Hematite Photoanodes for Solar Water Splitting: A Detailed

Aug 5, 2019 - Photoelectrochemical (PEC) water splitting has gained momentum over .... For linear scanning voltammetry the voltage was swept at 20 mV/...
0 downloads 0 Views 762KB Size
Subscriber access provided by Mount Allison University | Libraries and Archives

Article

Hematite Photoanodes for Solar Water Splitting: a Detailed Spectroelectrochemical Analysis on the pH-Dependent Performance Yongpeng Liu, Florian Le Formal, Florent Boudoire, and Nestor Guijarro ACS Appl. Energy Mater., Just Accepted Manuscript • DOI: 10.1021/acsaem.9b01261 • Publication Date (Web): 05 Aug 2019 Downloaded from pubs.acs.org on August 5, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

Hematite Photoanodes for Solar Water Splitting: a Detailed Spectroelectrochemical Analysis on the pH-Dependent Performance Yongpeng Liu, Florian Le Formal, Florent Boudoire and Néstor Guijarro*

Laboratory for Molecular Engineering of Optoelectronic Nanomaterials (LIMNO), École Polytechnique Fédérale de Lausanne (EPFL), Station 6, CH-1015 Lausanne, Switzerland

ABSTRACT

The effect of the electrolyte pH on the performance of metal oxide photoanodes for solar water oxidation has not been fully resolved and contrasting views have been presented in recent reports. Herein, a comprehensive set of spectroelectrochemical techniques (impedance spectroscopy, intensity-modulated photocurrent/photovoltage spectroscopy

ACS Paragon Plus Environment

1

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 40

(IMPS/IMVS) and operando UV-Vis spectroscopy) are deployed to clearly uncover the role of pH on the performance of hematite photoanodes. Our results reveal that, despite the presence of high-valent iron-oxo active sites over a wide pH range (7-13.6), the observed performance improvement with increasing pH is mainly driven by the reduction of surface accumulated charges, in the form of reactive intermediate species, that alleviates Fermi level pinning (FLP). Interestingly, IMPS data provides compelling evidence that the mitigation of FLP originates from changes in the reaction mechanism which boost the rate of charge transfer reducing, in turn, the surface charging. Additionally, we present a phenomenological analysis of the IMVS response which brings to light the additional impact of the electrolyte pH on the surface-related recombination dynamics. Our work identifies the pH-dependent kinetics of water oxidation as the key step governing the performance, defining not only the efficiency of charge transfer across the interface but also the degree of FLP that determines both the photocurrent magnitude and onset potential.

ACS Paragon Plus Environment

2

Page 3 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

KEYWORDS -Fe2O3, water splitting, surface states, photoelectrochemical impedance spectroscopy, IMPS, IMVS

INTRODUCTION Photoelectrochemical (PEC) water splitting has gained momentum over the last years as a potential technology capable of disrupting society’s strong reliance on fossil fuels while providing sustainable synthetic pathways for the production of feedstock chemicals, such as hydrogen.1–6 A leading design of PEC technology, the so-called tandem cell configuration wherein a photocathode and a photoanode are assembled to both harvest a large portion of the available solar illumination and drive the unassisted solar water splitting reactions, has delivered solar-to-hydrogen (STH) conversion efficiencies of ca. 7.7%.7 Despite being encouraging, this performance remains impractical for commercial application. The major performance-bottleneck of these devices lies in the photoanode, where due to material constraints the semiconductors typically employed deliver low photovoltage and poor photocurrent when performing the demanding water oxidation reaction.8

ACS Paragon Plus Environment

3

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 40

Hematite (-Fe2O3) has been extensively explored as a photoanode for water oxidation leveraging its relatively narrow band gap energy (1.9-2.2 eV), robustness, and low-cost fabrication using solution-processable routes.9,10 A wide range of bulk and surface engineering

treatments

involving

doping,11,12

nanostructuring,13,14

surface

reconstruction15 and electrocatalyst deposition,16,17 among others, have shown great promise at overcoming the relatively poor performance of hematite photoanodes either by improving bulk electronic features or by conditioning the semiconductor-liquid junction (SCLJ) characteristics (reducing surface states, adjusting energy bands, enhancing catalytic activities). Alternatively, the electrolyte composition also plays an active role on the PEC response, not only by modulating ion transport but also by directly affecting the interfacial characteristics of the photoanode.18,19 For instance, it has been reported that changes of the local pH can alter the relative alignment between the semiconductor energy bands and the water redox energy levels,19 modify the reaction mechanism20,21 or even change the surface chemistry.22 All these aspects affect the crucial kinetic competition between the turnover of the water oxidation reaction and the interfacial recombination of photogenerated charge carriers that governs the PEC behavior.

ACS Paragon Plus Environment

4

Page 5 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

A deeper understanding of the interplay of these effects was first reported by Klahr et

al. using PEC impedance spectroscopy (PEIS), which revealed that the electronic properties of the SCLJ are clearly linked to the electrolyte pH.23 Despite the limitations of the study, restricted to two conditions (basic and neutral electrolyte), a noticeable mitigation of Fermi level pinning (FLP) and a concurrent decrease of the accumulated surface charges were considered behind the improved response at higher pH. In fact, later studies by operando infrared and UV-Vis spectroscopy confirmed that these accumulated charges correspond to reactive intermediate species, high valent iron species, such as –FeIV=O, which are actively involved in the water oxidation reaction.24,25 Alternatively, Iandolo et al. combined theoretical calculations with PEIS studies to conclude that the FLP that deteriorates the performance with decreasing pH resulted from the accumulation of charges at surface trap states generated with the protonation of the metal oxide from where water oxidation reaction cannot proceed.22 Aside from this purely electronic view, Zhao and co-workers suggested that pH could affect the mechanism of the oxygen evolution reaction (OER) occurring at the SCLJ in hematite photoanodes.20 However, although both the phenomena of surface charge accumulation and reaction

ACS Paragon Plus Environment

5

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 40

kinetics have been argued to change with pH, to the best of our knowledge no study has systematically explored these competing factors to construct a clear view on how the electrolyte pH affects the performance, and more importantly, which factor is the dominant. Indeed, insight into this aspect could reframe strategies for photoanode optimization. Therefore, we present herein a set of spectroelectrochemical techniques including frequency-domain electrochemical measurements, namely PEIS, intensitymodulated photocurrent/photovoltage spectroscopy (IMPS/IMVS) and operando UV-Vis spectroscopy to investigate -Fe2O3 photoanodes over a wide range of pH values. Results clearly evidence an inverse correlation between the steady-state surface accumulated charge and the rate of water oxidation reaction, as well as a progressive suppression of FLP with increasing pH. EXPERIMENTAL SECTION Preparation of Hematite Thin Films. Fluorine doped tin oxide (FTO) coated glass substrates (Solaronix TCO10-10, 8 Ω/sq.) were positioned vertically in an aqueous solution containing 0.15 M iron chloride hexahydrate (FeCl3·6H2O, 99+%, Acros) and 1 M sodium nitrate (NaNO3, 99+%, Acros). A precursor film of β-FeOOH nanowires was

ACS Paragon Plus Environment

6

Page 7 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

grown on FTO by chemical bath deposition at 100 ℃ for 3 hours in a universal oven (Memmert UF 30 plus). After rinsing with deionized water (18.2 MΩ·cm, Synergy), these β-FeOOH thin films were introduced into a tube furnace (MTI OTF-1200X-S) preheated at 800 ℃ for 15 minutes to be transformed into hematite nanowires (see the complete morphological and structural characterization in the Supporting Information Figure S1). Preparation of Electrolytes. 1 M NaOH (pH 13.6, Reactolab SA) and 1 M potassium phosphate buffer (pH 7-12.7) were used as electrolyte for the electrochemical measurements performed on hematite. The buffer solution consists of potassium phosphate tribasic (K3PO4, >98%, Sigma-Aldrich), dibasic trihydrate (K2HPO4·3H2O, 99+%, Acros) and monobasic (KH2PO4, 99+%, Acros). The composition is displayed in Table S2. The pH of the electrolytes was determined by using a pH meter (VMR pH1000L). For experiments performed in the presence of a sacrificial agent, 0.5 mL of hydrogen peroxide (H2O2, 30%, Reactolab SA) was added into 9.5 mL of electrolyte. Photoelectrochemical Characterization. Photoelectrochemical measurements were carried out with a computer controlled (EC-LAB V11.12) potentiostat (Bio-Logic SP-300) with a 3-electrode configuration: a hematite working electrode, a platinum counter

ACS Paragon Plus Environment

7

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 40

electrode (0.25 mm diameter, 99.99%, chemPUR) and a Ag/AgCl reference electrode (ALS RE-1CP). A Cappuccino-type cell with an active geometric area of 0.238 cm2 for the working electrode was used. A xenon arc lamp (Newport 66921, 450 W), calibrated to provide simulated AM 1.5G solar irradiation. In all cases, the working electrodes were illuminated from the substrate side. For linear scanning voltammerty the voltage was swept at 20 mV/s. The applied potentials in all the measurements were converted to the reversible hydrogen electrode (RHE) using the following equation: 𝑉𝑅𝐻𝐸 = 𝑉𝐴𝑔/𝐴𝑔𝐶𝑙 + 0.059 × 𝑝𝐻 + 0.197#(1) For PEIS measurements, the frequency was scanned from 100 kHz to 50 mHz with a sinusoidal amplitude of 25 mV. An equivalent circuit modeling software ZView (Scribner Associates) was used to fit the PEIS data. All the impedance data were corrected with the actual surface area which was determined by adsorption of the organic dye orange II (Figure S2). IMPS/IMVS Measurement and Data Analysis. A white LED (Cree XLamp MC-E White, see Figure S3 for the irradiation spectrum) running at a DC background of 2.9 V (100 mA) powered by a triple power supply (HAMEG HM8040-2) was employed as light source.

ACS Paragon Plus Environment

8

Page 9 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

The sinusoidal modulation of the light intensity was controlled by an arbitrary/function generator (Tektronix AFG3021C), with a modulation depth of 7% and a frequency range from 100 kHz to 50 mHz. IMPS/IMVS measurements were performed in the 3-electrode Cappuccino-type cell described above. In IMPS experiments, the applied potential on working electrode was controlled by a potentiostat (Thompson Electrochem model 251) and the photocurrent response was recorded by an oscilloscope (Tektronix DPO7254C)

via a 50  series resistor (Velleman E6/E12). In IMVS measurements, the open circuit photovoltage response under modulated light intensity was directly probed with the oscilloscope through a BNC to banana adapter (Pomona Electronics). IMPS/IMVS data were processed and fitted using a custom program written with the Python programming language. The lmfit package for Python was used to perform curve fitting with nonlinear regression.

Operando UV-Vis Measurements. A 3-electrode configuration was set up in a cuvette (Hellma Analytics OS, 10 mm) fixed with a homemade 3D printed cuvette extender that provides larger volume to accommodate the counter and reference electrodes. For this experiments, hematite thin films were specially cut into a size of 9 × 50 mm to fit the

ACS Paragon Plus Environment

9

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 40

cuvette and were positioned in the light path. A mini Ag/AgCl reference electrode (ALS RE-1S) and a platinum counter electrode (0.25 mm diameter, 99.99%, chemPUR) were assembled in the cuvette extender. The 3-electrode setup was built inside a UV-Vis spectrophotometer (Shimadzu UV-3600) and connected to a Bio-Logic SP-50 potentiostat. A 405 nm laser diode (Thorlabs L405P150), working at 66.7% of its maximum power (100 mW) using a DC power supply (Tektronix PWS2721), was placed in front of the cuvette to photoexcite the sample while recording the absorption spectrum. To prevent the saturation of the detector by the laser diode, a 450 nm longpass filter (Thorlabs FEL0450, 21 mm diameter) was positioned in front of the UV-Vis detector in the optical path (see more details of the setup in Figure S4). The baseline was recorded under open circuit condition with external illumination (laser diode). RESULTS AND DISCUSSION Figure 1 displays the linear sweep voltammetry of nanostructured hematite photoanodes at different electrolyte pH and intermittent solar illumination. Note that in all cases a high concentration of electrolyte, 1 M potassium phosphate buffer (pH 7-12.7) or 1 M NaOH (pH 13.6), was employed to prevent dynamic changes of pH in the electrical double layer

ACS Paragon Plus Environment

10

Page 11 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

during water oxidation. At first glance, the photoanode response certainly displays a progressive enhancement with increasing pH, which is similar to that shown in previous reports.20,22 Likewise, the dynamic potential–pH diagram (DPPD, Figure 2) developed by Bard and co-workers as an extended version of the Pourbaix diagram further evidences the evolution of the j–V profile with a noticeable change in photocurrent value and onset potential.26 On the one hand, increasing the electrolyte pH from 7.0 to 13.6 causes the photocurrent values to increase from 0.15 to 1.13 mA/cm2 at 1.23 V vs RHE. Note that the highest photocurrent is close to the state-of-the-art value reported for this kind of nanostructured hematite electrode.27 On the other hand, this same increase in pH shifts the photocurrent onset cathodically from 1.1 V to 0.7 V vs RHE (when the onset potential is defined as where photocurrent density reaches 10 A/cm2). This decrease of the onset potential has been reported to be linked to the gradual mitigation of the FLP with pH.23 This phenomenon, whose origin remain unclear, will be subject of scrutiny in this study. It is worth mentioning that whereas Iandolo et al. reported a linear relationship between the onset and pH (quasi-Nernstian behaviour with a slope of 49 mV/pH),22 Zhang et al. observed two regimes, viz. the onset potential remains constant below pH 12 and

ACS Paragon Plus Environment

11

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 40

decreased in a Nernstian fashion above pH 12.20 Likewise, we observe two regimes in our experiments albeit different from that previously reported. From pH 7 to 10, the onset potential increases in a Nernstian fashion (59 mV/pH) but at higher pH this increase turned quasi-Nernstian (36 mV/pH, see Figure S5). We hypothesize that minor changes in the reactive interface (surface chemistry and exposed facets) among differentlyprepared hematite electrodes could account for the different trends on the onset potential–pH dependency.28,29 Overall, although the j–V curves support the known view of the impact of the pH on the performance they lack specificity when it comes to identify the parameters, either energetic or kinetic, controlling both the photocurrent values and the onset potential.

ACS Paragon Plus Environment

12

Page 13 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

Figure 1. Linear sweep voltammetry (LSV) curves of hematite photoanode under intermittent AM 1.5G illumination in different pH. The gray solid line represents zero current at each pH.

ACS Paragon Plus Environment

13

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 40

Figure 2. Dynamic potential–pH diagram (DPPD) of hematite photoanode under AM 1.5G illumination.

To elucidate the impact of the pH on the interfacial energetic characteristics under operation conditions PEIS experiments were performed under illumination. Figure S6 displays a representative Nyquist plot, which features two semicircles: one at high frequency assigned to the electrical response of the space charge region and a second at low frequency attributed to the interfacial charge transfer processes. A quantitative description of the electrical properties of this system is obtained by fitting the impedance spectra to a two-time-constant equivalent circuit.30,31 The equivalent circuit employed,23 contains a system series resistance (𝑅𝑆), the bulk capacitance of the semiconductor

ACS Paragon Plus Environment

14

Page 15 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

electrode (𝐶𝑏𝑢𝑙𝑘) due to the space charge region and the resistance associated with hole transport and trapping at surface states (𝑅𝑡𝑟𝑎𝑝) as well as the built-up capacitance due to the accumulation of charges at the interface (𝐶𝑆𝑆) and the charge transfer resistance corresponding to the water oxidation reaction considered to occur from the surface states (𝑅𝑐𝑡,𝑠𝑠). Figure 3a shows the Mott–Schottky (MS) plots constructed with the values of 𝐶𝑏𝑢𝑙𝑘 obtained from the impedance spectra at different pH. From 0.6 V to 0.9 V vs RHE the MS plots are similar for all pH showing the expected linear behavior. The extrapolated flat band potential (𝑉𝑓𝑏 = 0.57 ± 0.02 V vs RHE) shifts in a Nernstian fashion with pH,22,23 hence remaining unchanged with respect to the RHE (Table S3). We note that the extracted donor density 𝑁𝐷 of (2.90 ± 0.14) × 1018 cm–3 (Table S3) under illumination (considering a dielectric constant of 32,23 and a roughness factor of 20, see Figure S2) is similar to values reported in literature.23 Interestingly, decreasing the pH disrupts the linearity of the MS plot causing the appearance of a plateau at intermediate potentials, wherein 𝐶𝑏𝑢𝑙𝑘 remains constant over a range of applied bias. The drastic decrease of the slope indicates that band bending barely develops in this potential range, a sign of severe

ACS Paragon Plus Environment

15

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 40

FLP.32 We note that Klahr et al. briefly advanced the impact of pH on the FLP reporting the transition from a single- to a multiple-slope MS plot when comparing measurements at pH 13.3 and pH 6.9.23

Figure 3. (a) Mott–Schottky plots. (b) Energetic distribution of density of surface states (DOS) and surface states capacitance (𝐶𝑆𝑆) extracted from the impedance response

ACS Paragon Plus Environment

16

Page 17 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

measured at different pH conditions under 1 sun illumination. After the removal of the background capacitance, DOS profiles were fitted with Gaussian functions. A roughness factor of 20 ± 4 (Figure S2) is used for correcting the real surface area. (c) 𝑉𝑝𝑒𝑎𝑘 (black sphere) and 𝑁𝑆𝑆 (red sphere) determined from Gaussian fit as a function of pH.

Continuing the method of Klahr et al., the steady-state surface accumulated charge, expressed in the form of density of surface states (DOS) calculated by the expression23,33 𝐶𝑆𝑆(𝐸) = 𝑞 × 𝐷𝑂𝑆, is plotted as a function of the applied potential at different pH conditions (Figure 3b). The DOS can be fit to gaussian functions whose integration gives the total density of surface states (𝑁𝑆𝑆) and the peak position (𝑉𝑝𝑒𝑎𝑘) defines their energetic location.23,33 Figure 3c provides an overview on how the pH affects both 𝑁𝑆𝑆 and 𝑉𝑝𝑒𝑎𝑘. As observed, 𝑁𝑆𝑆 drops with increasing pH which is consistent with FLP mitigation and the shortening of the plateau region in the MS plot. Note that although 𝑁𝑆𝑆 values are in all the cases above 1013 cm–2 (the threshold above which FLP would manifest based on the doping density of the bulk),32 the pinning is barely visible in the MS plot at high pH. Likewise, 𝑉𝑝𝑒𝑎𝑘 shifts towards more negative potentials with increasing pH, which agrees

ACS Paragon Plus Environment

17

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 40

with the observed changes in the photocurrent onset potential. Identifying the underlying phenomena whereby the pH dictates the 𝑁𝑆𝑆 and 𝑉𝑝𝑒𝑎𝑘 values is key to rationalize the pHdependence performance. It is well accepted that the DOS extracted from PEIS corresponds to the intermediate species involved in the water oxidation reaction, rather than surface trap states,34 and that their location, just before the onset of photocurrent, unequivocally demonstrates the build-up of surface holes required to trigger the rather sluggish interfacial charge transfer.22,23,33 Note that the applied potential at which DOS are observed (Figure 3b) does not correspond their actual energetic location. Given the photovoltage generated at the SCLJ (due to quasi-Fermi levels splitting) the DOS will certainly be more oxidizing and obviously below the redox potential for water oxidation (>1.23 V vs RHE). To further corroborate the identity of DOS, operando UV-Vis spectroscopy measurements were performed as a function of the pH. As shown in Figure S7, a distinct broad absorption band centered at around 570 nm was observed in all cases. This has been previously ascribed to the formation of –FeIV=O, which is actively involved in the water oxidation reaction.25,35 These results suggests that the surface accumulated holes tracked by PEIS

ACS Paragon Plus Environment

18

Page 19 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

correspond to reactive intermediates, and thus, the surface concentration of these interfacial active species dictate the FLP. It is worth noting that Iandolo et al. explained the pH-dependent performance of hematite photoanodes by invoking a gradual increase of the density of surface trap states with decreasing pH.22 However, in that study, the authors did not observe any evolution in 𝑁𝑆𝑆 nor in the MS plot but a shift in 𝑉𝑝𝑒𝑎𝑘, which indeed does not directly support FLP mitigation since charge accumulation still occurs. Alternatively to the purely energetic proposition of Iandolo et al.,22 changes in the relative kinetics of surface recombination and charge transfer triggered by the electrolyte pH could possibly alter the balance between these competitive processes rendering higher or lower surface accumulated charges that in turn would govern the FLP. However, to support this view it is necessary decouple the kinetics of the surface recombination and charge transfer, which can be accomplished with IMPS measurements. The IMPS response 𝑗(𝜔) of a photoelectrode can be expressed within the theoretical framework constructed by Peter and co-coworkers36 and considering a small modulation depth regime (7%) where:

ACS Paragon Plus Environment

19

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

𝑘𝑡𝑟𝑎𝑛 + 𝑖𝜔

𝑗(𝜔) = 𝑗ℎ

(

)

𝐶𝐻 𝐶 1 + 𝐻 𝐶𝑆𝐶

𝑘𝑡𝑟𝑎𝑛 + 𝑘𝑟𝑒𝑐 + 𝑖𝜔

Page 20 of 40

(1 +1𝑖𝜔𝜏)#(2)

where 𝑘𝑡𝑟𝑎𝑛 and 𝑘𝑟𝑒𝑐 are the pseudo-first order charge transfer and recombination rate constant, respectively. 𝐶𝑆𝐶 and 𝐶𝐻 represent capacitance of the space charge region and the Helmholtz layer, 𝑗ℎ denotes the hole current towards the SCLJ, and 𝜏 corresponds to the RC time constant of the high frequency attenuation by the electrochemical cell. The IMPS spectra of a hematite photoanode recorded at 1.4 V vs RHE where the effect of FLP is minimized (Figure 3b) under different pH conditions is displayed in Figure 4a. As observed, the IMPS response consists of two well-defined semicircles with characteristic high- and low-frequency intersects (HFI and LFI, respectively) with the real axis. It is commonly accepted that the HFI directly relates to the flux of holes arriving at the SCLJ while the LFI accounts for those holes transferred to the electrolyte.36–38 Interestingly, as shown in Figure 4b, the current corresponding to the HFI barely changes under neutral and moderately alkaline conditions, whereas the LFI increases steadily with pH. On the one hand, the sluggish shift of the HFI reflects that the built-in electric field within the hematite photoanode, which controls the collection of holes at the interface,

ACS Paragon Plus Environment

20

Page 21 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

barely changes over a wide range of pH. Note that the HFI slightly increases under strongly alkaline conditions. This could originate from the mitigation of the FLP as previously discussed. On the other hand, the direct relation between LFI and pH clearly shows the gradual suppression of surface recombination with increasing pH. Thus the IMPS data are consistent with the view that the suppression of surface recombination with increasing pH does not result from an enlarged band bending, as commonly accepted,37,39 but rather from changes in the delicate balance between the surface recombination and charge transfer.

ACS Paragon Plus Environment

21

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 40

Figure 4. (a) Representative Nyquist plots of IMPS response measured at 1.4 V vs RHE at 7 different pH conditions. Corresponding dash lines represent the fitting results. (b) Low- and high- frequency intersects (LFI and HFI) and HFI–LFI as a function of pH. (c) Charge transfer (𝑘𝑡𝑟𝑎𝑛) and (d) surface recombination (𝑘𝑟𝑒𝑐) rate constant as a function of pH recorded at 1.0 V (magenta), 1.2 V (olive) and 1.4 V vs RHE (cyan).

The development of the extracted rate constants 𝑘𝑡𝑟𝑎𝑛 and 𝑘𝑟𝑒𝑐 as a function of pH (see Figure 4c and d, respectively) at 1.4 V vs RHE and at lower applied potentials (1.0 and 1.2 V, see Figure S8 for raw IMPS spectra) further support this view. Firstly, 𝑘𝑡𝑟𝑎𝑛 increases with the applied bias regardless of the pH. This is commonly assumed to be a sign of FLP33,39, but it could also be the result of changes in the reaction mechanism. When examining the evolution of 𝑘𝑡𝑟𝑎𝑛 with the applied potential, it is evident that it displays a larger increase at high pH (>10). Interestingly, this behavior contradicts the one expected from the pH-dependence of the FLP described before, wherein the progressive mitigation of FLP with pH would lead to a constant 𝑘𝑡𝑟𝑎𝑛 value. However, this unexpected behavior of 𝑘𝑡𝑟𝑎𝑛 can be explained by invoking changes in the reaction

ACS Paragon Plus Environment

22

Page 23 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

mechanism, which would plausibly affect 𝑘𝑡𝑟𝑎𝑛. In fact, Zhang et al. reported a gradual change in the reaction mechanism (from first- to second-order) for the OER on hematite photoanodes when the pH increased above 10, although the effect of the pH on the rate constant of the interfacial reaction was not discussed.20,21 Secondly, 𝑘𝑟𝑒𝑐 displays a much less obvious dependence with applied bias and pH compared to 𝑘𝑡𝑟𝑎𝑛. Regardless of the pH, 𝑘𝑟𝑒𝑐 decreases with increasing applied potential although the magnitude of the decrease becomes more significant at high pH (>10). Bearing in mind that 𝑘𝑟𝑒𝑐 is governed by the built-in electric field at the space charge region, it is not unexpected that increasing the applied bias would drive electrons away from the interface mitigating the surface recombination and hence reducing 𝑘𝑟𝑒𝑐. Interestingly, the decrease in 𝑘𝑟𝑒𝑐 observed at pH>10 is directly correlated with the drastic suppression of the FLP that allows building a larger band bending for the same applied potential. Having decoupled the surface-related processes of charge recombination and transfer using IMPS, we can now rationalize the evolution of 𝑁𝑆𝑆 with pH (as seen in Figure 3c). Under illumination and at the photocurrent onset potential, the quasi-Fermi level of holes is sufficiently oxidizing to form the high valence iron species (reactive sites) tracked by

ACS Paragon Plus Environment

23

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 40

PEIS. The surface concentration of these intermediates, 𝑁𝑆𝑆, in steady-state conditions will depend upon the balance between the rate constants of surface charging, recombination (𝑘𝑟𝑒𝑐) and interfacial reaction (𝑘𝑡𝑟𝑎𝑛). Given that the surface charging rate remains unaltered by changes in pH (according to the IMPS data and the extracted HFI), the changing accumulation of holes at the SCLJ must be controlled by the active competition between recombination and charge transfer. On one hand, the reduction of 𝑁𝑆𝑆 with increasing pH clearly is correlated to the increase in 𝑘𝑡𝑟𝑎𝑛 with constant 𝑘𝑟𝑒𝑐, especially with a drastic change at around pH 9-10. On the other hand, the shift of 𝑉𝑝𝑒𝑎𝑘 and therefore, of the apparent location of the intermediate species would result from the unbalanced surface kinetics. At low pH, the large build-up of surface charges as the potential is scanned anodically causes the potential to drop across the Helmholtz layer, shifting the bands, instead of across the space charge region, delaying the development of band bending and therefore the photocurrent onset. At high pH, the favorable kinetics for water oxidation prevents the large accumulation of holes at the interface, minimizing the potential region wherein band bending is not developed, and shifting, therefore, the photocurrent onset.

ACS Paragon Plus Environment

24

Page 25 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

Results provided so far unveiled the effect of the pH on the kinetics of water oxidation as the prime agent programming the photoanodic response but could not elucidate whether the pH impacted as well the interfacial electronic structure of the hematite electrodes. We note that this specific information is important when considering other PEC reactions different from water oxidation performed in aqueous solutions.40 With the aim of the resolving specifically the impact of pH on the photogenerated charge carrier behavior in the absence of the water oxidation reaction, IMVS measurements were performed. One of the distinct features of this technique is that it is only sensitive to the recombination processes occurring within the photoelectrode since, undertaken at open circuit conditions, the ultimate fate of all photogenerated carries is recombination. Therefore, IMVS not only uncovers the characteristic time constants associated with recombination processes, but more importantly, untangles recombination pathways that occur at different timescale. Contrarily to the vast literature on IMVS measurements to characterize photovoltaic devices such as dye-sensitized solar cells (DSSCs)41 and organic solar cells,42 the use of this technique for PEC water splitting is relatively novel and its interpretation remains, to some extent, unclear.43 In fact, Wang and co-workers

ACS Paragon Plus Environment

25

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 40

utilized this technique to qualitatively analyze the effect of a series of surface treatments on the response of hematite photoanodes,38 whereas Rothschild and co-workers demonstrated its advantages over conventional PEIS.44 Nyquist plots of IMVS spectra for hematite photoanodes at different pH conditions are shown in Figure 5a, whereas those including H2O2 as sacrificial hole scavenger are displayed in Figure S10. Interestingly, the IMVS response features two distinct semicircles in the quadrant I (low-frequency) and quadrant IV (high-frequency), albeit the latter appears to be noticeably flattened as a consequence of the RC attenuation. Note as well that the low-frequency response does not appear in the presence of H2O2, suggesting that one of the recombination pathways is drastically suppressed in these measurement conditions. Likewise, in the Bode magnitude plots (Figure 5b and S10), two processes are clearly resolved in the absence of H2O2, i.e. a slow process in the frequency range of 0.1-1 Hz) and a fast process at around 1-1000 Hz, although in the presence of the sacrificial reagent only the faster remains. Generally speaking, three main recombination pathways coexist in the illuminated photoelectrode under open circuit conditions, namely, bulk, space charge region recombination (recombination of holes at the SCLJ with electrons at the edge of

ACS Paragon Plus Environment

26

Page 27 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

the space charge region) and interfacial recombination (involving long-lived reactive intermediate species).38 In the specific case of hematite photoanodes, transient absorption spectroscopy measurements have revealed that the bulk recombination extends over a wide range of timescales, from picoseconds to milliseconds, whereas the space charge region recombination, occurs in the millisecond to second timescale.45 With all this in mind, and taking into account that the bulk recombination is likely out of the scope of our technique (sub-microsecond timescale), it would not be unreasonable to propose that the faster component is associated to the space charge region recombination, and therefore, present regardless of the electrolyte conditions whereas the slower one would be linked to the interfacial recombination involving water oxidation intermediates. It is worth noting that similar observations have been reported on a recent impedance study by Klotz et al., wherein the disappearance of the slow process in the presence of H2O2 was ascribed to the suppression of surface recombination.46

ACS Paragon Plus Environment

27

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 40

Figure 5. (a) Nyquist plots and (b) Bode magnitude plots of IMVS response with the corresponding fitting curves (solid line). (c) 7% (modulation depth) of steady-state photovoltage and 𝑀1 ― 𝑀2 as a function of pH. (d) Characteristic lifetimes 𝜏𝑆𝐶𝑅 (violet) and 𝜏𝑖𝑛𝑡 (orange) as a function of the pH. It worth noting that a strong Fermi level pinning at pH 7 and 8 prevents the photovoltage modulation to occur to a noticeable extent, therefore IMVS response at these two pH conditions are not included.

To further advance a more quantitative study of the IMVS spectra, we undertake a phenomenological analysis by using the analytical expression previously reported by Frank and co-workers:47

ACS Paragon Plus Environment

28

Page 29 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

𝑅𝑒[𝑉𝑂𝐶](𝜔) =

𝐼𝑚[𝑉𝑂𝐶](𝜔) =

𝑀1 1 + 𝜔2𝜏2𝑖𝑛𝑡 ― 𝑀1𝜔𝜏𝑖𝑛𝑡 1 + 𝜔2𝜏2𝑖𝑛𝑡



+

𝑀2 1 + 𝜔2𝜏2𝑆𝐶𝑅

#(3)

𝑀2𝜔𝜏𝑆𝐶𝑅 1 + 𝜔2𝜏2𝑆𝐶𝑅

#(4)

where 𝑀1 and 𝑀2 are scaling factors while 𝜏𝑆𝐶𝑅 and 𝜏𝑖𝑛𝑡 correspond to the time constant associated to the recombination occurring at the space charge region and involving surface intermediate species, respectively. Regarding the physical meaning of the scaling factors of eq 3 and eq 4, we note that after normalizing the steady-state photovoltage ( 𝑉𝑝ℎ) by the modulation depth (7%), the values obtained match with the low-frequency intersect (𝑀1 ― 𝑀2) (Figure 5c and S10c) unveiling their direct link to the photovoltage developed by the electrode. As shown in Figure 5a,b and S10a,b these expressions describe quite well the IMVS spectra allowing for the extraction of the characteristic time constants, plotted in Figure 5d and S10d as a function of the pH. 𝜏𝑆𝐶𝑅 is observed to increase with pH from about 8 ms to approximately 16 ms, which is in the order of magnitude of that reported for other iron-based metal oxide photoanodes.33 Surprisingly, 𝜏𝑖𝑛𝑡 exhibit much longer characteristic times and a different behavior, since it drops from 525 ms to 300 ms with increasing pH. Interestingly, the charge carrier behavior changes

ACS Paragon Plus Environment

29

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 40

in the presence of H2O2, with 𝜏𝑆𝐶𝑅 displaying slightly longer values (around 25 ms) that remain virtually unaltered with pH. To better interpret these trends, it is worth including another piece of information, i.e., 𝑉𝑝ℎ achieved under standard illumination conditions (Figure 5c). which ultimately relates to the concentration of electrons within the semiconductor under illumination. As observed, the 𝑉𝑝ℎ appears to be mainly controlled by 𝜏𝑆𝐶𝑅 since it increases with pH and further ameliorates in the presence of H2O2. Intuitively, 𝜏𝑆𝐶𝑅 is expected to remain constant regardless of the pH since it represents the recombination behavior of carriers located within the space charge region, and this agrees quite well with the observations in the presence of H2O2 (Figure S10d). However, the pH-dependent nature of 𝜏𝑆𝐶𝑅 clearly indicates the occurrence of a different recombination pathway that is clearly influenced by the pH. Indeed, the increase of 𝜏𝑆𝐶𝑅 with pH matches well with the decrease of the density of surface traps proposed by Iandolo et al., which could act as recombination centers.22 The shortening of 𝜏𝑖𝑛𝑡 reflects a faster recombination between the holes accumulated at the interface (in the form of intermediate species of water oxidation) and photogenerated electrons at the space charge region. Note that this same recombination pathway is monitored by IMPS as 𝑘𝑟𝑒𝑐

ACS Paragon Plus Environment

30

Page 31 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

although in different experimental conditions, what accounts for the comparable time constants but different trend. As described before, 𝑘𝑟𝑒𝑐 is mainly controlled by the degree of band bending at a certain applied potential when comparing different pH conditions. Bearing in mind that the increased 𝑉𝑝ℎ with pH indicates a reduction in the built-in electric field existing in the semiconductor under illumination, i.e. the bands are flatten to a larger degree, it would not be unreasonable to observe a faster recombination kinetics of the intermediate species with photogenerated electrons. Likewise, 𝜏𝑖𝑛𝑡 could be also influenced by the identity of the intermediate species, being those formed at higher pH more reactive showing faster recombination kinetics. It is worth noting that Wang and coworkers assigned both the fast and slow processes to surface-related charge recombination pathways involving adsorbed water molecules and intermediate species of the water oxidation reaction, respectively.38 However, our analysis of the IMVS data as a function of the pH suggests that surface recombination pathways involving reactive species and trap states could be resolved separately. CONCLUSIONS

ACS Paragon Plus Environment

31

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

In

this

work,

conventional

PEC

measurements

Page 32 of 40

alongside

a

series

of

spectroelectrochemical techniques that interrogate the bulk/interfacial electronics and the identity of the surface reactive sites have been deployed to shed some light into the longrunning debate on which factor, either the semiconductor surface energetics or the interfacial reaction kinetics, control the pH-dependent performance on metal oxide photoanodes, particularly, hematite. Our studies confirmed that the degree of FLP increases with decreasing pH affecting adversely to the performance as reported elsewhere but challenged the current proposition that the changes in the interfacial structure are the ones governing the FLP. We observed that the surface accumulation of charges, which originates the FLP, links directly to the kinetic balance between surface recombination and charge transfer, concluding that changes in the water oxidation mechanism ultimately control the FLP. Likewise, we introduced a phenomenological analysis for the IMVS response of photoelectrodes that allowed to decouple the recombination pathways involving surface trap states and intermediate reactive species. In a more general vein, we pin down the interplay between the pH and the solar water oxidation on hematite photoanodes within an electrochemical framework, providing

ACS Paragon Plus Environment

32

Page 33 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

support and complementing recent insights gained on the reaction mechanism. We hope that these results will transcend beyond hematite and extend over the broad family of metal oxide photoanodes, bringing the attention not only to the delicate relationship that exists between the semiconductor surface and the electrolyte pH but also to the need for specific optimization. ASSOCIATED CONTENT

Supporting Information

The Supporting Information is available free of charge on the ACS Publications Website at

DOI:

******.

Hematite morphology, Raman spectrum, electrolyte preparation, roughness factor determination, irradiation spectrum of the white LED, operando UV-Vis setup and spectrum, onset potential vs pH, impedance fitting, 𝑁𝐷 and 𝑉𝑓𝑏 in different pH, IMPS response at other potential, photocurrent response with the white LED, IMVS response in sacrificial conditions (PDF)

AUTHOR INFORMATION

ACS Paragon Plus Environment

33

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 40

Corresponding Author *[email protected]

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT This work was supported by the Swiss National Science Foundation (SNSF) under the Ambizione Energy grant (PZENP2_166871). The authors want to thank Prof. Kevin Sivula for the helpful discussions as well as for his insightful comments regarding the experimental design and data interpretation.

REFERENCES (1)

Dalle, K. E.; Warnan, J.; Leung, J. J.; Reuillard, B.; Karmel, I. S.; Reisner, E. Electro- and Solar-Driven Fuel Synthesis with First Row Transition Metal Complexes. Chem. Rev. 2019, 119 (4), 2752–2875.

(2)

Cherevan, A. S.; Deilmann, L.; Weller, T.; Eder, D.; Marschall, R. Mesoporous Semiconductors: A New Model to Assess Accessible Surface Area and Increased Photocatalytic Activity? ACS Appl. Energy Mater. 2018, 1 (11), 5787–5799.

(3)

Yang, X.; Wang, D. Photocatalysis: From Fundamental Principles to Materials and Applications. ACS Appl. Energy Mater. 2018, 1 (12), 6657–6693.

ACS Paragon Plus Environment

34

Page 35 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

(4)

Rahmanudin, A.; Yao, L.; Sekar, A.; Cho, H.-H.; Liu, Y.; Lhermitte, C. R.; Sivula, K. Fully Conjugated Donor–Acceptor Block Copolymers for Organic Photovoltaics via Heck–Mizoroki Coupling. ACS Macro Lett. 2019, 8 (2), 134–139.

(5)

Rahmanudin, A.; Yao, L.; A. Jeanbourquin, X.; Liu, Y.; Sekar, A.; Ripaud, E.; Sivula, K. Melt-Processing of Small Molecule Organic Photovoltaics via Bulk Heterojunction Compatibilization. Green Chem. 2018, 20 (10), 2218–2224.

(6)

Xia, M.; Jia, L.; Li, J.; Liu, Y.; Wang, X.; Chi, B.; Pu, J.; Jian, L. Effects of CoDoped Barium Cerate Additive on Morphology, Conductivity and Electrochemical Properties of Samarium Doped Ceria Electrolyte for Intermediate Temperature Solid Oxide Fuel Cells. Int. J. Hydrog. Energy 2018, 43 (33), 16293–16301.

(7)

Higashi, T.; Nishiyama, H.; Suzuki, Y.; Sasaki, Y.; Hisatomi, T.; Katayama, M.; Minegishi, T.; Seki, K.; Yamada, T.; Domen, K. Transparent Ta3N5 Photoanodes for Efficient Oxygen Evolution toward the Development of Tandem Cells. Angew.

Chem. Int. Ed. 2019, 58 (8), 2300–2304. (8)

Zhu, X.; Guijarro, N.; Liu, Y.; Schouwink, P.; Wells, R. A.; Formal, F. L.; Sun, S.; Gao, C.; Sivula, K. Spinel Structural Disorder Influences Solar-Water-Splitting Performance of ZnFe2O4 Nanorod Photoanodes. Adv. Mater. 2018, 30 (34), 1801612.

(9)

Ferraz, L. C. C.; Carvalho, W. M.; Criado, D.; Souza, F. L. Vertically Oriented Iron Oxide Films Produced by Hydrothermal Process: Effect of Thermal Treatment on the Physical Chemical Properties. ACS Appl. Mater. Interfaces 2012, 4 (10), 5515–5523.

(10) Miyamoto, H.; Matsuzaki, K.; Yamaguchi, A.; Miyauchi, M. Visible-Light-Active Photoelectrochemical Z-Scheme System Based on Top 5 Clarke-Number Elements. ACS Appl. Energy Mater. 2018, 1 (11), 5954–5959. (11) Xie, J.; Yang, P.; Liang, X.; Xiong, J. Self-Improvement of Ti:Fe2O3 Photoanodes: Photoelectrocatalysis Improvement after Long-Term Stability Testing in Alkaline Electrolyte. ACS Appl. Energy Mater. 2018, 1 (6), 2769–2775. (12) Sahu, T. K.; Shah, A. K.; Banik, A.; Qureshi, M. Enhanced Surface and Bulk Recombination Kinetics by Virtue of Sequential Metal and Nonmetal Incorporation

ACS Paragon Plus Environment

35

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 40

in Hematite-Based Photoanode for Superior Photoelectrochemical Water Oxidation. ACS Appl. Energy Mater. 2019, 2 (6), 4325–4334. (13) Li, M.; Qiu, Y.; Qiu, W.; Tang, S.; Xiao, S.; Ma, S.; Ouyang, G.; Tong, Y.; Yang, S. Boosting the Photoelectrochemical Water Oxidation at Hematite Photoanode by Innovating a Hierarchical Ball-on-Wire-Array Structure. ACS Appl. Energy Mater. 2018, 1 (11), 5836–5841. (14) Yang, L.; Xiong, Y.; Li, L.; Yang, Y.; Gao, D.; Liu, L.; Dong, H.; Xiao, P.; Zhang, Y. Manipulation of Charge Transfer in FeP@Fe2O3 Core–Shell Photoanode by Directed Built-In Electric Field. ACS Appl. Energy Mater. 2018, 1 (9), 4591–4598. (15) Jia, L.; Bogdanoff, P.; Schmid, M.; Bloeck, U.; Fiechter, S.; Wang, H. Reconstruction of Solar Fuel Ultrathin Films via Periodically Microbending for Efficient Photoelectrochemical Water Splitting. ACS Appl. Energy Mater. 2018, 1 (12), 6748–6757. (16) Walsh, D.; Zhang, J.; Regue, M.; Dassanayake, R.; Eslava, S. Simultaneous Formation of FeOx Electrocatalyst Coating within Hematite Photoanodes for Solar Water Splitting. ACS Appl. Energy Mater. 2019, 2 (3), 2043–2052. (17) Liu, Y.; Wei, C.; Ngaw, C. K.; Zhou, Y.; Sun, S.; Xi, S.; Du, Y.; Loo, J. S. C.; Ager, J. W.; Xu, Z. J. Operando Investigation of Mn3O4+δ Co-Catalyst on Fe2O3 Photoanode: Manganese-Valency-Determined Enhancement at Varied Potentials.

ACS Appl. Energy Mater. 2018, 1 (2), 814–821. (18) Shinagawa, T.; Takanabe, K. Towards Versatile and Sustainable Hydrogen Production through Electrocatalytic Water Splitting: Electrolyte Engineering.

ChemSusChem 2017, 10 (7), 1318–1336. (19) Guijarro, N.; Prévot, M. S.; Sivula, K. Surface Modification of Semiconductor Photoelectrodes. Phys. Chem. Chem. Phys. 2015, 17 (24), 15655–15674. (20) Zhang, Y.; Zhang, H.; Ji, H.; Ma, W.; Chen, C.; Zhao, J. Pivotal Role and Regulation of Proton Transfer in Water Oxidation on Hematite Photoanodes. J.

Am. Chem. Soc. 2016, 138 (8), 2705–2711. (21) Zhang, Y.; Zhang, H.; Liu, A.; Chen, C.; Song, W.; Zhao, J. Rate-Limiting O–O Bond Formation Pathways for Water Oxidation on Hematite Photoanode. J. Am.

Chem. Soc. 2018, 140 (9), 3264–3269.

ACS Paragon Plus Environment

36

Page 37 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

(22) Iandolo, B.; Hellman, A. The Role of Surface States in the Oxygen Evolution Reaction on Hematite. Angew. Chem. Int. Ed. 2014, 53 (49), 13404–13408. (23) Klahr, B.; Gimenez, S.; Fabregat-Santiago, F.; Hamann, T.; Bisquert, J. Water Oxidation at Hematite Photoelectrodes: The Role of Surface States. J. Am. Chem.

Soc. 2012, 134 (9), 4294–4302. (24) Zandi, O.; Hamann, T. W. Determination of Photoelectrochemical Water Oxidation Intermediates on Haematite Electrode Surfaces Using Operando Infrared Spectroscopy. Nat. Chem. 2016, 8 (8), 778–783. (25) Klahr, B.; Hamann, T. Water Oxidation on Hematite Photoelectrodes: Insight into the Nature of Surface States through In Situ Spectroelectrochemistry. J. Phys.

Chem. C 2014, 118 (19), 10393–10399. (26) Minguzzi, A.; F. Fan, F.-R.; Vertova, A.; Rondinini, S.; J. Bard, A. Dynamic Potential–pH Diagrams Application to Electrocatalysts for Water Oxidation. Chem.

Sci. 2012, 3 (1), 217–229. (27) Jeon, T. H.; Moon, G.; Park, H.; Choi, W. Ultra-Efficient and Durable Photoelectrochemical Water Oxidation Using Elaborately Designed Hematite Nanorod Arrays. Nano Energy 2017, 39, 211–218. (28) Kallay, N.; Dojnović, Z.; Čop, A. Surface Potential at the Hematite–Water Interface. J. Colloid Interface Sci. 2005, 286 (2), 610–614. (29) Chatman, S.; Zarzycki, P.; Rosso, K. M. Surface Potentials of (001), (012), (113) Hematite (α-Fe2O3) Crystal Faces in Aqueous Solution. Phys. Chem. Chem. Phys. 2013, 15 (33), 13911–13921. (30) Xing, C.; Liu, Y.; Zhang, Y.; Liu, J.; Zhang, T.; Tang, P.; Arbiol, J.; Soler, L.; Sivula, K.; Guijarro, N.; Wang, X.; Li, J.; Du, R.; Zuo, Y.; Cabot, A.; Llorca, J. Porous NiTiO3/TiO2 Nanostructures for Photocatatalytic Hydrogen Evolution. J.

Mater. Chem. A 2019, 7 (28), 17053–17059. (31) Zuo, Y.; Liu, Y.; Li, J.; Du, R.; Yu, X.; Xing, C.; Zhang, T.; Yao, L.; Arbiol, J.; Llorca, J.; Sivula, K.; Guijarro, N.; Cabot, A. Solution-Processed Ultrathin SnS2–Pt Nanoplates for Photoelectrochemical Water Oxidation. ACS Appl. Mater.

Interfaces 2019, 11 (7), 6918–6926.

ACS Paragon Plus Environment

37

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 38 of 40

(32) Bard, A. J.; Bocarsly, A. B.; Fan, F. R. F.; Walton, E. G.; Wrighton, M. S. The Concept of Fermi Level Pinning at Semiconductor/Liquid Junctions. Consequences for Energy Conversion Efficiency and Selection of Useful Solution Redox Couples in Solar Devices. J. Am. Chem. Soc. 1980, 102 (11), 3671–3677. (33) Liu, Y.; Formal, F. L.; Boudoire, F.; Yao, L.; Sivula, K.; Guijarro, N. Insights into the Interfacial Carrier Behaviour of Copper Ferrite (CuFe2O4) Photoanodes for Solar Water Oxidation. J. Mater. Chem. A 2019, 7 (4), 1669–1677. (34) Zandi, O.; Hamann, T. W. Enhanced Water Splitting Efficiency Through Selective Surface State Removal. J. Phys. Chem. Lett. 2014, 5 (9), 1522–1526. (35) Takashima, T.; Ishikawa, K.; Irie, H. Detection of Intermediate Species in Oxygen Evolution on Hematite Electrodes Using Spectroelectrochemical Measurements.

J. Phys. Chem. C 2016, 120 (43), 24827–24834. (36) Ponomarev, E. A.; Peter, L. M. A Generalized Theory of Intensity Modulated Photocurrent Spectroscopy (IMPS). J. Electroanal. Chem. 1995, 396 (1), 219– 226. (37) Zachäus, C.; Abdi, F. F.; Peter, L. M.; Krol, R. van de. Photocurrent of BiVO4 is Limited by Surface Recombination, Not Surface Catalysis. Chem. Sci. 2017, 8 (5), 3712–3719. (38) Thorne, J. E.; Jang, J.-W.; Liu, E. Y.; Wang, D. Understanding the Origin of Photoelectrode Performance Enhancement by Probing Surface Kinetics. Chem.

Sci. 2016, 7 (5), 3347–3354. (39) Peter, L. M.; Wijayantha, K. G. U.; Tahir, A. A. Kinetics of Light-Driven Oxygen Evolution at α-Fe2O3 Electrodes. Faraday Discuss. 2012, 155 (0), 309–322. (40) Zhang, L.; Liardet, L.; Luo, J.; Ren, D.; Grätzel, M.; Hu, X. Photoelectrocatalytic Arene C–H Amination. Nat. Catal. 2019, 2 (4), 366–373. (41) Langmar, O.; Ganivet, C. R.; Scharl, T.; de la Torre, G.; Torres, T.; Costa, R. D.; Guldi, D. M. Modifying the Semiconductor/Electrolyte Interface in CuO p-Type Dye-Sensitized Solar Cells: Optimization of Iodide/Triiodide-Based Electrolytes.

ACS Appl. Energy Mater. 2018, 1 (11), 6388–6400. (42) Guijarro, N.; Yao, L.; Le Formal, F.; Wells, R. A.; Liu, Y.; Darwich, B. P.; Navratilova, L.; Cho, H.-H.; Yum, J.-H.; Sivula, K. Lead Halide Perovskite

ACS Paragon Plus Environment

38

Page 39 of 40 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

Quantum Dots to Enhance the Power Conversion Efficiency of Organic Solar Cells. Angew. Chem. Int. Ed. 2019, 58. doi:10.1002/anie.201906803 (43) Tian, Z.; Shao, F.; Zhao, W.; Qin, P.; He, J.; Huang, F. Enhanced Charge Injection and Collection of Niobium-Doped TiO2/Gradient Tungsten-Doped BiVO4 Nanowires for Efficient Solar Water Splitting. ACS Appl. Energy Mater. 2018, 1 (3), 1218–1225. (44) Klotz, D.; Ellis, D. S.; Dotan, H.; Rothschild, A. Empirical in Operando Analysis of the Charge Carrier Dynamics in Hematite Photoanodes by PEIS, IMPS and IMVS.

Phys. Chem. Chem. Phys. 2016, 18 (34), 23438–23457. (45) Le Formal, F.; Pendlebury, S. R.; Cornuz, M.; Tilley, S. D.; Grätzel, M.; Durrant, J. R. Back Electron–Hole Recombination in Hematite Photoanodes for Water Splitting. J. Am. Chem. Soc. 2014, 136 (6), 2564–2574. (46) Klotz, D.; Grave, D. A.; Dotan, H.; Rothschild, A. Empirical Analysis of the Photoelectrochemical Impedance Response of Hematite Photoanodes for Water Photo-Oxidation. J. Phys. Chem. Lett. 2018, 9 (6), 1466–1472. (47) Schlichthörl, G.; Huang, S. Y.; Sprague, J.; Frank, A. J. Band Edge Movement and Recombination Kinetics in Dye-Sensitized Nanocrystalline TiO2 Solar Cells:  A Study by Intensity Modulated Photovoltage Spectroscopy. J. Phys. Chem. B 1997,

101 (41), 8141–8155.

ACS Paragon Plus Environment

39

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 40 of 40

TOC GRAPHICS

ACS Paragon Plus Environment

40