2.1. Rate law analysis of surface holes
Physical characterizations and PEC performances of four typical photoanodes were illustrated in Figures S1 − S4 and Figure S5, respectively. The surface hole density was measured by the EIS method according to our previous works (detailed in the Supporting Information), as shown in Figures S6 and S7. The rate law analysis of surface holes was conducted on the four emblematic metal oxide photoanodes: hematite (α-Fe2O3), titanium dioxide (TiO2), tungsten oxide (WO3), and bismuth vanadate (BiVO4), as depicted in Fig. 1. The reaction order of surface holes can be determined from the slope of log(Jph) − log([h+]) curve according to Eq. 1. For these four photoanodes with low surface hole densities (under low light intensities), α-Fe2O3 (Fig. 1a) and TiO2 (Fig. 1b) photoanodes exhibited a second-order kinetics, and WO3 (Fig. 1c) and BiVO4 (Fig. 1d) photoanodes exhibited a third-order kinetics. As surface hole densities increased, all four photoanodes showed a fourth-order kinetics of surface holes for WOR. The light intensity of the transition from second- or third-order to fourth-order was 260 mW·cm− 2 for α-Fe2O3 and WO3, 190 mW·cm− 2 for BiVO4, and 90 mW·cm− 2 for TiO2. The fourth-order kinetics of surface holes implied that the four photoanodes can accumulate four surface holes at the active site before the RDS of WOR, the O − O bond formation step.52
In the following section, we specifically focused on α-Fe2O3 photoanodes to further investigate the detailed reaction mechanism. The fourth-order kinetics of surface holes were also observed under white LED illumination (Figure S8) or in near neutral conditions (pH = 8 ~ 10) (Figure S9a) on the α-Fe2O3 photoanode. More importantly, it can be seen from Fig. 1e that the reaction orders of surface holes could be further divided into three distinct regions with the surface hole density ranging from 0.4 to 1.8 h+·nm− 2. In addition to the second-order kinetics at surface hole densities below 1 h+·nm− 2 and fourth-order kinetics at surface hole densities between 1 h+·nm− 2 and 1.7 h+·nm− 2, an unexpected eighth-order kinetics of surface holes appeared with surface hole densities exceeding 1.7 h+·nm− 2. Accordingly, the Jph increased from 0.02 ~ 0.09 mA·cm− 2 to 0.12 ~ 0.74 mA·cm− 2 and was further enhanced to 1.10 mA·cm− 2 for the third region, which was nearly two orders of magnitude enhancement. Besides, an ultra-high reaction order (~ seventh-order) was also observed in 1 M NaOH solution at high surface hole densities on the α-Fe2O3 photoanode (Figure S9b). Similar cases of reaction orders exceeding fourth-order have been documented in previous studies and were explicitly mentioned by Cowan et al.,48–50 but no further understanding for this ultra-high reaction order of surface holes was provided.
The theoretical maximum reaction order of surface holes for WOR should be four, so we believe that the ultra-high reaction order is an illusion derided from the over-simplified application of Eq. 1. It is important to recognize that the apparent reaction order obtained from Eq. 1 may be influenced by other factors. For instance, kwo may be not constant and impact the apparent reaction order. Recently, Teschner et al. discovered that the Ea of WOR on metal oxide electrocatalysts was not constant but decreased linearly with the increase of surface hole density.53 The relationship between kwo and Ea is described by:
$$\:{k}_{wo}={k}_{0}\text{exp}\left(-\frac{{E}_{a}}{{k}_{B}T}\right)$$
2
where k0 is the pre-exponential factor and independent of surface holes. kB and T are the Boltzmann constant and temperature, respectively. Combining Eq. 1 and Eq. 2 with Eyring theory, a new exponential linear formula can be derived (for detailed discussions, see Supporting Information Note 1), as shown in Eq. 3:
$$\:ln\left({J}^{ph}\right)=\beta\:\text{l}\text{n}\left(\right[{h}^{+}\left]\right)-\frac{\xi\:\left[{h}^{+}\right]}{{k}_{B}T}+{\text{l}\text{n}(k}_{0})-\frac{{E}_{0}}{{k}_{B}T}$$
3
where Ea = ζ[h+] + E0 and ζ is the constant determining the change of Ea with surface hole densities. E0 is the constant to determine the Ea at surface hole density of 0 h+·nm− 2.53, 54
This relationship provides a more accurate description of the surface holes involved in the WOR. It can be seen that the appearent reaction order of surface holes is determined not only by β, but also by the variation of Ea that is related to the surface hole density. When β was fixed at 2 for the second-order kinetics and at 4 for the fourth-order kinetics, the values of ζ were calculated to be 1.8×10− 3 and − 3.8×10− 3, respectively (Figure S10). This suggests that the term ζ[h+]/kBT (0.07 and − 0.15) can be neglected under these conditions, thus making the Ea almost independent on surface hole density for both second- and fourth-order pathways. Hence, Eq. 1 can be directly employed to estimate the reaction order of surface holes under these conditions. Piccinin et al. proposed that the Ea for WOR was weakly dependent on the surface hole density and thus lead to the observed power law on the α-Fe2O3 photoanode.44 However, when the reaction order was over fourth-order and the value of β was fixed at 4, the the term ζ[h+]/kBT (− 9.34) can not be neglected and the value of ζ was calculated to be − 0.24 (Figure S10). By this theory, we could infer that the over fourth-order kinetics was still essentially fourth-order kinetics, but the Ea exhibited an obvious negative dependency on the surface hole density, resulted in that the slope of log(Jph) − log([h+]) curve can exceed 4 when the surface hole density was greater than 1.7 h+·nm− 2 (Figure S10).
2.2 Activation energy of WOR on α-Fe2O3
To obtain the apparent Ea for WOR with fourth-order kinetics, we initially examined the relationship of WOR rates with temperatures and light intensities, as shown in Fig. 2a. The WOR rate under low light intensities was more significantly affected by temperature compared to that under high light intensities. For example, as the temperature increased from 20°C to 48°C, the WOR rate increased by 40% under 100 mW·cm− 2 illumination. Comparatively, it exhibited an almost temperature-independent reaction rate (around 0.21 mA·cm− 2) for WOR under 700 mW·cm⁻2 illumination. Moreover, surface hole densities obtained by EIS and TPD methods at different temperatures remained almost unchanged under high light intensities (Figure S11). Based on the above results, we consider the temperature dependences of the second- and fourth-order kinetics of surface holes to obtain the apparent Ea for WOR according to Eq. 2, as shown in the rate law analysis plotted in Fig. 2b. The rate constants of WOR at different temperatures could be obtained from the intercepts of log(Jph) − log([h+]) curves with the y-axis. Figure 2c shows the Arrhenius plots for the second- and fourth-order rate constants. The apparent Ea for WOR with second-order kinetics was 0.16 ± 0.02 eV. More importantly, a much lower Ea of only 0.03 ± 0.01 eV was observed for WOR with fourth-order kinetics and the same results were obtained by the TPD method (Figure S11), indicating that the α-Fe2O3 photoanode was able to access a fourth-order kinetics reaction pathway to achieve a nearly barrierless WOR. According to the analysis in Supporting Information Note 1, it should be noted that the variations of Ea with surface hole density (i.e., ζ in Eq. 3) was negligible for the fourth-order kinetics pathway. Therefore, the apparent Ea obtained in Fig. 2c corresponds to the E0 in Eq. 3. Furthermore, the relationship between activation energy and surface hole density can be obtained. As shown in Fig. 2d, the slope of Ea−[h+] curve was only − 3.8×10− 3 with surface hole densities ranging from 1.0 h+·nm− 2 to 1.7 h+·nm− 2, indicating that the Ea of the fourth-order kinetics was almost independent of surface hole densities. In contrast, the slope of Ea−[h+] curve was − 0.24 at surface hole densities exceeding 1.7 h+·nm− 2, implying that the Ea of WOR could be further reduced by increasing surface hole densities. Therefore, α-Fe2O3 photoanodes can exhibit superior WOR performance at high surface hole densities. The reason for this transition under high-intensity irradiations was further investigated by kinetic isotope effect (KIE) experiments, operando spectroscopies and theoretical simulations.
H/D KIE experiments were conducted to illuminate the impact of the proton-coupled electron transfer (PCET) process on the reaction order of surface holes for WOR. As shown in Fig. 2e, a second-order kinetics of surface holes at low surface hole densities of 0.2 − 0.8 h+·nm− 2 and a third-order kinetics at high surface hole densities of 0.8 − 1.0 h+·nm− 2 were observed for D2O oxidation on the α-Fe2O3 photoanode (Fig. 2e). In addition to the difference in the reaction order, it should be noted that under the same light intensity, the surface hole density for D2O oxidation was lower than that for H2O oxidation. These results suggested that it is more difficult to accumulate four holes on the α-Fe2O3 surface in D2O solution compared to that in H2O solution. Therefore, the surface hole accumulation process has the proton-coupled characterisitcs as show in the below density functional theory (DFT) calculations. As shown in Fig. 2f, the primary KIE values ranging from 2.9 to 3.7 were observed within the entire range of light intensity illumination, indicating that O–H bond cleavage of H2O molecules was involved in the RDS of WOR. Thus, the Ea measured in this work pertained to the oxidation of H2O molecules rather than hydroxide (OH–) ions, consistent with our previous findings.26 The theoretical results of Piccinin et al. suggested that the Ea for reactions involving H2O molecules (~ 0.40 eV) was substantially higher than for reactions involving OH– (~ 0.16 eV).44 However, our results emphasize that the Ea for oxidizing H2O molecules could be very low (~ 0.03 eV) at high surface hole densities exceeding 1 h+·nm− 2.
2.3 Theoratical simulations
The WOR pathways on α-Fe2O3 with second- and fourth-order kinetics were further compared by DFT calculations. According to our experiments and previous research, the (110) facet of α-Fe2O3 saturated by water was constructed as the catalytic surface. The potential energy surfaces of WOR were shown in Fig. 3. For the second-order pathway (Fig. 3a), two FeIII−OH species (I) undergo two concerted proton-coupled electron transfer (CPET) processes, forming adjacent surface-trapped holes [FeIV=O sites (III)] with energy inputs of ~ 1.5 eV per step. Then the O − O bond formed via the nucleophilic attack of an H2O molecule to the FeIV=O sites, a potential-independent chemical step, which is usually deemed the RDS of WOR.27, 44, 55 The activation free energy for the O − O bond formation via second-order kinetics was 0.39 eV, which was more favorable than the first-order kinetics (0.54 eV) as previously reported by Durrant’s group.45 The formed FeIII−OOH intermediate (IV) can be further oxidized and finally release O2 (V). The oxidation of another incoming H2O molecule on α-Fe2O3 subsequently occurs, leading to the regeneration of the FeIII−OH species (I) and thereby completing one catalytic cycle.
For the fourth-order kinetics (Fig. 3b), there were four surface-trapped holes formed before the O − O bond formation, which can only occur under high-intensity illumination. The produced adjacent FeV=O intermediate (E) gains enough energy to enable the following oxidation reactions. In this situation, the activation free energy for O − O bond formation via the nucleophilic attack pathway is 0.31 eV, implying more facile kinetics than second-order kinetics (0.39 eV) and third-order kinetics (0.34 eV). Notably, the O − O bond could also be formed via the coupling of adjacent FeV=O intermediate (E) under high surface hole density, which exhibited a faster kinetics with Ea of 0.19 eV than the nucleophilic attack pathway. Both pathways occur under high-intensity illumination.
Based on the analysis above, the WOR kinetics on α-Fe2O3 will be accelerated with the surface hole accumulation, following the order: first-order < second-order < third-order < fourth-order (Figure S12). These results are in accordance with the experimental observation. We further dissected the structure of the transition state (TS) during the O − O bond formation via these four kinetics processes. The TS of first-order kinetics showed an H2O2-like structure [HO·---O(H) − FeIII], while the TS in the other three pathways involved an incompletely dissociated H2O molecule (Figure S12a). Specifically, the H − O−H angle of the H2O molecule in the TS of second-order kinetics was 108.9°, which was close to the value of 104.5° for the free H2O molecule in the ground state but possessed elongated O − H bonds (1.07 Å, versus 0.96 Å in water), indicating of the partly dissociated O − H bonds. The other two TSs (third- and fourth-order kinetics) boasted similar structures of the incoming H2O molecule. Compared with the TS in second-order kinetics, the H − O−H angle of the H2O decreased to 92.6° and 94.0° for third- and fourth-order pathways, respectively, illustrating a greater degree of water activation under high surface hole density. The mean length of the O − H bond in both TSs was 1.04 Å, which remained longer than that in free water but shorter than that in the TS of second-order kinetics. This observation indicates that the dissociation of the O − H bond in H2O tends to occur after the TS for O − O bond formation when the reaction order of surface holes > 2, as the activated H2O molecule has enough potential energy to break the O − H bond. In other words, the higher density of surface holes on the α-Fe2O3 photoanode results in enhanced H2O activation that subsequently promotes the kinetics of O − O bond formation, thereby accounting for the superior performance of α-Fe2O3 under high-intensity illumination.
2.4 Unraveling the Interfacial H2O Structure
The structures of interfacial H2O during WOR were probed by operando Raman measurements on α-Fe2O3 photoanodes (Fig. 4a). The broad band extending from 3000 to 3700 cm− 1 was assigned to the O − H stretching mode of H2O molecules. It can be further deconvoluted into three distinct components of O − H stretching vibration bands via Gaussian fitting, which corresponded to three types of H2O molecules at the interface.56–58 Specifically, the peak centered at the highest wavenumbers of ∼3610 cm− 1 was assigned to Na+ ion hydrated water (Na·H2O) with weak hydrogen-bond interactions. The peak at the wavenumbers of ∼3450 cm− 1 and the peak at the lowest wavenumbers of ∼3280 cm− 1 were attributed to 2-coordinated hydrogen-bonded water (2-HB·H2O) and 4-coordinated hydrogen-bonded water (4-HB·H2O), respectively. Each of these O − H associated peaks have been validated through a deuterium substitution experiment by using D2O (Figures S13 and S14). The areas of these three Raman peaks showed a slowly increasing trend in the second-order kinetics region, while they showed a rapidly increasing trend in the fourth-order kinetics regions (Figs. 4a and S15). The increase in peak intensity indicates that highly accumulated surface holes can promotes the adsorption of interfacial H2O molecules. To better describe the three types of interfacial H2O molecules on the α-Fe2O3 photoanode surface, the proportions of the three O − H stretching Raman bands for different reaction orders were compared. Figure 4b showed that the proportions of 4-HB·H2O were both more than 50% at the surface hole densities of 0.8 h+·nm− 2 and 1.7 h+·nm− 2, indicating that 4-HB·H2O dominated the interfacial H2O structure for WOR with the second-order kinetics and fourth-order kinetics. Furthermore, as the reaction order of surface holes changed from the second-order to the fourth-order, the proportion of 4-HB·H2O increased from 57–61% and the proportion of 2-HB·H2O decreased from 36–31%, while the proportions of Na·H2O remained almost unchanged (Fig. 4b). These results suggested that the dominance of 4-HB·H2O molecules became more pronounced as the increase of surface hole density, leading to a more ordered interfacial H2O structure for WOR with fourth-order kinetics. In electrochemical Raman spectroscopy, the vibrational frequency of an adsorbate species varied with electrode potential has been attributed to the vibrational Stark effect.59 In this case, it can be seen from Fig. 4c that the vibrational frequencies of these O − H stretching Raman bands showed a red shift with the increase of surface hole density, which could also be attributed to the vibrational Stark effect. At the surface hole density ranging from 0.6 to 1.8 h+·nm− 2, the Stark slopes for Na·H2O and 2-HB·H2O were ~ 4.5 and ~ 3.7, respectively. In contrast, for 4-HB·H2O, the dependency of frequency on surface hole density could be divided into two segments. The Stark slope within the surface hole density of 0.6 − 1.0 h+·nm− 2 was ~ 3.9, and it increased dramatically to ~ 37.3 within the surface hole density of 1.0–1.7 h+·nm− 2. These results showed that the stark shift of O − H stretching in 4-HB-H2O was more pronounced for fourth-order kinetics, indicating that it was highly susceptible to the impact of surface-trapped holes. This was attributed to the formation of hydrogen bonds between reaction intermediates and interfacial H2O, bringing the 4-HB·H2O closest to the photoanode surface.60 The more ordered interfacial structure of H2O molecules for fourth-order kinetics possessed a stronger hydration ability and resulted in the formation of more polarized H2O molecules. According to previous works, this structure with stronger polarized H2O molecules enabled the formation of a robust absorbate − H2O interaction network,54 which facilitates the formation of a higher-valent Fe − O species (i.e. adjacent FeV=O intermediates) through accumulating four holes on the α-Fe2O3 photoanode surface.25
We notice that previous reports on electrocatalytic system showed a lower Ea for water dissociation with decreasing degree of hydrogen bonding of H2O during the operando interfacial H2O structure evaluation using Raman spectroscopy,56, 61 which is in contrast with our results. It is noteworthy that all these electrocatalytic systems follow the Butler–Volmer theory that change the Fermi level of electrocatalysts via applied bias,18, 62 while in this work we modulate the surface hole accumulation via irradiation light intensity, which follow the population model.45 This operando interfacial H2O structure evaluation in the PEC system is reported for the first time, thus we further validate the surface-hole-induced H2O structure variation by using ab initio molecular dynamics (AIMD) simulations.
The model was constructed by using the slab model of α-Fe2O3 (110) and incorporating 69 H2O molecules, and surface holes were introduced by removing hydrogen atoms from the adsorbed H2O molecules on the surface. The results analysis was conducted based on the AIMD trajectories of 10 ps after the pre-equilibration of 3 ps. As shown in Fig. 4d and e, the presence of the surface hole resulted in the more ordered interfacial H2O molecules, exhibiting the H-atom-down configuration. The distribution of H atoms along the surface normal direction was shown in Figs. 4f and S16, where the pronounced accumulation of H atoms in the Helmholtz layer (2 − 3.5 Å to the electrode surface) was observed with the increasing of surface hole density. Besides, the average numbers of hydrogen bonds within the interface (1 − 4 Å to the electrode surface) were 6, 7, and 9 for surfaces with surface hole densities of 0, 0.8, and 2.5 h+·nm− 2 (Figure S17), respectively, which agreed with the enhanced hydrogen bonds induced by surface hole accumulation as revealed by operando Raman spectra. Additionally, the AIMD simulations also revealed the polarization of H2O molecules on the surface exhibiting a high density of surface holes (Figure S18), elucidating decreased average H − O−H angles of 105.1° and 104.5° for interfacial water molecules on surfaces with surface hole densities of 0.8, and 2.5 h+·nm− 2, respectively, compared with that of 107.0° for the H2O on the hole-free surface. These polarized H2O molecules on the surface with high hole density exhibit structures that are closer to the transition state for O − O bond formation (Figs. 3 and S12), thereby significantly promoting the WOR kinetics.
Intensity-modulated photocurrent spectroscopy (IMPS) experiments were performed to investigate the charge-transfer kinetics at the interface between α-Fe2O3 and electrolyte. The steady-state surface hole density depends on the total holes arriving on the surface, surface hole recombination, and interfacial hole transfer. Therefore, IMPS measurements were used to quantitatively describe the variations in surface hole recombination and interfacial hole transfer with surface hole densities. The IMPS spectra of α-Fe2O3 photoanodes under the illumination of different light intensities were displayed in Fig. 5a, from which the surface charge-transfer efficiency (ηct), the charge transfer rate constant (kct), and the surface recombination rate constant (krec) were extracted (Figure S19).43, 63 The extracted kct and krec as a function of surface hole density were shown in Fig. 5b. Within surface hole density of 0.4 − 1.0 h+·nm− 2 (i.e., second-order kinetics region), the values of kct slowly increased from 0.07 s− 1 to 0.16 s− 1 and the values of krec also increased from 0.72 s− 1 to 1.03 s− 1. This same trend in kct and krec with the surface hole density caused the very low value of ηct (< 15%) (Figure S19), resulted in only the accumulation of two holes on the surface to participate in the O − O bond formation step of WOR. Within surface hole density of 1.0 − 1.8 h+·nm− 2 (i.e., fourth-order kinetics region), however, the values of kct rapidly increased from 0.16 s− 1 to 0.51 s− 1 and the values of krec decreased from 1.03 s− 1 to 0.88 s− 1. As a result, this opposite trend in kct and krec caused a significant increase in the ηct, ranging from 15–36% (Figure S19). This facilitates the accumulation of multiple holes on the surface, leading to a transition from second-order kinetics to fourth-order kinetics of surface holes.