Muhammad
Qureshi
a,
Angel T.
Garcia-Esparza‡
b,
Tatsuya
Shinagawa§
a,
Philippe
Sautet
cd,
Tangui
Le Bahers
b and
Kazuhiro
Takanabe
*a
aKing Abdullah University of Science and Technology (KAUST), KAUST Catalysis Center (KCC), Physical Sciences and Engineering Division, 4700 KAUST, Thuwal, 23955-6900, Saudi Arabia
bUniv Lyon, ENS de Lyon, CNRS, Université Claude Bernard Lyon 1, Laboratoire de Chimie UMR 5182, F-69342 Lyon, France
cDepartment of Chemical and Biomolecular Engineering, University of California Los Angeles (UCLA), Los Angeles, CA 90095, USA
dDepartment of Chemistry and Biochemistry, University of California Los Angeles, Los Angeles, CA 90095, USA
First published on 29th June 2018
Photocatalysis using semiconductor powders in suspension performs reduction and oxidation reactions at nanometer-scale distances. Such short distances between the reduction (cathode) and the oxidation (anode) sites enable photocatalytic water splitting to generate H2 and O2 from pure water without a supporting electrolyte, which is otherwise impossible in conventional electrode systems due to the high solution resistance. A CrOx/Pt/SrTiO3 model photocatalyst achieves high efficiency under UV irradiation in ultra-pure water splitting at rates (>1 μmol-H2 per cm2 per h) corresponding to electrocatalysis on the order of mA cm−2. The introduction of an unbuffered supporting electrolyte did not improve the photocatalytic rates, consistent with the negligible ohmic losses (<1 mV) numerically calculated using the Poisson–Nernst–Planck equations. The Nernstian potential loss resulting from pH gradients became apparent at high photocatalytic rates (>100 mV when rate >1 μmol-H2 per cm2 per h) even when the distance between redox sites was below 10 nm. Substantial improvements in photocatalytic rates were observed when buffer ions were introduced into near-neutral pH media by not only circumventing pH gradients but introducing kinetically facile H+ reduction to H2 instead of the kinetically sluggish direct reduction of H2O to H2. Herein, the quantitative descriptions of the electric potential, concentration gradients, and catalytic performance in nanoscale water electrolysis are presented with emphasis on (1) the advantages of performing redox reactions at the nanoscale, (2) the use of electrolyte engineering at near-neutral pH as a universal and effective strategy, and (3) the effectiveness of transferring knowledge from electrocatalysis to photocatalysis, where the potential is quantitatively defined regarding the former and poorly quantified regarding the latter.
Comprehensive studies dedicated to the fundamental understanding of the photocatalytic system are essential to identify the factors limiting the performance for the following complex physicochemical processes: (1) photon absorption, (2) exciton separation, (3) charge diffusion, (4) charge transfer, (5) surface catalysis, and (6) mass transport of reactants.10,11 Successful photocatalysis requires all these processes to function simultaneously and harmoniously for both cathodic and anodic reactions, which makes the realization of such a system challenging. A promising approach is the decoupling of the physicochemical processes to study each involved step separately. In particular, when multielectron processes with sufficiently long timescales are involved in the surface catalysis, such as H2 and O2 evolution, it is effective to utilize electrochemical measurements to separately elucidate the surface redox reactions. While an electrochemical configuration can be effectively utilized to pin down the potential of photon absorber and its contact with electrode,12,13 the potential of powder photocatalysts in a suspension form is difficult to quantify because some surfaces must be charged up negatively and/or positively to achieve steady-state redox reactions.14,15 In this way, the knowledge obtained from the quantitative study of electrocatalysis can be transferred to photocatalysis, which provides key insights for the photocatalytic systems.
Recently, theoretical multiphysics models were used to quantitatively drive the design and optimization of macroscale water-splitting devices on the order of the micro- to centimeter scale,16–23 which indicated the following as limiting factors: ohmic losses, potential losses from pH gradients, mass transport, and product crossover.18,19,23–25 These findings led to the proposal of new concepts for improving system efficiency, for example, the optimization of the reactor's geometry18 and the design of dynamic flow systems for on-site gas separation.26 In contrast, nanoscale electrolysis is expected not to suffer from the aforementioned macroscale limitations.27 Theoretical models for such illuminated semiconductor particles impregnated with electrocatalysts are scarce but crucial to drive the research in the field.28–32
In this context, we consider photocatalytic water splitting as nanoscale water electrolysis driven by photocatalyst particulates. Photo- and electrocatalytic measurements combined with theoretical simulations are performed to quantify the extent to which nanoscale electrolysis is advantageous over conventional macroscale electrolysis. More specifically, potential losses associated with surface kinetics, diffusion, and migration phenomena in nanometer-sized domains are quantitatively described. The nanoelectrochemistry disclosed in this study is correlated with the development of applications in fields ranging from energy to bioanalysis.33–37
:
5 were ground up with a mortar and pestle for 25 min followed by a flux treatment at 1273 K for 10 h to generate highly crystalline SrTiO3. The powder was then washed with water until no more chloride salt was detected by the AgNO3 test. Pt nanoparticles were deposited on the surface of SrTiO3via wet impregnation for a calculated metal loading of 0.5 wt%. Na2PtCl6·6H2O (0.1 M solution) was mixed with 100 mg of SrTiO3 in 5 mL of water over a water bath until all the water evaporated. The powder was collected by filtration, calcined at 573 K for 1 h and then used for the photocatalytic reaction. CrOx was deposited on Pt/SrTiO3 by photodeposition, where 50 mg of Pt/SrTiO3 was dispersed by sonication in 100 mL of 2 mM K2CrO4 solution, which was then placed in a photoreactor and irradiated for 6 h. The error bars were calculated based on the standard deviation for three separate photocatalytic tests utilizing a pristine sample with a new solution each time. The photocatalytic reactions were conducted at 100 Torr Ar gas in a circulating batch reactor (total volume of 515 mL) equipped with an online gas chromatograph (GC: Bruker 450 GC, TCD, Ar gas, molecular sieve 13X) connected to a vacuum line. Two Xe lamps (CERMAX PE300-BF and PE300-BUV, 300 W) were used as the light source, and the irradiation wavelength was controlled with a combination of a cold mirror and a water filter (300 < λ < 800 nm). To further control the photon flux, the neutral density filters (5 × 5 cm2, 2.5 mm thick, HOYA Optics, ND-13, -25, -50, -70) were used. The spectral area of the photocatalytic reactor was 38.5 cm2, and the photon flux was measured using the AvaSpec-3648 spectrometer, an AvaLight DHS calibration light source, and a FC-UV200-2 fiber-optic cable. The AQE measurements were conducted with a 350 nm bandpass filter at different time points in the experiment.
In general, an electric field originating from the charge density may be described using Gauss's law (∇·D = ρc). In electrochemistry, we may assume a constant permittivity of the electrolyte (D = ε0εrE = −ε0εr∇ϕ). In the electrolyte, the charge density (ρc) is the summation of ionic charge carriers. The Poisson equation (eqn (1)) describes the potential (ϕ) as a function of the distribution and concentration (ci) of the ionic charge carriers in the electrolyte of constant isotropic permittivity (εr):
![]() | (1) |
![]() | (2) |
![]() | (3) |
For the kinetic contribution, the Tafel relation was utilized to represent the overpotential (η) as a function of the overall current density in the system (jT) for the HER and OER charge transfer reactions.
η = a + b log jT | (4) |
A constant Tafel slope (b) of 120 mV dec−1 was used for the kinetics of both electrocatalytic reactions.41 The HER exchange current density parameters (10−a/b) under unbuffered and buffered conditions were set at 1 × 10−2 and 3 mA cm−2, respectively, derived from the Tafel analysis of the CrOx/Pt RDE via CV in a H2 atmosphere (Fig. S8†). The OER exchange current density was obtained from the literature (2 × 10−3 mA cm−2).42 Table S1† summarizes the simulation parameters used for the numerical simulations.
With respect to the solution resistance, in a conventional macroscale electrochemical system, the splitting of ultra-pure water requires a significantly large overpotential because of the low conductivity of the solution, and thus, the use of supporting inert ions to increase the conductivity of the solution is required. In stark contrast, the photocatalytic powder system successfully achieves the splitting of ultra-pure water (18.2 MΩ cm, corresponding to 5.5 × 10−6 S m−1) as demonstrated in Fig. 2, stably generating H2 and O2 with a 2 to 1 ratio for 10 h. Furthermore, the photocatalytic performance in a 0.5 M K2SO4 solution (pH 6, unbuffered) was similar to that in ultra-pure water (Fig. 2), consistent with negligible solution resistance in the photocatalytic system.
To complement the experimental observation on the negligible ohmic loss in nanoscale electrolysis, simulations of the electrolysis of ultra-pure water (i.e., photocatalytic overall water splitting) solving the Poisson–Nernst–Planck (PNP) equations for the concentration of H+ and OH−, the flux of the ionic species, and the potential and current distributions in the solution were performed by controlling the gas evolution rates (Fig. S4, Table S1†). Fig. 3A–C describe the two-dimensional diagrams of the potential drop due to solution resistance, and Fig. 3D provides a summary of the ohmic drop for distances between 5 and 110 nm. The trend indicates a minimal potential drop (typically <1 mV) when the electrolysis of ultra-pure water occurred at distances shorter than 40 nm at rates of 1–100 μmol-H2 per cm2 per h. Ultra-pure water (σl < 10−5 S m−1) roughly exhibits a conductivity four to five orders of magnitude smaller than conductive supporting electrolytes (e.g., 0.5 M Na2SO4, σl ≈ 10−1 S m−1).46 Nevertheless, it appears that improving the conductivity of the solution does not significantly contribute to improvements in the overall efficiency. Furthermore, on larger scales (>μm), the ohmic drop could easily reach values larger than 100 mV in Milli-Q water (σl = 5.5 × 10−6 S m−1) when using simplified simulations performed to validate our results following the work of Haussener et al. (Fig. S5†),18 which emphasizes the benefit of nanoscale water electrolysis.
Using the same computational framework, the pH gradient caused by redox reactions is also addressed. Potential loss due to pH gradients causes an additional overpotential of 59 mV per pH unit (derived from the Nernst equation). The resulting pH trends from water electrolysis at a 5 nm distance are presented in Fig. 4A–C over three orders of magnitude in photocatalytic rates (1.5–150 μmol-H2 per cm2 per h). The computations showed that ultra-pure water splitting or, more generally, water splitting under unbuffered conditions at near neutral pH suffered from pH gradients close to the redox-active surfaces when the reaction occurred at higher rates (Fig. 4D). At relevant photocatalytic rates, such as achieving 10% STH efficiency or ≈153 μmol-H2 per cm2 per h (see Fig. S1†), Nernstian potential losses due to the pH gradients could easily reach 100 mV or more, even if a 5 nm distance is used. Therefore, photocatalytic water splitting under conventional near-neutral pH conditions even on the nanoscale is largely hindered by the accumulation of generated H+ and OH− at the proximity of the active sites, which adds substantial overpotential to the system.
To confirm the effects of buffer ions on pH alteration and the resultant kinetics, photocatalytic tests were performed using hole or electron scavengers to separately investigate the HER and OER half-reactions. If the buffer exclusively minimizes pH alteration, both HER and OER half-reactions should be equally improved in the presence of a buffer. If a buffer contributes to the kinetics of the reactions, a significant improvement in the performance in such half-reactions should be observed using a buffer. Fig. 6 presents the rate of gas evolution over Pt/SrTiO3 using a hole scavenger of CH3OH or an electron scavenger of IO3− in 0.5 M K-sulfate and K-phosphate solutions (see Fig. S6† for the photocatalytic activity time course). In the presence of CH3OH, a rate of 5.5 μmol-H2 per cm2 per h for the HER was achieved in sulfate, corresponding to an AQE of 48 ± 2% at 350 nm, which was significantly improved to 7.6 μmol-H2 per cm2 per h in phosphate solution, corresponding to an AQE of 67 ± 2% (Fig. 6). In contrast, such an improvement by the introduction of buffer was minimal when IO3− was used as the electron donor to study OER (1.9 vs. 2.2 μmol-O2 per cm2 per h in sulfate and phosphate solutions, respectively). Hence, it can be concluded that the buffer ions do not simply minimize the pH gradient but also improve the HER process.
The electrochemical performance of a model CrOx/Pt electrode was used to investigate the HER catalytic performance in various electrolytes. Although the model electrode system used here does not fully replicate the complicated and undefined environment of the complex photocatalytic system, the results would highlight the role of electrolyte in an extreme convection mode at near-neutral pH conditions. We performed cyclic voltammetry (CV) utilizing the rotating disk electrode (RDE) to regulate the diffusion flux in the system. Fig. 7 shows the absolute current density as a function of overpotential at pH 4. Fig. 7 shows that the unbuffered conditions showed a two-step reduction. The first reduction is attributable to proton reduction, which became diffusion limited, consistent with a macroscopic calculation of the theoretical diffusion flux of protons (Fig. S7†) and the previous literature.41,48 The second reduction process observed at a more negative potential was ascribed to the reduction of the water molecule, consistent with the constant overpotential among different pH on the standard hydrogen electrode scale (SHE), as shown in Fig. S8.†41 The HER currents in the presence of phosphate buffer (Fig. 7) show a monotonic increase with overpotential, indicating that protons were successfully supplied by the buffering ions.41 These electrocatalytic measurements confirm that the introduction of buffer species to the system improved the performance not only by mitigating the local pH alteration but also by maintaining kinetically facile proton reduction instead of the kinetically slow H2O reduction (O–H dissociation).
![]() | ||
| Fig. 7 Cyclic voltammograms in 0.5 M potassium sulfate solution (unbuffered) and 0.5 M potassium phosphate solution (buffered) using the CrOx/Pt RDE (at 50 mV s−1, 3600 rpm, with H2 bubbling, 298 K). | ||
The ion mass transport limitations depend on the current density and extent of convection in the electrochemical configuration (Fig. S7†). Likewise, the mass transport contributions in the photocatalytic experiments should be only observable when highly efficient photocatalysts are employed at high rates of reaction49,50 and/or high photon flux is used until the rate is exclusively determined by the upper limit of quantum yields. The photocatalytic rate of CrOx/Pt/SrTiO3 was correlated with the light intensity under unbuffered and buffered conditions at pH 4. The results are shown in Fig. 8. The light intensity was controlled using a combination of two different types of Xe bulbs and neutral density filters (Fig. S9† for diffuse reflectance spectra of SrTiO3 samples). Below 2 μmol-H2 per cm2 per h, similar rates were observed under unbuffered and buffered conditions (inset in Fig. 8), where the rates increased monotonically with increasing photon flux. When the photon flux was ∼100 μmol cm−2 h−1, the photocatalytic HER rate under the unbuffered conditions was found to be almost insensitive to the light intensity until ∼400 μmol cm−2 h−1, which indicates that the rate is not limited by photon-induced processes in this flux range. This observation is analogous to the potential-independent HER rate shown above under unbuffered near-neutral pH conditions caused by the mass transport limitation of H+ to the CrOx/Pt cathodic sites (Fig. 7). When the photon flux was further increased above ∼400 μmol cm−2 h−1, the photocatalytic HER rate again increased monotonically by the reactant switching from H+ to H2O. Under the buffered conditions, only a continuous monotonic increase of the HER rate was observable with increasing photon flux, consistent with the behavior of the HER with increasing overpotential in the electrochemical system and with the absence of mass transport limitation of H+.
The extent of ion mass transport purely depends on the overall rate of redox reactions. Therefore, the extrapolation of such information to other photocatalytic systems is possible. This study suggests that a significant contribution originating from the ion mass transport is expected for viable solar fuel technology (approximately 10% STH efficiency ≈ 153 μmol-H2 per cm2 per h; see Fig. S10† for a description of the theoretical potential distribution of an electrochemical system with 5 nm distance between anode and cathode). Hence, it is becoming increasingly important to fine-tune the electrolyte properties, i.e., electrolyte engineering, for further improvement in the overall rates of photocatalytic water splitting. Fundamentally, the findings indicate that an electrolyte that maximizes the flux of proton carriers enables efficient water splitting at the nanoscale, i.e., low viscosity of the solution as well as high molarity of a small-sized proton carrier, which has to be taken into account in the future development of nanoscale electrolysis. Remarkably, the results of this study can be generalized to advance nanoelectrochemical systems with wide-ranging applications in energy, nanotechnology, materials synthesis, analytical chemistry, and biotechnology. In particular, photocatalytic particulate systems for CO2 reduction and organic transformation involving protons are directly correlated with the findings obtained in this study.
Footnotes |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c8se00272j |
| ‡ Present address: Stanford Synchrotron Radiation Light Source, SLAC National Accelerator Laboratory, Menlo Park, California 94025, USA; Department of Mechanical Engineering, Stanford University, Stanford, USA. |
| § Present address: Department of Chemistry and Applied Biosciences, Institute for Chemical and Bioengineering, ETH Zurich, Vladimir-Prelog-Weg 1, CH-8093 Zurich, Switzerland. |
| This journal is © The Royal Society of Chemistry 2018 |