F. R. F.; Bard, A. J. Semiconductor Electrodes: LVI. Principles of Multijunction Electrodes and Photoelectrosynthesis at Texas Instruments’ p/n-Si Solar Arrays. J. Electrochem. Soc. 1985, 132 (3), 544550,  DOI: 10.1149/1.2113884
  29. 29
    Hankin, A.; Bedoya-Lora, F. E.; Alexander, J. C.; Regoutz, A.; Kelsall, G. H. Flat band potential determination: avoiding the pitfalls. J. Mater. Chem. A 2019, 7 (45), 2616226176,  DOI: 10.1039/C9TA09569A
  30. 30
    Amano, F.; Tian, M.; Wu, G.; Ohtani, B.; Chen, A. Facile preparation of platelike tungsten oxide thin film electrodes with high photoelectrode activity. ACS Appl. Mater. Interfaces 2011, 3 (10), 404752,  DOI: 10.1021/am200897n
  31. 31
    Cai, M.; Fan, P.; Long, J.; Han, J.; Lin, Y.; Zhang, H.; Zhong, M. Large-Scale Tunable 3D Self-Supporting WO3Micro-Nano Architectures as Direct Photoanodes for Efficient Photoelectrochemical Water Splitting. ACS Appl. Mater. Interfaces 2017, 9 (21), 1785617864,  DOI: 10.1021/acsami.7b02386
  32. 32
    Santato, C.; Ulmann, M.; Augustynski, J. Photoelectrochemical Properties of Nanostructured Tungsten Trioxide Films. J. Phys. Chem. B 2001, 105 (5), 936940,  DOI: 10.1021/jp002232q
  33. 33
    Monllor-Satoca, D.; Borja, L.; Rodes, A.; Gomez, R.; Salvador, P. Photoelectrochemical behavior of nanostructured WO3 thin-film electrodes: The oxidation of formic acid. ChemPhysChem 2006, 7 (12), 254051,  DOI: 10.1002/cphc.200600379
  34. 34
    Yagi, M.; Maruyama, S.; Sone, K.; Nagai, K.; Norimatsu, T. Preparation and photoelectrocatalytic activity of a nano-structured WO3 platelet film. J. Solid State Chem. 2008, 181 (1), 175182,  DOI: 10.1016/j.jssc.2007.11.018
  35. 35
    Butler, M. A. Photoelectrolysis and physical properties of the semiconducting electrode WO2. J. Appl. Phys. 1977, 48 (5), 19141920,  DOI: 10.1063/1.323948
  36. 36
    Knöppel, J.; Kormányos, A.; Mayerhöfer, B.; Hofer, A.; Bierling, M.; Bachmann, J.; Thiele, S.; Cherevko, S. Photocorrosion of WO3 Photoanodes in Different Electrolytes. ACS Phys. Chem. Au 2021,  DOI: 10.1021/acsphyschemau.1c00004
  37. 37
    Samu, G. F.; Janaky, C. Photocorrosion at Irradiated Perovskite/Electrolyte Interfaces. J. Am. Chem. Soc. 2020, 142 (52), 2159521614,  DOI: 10.1021/jacs.0c10348
  38. 38
    Nandjou, F.; Haussener, S. Degradation in photoelectrochemical devices: review with an illustrative case study. J. Phys. D: Appl. Phys. 2017, 50 (12), 124002,  DOI: 10.1088/1361-6463/aa5b11
  39. 39
    Geiger, S.; Kasian, O.; Ledendecker, M.; Pizzutilo, E.; Mingers, A. M.; Fu, W. T.; Diaz-Morales, O.; Li, Z.; Oellers, T.; Fruchter, L.; Ludwig, A.; Mayrhofer, K. J. J.; Koper, M. T. M.; Cherevko, S. The stability number as a metric for electrocatalyst stability benchmarking. Nature Catalysis 2018, 1 (7), 508515,  DOI: 10.1038/s41929-018-0085-6
  40. 40
    Mi, Q.; Zhanaidarova, A.; Brunschwig, B. S.; Gray, H. B.; Lewis, N. S. A quantitative assessment of the competition between water and anion oxidation at WO3 photoanodes in acidic aqueous electrolytes. Energy Environ. Sci. 2012, 5 (2), 5694,  DOI: 10.1039/c2ee02929d
  41. 41
    Gregoire, J. M.; Xiang, C.; Liu, X.; Marcin, M.; Jin, J. Scanning droplet cell for high throughput electrochemical and photoelectrochemical measurements. Rev. Sci. Instrum. 2013, 84 (2), 024102,  DOI: 10.1063/1.4790419

Cited By

Click to copy section linkSection link copied!

This article is cited by 15 publications.

  1. Ina Reichmann, Vicent Lloret, Konrad Ehelebe, Pascal Lauf, Ken Jenewein, Karl J. J. Mayrhofer, Serhiy Cherevko. Scanning Gas Diffusion Electrode Setup for Real-Time Analysis of Catalyst Layers. ACS Measurement Science Au 2024, Article ASAP.
  2. Krista M. Kulesa, Erin A. Hirtzel, Vinh T. Nguyen, Dallas P. Freitas, Madison E. Edwards, Xin Yan, Lane A. Baker. Interfacing High-Throughput Electrosynthesis and Mass Spectrometric Analysis of Azines. Analytical Chemistry 2024, 96 (21) , 8249-8253. https://doi.org/10.1021/acs.analchem.4c01110
  3. Victoria Benavente Llorente, Ken J. Jenewein, Markus Bierling, Andreas Körner, Andreas Hutzler, Attila Kormányos, Serhiy Cherevko. Photocorrosion of Hematite Photoanodes in Neutral and Alkaline Electrolytes. The Journal of Physical Chemistry C 2023, 127 (39) , 19687-19697. https://doi.org/10.1021/acs.jpcc.3c02969
  4. Svenja Speldrich, Michael Wark, Gunther Wittstock. Metal Oxide Protection Layers for Enhanced Stability and Activity of WO3 Photoanodes in Alkaline Media. ACS Applied Energy Materials 2023, 6 (18) , 9602-9614. https://doi.org/10.1021/acsaem.3c01642
  5. Serhiy Cherevko. Stabilization of non-noble metal electrocatalysts for acidic oxygen evolution reaction. Current Opinion in Electrochemistry 2023, 38 , 101213. https://doi.org/10.1016/j.coelec.2023.101213
  6. Fredrik Forr, Erlind Mysliu, Jan Halvor Nordlien, Andreas Erbe. The effect of bird droppings on the corrosion of steel and aluminum used in offshore applications. Materials and Corrosion 2023, 74 (3) , 464-481. https://doi.org/10.1002/maco.202213533
  7. Ken J. Jenewein, Yuanxing Wang, Tianying Liu, Tara McDonald, Matej Zlatar, Nadiia Kulyk, Victoria Benavente Llorente, Attila Kormányos, Dunwei Wang, Serhiy Cherevko. Electrolyte Engineering Stabilizes Photoanodes Decorated with Molecular Catalysts. ChemSusChem 2023, 11 https://doi.org/10.1002/cssc.202202319
  8. Sang Youn Chae, Noyoung Yoon, Oh shim Joo, Eun Duck Park. Monitoring Transformations of Catalytic Active States in Photocathodes Based on MoS x Layers on CuInS 2 Using In Operando Raman Spectroscopy. Angewandte Chemie 2023, 135 (7) https://doi.org/10.1002/ange.202215227
  9. Sang Youn Chae, Noyoung Yoon, Oh shim Joo, Eun Duck Park. Monitoring Transformations of Catalytic Active States in Photocathodes Based on MoS x Layers on CuInS 2 Using In Operando Raman Spectroscopy. Angewandte Chemie International Edition 2023, 62 (7) https://doi.org/10.1002/anie.202215227
  10. Ken J. Jenewein, Julius Knöppel, André Hofer, Attila Kormányos, Britta Mayerhöfer, Florian D. Speck, Markus Bierling, Simon Thiele, Julien Bachmann, Serhiy Cherevko. Dissolution of WO 3 modified with IrO x overlayers during photoelectrochemical water splitting. SusMat 2023, 3 (1) , 128-136. https://doi.org/10.1002/sus2.107
  11. Ken J. Jenewein, Sigurd Thienhaus, Attila Kormányos, Alfred Ludwig, Serhiy Cherevko. High-throughput exploration of activity and stability for identifying photoelectrochemical water splitting materials. Chemical Science 2022, 13 (46) , 13774-13781. https://doi.org/10.1039/D2SC05115J
  12. Ken J. Jenewein, Gun D. Akkoc, Attila Kormányos, Serhiy Cherevko. Automated high-throughput activity and stability screening of electrocatalysts. Chem Catalysis 2022, 2 (10) , 2778-2794. https://doi.org/10.1016/j.checat.2022.09.019
  13. Håvard Wilson, Arnoul Van Rooij, Ken Jenewein, Julius Knöppel, Attila Kormányos, Serhiy Cherevko, Svein Sunde, Andreas Erbe. Photocorrosion of n‐ and p‐Type Semiconducting Oxide‐Covered Metals: Case Studies of Anodized Titanium and Copper. physica status solidi (a) 2022, 219 (11) , 2100852. https://doi.org/10.1002/pssa.202100852
  14. A. Kormányos, K.J. Jenewein, S. Cherevko. High-throughput workflows in the service of (photo)electrocatalysis research. Trends in Chemistry 2022, 4 (6) , 475-478. https://doi.org/10.1016/j.trechm.2022.01.004
  15. Guillaume Delaittre, Johanna Heine, Alexander Hinz, Ullrich Jahn, Hajo Kries, Björn Meermann, Carl Christoph Tzschucke, Markus Zegke. Notizen aus der Chemie. Nachrichten aus der Chemie 2021, 69 (12) , 54-57. https://doi.org/10.1002/nadc.20214121060
Open PDF

ACS Measurement Science Au

Cite this: ACS Meas. Sci. Au 2021, 1, 2, 74–81
Click to copy citationCitation copied!
https://doi.org/10.1021/acsmeasuresciau.1c00016
Published August 19, 2021

Copyright © 2021 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

2592

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Scheme 1

    Scheme 1. Schematic Illustration of PEC-ICP-MS Setup Employed in This Studya

    aThe optical components consist of: (I) 300 W ozone-free Xe lamp, (II) focusing lens, (III) AM 1.5 G filter, (IV) monochromator, (V) filter wheel with different cutoff filters, (VI) focusing assembly, and (VII) liquid light guide.

    Figure 1

    Figure 1. Cross-sectional 3D design of a PEC-SFC. (I) Electrolyte outlet channel, (II) reference electrode channel, (III) inert gas pocket, (IV) liquid light guide channel, (V) electrolyte inlet channel, and (VI) inert gas inlet.

    Figure 2

    Figure 2. Output spectrum of PEC-SFC setup operated in (a) monochromatic mode and (b) bypass mode. All spectra were recorded at the PEC-SFC opening after the light has traveled through a 0.1 M HClO4 solution.

    Figure 3

    Figure 3. (a) Photopotential under dark and illuminated state. (b) Photovoltammogram recorded with a scan rate of 10 mV s–1 and a chopping frequency of 0.2 Hz. Inset showing magnified potential region around onset potential. (c) IPCE measurement at 1.23 VRHE. (d) (IPCE hν)0.5 vs hν (gray) with linear fit (red). All measurements were conducted on spray-coated WO3 in Ar-saturated 0.1 M HClO4. (a, b) were measured with the monochromator in bypass mode to access the polychromatic spectrum of the solar simulator calibrated to 1 Sun and (c) was measured using the monochromatic mode.

    Figure 4

    Figure 4. Current density and dissolution profile of spray-coated WO3 under dark and 1 Sun AM 1.5 G illuminated condition in 0.1 M HClO4. (a) Potential protocol with a scan rate of 10 mV s–1 for the LSVs. CA holds were performed for two minutes. Upper potential limit for all LSVs and CA holds was 1.5 VRHE. (b) Measured current density. (c) Dissolution rate of W (gray, measured signal; dashed red, linear fit of dissolution rate before and after dissolution rate increase; dashed gray, transition in dissolution rate). Yellow bars indicate the period of illumination.

  • References


    This article references 41 other publications.

    1. 1
      Gratzel, M. Photoelectrochemical cells. Nature 2001, 414 (6861), 33844,  DOI: 10.1038/35104607
    2. 2
      Tilley, S. D. Recent Advances and Emerging Trends in Photo-Electrochemical Solar Energy Conversion. Adv. Energy Mater. 2019, 9 (2), 1802877,  DOI: 10.1002/aenm.201802877
    3. 3
      Eichhorn, J.; Liu, G.; Toma, F. M. Degradation of Semiconductor Electrodes in Photoelectrochemical Devices: Principles and Case Studies. In Integrated Solar Fuel Generators; Royal Society of Chemistry, 2018; Chapter 8, pp 281303. DOI: 10.1039/9781788010313-00281
    4. 4
      Chen, S.; Huang, D.; Xu, P.; Xue, W.; Lei, L.; Cheng, M.; Wang, R.; Liu, X.; Deng, R. Semiconductor-based photocatalysts for photocatalytic and photoelectrochemical water splitting: will we stop with photocorrosion?. J. Mater. Chem. A 2020, 8 (5), 22862322,  DOI: 10.1039/C9TA12799B
    5. 5
      Topalov, A. A.; Cherevko, S.; Zeradjanin, A. R.; Meier, J. C.; Katsounaros, I.; Mayrhofer, K. J. J. Towards a comprehensive understanding of platinum dissolution in acidic media. Chem. Sci. 2014, 5 (2), 631638,  DOI: 10.1039/C3SC52411F
    6. 6
      Klemm, S. O.; Topalov, A. A.; Laska, C. A.; Mayrhofer, K. J. J. Coupling of a high throughput microelectrochemical cell with online multielemental trace analysis by ICP-MS. Electrochem. Commun. 2011, 13 (12), 15331535,  DOI: 10.1016/j.elecom.2011.10.017
    7. 7
      Knöppel, J.; Zhang, S.; Speck, F. D.; Mayrhofer, K. J. J.; Scheu, C.; Cherevko, S. Time-resolved analysis of dissolution phenomena in photoelectrochemistry - A case study of WO3 photocorrosion. Electrochem. Commun. 2018, 96, 5356,  DOI: 10.1016/j.elecom.2018.09.008
    8. 8
      Dworschak, D.; Brunnhofer, C.; Valtiner, M. Photocorrosion of ZnO Single Crystals during Electrochemical Water Splitting. ACS Appl. Mater. Interfaces 2020, 12 (46), 5153051536,  DOI: 10.1021/acsami.0c15508
    9. 9
      Zhang, S.; Rohloff, M.; Kasian, O.; Mingers, A. M.; Mayrhofer, K. J. J.; Fischer, A.; Scheu, C.; Cherevko, S. Dissolution of BiVO4 Photoanodes Revealed by Time-Resolved Measurements under Photoelectrochemical Conditions. J. Phys. Chem. C 2019, 123 (38), 2341023418,  DOI: 10.1021/acs.jpcc.9b07220
    10. 10
      Sliozberg, K.; Schafer, D.; Erichsen, T.; Meyer, R.; Khare, C.; Ludwig, A.; Schuhmann, W. High-throughput screening of thin-film semiconductor material libraries I: system development and case study for Ti-W-O. ChemSusChem 2015, 8 (7), 12708,  DOI: 10.1002/cssc.201402917
    11. 11
      Choi, Y.; Yim, C.; Baek, S.; Choi, M.; Jeon, S.; Yong, K. In situ measurement of photostability of CdSe/CdS/ZnO nanowires photoelectrode for photoelectrochemical water splitting. Sens. Actuators, B 2015, 221, 113119,  DOI: 10.1016/j.snb.2015.06.092
    12. 12
      Wang, Y.; Tian, W.; Chen, C.; Xu, W.; Li, L. Tungsten Trioxide Nanostructures for Photoelectrochemical Water Splitting: Material Engineering and Charge Carrier Dynamic Manipulation. Adv. Funct. Mater. 2019, 29 (23), 1809036,  DOI: 10.1002/adfm.201809036
    13. 13
      Zheng, G.; Wang, J.; Liu, H.; Murugadoss, V.; Zu, G.; Che, H.; Lai, C.; Li, H.; Ding, T.; Gao, Q.; Guo, Z. Tungsten oxide nanostructures and nanocomposites for photoelectrochemical water splitting. Nanoscale 2019, 11 (41), 1896818994,  DOI: 10.1039/C9NR03474A
    14. 14
      Schuppert, A. K.; Topalov, A. A.; Katsounaros, I.; Klemm, S. O.; Mayrhofer, K. J. J. A Scanning Flow Cell System for Fully Automated Screening of Electrocatalyst Materials. J. Electrochem. Soc. 2012, 159 (11), F670F675,  DOI: 10.1149/2.009211jes
    15. 15
      Grote, J. P.; Zeradjanin, A. R.; Cherevko, S.; Mayrhofer, K. J. Coupling of a scanning flow cell with online electrochemical mass spectrometry for screening of reaction selectivity. Rev. Sci. Instrum. 2014, 85 (10), 104101,  DOI: 10.1063/1.4896755
    16. 16
      Shkirskiy, V.; Speck, F. D.; Kulyk, N.; Cherevko, S. On the Time Resolution of Electrochemical Scanning Flow Cell Coupled to Downstream Analysis. J. Electrochem. Soc. 2019, 166 (16), H866H870,  DOI: 10.1149/2.1401915jes
    17. 17
      Benn, E. E.; Gaskey, B.; Erlebacher, J. D. Suppression of Hydrogen Evolution by Oxygen Reduction in Nanoporous Electrocatalysts. J. Am. Chem. Soc. 2017, 139 (10), 36633668,  DOI: 10.1021/jacs.6b10855
    18. 18
      Geiger, S.; Cherevko, S.; Mayrhofer, K. J. J. Dissolution of Platinum in Presence of Chloride Traces. Electrochim. Acta 2015, 179, 2431,  DOI: 10.1016/j.electacta.2015.03.059
    19. 19
      Shinozaki, K.; Zack, J. W.; Richards, R. M.; Pivovar, B. S.; Kocha, S. S. Oxygen Reduction Reaction Measurements on Platinum Electrocatalysts Utilizing Rotating Disk Electrode Technique. J. Electrochem. Soc. 2015, 162 (10), F1144F1158,  DOI: 10.1149/2.1071509jes
    20. 20
      Klemm, S. O.; Karschin, A.; Schuppert, A. K.; Topalov, A. A.; Mingers, A. M.; Katsounaros, I.; Mayrhofer, K. J. J. Time and potential resolved dissolution analysis of rhodium using a microelectrochemical flow cell coupled to an ICP-MS. J. Electroanal. Chem. 2012, 677–680, 5055,  DOI: 10.1016/j.jelechem.2012.05.006
    21. 21
      Chen, Z.; Jaramillo, T. F.; Deutsch, T. G.; Kleiman-Shwarsctein, A.; Forman, A. J.; Gaillard, N.; Garland, R.; Takanabe, K.; Heske, C.; Sunkara, M.; McFarland, E. W.; Domen, K.; Miller, E. L.; Turner, J. A.; Dinh, H. N. Accelerating materials development for photoelectrochemical hydrogen production: Standards for methods, definitions, and reporting protocols. J. Mater. Res. 2010, 25 (1), 316,  DOI: 10.1557/JMR.2010.0020
    22. 22
      Shi, X.; Cai, L.; Ma, M.; Zheng, X.; Park, J. H. General Characterization Methods for Photoelectrochemical Cells for Solar Water Splitting. ChemSusChem 2015, 8 (19), 3192203,  DOI: 10.1002/cssc.201500075
    23. 23
      Bedoya-Lora, F. E.; Holmes-Gentle, I.; Hankin, A. Electrochemical techniques for photoelectrode characterisation. Current Opinion in Green and Sustainable Chemistry 2021, 29, 100463,  DOI: 10.1016/j.cogsc.2021.100463
    24. 24
      Nanba, T.; Takano, S.; Yasui, I.; Kudo, T. Structural study of peroxopolytungstic acid prepared from metallic tungsten and hydrogen peroxide. J. Solid State Chem. 1991, 90 (1), 4753,  DOI: 10.1016/0022-4596(91)90170-M
    25. 25
      Kwong, W. L.; Savvides, N.; Sorrell, C. C. Electrodeposited nanostructured WO3 thin films for photoelectrochemical applications. Electrochim. Acta 2012, 75, 371380,  DOI: 10.1016/j.electacta.2012.05.019
    26. 26
      Vidmar, T.; Topič, M.; Dzik, P.; Opara Krašovec, U. Inkjet printing of sol-gel derived tungsten oxide inks. Sol. Energy Mater. Sol. Cells 2014, 125, 8795,  DOI: 10.1016/j.solmat.2014.02.023
    27. 27
      Krishnan, R. Fundamentals of Semiconductor Electrochemistry and Photoelectrochemistry. In Encyclopedia of Electrochemistry; Bard, A.J., Ed.; Wiley, 2007.
    28. 28
      White, J. R.; Fan, F. R. F.; Bard, A. J. Semiconductor Electrodes: LVI. Principles of Multijunction Electrodes and Photoelectrosynthesis at Texas Instruments’ p/n-Si Solar Arrays. J. Electrochem. Soc. 1985, 132 (3), 544550,  DOI: 10.1149/1.2113884
    29. 29
      Hankin, A.; Bedoya-Lora, F. E.; Alexander, J. C.; Regoutz, A.; Kelsall, G. H. Flat band potential determination: avoiding the pitfalls. J. Mater. Chem. A 2019, 7 (45), 2616226176,  DOI: 10.1039/C9TA09569A
    30. 30
      Amano, F.; Tian, M.; Wu, G.; Ohtani, B.; Chen, A. Facile preparation of platelike tungsten oxide thin film electrodes with high photoelectrode activity. ACS Appl. Mater. Interfaces 2011, 3 (10), 404752,  DOI: 10.1021/am200897n
    31. 31
      Cai, M.; Fan, P.; Long, J.; Han, J.; Lin, Y.; Zhang, H.; Zhong, M. Large-Scale Tunable 3D Self-Supporting WO3Micro-Nano Architectures as Direct Photoanodes for Efficient Photoelectrochemical Water Splitting. ACS Appl. Mater. Interfaces 2017, 9 (21), 1785617864,  DOI: 10.1021/acsami.7b02386
    32. 32
      Santato, C.; Ulmann, M.; Augustynski, J. Photoelectrochemical Properties of Nanostructured Tungsten Trioxide Films. J. Phys. Chem. B 2001, 105 (5), 936940,  DOI: 10.1021/jp002232q
    33. 33
      Monllor-Satoca, D.; Borja, L.; Rodes, A.; Gomez, R.; Salvador, P. Photoelectrochemical behavior of nanostructured WO3 thin-film electrodes: The oxidation of formic acid. ChemPhysChem 2006, 7 (12), 254051,  DOI: 10.1002/cphc.200600379
    34. 34
      Yagi, M.; Maruyama, S.; Sone, K.; Nagai, K.; Norimatsu, T. Preparation and photoelectrocatalytic activity of a nano-structured WO3 platelet film. J. Solid State Chem. 2008, 181 (1), 175182,  DOI: 10.1016/j.jssc.2007.11.018
    35. 35
      Butler, M. A. Photoelectrolysis and physical properties of the semiconducting electrode WO2. J. Appl. Phys. 1977, 48 (5), 19141920,  DOI: 10.1063/1.323948
    36. 36
      Knöppel, J.; Kormányos, A.; Mayerhöfer, B.; Hofer, A.; Bierling, M.; Bachmann, J.; Thiele, S.; Cherevko, S. Photocorrosion of WO3 Photoanodes in Different Electrolytes. ACS Phys. Chem. Au 2021,  DOI: 10.1021/acsphyschemau.1c00004
    37. 37
      Samu, G. F.; Janaky, C. Photocorrosion at Irradiated Perovskite/Electrolyte Interfaces. J. Am. Chem. Soc. 2020, 142 (52), 2159521614,  DOI: 10.1021/jacs.0c10348
    38. 38
      Nandjou, F.; Haussener, S. Degradation in photoelectrochemical devices: review with an illustrative case study. J. Phys. D: Appl. Phys. 2017, 50 (12), 124002,  DOI: 10.1088/1361-6463/aa5b11
    39. 39
      Geiger, S.; Kasian, O.; Ledendecker, M.; Pizzutilo, E.; Mingers, A. M.; Fu, W. T.; Diaz-Morales, O.; Li, Z.; Oellers, T.; Fruchter, L.; Ludwig, A.; Mayrhofer, K. J. J.; Koper, M. T. M.; Cherevko, S. The stability number as a metric for electrocatalyst stability benchmarking. Nature Catalysis 2018, 1 (7), 508515,  DOI: 10.1038/s41929-018-0085-6
    40. 40
      Mi, Q.; Zhanaidarova, A.; Brunschwig, B. S.; Gray, H. B.; Lewis, N. S. A quantitative assessment of the competition between water and anion oxidation at WO3 photoanodes in acidic aqueous electrolytes. Energy Environ. Sci. 2012, 5 (2), 5694,  DOI: 10.1039/c2ee02929d
    41. 41
      Gregoire, J. M.; Xiang, C.; Liu, X.; Marcin, M.; Jin, J. Scanning droplet cell for high throughput electrochemical and photoelectrochemical measurements. Rev. Sci. Instrum. 2013, 84 (2), 024102,  DOI: 10.1063/1.4790419
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsmeasuresciau.1c00016.

    • Electrolyte flow profile simulation, WO3 synthesis procedure, photograph of light components, photograph of transparent PEC-SFC with flowing electrolyte, photograph of UV sensitive paper to determine illuminated area, CV and dissolution profile of polycrystalline Pt, reproducibility of WO3 dissolution experiments (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.