WO2011112620A2 - Composite photoanodes - Google Patents

Composite photoanodes Download PDF

Info

Publication number
WO2011112620A2
WO2011112620A2 PCT/US2011/027603 US2011027603W WO2011112620A2 WO 2011112620 A2 WO2011112620 A2 WO 2011112620A2 US 2011027603 W US2011027603 W US 2011027603W WO 2011112620 A2 WO2011112620 A2 WO 2011112620A2
Authority
WO
WIPO (PCT)
Prior art keywords
semiconductor
photoanode
deposition
catalyst
photoanodes
Prior art date
Application number
PCT/US2011/027603
Other languages
French (fr)
Other versions
WO2011112620A3 (en
Inventor
Daniel R. Gamelin
Diane K. Zhong
Original Assignee
University Of Washington
Priority date (The priority date is an assumption and is not a legal conclusion. Google has not performed a legal analysis and makes no representation as to the accuracy of the date listed.)
Filing date
Publication date
Application filed by University Of Washington filed Critical University Of Washington
Publication of WO2011112620A2 publication Critical patent/WO2011112620A2/en
Publication of WO2011112620A3 publication Critical patent/WO2011112620A3/en
Priority to US13/606,439 priority Critical patent/US20130240364A1/en

Links

Classifications

    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01GCAPACITORS; CAPACITORS, RECTIFIERS, DETECTORS, SWITCHING DEVICES, LIGHT-SENSITIVE OR TEMPERATURE-SENSITIVE DEVICES OF THE ELECTROLYTIC TYPE
    • H01G9/00Electrolytic capacitors, rectifiers, detectors, switching devices, light-sensitive or temperature-sensitive devices; Processes of their manufacture
    • H01G9/20Light-sensitive devices
    • CCHEMISTRY; METALLURGY
    • C25ELECTROLYTIC OR ELECTROPHORETIC PROCESSES; APPARATUS THEREFOR
    • C25DPROCESSES FOR THE ELECTROLYTIC OR ELECTROPHORETIC PRODUCTION OF COATINGS; ELECTROFORMING; APPARATUS THEREFOR
    • C25D3/00Electroplating: Baths therefor
    • C25D3/02Electroplating: Baths therefor from solutions
    • C25D3/56Electroplating: Baths therefor from solutions of alloys
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01GCAPACITORS; CAPACITORS, RECTIFIERS, DETECTORS, SWITCHING DEVICES, LIGHT-SENSITIVE OR TEMPERATURE-SENSITIVE DEVICES OF THE ELECTROLYTIC TYPE
    • H01G9/00Electrolytic capacitors, rectifiers, detectors, switching devices, light-sensitive or temperature-sensitive devices; Processes of their manufacture
    • H01G9/20Light-sensitive devices
    • H01G9/2027Light-sensitive devices comprising an oxide semiconductor electrode
    • YGENERAL TAGGING OF NEW TECHNOLOGICAL DEVELOPMENTS; GENERAL TAGGING OF CROSS-SECTIONAL TECHNOLOGIES SPANNING OVER SEVERAL SECTIONS OF THE IPC; TECHNICAL SUBJECTS COVERED BY FORMER USPC CROSS-REFERENCE ART COLLECTIONS [XRACs] AND DIGESTS
    • Y02TECHNOLOGIES OR APPLICATIONS FOR MITIGATION OR ADAPTATION AGAINST CLIMATE CHANGE
    • Y02EREDUCTION OF GREENHOUSE GAS [GHG] EMISSIONS, RELATED TO ENERGY GENERATION, TRANSMISSION OR DISTRIBUTION
    • Y02E10/00Energy generation through renewable energy sources
    • Y02E10/50Photovoltaic [PV] energy
    • Y02E10/542Dye sensitized solar cells

Definitions

  • Hematite meets many of the target photoanode requirements: It is inexpensive, oxidatively robust, environmentally benign, and it absorbs visible light (E g ⁇ 2.1 eV). Although the cc-Fe 2 03 valence band edge potential is about 1 V or higher more positive than required for Equation 1 thermodynamically, water oxidation by photogenerated valence-band holes in cc-Fe 2 03 is kinetically inefficient, and additional anodic overpotentials are typically required before significant PEC water splitting is observed. A remaining fundamental limitation of cc-Fe 2 03 is that its conduction band edge potential resides -200 mV below that required to drive the cathodic half reaction (Equation 2).
  • Tandem PEC/photovoltaic (PV) configurations have been envisioned to provide the bias needed to meet these demands.
  • Recent advances in controlled growth and doping of a-Fe 2 03 nanostructures attempt to overcome many of the limitations associated with the short hole-diffusion length (-2-4 nm), low electron mobility (-10 "1 cm 2 V “1 s “1 ), and efficient charge carrier recombination characteristics of bulk a-Fe 2 03 yielding promising PEC performance.
  • an overall solar-to-hydrogen power conversion efficiency of -2.1% has been estimated for one set of mesostructured cc-Fe 2 03 photoanodes when powered by a PV device providing 1.4 V in a tandem configuration.
  • many low-cost PV devices such as dye-sensitized solar cells or organic PVs typically provide about 1 V or lower, and two such PVs in series would thus be required to provide the necessary 1.4 V.
  • Co-Pi amorphous cobalt/phosphate catalyst
  • ITO or FTO electrodes amorphous cobalt/phosphate catalyst
  • Remaining uncertainties about the catalyst's precise microscopic identity do not diminish its attractiveness for water- splitting PECs.
  • Co-Pi requires 0.41 V overpotential at pH 7 to oxidize water with a current density of 1 mA/cm 2
  • the cc-Fe 2 03 valence band edge potential provides about 1.2 V or higher. Photogenerated holes in cc-Fe 2 03 should thus be amply capable of driving water oxidation by this electrocatalyst.
  • the composite photoanodes comprise a semiconductor and an electrocatalyst.
  • a method of forming a composite photoanode by photoelectrodeposition is provided.
  • the method comprises photoelectrodepositing a solid conformal layer of an electrocatalyst from an electrolyte solution on a surface of a semiconductor submerged in the electrolyte solution by simultaneously:
  • a method for making an electrode is provided.
  • the electrode is formed by deposition of a competent electrocatalyst onto a photoanode from an electrolyte, wherein the deposition can be carried out by photodeposition, electrochemical deposition or a combination thereof.
  • the electrolyte may comprise inorganic and organic ions, such as phosphate anion, acetate anion, sulfate anion, chloride, nitrate, sodium, potassium, or any combination thereof.
  • an electrode comprising:
  • a photoanode having a first onset potential when incorporated into a photoelectrochemical cell for electrolysis of water into oxygen
  • a layer of an electrocatalyst conformally formed on a surface of the photoanode wherein said electrocatalyst causes a cathodic shift in an onset potential of the electrode such that the electrode has a second onset potential when incorporated into a photoelectrochemical cell for electrolysis of water into oxygen, and wherein said second onset potential is less than said first onset potential.
  • FIGURES 1A-1D SEM images of mesostructured a-Fe 2 03 photoanode before (FIGURES 1A, IB), and after (FIGURES 1C, ID) 1 hour electrochemical deposition of Co-Pi catalyst.
  • the catalyst layer cracking occurs upon drying for the SEM measurement and in some cases allows inspection of the catalyst underside:
  • FIGURE ID shows that the Co-Pi underside topology conforms to the cc-Fe203 mesostructure.
  • FIGURES 2A-2C Dark (dashed) and photocurrent (solid) densities for cc-Fe203 and Co-Pi/cc-Fe203 photoanodes collected using simulated AM 1.5 illumination (1 sun, backside) at a scan rate of 50 mV/s.
  • FIGURE 2B Electronic absorption and
  • FIGURE 2C IPCE spectra for a-Fe 2 0 3 and Co-Pi/a-Fe 2 0 3 (at 1.23 and 1 V vs RHE, respectively). The absorption spectrum of Co-Pi on FTO without a-Fe 2 03 is included in FIGURE 2B, but no photocurrent was detected for these anodes.
  • FIGURE 3 Dark (dashed) and photocurrent (solid) densities for mesoscopic CC-Fe203 photoanodes used in this study under backside and frontside illumination, collected using simulated AM 1.5 sunlight (1 sun). Scan rate 50 mV/s.
  • FIGURE 4 Absorption spectrum of Co(OH)4 2 ⁇ at pH - 13 measured -30 min after preparation of the solution.
  • FIGURE 5 Summary of solar water- splitting PECs using Co-Pi/a-Fe 2 03 composite photoanodes and Pt counter electrode.
  • FIGURE 6 Co-Pi electrochemical deposition on a-Fe 2 03 photoanode in pH 7,
  • FIGURE 7 Co-Pi electrochemical deposition on a-Fe 2 03 photoanode in pH 7,
  • FIGURE 8 Photocurrent decay of a-Fe 2 03 photoanode measured in pH 7, 0.1 M
  • FIGURES 9A and 9B Dark current (dotted) and photocurrent (solid and dashed) densities of cc-Fe203 photoanodes before and after 30 min of Co-Pi deposition, measured in pH 13.6 NaOH (FIGURE 9A), and pH 8 KPi (FIGURE 9B) at 50 mV/s (thick line) and 10 mV/s (dashed line).
  • the a-Fe 2 03 data were collected at 10 mV/s.
  • the circles denote steady state photocurrent densities after 200 s of continuous illumination under 1 sun, AM 1.5 simulated sunlight.
  • FIGURES 10A and 10B The circles denote steady state photocurrent densities after 200 s of continuous illumination under 1 sun, AM 1.5 simulated sunlight.
  • FIGURE 10A Dark current (dotted) and photocurrent (solid) densities of an cc-Fe203 photoanode before (thin black) and after 30 min of Co-Pi deposition (thick lines) in 0.1 M KPi electrolyte at pH 8.
  • the cc-Fe 2 03 data black curves were collected at 10 mV/s.
  • FIGURE 10B Power dependence of photocurrent density for an cc-Fe 2 03 photoanode before (x) and after (+) Co-Pi deposition, measured at +1.0 V vs RHE.
  • FIGURE 11 0 2 generation and photocurrent density over time measured for a Co-Pi/a-Fe 2 0 3 composite photoanode at +1.0 V vs RHE in 0.1 M KPi at pH 8. Co-Pi was electrodeposited on cc-Fe203 for 30 min. The top panel showed an initial spike in the rate of 0 2 evolution before relaxation to a steady state rate.
  • FIGURE 12 Linear sweep voltammetry of Co-Pi on FTO at various scan rates in pH 7, 0.1 M KPi electrolyte. Inset: Decay of the bulk electrolysis current density over time under these conditions, measured at +1.1 V vs Ag/AgCl (+1.3 V vs NHE).
  • FIGURES 13A-13D Scanning electron micrographs of Co-Pi/a-Fe 2 03 composite photoanodes after 15 minutes of Co-Pi deposition showing (FIGURE 13 A) ring-like deposition of Co-Pi in selective areas of the cc-Fe 2 03 surface, and
  • FIGURE 13D shows Co-Pi conforming to the topology of the underlying cc-Fe 2 03 mesostructure. The cracks in the Co-Pi result from drying.
  • FIGURES 14A and 14B Dark current (dotted) and photocurrent (solid) densities of an cc-Fe 2 03 photoanode before and after 15 min of Co-Pi deposition, measured in
  • FIGURE 14A pH 13.6 NaOH
  • FIG. 14B pH 8 KPi with (black) and without 0.1 M NaCl at 50 mV/s (thin line) and 10 mV/s (thick line).
  • the cc-Fe 2 0 3 data were collected at 10 mV/s.
  • the circles denote steady state photocurrent densities after 200 s of continuous illumination under 1 sun, AM 1.5 simulated sunlight.
  • FIGURES 15A and 15B Photocurrent decay curves measured under 1 sun, AM 1.5 simulated sunlight at various applied potentials for Co-Pi/cc-Fe 2 03 composite photoanodes in pH 8 KPi, pH 8 buffered salt water, and pH 13.6 NaOH (black) electrolytes.
  • FIGURE 15A Data collected following 30 min of Co-Pi deposition
  • FIGURE 15B Data collected following 15 min of Co-Pi deposition.
  • Photocurrent decay curves measured for cc-Fe 2 03 in pH 8, 0.1 M KPi electrolyte (grey) are included in
  • FIGURE 15B for comparison.
  • FIGURES 16A-16D SEM images of Co-Pi/a-Fe 2 0 3 composite photoanode after 30 min of Co-Pi deposition.
  • FIGURES 17A and 17B Dark current (dotted) and photocurrent (solid) densities measured for an cc-Fe203 photoanode before and after 15 min of Co-Pi deposition. Data collected at 10 mV/s, for both front- and backside illumination.
  • FIGURE 17B Photocurrent density and (3 ⁇ 4 generation measured for the above photoanodes vs time: Co-Pi/cc-Fe203 (black) and cc-Fe203 (grey). The numbers in
  • FIGURE 17B indicate the photocurrent and (3 ⁇ 4 enhancement factors (see text). Bubbles adhering to and releasing from the photoanode surface cause disruptions in the current density. All PEC data were collected under 1 sun, AM 1.5 simulated solar irradiation.
  • FIGURE 18 Dark current (dotted) and photocurrent (solid) densities measured for an a-Fe 2 03 photoanode before and after Co-Pi photoelectrochemical deposition.
  • FIGURE 19 Dark current (dotted) and photocurrent (solid) densities measured for an a-Fe 2 03 photoanode before and after CoO x electrochemical deposition. Data was collected at 10 mV/s with frontside illumination under 1 sun AM 1.5 simulated sunlight. CoO x was deposited for 15 min at 50 ⁇ / ⁇ 2 and -1.1-1.3 V vs Ag/AgCl.
  • FIGURE 20 Dark current (dotted) and photocurrent (solid) densities measured for an cc-Fe203 photoanode before and after CoO x photoelectrochemical deposition.
  • FIGURE 21A-21C illustrate three potential methods for adsorbing cobalt to a photoanode.
  • FIGURE 22A-22H Wide- and narrow-angle SEM images of an unmodified
  • CC-Fe203 mesostructured photoanode and similar cc-Fe203 photoanodes following Co-Pi electrochemical deposition, photo-assisted electrochemical deposition of Co-Pi, and Co 2+ adsorption.
  • FIGURE 23A-23C illustrate ark-current (dotted) and photocurrent (solid) densities of (23A) a Co-Pi/a-Fe 2 03 electrode prepared by photo-assisted electrochemical deposition, (23B) a Co-Pi/cc-Fe203 electrode prepared by electrochemical deposition, and (23C) a Co 2+ /a-Fe 2 03 electrode prepared by surface adsorption, compared to the parent cc-Fe203 photoanodes.
  • FIGURE 24A-24C illustrate dark-current (dotted) and photocurrent (solid) densities of (24A) a Co-Pi/cc-Fe203 electrode prepared by photo-assisted electrochemical deposition, (24B) a Co-Pi/a-Fe 2 03 electrode prepared by electrochemical deposition, and (24C) a Co 2+ /cc-Fe203 electrode prepared by surface adsorption, compared to the parent a-Fe 2 03 photoanodes.
  • FIGURE 25A-25C illustrate the best (filled bars) and average (empty line) cathodic shifts, photocurrent density increases, and onset potentials for Co-Pi/cc-Fe203 photoanodes prepared by photo-assisted electrochemical deposition (P-Dep) and electrochemical deposition (E-Dep) of Co-Pi, and for Co 2+ /cc-Fe203 photoanodes prepared by surface adsorption of Co 2+ (Co-dip).
  • FIGURE 26 illustrates the average cathodic shifts plotted vs average onset potentials for Co-Pi/a-Fe203 photoanodes prepared by photo-assisted electrochemical deposition (P-Dep) and electrochemical deposition (E-dep) of Co-Pi, and for Co2+/a- Fe203 photoanodes prepared by surface adsorption of Co2+ (Co-dip) for the 12 films of Figure 5.
  • the open symbols represent the parent a-Fe203 photoanodes.
  • FIGURE 27 illustrates the Average photocurrent density increase vs photocurrent at 1.1 V vs RHE (one-sun) for the 12 films of Figure 5.
  • Photo-assisted electrochemical deposition of Co-Pi yields the largest photocurrent density increases and the highest absolute photocurrent densities.
  • the open symbols (grouped at the base of the dashed line) represent the parent a-Fe203 photoanodes.
  • FIGURE 28 illustrates the time dependence of the photocurrent density of a Co- Pi/a-Fe203 photoanode prepared by photo-assisted electrochemical deposition, measured at 1.0 V vs RHE in 1 M NaOH under continuous 1 sun, AM 1.5 simulated solar irradiation.
  • the electrolyte was not stirred.
  • the electrolyte was replaced after 75 hrs (dashed line), resulting in recovery of photocurrent density.
  • FIGURE 29 illustrates current density-voltage curves of a Ti0 2 nanowire photoanode before and after Co-Pi photoelectrochemical deposition, under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dashed) in 0.1M potassium phosphate buffer at pH 7.
  • FIGURE 30 illustrates current density-voltage curves of a Ti0 2 nanowire photoanode sensitized with CdS and coated with a thin amorphous Ti0 2 protective layer, before and after Co-Pi photoelectrochemical deposition, under 1 sun, AM 1.5 simulated solar irradiation in 0.5M sodium thiosulfate.
  • FIGURE 31 illustrates current density-voltage curves of a Co 2+ :Zno photoanode before and after Co-Pi photoelectrochemical deposition, under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dotted) in 0.1M potassium phosphate buffer at pH 11.
  • FIGURE 32 illustrates current density- voltage curves of a W doped BiV0 4 photoanode before and after Co-Pi photoelectrochemical deposition, under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dashed).
  • FIGURE 33 illustrates current density-voltage curves of an a-Fe 2 0 3 photoanode before and after cobalt methyl-phosphonate (Co-MePi) photo-assisted electrochemical deposition, under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dotted) in 1 M NaOH, pH 13.6.
  • Co-MePi cobalt methyl-phosphonate
  • FIGURE 34 illustrates current density- voltage curves of an -Fe 2 03 photoanode before and after nickel borate (Ni-Bi) electrochemical deposition, under backside illumination with 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dotted) in 1 M NaOH, pH 13.6.
  • Ni-Bi nickel borate
  • the composite photoanodes provide enhanced performance compared to known photoanodes when incorporated into photoelectrochemical (PEC) systems (e.g., a PEC cell for splitting water into hydrogen).
  • PEC photoelectrochemical
  • One particular benefit of the composite photoanodes is a cathodic shift in the onset potential of the photoanode when used in a PEC system. Such a cathodic shift allows for reduced electrical requirements to drive the PEC process, thereby increasing the efficiency of such systems.
  • the methods for forming composite photoanodes are light-enhanced deposition methods, referred to herein as photoelectrochemical deposition methods.
  • Photoelectrochemical deposition of an electrocatalyst onto a semiconductor to form a composite photoanode provides enhanced photoanode performance in PEC systems, including increased cathodic shift, compared to composite photoanodes fabricated using traditional electrochemical methods. Such improvements are attributed to the conformal nature of electrocatalyst layers formed using photoelectrodeposition.
  • a composite photoanode comprising a semiconductor having a solid conformal layer of an electrocatalyst formed on its surface.
  • the electrocatalysts referred to are competent electrocatalyst that produce a cathodic shift in the onset potential of the photoanode when used within a PEC.
  • the semiconductor acts as a photoanode.
  • the semiconductor is made of a photon-absorbing material.
  • the semiconductor is a-Fe 2 03
  • the cc-Fe203 semiconductor is meso structured.
  • the semiconductor is a high- surface-area a-Fe 2 03 photoanode. The use of hematite as a semiconductor is disclosed extensively herein, including in Examples 1-4 and 10.
  • Examples 5 and 6 disclose the use of titanium dioxide (particularly in nanowire form) as a semiconductor;
  • Example 7 discloses the use of a cobalt-ion:zinc oxide semiconductor; and
  • Example 8 discloses the use of a W:BiV04 semiconductor. These listed examples are generally inorganic in character.
  • the semiconductor comprises a group IV semiconductor having a formula selected from the group consisting of binary, ternary, and quaternary.
  • the group IV semiconductor further comprises ions selected from the group consisting of cations and anions.
  • the semiconductor is an n-type semiconductor.
  • photoanode materials useful in composite electrodes include a material selected from the group consisting of an iron oxide, a zinc oxide, a titanium oxide, a tungsten-bismuth-vanadium oxide, a tungsten oxide, a gallium-zinc-oxide-nitride, or these materials also containing additional cations or anions.
  • the semiconductor comprises a sensitizer having a sensitizer absorbance wavelength, said sensitizer absorbance being different from a semiconductor absorbance wavelength.
  • a sensitizer e.g., embedded within, or coating the surface of the semiconductor
  • Representative sensitizers include cadmium selenium, as described below in Example 6. Additional sensitizers include cationic or anionic impurities
  • the semiconductor has a physical shape selected from the group consisting of dendrites, wires, belts, rods, mesostructures, nanotubes, and thin films.
  • a high semiconductor surface area and/or a high composite photoanode surface area produces improved results for PEC reactions.
  • Nanoscopic high-surface area shapes are accordingly preferred. Therefore, in another embodiment, the physical shape has nanoscopic dimensions.
  • nanoscopic dimensions refers to a shape having at least one feature (e.g., dendrites) having a smallest size of 100 nm or smaller.
  • Semiconductors can be deposited on substrates, or otherwise formed, according to methods known to those of skill in the art, including those provided below in the Examples.
  • hematite can be grown using chemical vapor deposition (see
  • the electrocatalyst is formed on a surface (e.g., a surface that will face a light source during PEC) of the semiconductor.
  • the electrocatalyst produces a cathodic shift in the onset potential of a PEC process incorporating a composite electrode (semiconductor and electrocatalyst) when compared to the semiconductor alone. This comparison of electrodes can be found throughout the data provided herein so as to illustrate the efficacy of the disclosed materials, devices, and methods, in improving the PEC performance of semiconductors.
  • the electrocatalyst is selected from the group consisting of a cobalt-containing catalyst, an iridium-containing catalyst, a manganese-containing catalyst, a ruthenium-containing catalyst, a nickel-containing catalyst (e.g., nickel borate, see Example 10), a cobalt-containing oxygen evolving catalyst, a cobalt oxide/hydroxide catalyst, and a cobalt oxide catalyst.
  • the electrocatalyst is cobalt phosphate (Co-Pi).
  • Co-Pi is used extensively in the examples provided herein. Examples 1-3 and 5-9 describe the use of Co-Pi to improve PEC performance.
  • the layer of the electrocatalyst has a thickness of from 0.5 nm to 30 nm.
  • a thick electrocatalyst layer will inhibit performance of composite photoanodes. Accordingly, nanoscale-thick electrocatalyst films are preferred. Electrochemical deposition does not allow for quality films of such a thickness to be deposited. Accordingly, photoelectrodeposition is preferred for forming nanoscale-thick films of electrocatalyst.
  • the competent electrocatalyst examples include, without limitation, a cobalt catalyst, iridium catalyst (e.g. ) 2 ), manganese catalyst (e.g. Mn-oxo complexes), ruthenium catalyst (e.g. [Ru(L)2(OH)2] 2+ complexes, where L denotes ligand).
  • the cobalt catalyst was selected from the group consisting of cobalt based oxygen evolving catalyst and cobalt oxide catalyst (referred herein as "CoO x ", see Example 4).
  • the photoelectrochemical performance of composite the photoanodes provided herein is improved compared to "semiconductor-only" photoanodes.
  • the EXAMPLES describe these improvements extensively.
  • Co-Pi/cc-Fe 2 03 composite photoanodes for water oxidation are improved by optimization for front- side illumination in pH 8 electrolytes. Without being limited by theory, it is believed that a kinetic bottleneck appears to be related to the Co-Pi catalyst itself under these conditions. This kinetic bottleneck is overcome by more sparse deposition of Co-Pi onto cc-Fe203. Following these improvements, sustained water oxidation by Co-Pi/cc-Fe 2 03 composite photoanodes was demonstrated in both photocurrent and 0 2 evolution measurements.
  • Photoelectrochemical water oxidation by the Co-Pi/a-Fe 2 03 composite photoanodes was enhanced relative to that of cc-Fe 2 03 alone: Under these conditions, a five-fold enhancement in the photocurrent density and water oxidation rate was observed at +1.0 V vs RHE. This enhancement is even more substantial at about 1.0 V or lower vs RHE, where cc-Fe 2 03 alone does not exhibit significant photocurrent at all.
  • composite photoanodes can be anticipated, for example, by variation of the Co-Pi deposition conditions to optimize photocurrent densities at extremely low bias. More generally, these results emphasize that composite photoanode strategies offer promising prospects for sustainable, affordable, and distributed solar fuel technologies. This is equally applicable to include other catalysts such as &(3 ⁇ 4, Mn-oxo complexes, or [Ru(L)2(OH2)] 2+ complexes, that can be powered in part or entirely by light-harvesting electrodes.
  • catalysts such as &(3 ⁇ 4, Mn-oxo complexes, or [Ru(L)2(OH2)] 2+ complexes, that can be powered in part or entirely by light-harvesting electrodes.
  • the combination of the first photoenergy and the first electric bias are sufficient to deposit catalyst components from the electrolyte to form the solid conformal layer of the electrocatalyst.
  • the electrocatalyst is formed from a buffer solution in which the semiconductor is submerged.
  • the electrolyte solution can be any electrolyte solution known to those of skill in the art. Particularly, electrolyte solutions useful for traditional electrochemical deposition of an electrocatalyst onto a semiconductor are useful in the method.
  • the electrolyte solution is a buffer solution of potassium phosphate
  • the pH of the electrolyte is about 7 or higher. In one embodiment, the pH of the electrolyte was about 13 or higher. In another embodiment, the pH of the electrolyte was about 8.
  • the surface of the semiconductor is irradiated with electromagnetic radiation
  • the light having a first wavelength and a first irradiance.
  • the light can be a single wavelength or a broadband source. The only requirement is that the light provides a photoenergy sufficient to produce an electronic excited state in the semiconductor so as to provide a portion of the energy required to deposit the electrocatalyst from the electrolyte solution.
  • the electronic transition is a bandgap transition. By exciting the bandgap transition, photogenerated valence band holes can oxidize ions in the electrolyte to from an active catalyst at the surface. Because the first photoenergy is not sufficient to drive the deposition of the electrocatalyst from the electrolyte solution, a first electric bias is simultaneously applied to the semiconductor.
  • the first electric bias is significantly less than the bias required for electrochemical deposition.
  • the first electric bias is from 0.1 V to 0.4 V (e.g., versus Ag/AgCl). Therefore, the deposition (i.e., the photoelectrochemical deposition) of the electrocatalyst on the semiconductor is accomplished in the method by using energy from two sources (light and electricity) to facilitate the deposition reaction from the electrolyte. Neither of the two energy sources alone is sufficient to facilitate the deposition on their own.
  • the method according to this aspect utilizes light (e.g., sunlight or artificial sunlight) to assist in electrochemical deposition of an electrocatalyst onto a semiconductor.
  • light e.g., sunlight or artificial sunlight
  • the method is useful, for example, to fabricate a composite photoanode according to the other aspects and Examples provided herein.
  • Examples 1 and 2 below disclose composite photoanodes fabricated with the generally known technique of electrochemical deposition. These examples are contrasted, by further Examples 3-10 utilize photoelectrochemical deposition.
  • Example 3 below, provides an in-depth development of the theory and results of photoelectrodeposition. While Example 3 primarily describes composite photoanodes of Co-Pi and a-Fe 2 03, the method is not limited to these compounds. As illustrated in other Examples, photoelectrochemical deposition is compatible with any known semiconductors and electrocatalysts, particularly those used to make photoanodes using electrochemical deposition.
  • photogenerated holes can be used to oxidize an ion from an electrolyte.
  • Co 2+ can be deposited to form Co-Pi on the a-Fe 2 03 photoanode and the electron can be removed by water reduction.
  • photogenerated electrons in the conduction band of a-Fe 2 03 are below the energy needed to reduce protons to hydrogen, a very low bias is applied to assist in photoelectrochemical deposition.
  • electroactive to photoelectrochemical deposition.
  • photoelectrochemical deposition is lower than that required for electrochemical deposition of similar compounds (i.e., deposition without the assistance of light).
  • Any light source with sufficient energy to excite the band gap of the semiconductor can be used in photoelectrodeposition.
  • sunlight or artificial sunlight
  • photoelectrodeposition on a-Fe 2 03 was conducted in a three-electrode configuration from a solution of Co 2+ in potassium phosphate (KPi) buffer under 1 sun AM 1.5 simulated solar irradiation.
  • KPi potassium phosphate
  • a Pt mesh was used as the counter electrode and saturated Ag/AgCl was used as the reference electrode.
  • Typical current densities during deposition were -1-100 ⁇ /cm 2 .
  • a broad-spectrum light source e.g., sunlight
  • any light source capable of exciting the bandgap of the semiconductor is compatible with the method.
  • a single wavelength light source can be sufficient to excite the bandgap as long as
  • One impetus for the development of photoelectrochemical deposition was to develop an electrocatalyst deposition method that would allow for nanoscale-thick, conformal, continuous layers (films) of electrocatalyst to be deposited on a semiconductor. Particularly if the semiconductor is nano structured (e.g., dendritic). Traditional electrochemical deposition is insufficient in this regard. As demonstrated in the Examples (e.g., Example 3), thin, conformal electrocatalyst films are satisfactorily formed using photoelectrochemical deposition.
  • PEC reactions driven with photoanodes formed using photoelectrochemical deposition demonstrate improved absolute onset potential, cathodic shift of the onset potential, and maximum current density. All vital characteristics of, for example, PEC for water splitting, particularly in the context of solar-powered PEC devices.
  • the electrocatalyst is selected from the group consisting of a cobalt-containing catalyst, an iridium-containing catalyst, a manganese-containing catalyst, a ruthenium-containing catalyst, a nickel-containing catalyst, a cobalt-containing oxygen evolving catalyst, a cobalt oxide/hydroxide catalyst, and a cobalt oxide catalyst.
  • the electrocatalyst is cobalt phosphate.
  • the layer of the electrocatalyst has a thickness of from 0.5 nm to 30 nm.
  • the cathodic shift is from 50 mV to 400 mV.
  • the layer of the electrocatalyst has a thickness of from 0.5 nm to 30 nm. Such a thickness is indicative of the "thin" nature of the conformal electrocatalyst coating. As set forth herein, such a thin coating is essential, not only to allow light through the electrocatalyst to the semiconductor, but also due to the short charge diffusion lengths in the photoanode materials. Finally, as disclosed herein, thin films of electrocatalyst are less prone to defects (e.g., aggregates) than thicker electrocatalyst films are.
  • the range of 0.5 nm to 30 nm represents from about one molecular layer to about tens of molecular layers. Accordingly, it is preferred that a minimal number of molecular layers are used to conformally coat the semiconductor with electrocatalyst without pinhole defects (exposing the semiconductor) or aggregates (which diminish device performance).
  • the electrocatalyst is deposited from an electrolyte by photodeposition, electrochemical deposition, or combination thereof.
  • the electrode is formed by electrodepositing a conformal layer of an electrocatalyst from an electrolyte solution onto a surface of a photoanode.
  • the semiconductor is an n-type semiconductor.
  • the formed photoanode when incorporated into a photoelectrochemical cell for electrolysis of water into oxygen, reduces a water electrolysis onset voltage compared to a second photoanode comprising the semiconductor without the electrocatalyst.
  • the water electrolysis onset voltage is reduced by 50 mV to 400 mV.
  • the first wavelength of the electromagnetic radiation is from 300 nm to 800 nm.
  • the first irradiance of the electromagnetic radiation is from 0.1 W/m 2 to 1100 W/m 2 , or the equivalent in pulsed irradiation.
  • the electromagnetic radiation is selected from the group consisting of continuous radiation and pulsed radiation.
  • the first electric bias is applied to the semiconductor as part of an electrochemical deposition system comprising a power source in electrical communication with the semiconductor and a counter electrode.
  • the electrolyte solution comprises cations selected from the group consisting of cobalt, iridium, manganese, nickel, and ruthenium.
  • the electrolyte solution comprises anions selected from the group consisting of phosphate, methyl phosphonate, borate, acetate, sulfate, and hydroxide.
  • the semiconductor comprises a sensitizer having a sensitizer absorbance wavelength, said sensitizer absorbance being different from a semiconductor absorbance wavelength.
  • the semiconductor comprises a group IV semiconductor having a formula selected from the group consisting of binary, ternary, and quaternary.
  • the group IV semiconductor further comprises ions selected from the group consisting of cations and anions.
  • the semiconductor comprises a material selected from the group consisting of an iron oxide, a zinc oxide, a titanium oxide, a tungsten-bismuth- vanadium oxide, a tungsten oxide, a gallium-zinc-oxide-nitride, or these materials also containing additional cations or anions.
  • the combination of the first photoenergy and the first electric bias are sufficient to oxidize cations to deposit catalyst components from the electrolyte to form the solid conformal layer of the electrocatalyst.
  • an electrode was made by deposition of cobalt catalyst onto mesostructured cc-Fe 2 03 from an electrolyte of Co 2+ .
  • the deposition can be carried out by photodeposition or electrochemical deposition.
  • the electrolyte include, without limitation, cobalt phosphate, cobalt borate, cobalt methyl phosphonate, cobalt nitrate, cobalt acetate, cobalt sulfate, and any combination thereof.
  • the pH of the electrolyte is about 7 or higher. In one embodiment, the pH of the electrolyte was about 13 or higher. In another embodiment, the pH of the electrolyte was about 8.
  • an electrode was made by electrochemical deposition of cobalt/phosphate catalyst ("Co-Pi") onto mesostructured cc-Fe 2 03 and showed about
  • Co-Pi was electrodeposited onto a mesostructured a-Fe 2 03 photoanode.
  • the photoelectrochemical properties of the resulting composite photoanodes were optimized for solar water oxidation under front- side illumination in pH 8 electrolytes. Relative to cc-Fe 2 03 photoanodes, more sparse deposition of Co-Pi onto the cc-Fe 2 03 resulted in a sustained five-fold enhancement in the photocurrent density and 0 2 evolution rate at +1.0 V vs RHE.
  • the photoanode comprises a photoanode material selected from the group consisting of an iron oxide, a zinc oxide, a titanium oxide, a bismuth vanadium oxide.
  • the photoanode material has a physical shape selected from the group consisting of dendrites, wires, and belts.
  • said physical shape has nanoscopic dimensions.
  • the term nanoscopic dimensions refers to a shape having at least one feature (e.g., dendrites) having a smallest size of 100 nm or smaller.
  • the photoanode comprises hematite iron oxide dendrites.
  • the photoanode consists of hematite iron oxide dendrites conformally covered with a layer of cobalt phosphate.
  • the electrocatalyst is selected from the group consisting of a cobalt-containing catalyst, an iridium-containing catalyst, a manganese-containing catalyst, a ruthenium-containing catalyst, a cobalt-containing oxygen evolving catalyst and a cobalt oxide catalyst.
  • the electrocatalyst is cobalt phosphate.
  • the cathodic shift is from 50 mV to 400 mV.
  • the layer of the electrocatalyst has a thickness of from 0.5 nm to 30 nm. Such a thickness is indicative of the "thin" nature of the conformal electrocatalyst coating. As set forth herein, such a thin coating is essential, not only to allow light through the electrocatalyst to the semiconductor, but also due to exacerbated electron-hole recombination with thicker catalyst films.
  • the range of 0.5 nm to 30 nm represents from about one molecular layer to about tens of molecular layers. Accordingly, it is preferred that a minimal number of molecular layers are used to conformally coat the semiconductor with electrocatalyst without pinhole defects (exposing the semiconductor) or aggregates (which diminish device performance).
  • the electrocatalyst is deposited from an electrolyte by photodeposition, electrochemical deposition, or combination thereof.
  • the electrode is formed by electrodepositing a conformal layer of an electrocatalyst from an electrolyte solution onto a surface of a photoanode.
  • the formed photoanode when incorporated into a photoelectrochemical cell for electrolysis of water into oxygen, reduces a water electrolysis onset voltage compared to a second photoanode comprising the semiconductor without the electrocatalyst.
  • an electrode comprising: a photoanode;
  • an electrode comprising:
  • a competent electrocatalyst selected from the group consisting of cobalt catalyst, iridium catalyst, manganese catalyst, ruthenium catalyst, cobalt based oxygen evolving catalyst and cobalt oxide catalyst.
  • an electrode comprising:
  • a competent electrocatalyst comprising a cobalt catalyst deposited onto the a-Fe 2 0 3 photoanode from an electrolyte of Co 2+ by photodeposition, electrochemical deposition, or combination thereof,
  • the electrolyte comprises a composition selected from the group consisting of cobalt phosphate, cobalt nitrate, cobalt acetate, cobalt sulfate, and any combination thereof; and the electrode having about a several hundred millivolt cathodic shift of the onset potential for PEC water oxidation.
  • a system/device for converting water to hydrogen using only sunlight as an energy source includes a PEC comprising a photoanode formed using photoelectrochemical deposition and a photovoltaic cell.
  • a water- splitting PEC typically requires over 1 V to produce hydrogen and oxygen from water, which is an electrical requirement that cannot be met by present PV technology.
  • the cathodic shift achieved in improving present photoanodes for PEC e.g., Co- Pi/hematite
  • the Co-Pi catalyst was electrodepo sited onto the oxide anodes as known in the art.
  • the anode was submerged in a buffer solution of 0.1 M potassium phosphate (pH 7) containing 0.5 mM Co(N03)2 and a bias of 1.29 V (vs. NHE) was applied for 1 hr.
  • Composite Co-Pi/cc-Fe 2 03 anodes for which the mask was not used showed greater dark currents from the Co-Pi catalyst, but were otherwise very similar.
  • Electrochemical measurements were performed in a 3-electrode configuration using an aqueous hydroxide electrolyte (1 M NaOH, pH 13.6), a Pt counter electrode, and an Ag/AgCl reference electrode.
  • a titanium clasp was used to make contact with the upper 25% of the 5 cm long anode, where no cc-Fe203 had been deposited. The bottom -50% of the anode was submerged in the electrolyte solution in a home-built optical cell.
  • IPCE measurements were performed using a Xe arc lamp with an Oriel Cornerstone 74000 monochromator with slits set to -10 nm spectral bandwidth at the designated bias voltage provided by the potentiostat. The wavelength was scanned at 1 nm/s. Photon power densities were determined using a calibrated Si photodiode. Dark current measurements probe the entire submerged FTO + CC-Fe 2 03 (or Co-Pi/cc-Fe 2 03) surface, whereas photocurrents represent the response achieved from just the irradiated area normalized to 1 cm 2 . This area was circular with a diameter of 6 mm.
  • Typical monochromatic photon power densities in the IPCE measurements were ⁇ 0.50W/m 2 .
  • the a-Fe 2 0 3 photoanode data were collected first, then the Co-Pi catalyst was deposited onto the same cc-Fe 2 03 photoanode, and then the parallel data were collected on the Co-Pi/cc-Fe203 photoanode.
  • FIGURE 2A shows dark (dashed) and photocurrent (solid) densities for an CC-Fe203 photoanode with backside illumination. Whereas the dark response was negligible up to 1.5 V vs RHE, the photoresponse showed a rise and plateau with an onset voltage of ⁇ 1 V vs RHE that typifies cc-Fe 2 0 3 .
  • FIGURES 1C and ID showed SEM images of a representative cc-Fe203 photoanode following Co-Pi electrochemical deposition for 1 hr as known in the art. Extensive cracking of the -200 nm thick catalyst layer occurred upon drying for the SEM measurement.
  • FIGURE ID showed a portion of the catalyst layer that curled off of the cc-Fe 2 03 film upon drying, revealing its underside.
  • FIGURE 2A also showed the dark and photocurrent responses of the Co-Pi cc-Fe 2 03 composite photoanode prepared by electrochemical deposition of the Co-Pi catalyst on the same cc-Fe 2 03 photoanode.
  • cc-Fe 2 03 photocurrent densities with front-side illumination were approximately 2x greater than with backside illumination (FIGURE 3), without being bound by any theory, a common observation attributable to the greater surface area of the anode front.
  • front-side illumination did not greatly enhance the photocurrent, without being bound by any theory, likely because of non-productive absorption by the catalyst layer.
  • IPCE measurements of the cc-Fe 2 03 (1.23 V vs RHE) and Co-Pi/cc-Fe 2 03 (I V vs RHE) photoanodes using backside illumination showed essentially identical dispersion (FIGURE 2C), without being bound by any theory, in both cases deriving only from cc-Fe 2 03 excitation. Co-Pi thus behaved solely as a surface electrocatalyst.
  • the composite photoanode of FIGURE 2C showed IPCE about 15% or higher at 550 nm and 1 V vs RHE, conditions where cc-Fe 2 03 alone showed negligible photocurrent (FIGURE 2a). This IPCE maximized at 450 nm (18%) before decreasing again below -400 nm because of the decreasing light penetration depth (FIGURE 2B).
  • Co(OH) 4 2" was prepared by dissolving cobalt nitrate in a 50 wt% concentrated NaOH aqueous solution to make a -0.005M Co(OH)4 2 ⁇ solution, which was then added to distilled water to reach pH - 13. The final solution was added dropwise to the electrolyte of the PEC cell under operating conditions, where its influence on dark and photocurrent densities of various photoanodes could be monitored.
  • the conformal catalyst deposition facilitates interfacial hole transfer from cc-Fe203 to Co-Pi, allowing photon absorption and redox catalysis to be effectively decoupled while retaining photocurrent densities.
  • Efficient hole transfer from cc-Fe203 to Co-Pi should enhance the electron gradient in the cc-Fe 2 03 mesostructure under irradiation, also contributing to the driving force for electron diffusion to the FTO and reducing deleterious carrier recombination processes.
  • Catalyst electrochemical deposition onto a-Fe 2 03 may also passivate surface defects.
  • FIGURE 5 The experimental results for the Co-Pi/cc-Fe203 composite photoanodes may be summarized in FIGURE 5 (and this model can be used to describe the behavior of all composite photoanodes provided in the disclosed aspects and embodiments herein).
  • photoexcitation of cc-Fe 2 03 generates an electron-hole pair.
  • Photogenerated holes are trapped by the Co-Pi catalyst, which excels at water oxidation. Photogenerated electrons migrate to the FTO back contact and pass through the circuit to the Pt counter electrode, where water reduction occurs in the 3-electrode configuration.
  • Si doped cc-Fe 2 03 photoanodes were fabricated on fluorine doped tin oxide (FTO) glass (50 x 13 x 2.3 mm TEC 15 Hartford Glass Co.) at 470°C for 5 min by atmospheric pressure chemical vapor deposition (APCVD) following procedures known in the art.
  • the a-Fe 2 03 films investigated here were typically -400-500 nm thick.
  • electrical tape with an aperture that matched the irradiated area during photoelectrochemical (PEC) experiments (0 6 mm diameter) was applied onto the cc-Fe 2 03.
  • a-Fe203 was submerged into a solution of 0.5 mM cobalt nitrate in 0.1 M pH 7 potassium phosphate (KPi) buffer.
  • KPi potassium phosphate
  • a Pt mesh was used as the counter electrode and saturated Ag/AgCl was used as the reference electrode.
  • Co-Pi was electrodeposited at +1.1 V vs Ag/AgCl for 15 (FIGURE 6) or 30 (FIGURE 7) min. Typical current densities during deposition were -20-200 ⁇ /cm 2 (FIGURES 6 and 7)
  • Co-Pi was electrodeposited for 15 min using the above conditions and a mask of 1 cm x 1 cm.
  • Oxygen detection The detection of 0 2 was performed using a YSI 5000 dissolved oxygen meter equipped with a YSI 5010 self-stirring Clark-type probe in a three-neck flask with an optical window. Before use, the electrolyte (0.1 M KPi buffered at pH 8) was degassed and purged with argon gas. Measurements were conducted in argon in the same three-electrode configuration described for PEC experiments using the same light source. Again, the photoanodes were masked to illuminate a circular area of 6 mm in diameter. Consecutive measurements were taken at +1.0, 1.1, and 1.23 V vs RHE for two hours at each potential. While the light was off between voltages (-160 seconds), there was no increase and sometimes even a decrease in the 0 2 level due to consumption by the Clark electrode.
  • Co-Pi/cc-Fe203 photoanode performed under front-side illumination and mild pH conditions.
  • FIGURES 9A and 9B show current-voltage (J-V) curves collected for Co-Pi/cc-Fe203 composite photoanodes prepared with 30 min deposition of Co-Pi and measured in various electrolytes. Each data set was compared to analogous data collected for the same Fe 2 03 film measured before Co-Pi deposition.
  • FIGURE 9A shows the J-V curves collected using 1 M NaOH at pH 13.6 and
  • FIGURE 9B shows data collected using 0.1 M KPi at pH 8.
  • Co-Pi deposition yields a cathodic shift of about 350 mV or higher in the photocurrent onset potential relative to CC-Fe 2 0 3 (FIGURE 9 A).
  • FIGURES 9A and 9B also showed J-V curves of the same Co-Pi/cc-Fe 2 03 composite photoanodes measured at the slower scan rate of 10 mV/s (dashed line).
  • the open circles in FIGURES 9 A and 9B were the quasi- steady state photocurrent densities measured after 200 s of simulated solar irradiation at each applied potential.
  • FIGURES 10A and 10B The resulting data are summarized in FIGURES 10A and 10B. Expanding on FIGURES 9 A and 9B, photocurrents were measured at a greater variety of scan rates to show the evolving characteristics of the J-V curves. With faster scan rates, the first maximum shifted to higher potentials. The inset showed the photocurrent response vs time upon unblocking the light path, measured at +1.1 V vs RHE. A large initial spike in photocurrent upon illumination was followed by multi-exponential decay to a lower steady-state current density with an effective time constant on the order of 10 sec, i.e., comparable to the data collection timescale (10s of seconds).
  • FIGURE 10B plots the steady-state photocurrent density for a Co-Pi/cc-Fe 2 03 composite photoanode measured as a function of illumination power density between 0 and 1 sun. There was a marked saturation in the photocurrent as the light intensity was increased.
  • FIGURE 12 showed the J-V characteristics of Co-Pi at various scan rates, and the current density time dependence under typical electrolysis conditions of +1.1 V vs Ag/AgCl.
  • the bulk electrolysis by Co-Pi on FTO also showed a scan rate dependence and a decay in the current density in the region where water oxidation was normally observed, +1.3 V vs NHE, or +1.7 V vs RHE.
  • FIGURE 14B PEC measurements were also performed in 0.1 M NaCl buffered to pH 8 with KPi. The resulting J-V curves with NaCl added are essentially indistinguishable from those without NaCl (FIGURE 14B), demonstrating that the present of chloride did not interfere with PEC water oxidation with Co-Pi/cc-Fe 2 03 composite photoanodes.
  • FIGURES 15A and 15B compared the kinetic responses of Co-Pi/cc-Fe 2 03 photoanodes with thicker (FIGURE 15 A, 30 min deposition, see FIGURES 16A-16D) and thinner (FIGURE 15B, 15 min deposition, see FIGURES 13A-13D) Co-Pi coverage.
  • the photocurrent decay curves of FIGURE 15A showed a large initial spike in current density, followed by a multiexponential decrease with ⁇ - 10 sec to a small steady- state photocurrent density close to that of the underlying cc-Fe 2 03 photoanode.
  • the photoanodes with thinner Co-Pi coverage showed substantially more stable performance.
  • FIGURE 15B The steady-state photocurrent densities in FIGURE 15B were enhanced relative to those of the parent cc-Fe 2 03 photoanodes. The results showed a sustainable photocurrent density that was enhanced relative to cc-Fe203 by more than an order of magnitude at 0.83 V, where cc-Fe 2 03 alone did not exhibit significant photocurrent (FIGURE 15B).
  • FIGURES 14A, 14B, 15A, and 15B demonstrated that reduced Co-Pi deposition onto CC-Fe203 photoanodes circumvented the major kinetic limitations identified above, while still shifting the onset potential of cc-Fe 2 03 by -180 mV, and simultaneously facilitated front-side illumination for maximum photocurrent densities.
  • FIGURES 10A and 10B Decreased deposition of Co-Pi onto cc-Fe 2 03 largely overcame the kinetic limitations described in FIGURES 10A and 10B, but there was still some evidence of such kinetic effects in KPi electrolyte (FIGURES 14B, 15B) that were not observed in 1 M NaOH. For 1 M NaOH, there was a small initial spike in the photocurrent followed by a small gradual increase to steady state. Without being bound by any theory, it is possible that limited mobility of protons through the amorphous catalyst may contribute to the kinetic bottleneck described by FIGURES 9 A, 9B, 10A, 10B, 12A, and 12B, and that OH " is better able to overcome this limitation. Overall, FIGURES 14A, 14B, 15A, and 15B showed that this bottleneck was lessened by changing the electrolyte from pH 8
  • Oxygen evolution In addition to current density measurements, PEC 0 2 evolution by the Co-Pi/cc-Fe203 composite photoanodes was also examined. Oxygen evolution was measured at various applied potentials before and after 15 min of Co-Pi electrochemical deposition onto an cc-Fe203 photoanode. Measurements were performed in 0.1 M KPi electrolyte at pH 8. FIGURE 17 A showed the J-V characteristics of the CC-Fe203 photoanode used for these measurements, before and after Co-Pi deposition, and for both front and backside illumination. Photocurrent densities increased substantially with front-side illumination, particularly at low potentials.
  • FIGURE 17B plotted the photocurrent density vs time along with the 0 2 concentrations measured simultaneously using the Clark-type electrode. Sustained photocurrent was observed for the Co-Pi/cc-Fe 2 03 composite photoanode over the course of this ⁇ 6 hour experiment. This steady- state photocurrent was enhanced over that of the parent cc-Fe 2 03 film, even after several hours of illumination, and was accompanied by a correspondingly large enhancement in the 0 2 evolution rate.
  • each applied potential were indicated in FIGURE 17B.
  • the amount of dissolved 0 2 detected by the Clark-type electrode was lower than the theoretical maximum for the measured current densities, but without being bound by any theory, this difference may be attributable to the adherence of bubbles on the rough surface of the Co-Pi/cc-Fe 2 03 photoanode. Occasional jumps in the photocurrent density were observed for the composite photoanodes and may be related to release of these bubbles.
  • FIGURE 17B showed that PEC 0 2 evolution by the Co-Pi/cc-Fe 2 0 3 composite photoanode was enhanced over that of the same cc-Fe 2 03 photoanode without Co-Pi.
  • Example 3 Co-Pi/ 0C-Fe 2 O3 photoanodes prepared by photoelectrochemical deposition.
  • Co-Pi catalyst was photoelectrochemical deposited onto cc-Fe 2 03 photoanodes by using light and an external applied bias to deposit Co-Pi.
  • photogenerated holes can be used to oxidize Co 2+ from the electrolyte to form Co-Pi on the cc-Fe 2 03 photoanode and the electron can be removed by water reduction. Because photogenerated electrons in the conduction band of cc-Fe 2 03 are below the energy needed to reduce protons to hydrogen, a very low bias was applied to assist in photoelectrochemical deposition of Co-Pi.
  • any light source with sufficient energy to excite the band gap of cc-Fe 2 03 can be used in a photoelectrochemical deposition on cc-Fe 2 03.
  • photoelectrochemical deposition on CC-Fe 2 03 was conducted in a three-electrode configuration from a solution of Co 2+ in potassium phosphate (KPi) buffer under 1 sun AM 1.5 simulated solar irradiation.
  • KPi potassium phosphate
  • a Pt mesh was used as the counter electrode and saturated Ag/AgCl was used as the reference electrode.
  • Typical current densities during deposition were -1-100 ⁇ /cm 2 .
  • FIGURE 18 shows a -120 mV cathodic shift in the J-V curve of an a-Fe 2 03 photoanode after Co-Pi deposition measured in 0.1 M KPi at pH 8. Photoelectrochemical deposition and electrochemical deposition, of Co-Pi had similar effect of shifting the onset potential for water oxidation of a-Fe 2 03.
  • a photo-assisted electrochemical deposition approach i.e., photoelectrodeposition
  • a photoelectrodeposition approach was used to deposit a cobalt-phosphate water oxidation catalyst ("Co-Pi") onto dendritic mesostructures of cc-Fe203.
  • Co-Pi cobalt-phosphate water oxidation catalyst
  • a comparison between this approach, electrochemical deposition of Co-Pi, and Co 2+ wet impregnation showed that photo-assisted electrochemical deposition of Co-Pi yields superior cc-Fe203 photoanodes for photoelectrochemical water oxidation.
  • Stable photocurrent densities of 1.0 mA/cm2 at 1.0 V and 2.8 mA/cm 2 at 1.23 V vs RHE measured under standard illumination and basic conditions were achieved.
  • photo-assisted electrochemical deposition provides a more uniform distribution of Co-Pi onto a-Fe 2 03 than obtained by electrochemical deposition.
  • This approach of fabricating catalyst-modified metal-oxide photoelectrodes may be attractive for optimization in conjunction with tandem or hybrid photoelectrochemical cells.
  • Hematite (a-Fe 2 03) has emerged as a prototype photoanode for PEC water oxidation because of its balance of visible light absorption (bandgap of 2.1 eV), chemical stability, low cost, and large positive valence band edge potential. Low mobilities
  • Electrochemical deposition of Co-Pi forms an adequate junction between the catalyst and semiconductor for interfacial charge transfer, and the resulting Co-Pi/cc-Fe 2 03 composite photoanodes are stable under photolysis conditions.
  • a kinetic bottleneck was observed with thick layers of Co-Pi that hindered the steady-state turnover of the composite photoanodes, especially at low applied potentials. This kinetic limitation was remediated by reducing the Co-Pi coverage, but at the expense of overpotential.
  • Co-Pi was found to deposit preferentially at pinholes, scratches, or other imperfections in the cc-Fe 2 03 film, where more current can flow from the underlying conductive FTO substrate. This inhomogeneity affects the performance of Co-Pi/cc-Fe 2 03 photoanodes by creating areas where the catalyst layer is too thick (kinetic bottleneck), and it influences the reproducibility of the Co-Pi deposition itself.
  • a stable and efficient water oxidation photoanode is desired, and methods to apply a uniform thin catalyst layer onto highly mesostructured metal- oxide photoanodes, such as cc-Fe 2 03 are therefore needed.
  • photo-assisted electrochemical deposition (“photoelectrodeposition”) of Co-Pi onto mesostructured cc-Fe 2 03 photoanodes, and present a comparison between this approach, electrochemical deposition of Co-Pi and Co 2+ adsorption.
  • photoelectrodeposition photo-assisted electrochemical deposition
  • FIGURE 21A- FIGURE 21C photo-assisted electrochemical deposition of Co-Pi is found to yield superior PEC performance by all metrics, including absolute onset potential, cathodic shift of the onset potential, and maximum current density.
  • Co-Pi was electrodeposited onto cc-Fe 2 03 photoanodes by modification of published procedures.
  • a three-electrode cell was used with a-Fe 2 03 as the working electrode, Ag/AgCl as the reference electrode, and Pt mesh as the counter electrode.
  • 0.9 V vs Ag/AgCl was applied in a solution of 0.5 mM cobalt nitrate in 0.1 M potassium phosphate buffer at pH 7.
  • the amount of Co-Pi deposited was controlled by the deposition time, which ranged between 200-500s. Current densities were typically -2-10 ⁇ /cm 2 during deposition.
  • Photo-assisted electrochemical deposition of Co-Pi onto mesostructured cc-Fe 2 03 was performed from the same electrolyte composition used for electrochemical deposition, 0.5 mM cobalt nitrate in 0.1 M potassium phosphate buffer at pH 7, but with 1 sun AM 1.5 simulated sunlight illumination. Because conduction-band electrons in CC-Fe 2 03 do not have sufficient potential to reduce water, an external bias (-0.1-0.4 V) was applied. The amount of Co-Pi was again controlled by the deposition time, which ranged between 500-750 s. Current densities were typically -2-5 ⁇ /cm 2 during deposition.
  • Co 2+ adsorption onto mesostructured cc-Fe 2 03 photoanodes was achieved by dipping the photoanode in a solution of 0.1 M cobalt nitrate for 5 minutes.
  • the amount of Co 2+ adsorbed was optimized by repetition of this dipping process.
  • PEC enhancement reached its maximum after about three cycles. Subsequent cycles resulted in either no change or a decrease in the PEC performance.
  • PEC measurements were conducted in 1M NaOH (pH 13.6) using a three-electrode configuration, with the photoanode as the working electrode, Ag/AgCl as the reference electrode, and Pt as the counter electrode.
  • Photocurrent densities were measured with front-side illumination under 1 sun AM 1.5 simulated sunlight using an Oriel 96000 solar simulator equipped with a 150 W Xenon arc lamp and an Oriel AM 1.5 filter.
  • the amount of catalyst applied was optimized to give the largest sustainable cathodic shift and overall current density by controlling the amount of catalyst loading, either by adjusting the time of deposition for Co-Pi or the number of cobalt dipping cycles for Co 2+ adsorption.
  • Cathodic shifts were calculated as the average voltage shifts in the window where current densities range from 0.5-1.5 ⁇ /cm 2 .
  • reported photocurrent increases with catalyst deposition refer specifically to the difference in photocurrent at 1.1 V vs RHE.
  • Photocurrent onset potentials were calculated by extrapolation to zero current from the linear portion of the J-V curve where current densities range from 0.5-1.5 mA/cm 2 .
  • SEM Scanning electron microscopy
  • EDX energy dispersive X-ray
  • FIGURE 22A and 22B SEM images of a representative mesostructured cc-Fe203 photoanode are shown in FIGURE 22A and 22B.
  • the photoanode possesses the dendritic features typical of CC-Fe 2 03 grown by APCVD.
  • All catalyst-modified photoanodes show similar dendritic features but the images are slightly blurred (FIGURE 22 C-H), suggesting that the catalysts make the surfaces more insulating and hence more susceptible to charging effects from the electron beam.
  • All photoanode surfaces appear uniform except for the one involving electrodeposited Co-Pi (FIGURE 22 C,D), which shows patches of Co-Pi.
  • the catalyst itself is not resolved by the SEM measurement, but it can be detected by EDX.
  • EDX measurements on large and small areas of the films from FIGURE 22 E-H yield similar results, indicating uniform cobalt coverage on these length scales (TABLE 1).
  • PEC enhancement can be roughly estimated using the EDX results.
  • increasing the probe depth by increasing the electron acceleration voltage from 10 to 15 keV results in a substantial decrease in the relative cobalt peak intensity.
  • Approximating the probe depth of a 10 keV electron beam to be ⁇ 200 nm the assumption of a uniform flat surface would yield a Co-Pi thickness of -30 nm, but this value represents an upper limit because of the very high surface roughness of the a-Fe 2 03 mesostructure (roughness ⁇ 20).
  • the active Co-Pi cluster is believed to possess seven cobalt ions, with a volume of -700 A, from which an upper limit of 34 clusters thickness is obtained. In all likelihood, the actual thickness is substantially smaller.
  • FIGURE 23A-C compares current-voltage (J-V) characteristics of representative
  • Co-Pi/cc-Fe 2 03 and Co 2+ -modified cc-Fe 2 03 photoelectrodes All photoelectrodes have been optimized to give the largest steady-state cathodic shift and PEC enhancement compared to their parent cc-Fe 2 03 photoanodes.
  • Photo-assisted electrochemical deposition of Co-Pi onto cc-Fe 2 03 yields the greatest cathodic shift of the onset potential for PEC water oxidation, -170 mV. Similar results were described previously for electrochemical deposition of Co-Pi onto cc-Fe 2 03 following optimization.
  • FIGURE 28 illustrates the time dependence of the photocurrent density of a Co-Pi/a-Fe203 photoanode prepared by photo-assisted electrochemical deposition, measured at 1.0 V vs RHE in 1 M NaOH under continuous 1 sun, AM 1.5 simulated solar irradiation.
  • the electrolyte was not stirred.
  • the electrolyte was replaced after 75 hrs (dashed line), resulting in recovery of photocurrent density.
  • FIGURE 24A-24C shows the photocurrent responses of two quite different composite photoanodes in comparison with those of is their parent a-Fe 2 03 photoanodes.
  • the photoanode in FIGURE 24A shows large, stable photocurrent densities at high bias, whereas the one in FIGURE 24B excels at low bias. These differences are due to a small variation in the deposition temperature.
  • Co-Pi surface deposition has a similar effect on each parent cc-Fe 2 03 photoanode, despite their absolute performance differences. Both films show comparable cathodic shifts of their photocurrent onset potentials and small enhancements of their maximum photocurrent densities upon deposition of Co-Pi.
  • IPCE Incident-photon-to-current conversion efficiency
  • FIGURES 25A-25C summarizes the PEC results obtained from the investigation of a total of 12 catalyst-modified photoanodes, with particular care given to ensuring that they all involved very similar parent a-Fe 2 03 photoanodes as their starting points.
  • FIGURE 25B and absolute photocurrent onset potentials (FIGURE 25C) of the best photoanodes in each category are plotted as a bar graph.
  • the average performance in each category is indicated by a horizontal line in the top two graphs and by an empty bar in the bottom graph. Plotting one metric vs another confirms the linear relationship between cathodic shift and reduced photocurrent onset potential (FIGURE 26).
  • FIGURE 26 illustrates the average cathodic shifts plotted vs average onset potentials for Co-Pi/a-Fe203 photoanodes prepared by photo-assisted electrochemical deposition (P-Dep) and electrochemical deposition (E-dep) of Co-Pi, and for Co2+/a-Fe203 photoanodes prepared by surface adsorption of Co2+ (Co-dip) for the films used to generate the data of FIGURES 25A and 25C.
  • the open symbols represent the parent a-Fe203 photoanodes.
  • FIGURE 27 illustrates the Average photocurrent density increase vs photocurrent at 1.1 V vs RHE (one-sun) for the films used to generate the data of FIGURES 25B and 25C.
  • Photo-assisted electrochemical deposition of Co-Pi yields the largest photocurrent density increases and the highest absolute photocurrent densities.
  • the open symbols (grouped at the base of the dashed line) represent the parent a-Fe203 photoanodes.
  • Co-Pi/cc-Fe203 photoanodes are comparable with those of Ir02/oc-Fe203 photoanodes prepared by attachment of nanocrystals of the well-known water oxidation catalyst, IrC ⁇ , onto similar a-Fe 2 03 photoanodes.
  • IrC ⁇ water oxidation catalyst
  • the best Ir02/oc-Fe203 photoanode showed a 50 mV greater cathodic shift, a 60 mV lower onset potential, and a -13% larger photocurrent density at 1.23 V vs RHE.
  • Co-Pi/a-Fe 2 03 and Ir02/oc-Fe203 photoanodes are important differences between Co-Pi/a-Fe 2 03 and Ir02/oc-Fe203 photoanodes, however, is that the photocurrent responses of the Ir02/oc-Fe203 photoanodes appear to diminish on short (200 s) timescales because of detachment of the IrC ⁇ particles from the a-Fe 2 03 surface.
  • the Co-Pi/cc-Fe203 composite photoanodes show no similar instability (see supporting information).
  • the Co-Pi/a-Fe 2 03 photoanode of FIGURE 24B shows a relatively high photocurrent density of 1.0 mA/cm 2 at 1.0 V vs RHE, which constitutes a 500% improvement over a-Fe 2 03 alone at the same voltage (0.2 mA/cm 2 ).
  • photo-assisted electrochemical deposition of Co-Pi onto mesostructured cc-Fe203 yields better performing photoanodes than either electrochemical deposition of Co-Pi or simple Co 2+ wet impregnation.
  • a stable -170 mV cathodic shift was observed with photoelectrochemical deposition of Co-Pi, while the electrochemical deposition of Co-Pi gave cathodic shifts of -100 mV, and Co 2+ impregnation gave -80 mV cathodic shifts.
  • Photo-assisted electrochemical deposition provides a more uniform distribution of Co-Pi on cc-Fe 2 03 than obtained by electrochemical deposition by allowing deposition only where visible light generates oxidizing equivalents. Optimization of the photo-assisted electrochemical deposition conditions allowed elimination of all nodules and islands to yield thin uniform films of
  • Example 4 Deposition of cobalt oxide catalysts on oc-Fe2Q3 by deposition from an aqueous solution of Co 2+ , such as from cobalt nitrate, cobalt acetate or cobalt sulfate.
  • CoO x Electrochemical deposition and photoelectrochemical deposition of a cobalt oxide catalyst, referred to here as "CoO x ,” on cc-Fe 2 03 were produced by deposition from an aqueous solution of Co 2+ , such as from cobalt nitrate, cobalt acetate or cobalt sulfate. X-ray diffraction experiments showed that CoO x did not match the typical diffraction patterns of known cobalt oxides, CoO, C02O3, or C03O4. In one embodiment, CoO x was electrodeposited from an aqueous solution of lOmM cobalt nitrate (pH -4) at 0.7-1.4 V vs Ag/AgCl.
  • FIGURE 19 shows the J-V characteristics of a-Fe 2 03 and the composite CoO x /cc-Fe 2 03 photoanode after electrochemical deposition in 1 M NaOH. Dark current (dotted) and photocurrent (solid) are illustrated. A -100 mV cathodic shift of the onset potential for water oxidation was observed in CoO x -modified
  • a photo-assisted electrochemical deposition (photoelectrochemical) approach was employed to achieve selective deposition of Co-Pi onto Ti0 2 nanowires (NWs).
  • FIGURE 29 illustrates current density-voltage curves of a Ti0 2 nanowire photoanode before and after Co-Pi photoelectrochemical deposition, measured under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dashed) in 0.1M potassium phosphate buffer at pH 7. Co-Pi modification results in a -190 mV cathodic shift in the photocurrent.
  • cobalt-containing catalyst Co-Pi can be photoelectrochemically deposited onto other semiconductor materials of different shapes, such as Ti0 2 nanowires, as well as onto dendritic a-Fe 2 03 photoanodes.
  • Simple electrodeposition of Co-Pi onto the same Ti0 2 nanowire structures grown on conductive FTO substrates results in preferential catalyst deposition onto the exposed more- conductive FTO instead and does not improve PEC water oxidation performance. Direct photodeposition of Co-Pi did not result in successful application of the catalyst.
  • Co-Pi can be used to improve the PEC water oxidation of a semiconductor such as Ti0 2 with an already low onset potential towards PEC water oxidation.
  • the successful Co-Pi modification of Ti0 2 nanowires demonstrates the versatility of this photoelectrochemical deposition method to apply cobalt-containing water oxidation catalysts onto semiconductor materials of various shapes and sizes.
  • FIGURE 30 illustrates current density-voltage curves of a Ti0 2 nanowire photoanode sensitized with CdS nanoparticles coated with a thin amorphous Ti0 2 protective layer, before and after Co-Pi photoelectrochemical deposition, measured under 1 sun, AM 1.5 simulated solar irradiation in 0.5M sodium thiosulfate.
  • the catalyst Co-Pi can be deposited by photoelectrochemical deposition onto complex electrodes involving visible-light- absorbing sensitizers, such as CdS, integrated with UV light absorbing wide-bandgap semiconductors, such as Ti0 2 , via photoexcitation of the sensitizer and an applied potential.
  • CdS bandgap 2.4 eV
  • the PEC water oxidation electrode is made more sensitive to visible light (i.e., sunlight), as seen by the large photocurrent enhancement.
  • a cathodic shift is also observed after catalyst modification, demonstrating the compatibility of this catalyst deposition method with sensitizers such as CdS and with complex electrodes involving both sensitizers and wide- gap oxides.
  • Example 7 Composite Co-Pi/Co :ZnO Photoanodes.
  • FIGURE 31 illustrates current density- voltage curves of a Co 2+ :ZnO photoanode before and after Co-Pi photoelectrochemical deposition, measured under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dotted) in 0.1M potassium phosphate buffer at pH 11.
  • photoelectrochemical deposition can also be applied to deposite Co-Pi onto wide-gap semiconductors doped with cationic impurities (Co 2+ ) introduced to extend PEC water oxidation into the visible region and increase the solar photocurrent densities relative to undoped ZnO.
  • Photoelectrochemical deposition of catalysts onto such doped semiconductors can also be achieved via excitation of mid-gap electronic transitions arising from the dopants, demonstrating that the photoelectrochemical deposition method is not limited to bandgap excitation of semiconductors. Regardless of the electronic transition used for photoelectrochemical deposition, the result is an increase in the overall PEC water oxidation efficiency.
  • Example 8 Composite Co-Pi/W iBiVCy Photoanodes.
  • FIGURE 32 illustrates current density- voltage curves of a W-doped BiV0 4 photoanode before and after Co-Pi photoelectrochemical deposition, measured under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dashed).
  • Example 9 Composite Cobalt Methyl-Phosphonate/a-FeqO Photoanodes.
  • FIGURE 33 illustrates current density-voltage curves of an -Fe 2 03 photoanode before and after cobalt methyl-phosphonate (Co-MePi) photoelectrodeposition, measured under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dotted) in 1 M NaOH, pH 13.6.
  • Co-MePi cobalt methyl-phosphonate
  • FIGURE 33 illustrates current density-voltage curves of an -Fe 2 03 photoanode before and after cobalt methyl-phosphonate (Co-MePi) photoelectrodeposition, measured under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dotted) in 1 M NaOH, pH 13.6.
  • Co-MePi cobalt methyl-phosphonate
  • FIGURE 34 illustrates current density- voltage curves of an -Fe203 photoanode before and after nickel borate (Ni-Bi) electrodeposition, measured under backside illumination with 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dotted) in 1 M NaOH, pH 13.6.
  • Ni-Bi nickel borate

Landscapes

  • Engineering & Computer Science (AREA)
  • Power Engineering (AREA)
  • Chemical & Material Sciences (AREA)
  • Microelectronics & Electronic Packaging (AREA)
  • Chemical Kinetics & Catalysis (AREA)
  • Electrochemistry (AREA)
  • Materials Engineering (AREA)
  • Metallurgy (AREA)
  • Organic Chemistry (AREA)
  • Electrolytic Production Of Non-Metals, Compounds, Apparatuses Therefor (AREA)
  • Catalysts (AREA)

Abstract

The provided method includes photoelectrodeposition of an electrocatalyst onto a semiconductor to form a photoanode. The method yields composite photoanodes showing enhancement of photocurrent (water splitting rate) when incorporated into a photoelectrochemical cell for water electrolysis.

Description

COMPOSITE PHOTOANODES
CROSS-REFERENCE TO RELATED APPLICATION This application claims the benefit of U.S. Provisional Application No. 61/311,724, filed March 8, 2010, the disclosure of which is incorporated herein by reference in its entirety.
STATEMENT OF GOVERNMENT LICENSE RIGHTS
This invention was made with Government support under the Integrative Graduate Education and Research Traineeship (IGERT) awarded by the National Science Foundation (Award No. DGE-Q5Q-4573). The Government has certain rights in the invention.
BACKGROUND
The photoelectrochemical (PEC) conversion of photon power into chemical fuels offers an attractive approach to storing solar energy, but it poses many fundamental chemical challenges. Hematite (cc-Fe203) has emerged as a prototype photoanode material for testing strategies to overcome the challenging 4-electron oxidation of water, which under basic conditions is described by Equation 1: 40H-→ 02 + 4e- + 2H20 ( 1 )
Hematite meets many of the target photoanode requirements: It is inexpensive, oxidatively robust, environmentally benign, and it absorbs visible light (Eg ~ 2.1 eV). Although the cc-Fe203 valence band edge potential is about 1 V or higher more positive than required for Equation 1 thermodynamically, water oxidation by photogenerated valence-band holes in cc-Fe203 is kinetically inefficient, and additional anodic overpotentials are typically required before significant PEC water splitting is observed. A remaining fundamental limitation of cc-Fe203 is that its conduction band edge potential resides -200 mV below that required to drive the cathodic half reaction (Equation 2).
2H20 + 2e-→ H2 + 20H" (2) Tandem PEC/photovoltaic (PV) configurations have been envisioned to provide the bias needed to meet these demands. Recent advances in controlled growth and doping of a-Fe203 nanostructures attempt to overcome many of the limitations associated with the short hole-diffusion length (-2-4 nm), low electron mobility (-10"1 cm2 V"1 s"1), and efficient charge carrier recombination characteristics of bulk a-Fe203 yielding promising PEC performance. For example, an overall solar-to-hydrogen power conversion efficiency of -2.1% has been estimated for one set of mesostructured cc-Fe203 photoanodes when powered by a PV device providing 1.4 V in a tandem configuration. Unfortunately, many low-cost PV devices such as dye-sensitized solar cells or organic PVs typically provide about 1 V or lower, and two such PVs in series would thus be required to provide the necessary 1.4 V. The development of cc-Fe203 photoanodes that require smaller overpotentials to oxidize water, such that they could be powered by single low-cost PV cells, would thus be attractive for reducing solar hydrogen production costs.
Recently, electrochemical water oxidation with low overpotentials was demonstrated over a range of pH values using an amorphous cobalt/phosphate catalyst ("Co-Pi") electrodeposited onto ITO or FTO electrodes. Remaining uncertainties about the catalyst's precise microscopic identity do not diminish its attractiveness for water- splitting PECs. Co-Pi requires 0.41 V overpotential at pH 7 to oxidize water with a current density of 1 mA/cm2, whereas the cc-Fe203 valence band edge potential provides about 1.2 V or higher. Photogenerated holes in cc-Fe203 should thus be amply capable of driving water oxidation by this electrocatalyst.
What is desired, therefore, are new methods and materials for forming improved photoelectrodes for use in PECs.
SUMMARY
This summary is provided to introduce a selection of concepts in a simplified form that are further described below in the Detailed Description. This summary is not intended to identify key features of the claimed subject matter, nor is it intended to be used as an aid in determining the scope of the claimed subject matter.
Composite photoanodes and methods for making the composite photoanodes are provided. The composite photoanodes comprise a semiconductor and an electrocatalyst. In one aspect, a method of forming a composite photoanode by photoelectrodeposition is provided. In one embodiment, the method comprises photoelectrodepositing a solid conformal layer of an electrocatalyst from an electrolyte solution on a surface of a semiconductor submerged in the electrolyte solution by simultaneously:
(1) impinging the surface of the semiconductor with electromagnetic radiation having a first wavelength and a first irradiance, to provide a first photoenergy that is sufficient to excite an electronic transition of the semiconductor; and
(2) applying a first electric bias to the semiconductor, wherein the first electric bias is less than an electrochemical deposition bias, said electrochemical deposition bias being the minimum voltage required to electrodeposit the electrocatalyst onto the surface of the semiconductor without impinging the surface of the semiconductor with electromagnetic radiation having the first photoenergy.
In another aspect, a method for making an electrode is provided. In one embodiment, the electrode is formed by deposition of a competent electrocatalyst onto a photoanode from an electrolyte, wherein the deposition can be carried out by photodeposition, electrochemical deposition or a combination thereof. The electrolyte may comprise inorganic and organic ions, such as phosphate anion, acetate anion, sulfate anion, chloride, nitrate, sodium, potassium, or any combination thereof. In another aspect, an electrode is provided comprising:
a photoanode having a first onset potential when incorporated into a photoelectrochemical cell for electrolysis of water into oxygen; and
a layer of an electrocatalyst conformally formed on a surface of the photoanode, wherein said electrocatalyst causes a cathodic shift in an onset potential of the electrode such that the electrode has a second onset potential when incorporated into a photoelectrochemical cell for electrolysis of water into oxygen, and wherein said second onset potential is less than said first onset potential.
DESCRIPTION OF THE DRAWINGS
The foregoing aspects and many of the attendant advantages of this invention will become more readily appreciated as the same become better understood by reference to the following detailed description, when taken in conjunction with the accompanying drawings, wherein:
FIGURES 1A-1D. SEM images of mesostructured a-Fe203 photoanode before (FIGURES 1A, IB), and after (FIGURES 1C, ID) 1 hour electrochemical deposition of Co-Pi catalyst. The catalyst layer cracking occurs upon drying for the SEM measurement and in some cases allows inspection of the catalyst underside: FIGURE ID shows that the Co-Pi underside topology conforms to the cc-Fe203 mesostructure.
FIGURES 2A-2C. (FIGURE 2A) Dark (dashed) and photocurrent (solid) densities for cc-Fe203 and Co-Pi/cc-Fe203 photoanodes collected using simulated AM 1.5 illumination (1 sun, backside) at a scan rate of 50 mV/s. (FIGURE 2B) Electronic absorption and (FIGURE 2C) IPCE spectra for a-Fe203 and Co-Pi/a-Fe203 (at 1.23 and 1 V vs RHE, respectively). The absorption spectrum of Co-Pi on FTO without a-Fe203 is included in FIGURE 2B, but no photocurrent was detected for these anodes.
FIGURE 3. Dark (dashed) and photocurrent (solid) densities for mesoscopic CC-Fe203 photoanodes used in this study under backside and frontside illumination, collected using simulated AM 1.5 sunlight (1 sun). Scan rate 50 mV/s.
FIGURE 4. Absorption spectrum of Co(OH)42~ at pH - 13 measured -30 min after preparation of the solution.
FIGURE 5. Summary of solar water- splitting PECs using Co-Pi/a-Fe203 composite photoanodes and Pt counter electrode.
FIGURE 6. Co-Pi electrochemical deposition on a-Fe203 photoanode in pH 7,
0.1 M KPi buffer at +1.1 V vs Ag/AgCl for 30 min.
FIGURE 7. Co-Pi electrochemical deposition on a-Fe203 photoanode in pH 7,
0.1 M KPi buffer at +1.1 V vs Ag/AgCl for 15 min.
FIGURE 8. Photocurrent decay of a-Fe203 photoanode measured in pH 7, 0.1 M
KPi buffer at +1.3 V vs RHE, 1 sun AM 1.5.
FIGURES 9A and 9B. Dark current (dotted) and photocurrent (solid and dashed) densities of cc-Fe203 photoanodes before and after 30 min of Co-Pi deposition, measured in pH 13.6 NaOH (FIGURE 9A), and pH 8 KPi (FIGURE 9B) at 50 mV/s (thick line) and 10 mV/s (dashed line). The a-Fe203 data were collected at 10 mV/s. The circles denote steady state photocurrent densities after 200 s of continuous illumination under 1 sun, AM 1.5 simulated sunlight. FIGURES 10A and 10B. (FIGURE 10A) Dark current (dotted) and photocurrent (solid) densities of an cc-Fe203 photoanode before (thin black) and after 30 min of Co-Pi deposition (thick lines) in 0.1 M KPi electrolyte at pH 8. The cc-Fe203 data (black curves) were collected at 10 mV/s. Inset: Photocurrent density vs time at +1.1 V vs RHE. (FIGURE 10B) Power dependence of photocurrent density for an cc-Fe203 photoanode before (x) and after (+) Co-Pi deposition, measured at +1.0 V vs RHE.
FIGURE 11. 02 generation and photocurrent density over time measured for a Co-Pi/a-Fe203 composite photoanode at +1.0 V vs RHE in 0.1 M KPi at pH 8. Co-Pi was electrodeposited on cc-Fe203 for 30 min. The top panel showed an initial spike in the rate of 02 evolution before relaxation to a steady state rate.
FIGURE 12. Linear sweep voltammetry of Co-Pi on FTO at various scan rates in pH 7, 0.1 M KPi electrolyte. Inset: Decay of the bulk electrolysis current density over time under these conditions, measured at +1.1 V vs Ag/AgCl (+1.3 V vs NHE).
FIGURES 13A-13D. Scanning electron micrographs of Co-Pi/a-Fe203 composite photoanodes after 15 minutes of Co-Pi deposition showing (FIGURE 13 A) ring-like deposition of Co-Pi in selective areas of the cc-Fe203 surface, and
(FIGURES 13B-13D) magnified views of Co-Pi patches within this ring. FIGURE 13D showed Co-Pi conforming to the topology of the underlying cc-Fe203 mesostructure. The cracks in the Co-Pi result from drying.
FIGURES 14A and 14B. Dark current (dotted) and photocurrent (solid) densities of an cc-Fe203 photoanode before and after 15 min of Co-Pi deposition, measured in
(FIGURE 14A) pH 13.6 NaOH, and (FIGURE 14B) pH 8 KPi with (black) and without 0.1 M NaCl at 50 mV/s (thin line) and 10 mV/s (thick line). The cc-Fe203 data were collected at 10 mV/s. The circles denote steady state photocurrent densities after 200 s of continuous illumination under 1 sun, AM 1.5 simulated sunlight.
FIGURES 15A and 15B. Photocurrent decay curves measured under 1 sun, AM 1.5 simulated sunlight at various applied potentials for Co-Pi/cc-Fe203 composite photoanodes in pH 8 KPi, pH 8 buffered salt water, and pH 13.6 NaOH (black) electrolytes. (FIGURE 15A) Data collected following 30 min of Co-Pi deposition, and (FIGURE 15B) data collected following 15 min of Co-Pi deposition. Photocurrent decay curves measured for cc-Fe203 in pH 8, 0.1 M KPi electrolyte (grey) are included in
FIGURE 15B for comparison. FIGURES 16A-16D. SEM images of Co-Pi/a-Fe203 composite photoanode after 30 min of Co-Pi deposition.
FIGURES 17A and 17B. (FIGURE 17A) Dark current (dotted) and photocurrent (solid) densities measured for an cc-Fe203 photoanode before and after 15 min of Co-Pi deposition. Data collected at 10 mV/s, for both front- and backside illumination. (FIGURE 17B) Photocurrent density and (¾ generation measured for the above photoanodes vs time: Co-Pi/cc-Fe203 (black) and cc-Fe203 (grey). The numbers in
(FIGURE 17B) indicate the photocurrent and (¾ enhancement factors (see text). Bubbles adhering to and releasing from the photoanode surface cause disruptions in the current density. All PEC data were collected under 1 sun, AM 1.5 simulated solar irradiation.
FIGURE 18. Dark current (dotted) and photocurrent (solid) densities measured for an a-Fe203 photoanode before and after Co-Pi photoelectrochemical deposition.
Data was collected at 10 mV/s with frontside illumination under 1 sun AM 1.5 simulated sunlight. Co-Pi was deposited for 10 min at -100 μΑ/cm2.
FIGURE 19. Dark current (dotted) and photocurrent (solid) densities measured for an a-Fe203 photoanode before and after CoOx electrochemical deposition. Data was collected at 10 mV/s with frontside illumination under 1 sun AM 1.5 simulated sunlight. CoOx was deposited for 15 min at 50μΑ/αιι2 and -1.1-1.3 V vs Ag/AgCl.
FIGURE 20. Dark current (dotted) and photocurrent (solid) densities measured for an cc-Fe203 photoanode before and after CoOx photoelectrochemical deposition.
Data was collected at 10 mV/s with frontside illumination under 1 sun AM 1.5 simulated sunlight. CoOx was deposited for 180 sec at ΙΟμΑ/cm2 and -0.1-0.3 V vs Ag/AgCl.
FIGURE 21A-21C illustrate three potential methods for adsorbing cobalt to a photoanode.
FIGURE 22A-22H Wide- and narrow-angle SEM images of an unmodified
CC-Fe203 mesostructured photoanode, and similar cc-Fe203 photoanodes following Co-Pi electrochemical deposition, photo-assisted electrochemical deposition of Co-Pi, and Co2+ adsorption.
FIGURE 23A-23C illustrate ark-current (dotted) and photocurrent (solid) densities of (23A) a Co-Pi/a-Fe203 electrode prepared by photo-assisted electrochemical deposition, (23B) a Co-Pi/cc-Fe203 electrode prepared by electrochemical deposition, and (23C) a Co2+/a-Fe203 electrode prepared by surface adsorption, compared to the parent cc-Fe203 photoanodes. FIGURE 24A-24C illustrate dark-current (dotted) and photocurrent (solid) densities of (24A) a Co-Pi/cc-Fe203 electrode prepared by photo-assisted electrochemical deposition, (24B) a Co-Pi/a-Fe203 electrode prepared by electrochemical deposition, and (24C) a Co2+/cc-Fe203 electrode prepared by surface adsorption, compared to the parent a-Fe203 photoanodes.
FIGURE 25A-25C illustrate the best (filled bars) and average (empty line) cathodic shifts, photocurrent density increases, and onset potentials for Co-Pi/cc-Fe203 photoanodes prepared by photo-assisted electrochemical deposition (P-Dep) and electrochemical deposition (E-Dep) of Co-Pi, and for Co2+/cc-Fe203 photoanodes prepared by surface adsorption of Co2+(Co-dip).
FIGURE 26 illustrates the average cathodic shifts plotted vs average onset potentials for Co-Pi/a-Fe203 photoanodes prepared by photo-assisted electrochemical deposition (P-Dep) and electrochemical deposition (E-dep) of Co-Pi, and for Co2+/a- Fe203 photoanodes prepared by surface adsorption of Co2+ (Co-dip) for the 12 films of Figure 5. The open symbols represent the parent a-Fe203 photoanodes. These data show a strong correlation between the two performance metrics, with photo-assisted electrochemical deposition of Co-Pi leading to the lowest onset potentials and the greatest cathodic shifts.
FIGURE 27 illustrates the Average photocurrent density increase vs photocurrent at 1.1 V vs RHE (one-sun) for the 12 films of Figure 5. Photo-assisted electrochemical deposition of Co-Pi yields the largest photocurrent density increases and the highest absolute photocurrent densities. The open symbols (grouped at the base of the dashed line) represent the parent a-Fe203 photoanodes.
FIGURE 28 illustrates the time dependence of the photocurrent density of a Co- Pi/a-Fe203 photoanode prepared by photo-assisted electrochemical deposition, measured at 1.0 V vs RHE in 1 M NaOH under continuous 1 sun, AM 1.5 simulated solar irradiation. The electrolyte was not stirred. The electrolyte was replaced after 75 hrs (dashed line), resulting in recovery of photocurrent density.
FIGURE 29 illustrates current density-voltage curves of a Ti02 nanowire photoanode before and after Co-Pi photoelectrochemical deposition, under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dashed) in 0.1M potassium phosphate buffer at pH 7. FIGURE 30 illustrates current density-voltage curves of a Ti02 nanowire photoanode sensitized with CdS and coated with a thin amorphous Ti02 protective layer, before and after Co-Pi photoelectrochemical deposition, under 1 sun, AM 1.5 simulated solar irradiation in 0.5M sodium thiosulfate.
FIGURE 31 illustrates current density-voltage curves of a Co2+:Zno photoanode before and after Co-Pi photoelectrochemical deposition, under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dotted) in 0.1M potassium phosphate buffer at pH 11.
FIGURE 32 illustrates current density- voltage curves of a W doped BiV04 photoanode before and after Co-Pi photoelectrochemical deposition, under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dashed).
FIGURE 33 illustrates current density-voltage curves of an a-Fe203 photoanode before and after cobalt methyl-phosphonate (Co-MePi) photo-assisted electrochemical deposition, under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dotted) in 1 M NaOH, pH 13.6.
FIGURE 34 illustrates current density- voltage curves of an -Fe203 photoanode before and after nickel borate (Ni-Bi) electrochemical deposition, under backside illumination with 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dotted) in 1 M NaOH, pH 13.6.
DETAILED DESCRIPTION
The composite photoanodes provide enhanced performance compared to known photoanodes when incorporated into photoelectrochemical (PEC) systems (e.g., a PEC cell for splitting water into hydrogen). One particular benefit of the composite photoanodes is a cathodic shift in the onset potential of the photoanode when used in a PEC system. Such a cathodic shift allows for reduced electrical requirements to drive the PEC process, thereby increasing the efficiency of such systems.
The methods for forming composite photoanodes are light-enhanced deposition methods, referred to herein as photoelectrochemical deposition methods. Photoelectrochemical deposition of an electrocatalyst onto a semiconductor to form a composite photoanode provides enhanced photoanode performance in PEC systems, including increased cathodic shift, compared to composite photoanodes fabricated using traditional electrochemical methods. Such improvements are attributed to the conformal nature of electrocatalyst layers formed using photoelectrodeposition.
In one aspect, a composite photoanode is provided, said composite photoanode comprising a semiconductor having a solid conformal layer of an electrocatalyst formed on its surface. Herein, the electrocatalysts referred to are competent electrocatalyst that produce a cathodic shift in the onset potential of the photoanode when used within a PEC.
The semiconductor acts as a photoanode. The semiconductor is made of a photon-absorbing material. In one embodiment, the semiconductor is a-Fe203
("hematite"). In another embodiment, the cc-Fe203 semiconductor is meso structured. In another embodiment, the semiconductor is a high- surface-area a-Fe203 photoanode. The use of hematite as a semiconductor is disclosed extensively herein, including in Examples 1-4 and 10.
While hematite is a preferred semiconductor, other semiconductors are contemplated. For example, Examples 5 and 6 disclose the use of titanium dioxide (particularly in nanowire form) as a semiconductor; Example 7 discloses the use of a cobalt-ion:zinc oxide semiconductor; and Example 8 discloses the use of a W:BiV04 semiconductor. These listed examples are generally inorganic in character.
In one embodiment, the semiconductor comprises a group IV semiconductor having a formula selected from the group consisting of binary, ternary, and quaternary. In a further embodiment, the group IV semiconductor further comprises ions selected from the group consisting of cations and anions.
In one embodiment, the semiconductor is an n-type semiconductor.
Other photoanode materials useful in composite electrodes include a material selected from the group consisting of an iron oxide, a zinc oxide, a titanium oxide, a tungsten-bismuth-vanadium oxide, a tungsten oxide, a gallium-zinc-oxide-nitride, or these materials also containing additional cations or anions.
In another embodiment, the semiconductor comprises a sensitizer having a sensitizer absorbance wavelength, said sensitizer absorbance being different from a semiconductor absorbance wavelength. By incorporating a sensitizer with the semiconductor (e.g., embedded within, or coating the surface of the semiconductor), a broader spectrum of light can be usefully absorbed by the semiconductor and overall composite photoanode for use in the PEC process. Representative sensitizers include cadmium selenium, as described below in Example 6. Additional sensitizers include cationic or anionic impurities
In one embodiment, the semiconductor has a physical shape selected from the group consisting of dendrites, wires, belts, rods, mesostructures, nanotubes, and thin films. As described further in Example 2, a high semiconductor surface area and/or a high composite photoanode surface area produces improved results for PEC reactions.
Certain physical shapes, such as dendrites, etc. are known to create a relatively high surface area. Such physical shapes are preferred in the embodiments provided herein.
Nanoscopic high-surface area shapes are accordingly preferred. Therefore, in another embodiment, the physical shape has nanoscopic dimensions. As used herein, the term nanoscopic dimensions refers to a shape having at least one feature (e.g., dendrites) having a smallest size of 100 nm or smaller.
Semiconductors can be deposited on substrates, or otherwise formed, according to methods known to those of skill in the art, including those provided below in the Examples. For example, hematite can be grown using chemical vapor deposition (see
Example 1).
The electrocatalyst is formed on a surface (e.g., a surface that will face a light source during PEC) of the semiconductor. In certain embodiments, the electrocatalyst produces a cathodic shift in the onset potential of a PEC process incorporating a composite electrode (semiconductor and electrocatalyst) when compared to the semiconductor alone. This comparison of electrodes can be found throughout the data provided herein so as to illustrate the efficacy of the disclosed materials, devices, and methods, in improving the PEC performance of semiconductors.
In one embodiment, the electrocatalyst is selected from the group consisting of a cobalt-containing catalyst, an iridium-containing catalyst, a manganese-containing catalyst, a ruthenium-containing catalyst, a nickel-containing catalyst (e.g., nickel borate, see Example 10), a cobalt-containing oxygen evolving catalyst, a cobalt oxide/hydroxide catalyst, and a cobalt oxide catalyst.
In one embodiment, the electrocatalyst is cobalt phosphate (Co-Pi). Co-Pi is used extensively in the examples provided herein. Examples 1-3 and 5-9 describe the use of Co-Pi to improve PEC performance.
In one embodiment, the layer of the electrocatalyst has a thickness of from 0.5 nm to 30 nm. As will be described further below in Example 2, a thick electrocatalyst layer will inhibit performance of composite photoanodes. Accordingly, nanoscale-thick electrocatalyst films are preferred. Electrochemical deposition does not allow for quality films of such a thickness to be deposited. Accordingly, photoelectrodeposition is preferred for forming nanoscale-thick films of electrocatalyst.
Examples of the competent electrocatalyst include, without limitation, a cobalt catalyst, iridium catalyst (e.g. )2), manganese catalyst (e.g. Mn-oxo complexes), ruthenium catalyst (e.g. [Ru(L)2(OH)2]2+ complexes, where L denotes ligand). In one embodiment, the cobalt catalyst was selected from the group consisting of cobalt based oxygen evolving catalyst and cobalt oxide catalyst (referred herein as "CoOx", see Example 4).
The photoelectrochemical performance of composite the photoanodes provided herein is improved compared to "semiconductor-only" photoanodes. The EXAMPLES describe these improvements extensively.
For example, Co-Pi/cc-Fe203 composite photoanodes for water oxidation are improved by optimization for front- side illumination in pH 8 electrolytes. Without being limited by theory, it is believed that a kinetic bottleneck appears to be related to the Co-Pi catalyst itself under these conditions. This kinetic bottleneck is overcome by more sparse deposition of Co-Pi onto cc-Fe203. Following these improvements, sustained water oxidation by Co-Pi/cc-Fe203 composite photoanodes was demonstrated in both photocurrent and 02 evolution measurements. Photoelectrochemical water oxidation by the Co-Pi/a-Fe203 composite photoanodes was enhanced relative to that of cc-Fe203 alone: Under these conditions, a five-fold enhancement in the photocurrent density and water oxidation rate was observed at +1.0 V vs RHE. This enhancement is even more substantial at about 1.0 V or lower vs RHE, where cc-Fe203 alone does not exhibit significant photocurrent at all.
It is also interesting to compare these results with those obtained for bulk electrolysis by Co-Pi without a photon-absorbing substrate. By itself, Co-Pi electrolysis current densities reached -1.2 mA/cm2 at an applied bias of +1.29 V vs NHE (pH 7), or
-+1.7 V vs RHE. In conjunction with an inexpensive and robust photoanode such as CC-Fe203 under 1 sun, AM 1.5 illumination, the applied bias necessary to achieve the same current density can be reduced by over 0.5 V in buffered salt water at pH 8, the average pH of sea water. The results described here thus demonstrate that sustained 02 evolution in mild salt water conditions can be achieved with significantly reduced external power demands relative to Co-Pi alone, particularly in the low current density regime, by integrating this catalyst with a light-harvesting semiconductor substrate. The overall process, in which photogenerated holes in a-Fe203 are converted to oxidizing equivalents in Co-Pi, yielding (¾ evolution well below the Co-Pi bulk electrolysis threshold potential, is summarized schematically in FIGURE 5. Further improvement of the performance of these composite photoanodes can be anticipated, for example, by variation of the Co-Pi deposition conditions to optimize photocurrent densities at extremely low bias. More generally, these results emphasize that composite photoanode strategies offer promising prospects for sustainable, affordable, and distributed solar fuel technologies. This is equally applicable to include other catalysts such as &(¾, Mn-oxo complexes, or [Ru(L)2(OH2)]2+ complexes, that can be powered in part or entirely by light-harvesting electrodes.
The combination of the first photoenergy and the first electric bias are sufficient to deposit catalyst components from the electrolyte to form the solid conformal layer of the electrocatalyst.
The electrocatalyst is formed from a buffer solution in which the semiconductor is submerged. The electrolyte solution can be any electrolyte solution known to those of skill in the art. Particularly, electrolyte solutions useful for traditional electrochemical deposition of an electrocatalyst onto a semiconductor are useful in the method. In a preferred embodiment, the electrolyte solution is a buffer solution of potassium phosphate
(e.g., at pH 7) containing Co(N03)2 if Co-Pi is to be the formed electrocatalyst.
In one embodiment, the pH of the electrolyte is about 7 or higher. In one embodiment, the pH of the electrolyte was about 13 or higher. In another embodiment, the pH of the electrolyte was about 8.
The surface of the semiconductor is irradiated with electromagnetic radiation
("light") having a first wavelength and a first irradiance. The light can be a single wavelength or a broadband source. The only requirement is that the light provides a photoenergy sufficient to produce an electronic excited state in the semiconductor so as to provide a portion of the energy required to deposit the electrocatalyst from the electrolyte solution. In one embodiment, the electronic transition is a bandgap transition. By exciting the bandgap transition, photogenerated valence band holes can oxidize ions in the electrolyte to from an active catalyst at the surface. Because the first photoenergy is not sufficient to drive the deposition of the electrocatalyst from the electrolyte solution, a first electric bias is simultaneously applied to the semiconductor. The first electric bias is significantly less than the bias required for electrochemical deposition. In one embodiment, the first electric bias is from 0.1 V to 0.4 V (e.g., versus Ag/AgCl). Therefore, the deposition (i.e., the photoelectrochemical deposition) of the electrocatalyst on the semiconductor is accomplished in the method by using energy from two sources (light and electricity) to facilitate the deposition reaction from the electrolyte. Neither of the two energy sources alone is sufficient to facilitate the deposition on their own.
The method according to this aspect utilizes light (e.g., sunlight or artificial sunlight) to assist in electrochemical deposition of an electrocatalyst onto a semiconductor. The method is useful, for example, to fabricate a composite photoanode according to the other aspects and Examples provided herein.
Examples 1 and 2 below disclose composite photoanodes fabricated with the generally known technique of electrochemical deposition. These examples are contrasted, by further Examples 3-10 utilize photoelectrochemical deposition.
Example 3, below, provides an in-depth development of the theory and results of photoelectrodeposition. While Example 3 primarily describes composite photoanodes of Co-Pi and a-Fe203, the method is not limited to these compounds. As illustrated in other Examples, photoelectrochemical deposition is compatible with any known semiconductors and electrocatalysts, particularly those used to make photoanodes using electrochemical deposition.
Without being bound by theory, in principle, photogenerated holes can be used to oxidize an ion from an electrolyte. For example, with reference to Co-Pi deposition on hematite, Co2+ can be deposited to form Co-Pi on the a-Fe203 photoanode and the electron can be removed by water reduction. Because photogenerated electrons in the conduction band of a-Fe203 are below the energy needed to reduce protons to hydrogen, a very low bias is applied to assist in photoelectrochemical deposition. Hence the addition of "electro" to photoelectrochemical deposition. The bias required for photoelectrochemical deposition is lower than that required for electrochemical deposition of similar compounds (i.e., deposition without the assistance of light). Any light source with sufficient energy to excite the band gap of the semiconductor can be used in photoelectrodeposition. For example, sunlight (or artificial sunlight) can be used to drive photoelectrochemical deposition so as to test the possibility of a sunlight-driven reaction. In an exemplary embodiment (described further in Example 3) photoelectrodeposition on a-Fe203 was conducted in a three-electrode configuration from a solution of Co2+ in potassium phosphate (KPi) buffer under 1 sun AM 1.5 simulated solar irradiation. A Pt mesh was used as the counter electrode and saturated Ag/AgCl was used as the reference electrode. Typical current densities during deposition were -1-100 μΑ/cm2.
It will be appreciated that a broad-spectrum light source (e.g., sunlight) need not be used in the method, as any light source capable of exciting the bandgap of the semiconductor is compatible with the method. For example, a single wavelength light source can be sufficient to excite the bandgap as long as
One impetus for the development of photoelectrochemical deposition was to develop an electrocatalyst deposition method that would allow for nanoscale-thick, conformal, continuous layers (films) of electrocatalyst to be deposited on a semiconductor. Particularly if the semiconductor is nano structured (e.g., dendritic). Traditional electrochemical deposition is insufficient in this regard. As demonstrated in the Examples (e.g., Example 3), thin, conformal electrocatalyst films are satisfactorily formed using photoelectrochemical deposition.
PEC reactions driven with photoanodes formed using photoelectrochemical deposition demonstrate improved absolute onset potential, cathodic shift of the onset potential, and maximum current density. All vital characteristics of, for example, PEC for water splitting, particularly in the context of solar-powered PEC devices.
In one embodiment, the electrocatalyst is selected from the group consisting of a cobalt-containing catalyst, an iridium-containing catalyst, a manganese-containing catalyst, a ruthenium-containing catalyst, a nickel-containing catalyst, a cobalt-containing oxygen evolving catalyst, a cobalt oxide/hydroxide catalyst, and a cobalt oxide catalyst. In a preferred embodiment, the electrocatalyst is cobalt phosphate.
In one embodiment, the layer of the electrocatalyst has a thickness of from 0.5 nm to 30 nm.
In one embodiment, the cathodic shift is from 50 mV to 400 mV.
In one embodiment, the layer of the electrocatalyst has a thickness of from 0.5 nm to 30 nm. Such a thickness is indicative of the "thin" nature of the conformal electrocatalyst coating. As set forth herein, such a thin coating is essential, not only to allow light through the electrocatalyst to the semiconductor, but also due to the short charge diffusion lengths in the photoanode materials. Finally, as disclosed herein, thin films of electrocatalyst are less prone to defects (e.g., aggregates) than thicker electrocatalyst films are.
It will be appreciated that the range of 0.5 nm to 30 nm represents from about one molecular layer to about tens of molecular layers. Accordingly, it is preferred that a minimal number of molecular layers are used to conformally coat the semiconductor with electrocatalyst without pinhole defects (exposing the semiconductor) or aggregates (which diminish device performance).
In one embodiment, the electrocatalyst is deposited from an electrolyte by photodeposition, electrochemical deposition, or combination thereof.
In one embodiment, the electrode is formed by electrodepositing a conformal layer of an electrocatalyst from an electrolyte solution onto a surface of a photoanode.
In one embodiment, the semiconductor is an n-type semiconductor.
In one embodiment, the formed photoanode, when incorporated into a photoelectrochemical cell for electrolysis of water into oxygen, reduces a water electrolysis onset voltage compared to a second photoanode comprising the semiconductor without the electrocatalyst.
In one embodiment, the water electrolysis onset voltage is reduced by 50 mV to 400 mV. In one embodiment, the first wavelength of the electromagnetic radiation is from 300 nm to 800 nm.
In one embodiment, the first irradiance of the electromagnetic radiation is from 0.1 W/m2 to 1100 W/m2, or the equivalent in pulsed irradiation.
In one embodiment, the electromagnetic radiation is selected from the group consisting of continuous radiation and pulsed radiation.
In one embodiment, the first electric bias is applied to the semiconductor as part of an electrochemical deposition system comprising a power source in electrical communication with the semiconductor and a counter electrode.
In one embodiment, the electrolyte solution comprises cations selected from the group consisting of cobalt, iridium, manganese, nickel, and ruthenium.
In one embodiment, the electrolyte solution comprises anions selected from the group consisting of phosphate, methyl phosphonate, borate, acetate, sulfate, and hydroxide.
In one embodiment, the semiconductor comprises a sensitizer having a sensitizer absorbance wavelength, said sensitizer absorbance being different from a semiconductor absorbance wavelength.
In one embodiment, the semiconductor comprises a group IV semiconductor having a formula selected from the group consisting of binary, ternary, and quaternary. In a further embodiment, the group IV semiconductor further comprises ions selected from the group consisting of cations and anions.
In one embodiment, the semiconductor comprises a material selected from the group consisting of an iron oxide, a zinc oxide, a titanium oxide, a tungsten-bismuth- vanadium oxide, a tungsten oxide, a gallium-zinc-oxide-nitride, or these materials also containing additional cations or anions.
In one embodiment, the combination of the first photoenergy and the first electric bias are sufficient to oxidize cations to deposit catalyst components from the electrolyte to form the solid conformal layer of the electrocatalyst. In another embodiment, an electrode was made by deposition of cobalt catalyst onto mesostructured cc-Fe203 from an electrolyte of Co2+. The deposition can be carried out by photodeposition or electrochemical deposition. Examples of the electrolyte include, without limitation, cobalt phosphate, cobalt borate, cobalt methyl phosphonate, cobalt nitrate, cobalt acetate, cobalt sulfate, and any combination thereof.
In one embodiment, the pH of the electrolyte is about 7 or higher. In one embodiment, the pH of the electrolyte was about 13 or higher. In another embodiment, the pH of the electrolyte was about 8.
In another embodiment, an electrode was made by electrochemical deposition of cobalt/phosphate catalyst ("Co-Pi") onto mesostructured cc-Fe203 and showed about
350 mV or higher cathodic shift of the onset potential for PEC water oxidation while retaining substantial photocurrent densities.
In another embodiment, Co-Pi was electrodeposited onto a mesostructured a-Fe203 photoanode. The photoelectrochemical properties of the resulting composite photoanodes were optimized for solar water oxidation under front- side illumination in pH 8 electrolytes. Relative to cc-Fe203 photoanodes, more sparse deposition of Co-Pi onto the cc-Fe203 resulted in a sustained five-fold enhancement in the photocurrent density and 02 evolution rate at +1.0 V vs RHE.
In one embodiment, the photoanode comprises a photoanode material selected from the group consisting of an iron oxide, a zinc oxide, a titanium oxide, a bismuth vanadium oxide.
In one embodiment, the photoanode material has a physical shape selected from the group consisting of dendrites, wires, and belts. In another embodiment, said physical shape has nanoscopic dimensions. As used herein, the term nanoscopic dimensions refers to a shape having at least one feature (e.g., dendrites) having a smallest size of 100 nm or smaller. In one preferred embodiment, the photoanode comprises hematite iron oxide dendrites. In a further preferred embodiment, the photoanode consists of hematite iron oxide dendrites conformally covered with a layer of cobalt phosphate.
In one embodiment, the electrocatalyst is selected from the group consisting of a cobalt-containing catalyst, an iridium-containing catalyst, a manganese-containing catalyst, a ruthenium-containing catalyst, a cobalt-containing oxygen evolving catalyst and a cobalt oxide catalyst.
In one preferred embodiment, the electrocatalyst is cobalt phosphate.
In one embodiment, the cathodic shift is from 50 mV to 400 mV.
In one embodiment, the layer of the electrocatalyst has a thickness of from 0.5 nm to 30 nm. Such a thickness is indicative of the "thin" nature of the conformal electrocatalyst coating. As set forth herein, such a thin coating is essential, not only to allow light through the electrocatalyst to the semiconductor, but also due to exacerbated electron-hole recombination with thicker catalyst films.
It will be appreciated that the range of 0.5 nm to 30 nm represents from about one molecular layer to about tens of molecular layers. Accordingly, it is preferred that a minimal number of molecular layers are used to conformally coat the semiconductor with electrocatalyst without pinhole defects (exposing the semiconductor) or aggregates (which diminish device performance).
In one embodiment, the electrocatalyst is deposited from an electrolyte by photodeposition, electrochemical deposition, or combination thereof.
In one embodiment, the electrode is formed by electrodepositing a conformal layer of an electrocatalyst from an electrolyte solution onto a surface of a photoanode.
In one embodiment, the formed photoanode, when incorporated into a photoelectrochemical cell for electrolysis of water into oxygen, reduces a water electrolysis onset voltage compared to a second photoanode comprising the semiconductor without the electrocatalyst.
In another aspect, an electrode is provided, comprising: a photoanode; and
a competent electrocatalyst that causes a cathodic shift in the onset potential of the electrode.
In another aspect, an electrode is provided, comprising:
a a-Fe203 photoanode;
a competent electrocatalyst selected from the group consisting of cobalt catalyst, iridium catalyst, manganese catalyst, ruthenium catalyst, cobalt based oxygen evolving catalyst and cobalt oxide catalyst.
In another aspect, an electrode is provided, comprising:
a a-Fe203 photoanode;
a competent electrocatalyst comprising a cobalt catalyst deposited onto the a-Fe203 photoanode from an electrolyte of Co2+ by photodeposition, electrochemical deposition, or combination thereof,
wherein the electrolyte comprises a composition selected from the group consisting of cobalt phosphate, cobalt nitrate, cobalt acetate, cobalt sulfate, and any combination thereof; and the electrode having about a several hundred millivolt cathodic shift of the onset potential for PEC water oxidation.
A system/device for converting water to hydrogen using only sunlight as an energy source is provided. The system includes a PEC comprising a photoanode formed using photoelectrochemical deposition and a photovoltaic cell. As described elsewhere herein, a water- splitting PEC typically requires over 1 V to produce hydrogen and oxygen from water, which is an electrical requirement that cannot be met by present PV technology. However, using the photoelectrochemical deposition method provided herein, the cathodic shift achieved in improving present photoanodes for PEC (e.g., Co- Pi/hematite), makes efficient sub-1 V water splitting in a PEC possible. Accordingly, by combining a PV cell with a PEC system having a photoanode formed using a photoelectrochemically deposited electrocatalyst on a semiconductor results in a system for converting water to hydrogen using only sunlight. The following examples are provided to better illustrate the claimed invention and are not to be interpreted in any way as limiting the scope of the invention. All specific compositions, materials, and methods described below, in whole or in part, fall within the scope of the invention. These specific compositions, materials, and methods are not intended to limit the invention, but merely to illustrate specific embodiments falling within the scope of the invention. One skilled in the art may develop equivalent compositions, materials, and methods without the exercise of inventive capacity and without departing from the scope of the invention. It will be understood that many variations can be made in the procedures herein described while still remaining within the bounds of the invention. It is the intention of the inventors that such variations are included within the scope of the invention.
EXAMPLES
Example 1. oc-Fe2Q3 Photoanode Fabrication.
Mesostructured Si-doped cc-Fe203 photoanodes of 400 - 500 nm thickness were grown by atmospheric pressure chemical vapor deposition (APCVD) using Fe(CO)5 and tetraethoxysilane (TEOS) as precursors, delivered to an FTO substrate at 470°C using Ar carrier gas. SEM images of a representative cc-Fe203 photoanode are shown in FIGURES 1A and IB and revealed a highly structured electrode surface. PEC measurements were then performed in a 3-electrode configuration using an aqueous OH" electrolyte (1 M NaOH, pH 13.6), a Pt counter electrode, and Ag/AgCl as the reference electrode. Photocurrent densities were measured as a function of applied voltage under simulated 1 sun AM 1.5 solar irradiation. The cc-Fe203 PEC performance was found to depend strongly on surface morphology, Si doping level, and growth temperature, among other parameters.
Example 2. Fabrication of Co-Pi/oc-Fe2Q3 Photoanodes.
Mesostructured Si-doped cc-Fe203 photoanodes were grown on F:Sn02 (FTO)-coated glass substrates (TEC15, 15 Ω/cm2 Hartford Glass Co.) by atmospheric pressure chemical vapor deposition (APCVD) following procedures known in the art. The precursors, Fe(CO)5 (Aldrich 99.999%) and TEOS (Aldrich 99.999%), were delivered by bubbling Ar gas (Praxair, 5.0 Ultra High Purity) at 11.3 and 19.4 mL/min, respectively, controlled by mass flow controllers. The gas was then mixed with air flowing at 2L/min and directed by a glass tube onto the lower portion of a 50x13x2.3 mm3 FTO substrate kept at 470°C. The Co-Pi catalyst was electrodepo sited onto the oxide anodes as known in the art. The anode was submerged in a buffer solution of 0.1 M potassium phosphate (pH 7) containing 0.5 mM Co(N03)2 and a bias of 1.29 V (vs. NHE) was applied for 1 hr. For PEC measurements, the anode surface was masked during electrochemical deposition to yield a catalyst-covered area that matched the irradiated area (0 = 6 mm). Masking was achieved using electrical tape, which was then removed for PEC measurements. Composite Co-Pi/cc-Fe203 anodes for which the mask was not used showed greater dark currents from the Co-Pi catalyst, but were otherwise very similar.
Electronic absorption spectra were measured using a Cary 500 UV/vis/NIR spectrophotometer (Varian). SEM images were collected using a FEI Sirion scanning electron microscope operating at 5 kV. Electrochemical measurements were performed in a 3-electrode configuration using an aqueous hydroxide electrolyte (1 M NaOH, pH 13.6), a Pt counter electrode, and an Ag/AgCl reference electrode. In a typical measurement, a titanium clasp was used to make contact with the upper 25% of the 5 cm long anode, where no cc-Fe203 had been deposited. The bottom -50% of the anode was submerged in the electrolyte solution in a home-built optical cell. Cyclic voltammetry measurements were performed using a computer-controlled Eco Chemie μΑυίο^ II potentiostat. Potentials are reported vs both Ag/AgCl and RHE, the latter obtained using the formula ERHE= EAgC1+ 0.059 pH +0.1976V. Photocurrent densities were measured as a function of applied voltage under simulated AM 1.5 solar irradiation (1 sun), achieved using an Oriel 96000 solar simulator integrating a 150 W Xe arc lamp and Oriel 81094 filter, and delivered to the anode via fiber optic. Measurements were performed at a scan rate of 50 mV/s. IPCE measurements were performed using a Xe arc lamp with an Oriel Cornerstone 74000 monochromator with slits set to -10 nm spectral bandwidth at the designated bias voltage provided by the potentiostat. The wavelength was scanned at 1 nm/s. Photon power densities were determined using a calibrated Si photodiode. Dark current measurements probe the entire submerged FTO + CC-Fe203 (or Co-Pi/cc-Fe203) surface, whereas photocurrents represent the response achieved from just the irradiated area normalized to 1 cm2. This area was circular with a diameter of 6 mm.
Typical monochromatic photon power densities in the IPCE measurements were ~0.50W/m2. For the data shown in FIGURES 2A-2C, the a-Fe203 photoanode data were collected first, then the Co-Pi catalyst was deposited onto the same cc-Fe203 photoanode, and then the parallel data were collected on the Co-Pi/cc-Fe203 photoanode.
FIGURE 2A shows dark (dashed) and photocurrent (solid) densities for an CC-Fe203 photoanode with backside illumination. Whereas the dark response was negligible up to 1.5 V vs RHE, the photoresponse showed a rise and plateau with an onset voltage of ~1 V vs RHE that typifies cc-Fe203. FIGURES 1C and ID showed SEM images of a representative cc-Fe203 photoanode following Co-Pi electrochemical deposition for 1 hr as known in the art. Extensive cracking of the -200 nm thick catalyst layer occurred upon drying for the SEM measurement. FIGURE ID showed a portion of the catalyst layer that curled off of the cc-Fe203 film upon drying, revealing its underside.
This image showed the inverse mesostructure from the cc-Fe203 anode, demonstrating that the catalyst layer conformed to the topology of the cc-Fe203 surface. A high degree of interfacial contact between the cc-Fe203 and catalyst layers was achieved. FIGURE 2A also showed the dark and photocurrent responses of the Co-Pi cc-Fe203 composite photoanode prepared by electrochemical deposition of the Co-Pi catalyst on the same cc-Fe203 photoanode. The major phenomenological observation was that modification of cc-Fe203 with Co-Pi reduced the bias voltage required for solar PEC water oxidation by about 350 mV or higher, corresponding to a reduction from -1.2 to about -0.9 V or lower that would be required from the PV of a water splitting PEC/PV tandem cell.
At 1.4 V (RHE), cc-Fe203 photocurrent densities with front-side illumination were approximately 2x greater than with backside illumination (FIGURE 3), without being bound by any theory, a common observation attributable to the greater surface area of the anode front. In the Co-Pi/cc-Fe203 anodes, however, front-side illumination did not greatly enhance the photocurrent, without being bound by any theory, likely because of non-productive absorption by the catalyst layer. Co-Pi absorbed throughout the visible spectral region (FIGURE 2B) but generated no detectable photocurrent, either on CC-Fe203 or directly on FTO. IPCE measurements of the cc-Fe203 (1.23 V vs RHE) and Co-Pi/cc-Fe203 (I V vs RHE) photoanodes using backside illumination showed essentially identical dispersion (FIGURE 2C), without being bound by any theory, in both cases deriving only from cc-Fe203 excitation. Co-Pi thus behaved solely as a surface electrocatalyst. The composite photoanode of FIGURE 2C showed IPCE about 15% or higher at 550 nm and 1 V vs RHE, conditions where cc-Fe203 alone showed negligible photocurrent (FIGURE 2a). This IPCE maximized at 450 nm (18%) before decreasing again below -400 nm because of the decreasing light penetration depth (FIGURE 2B).
To test for the possibility that the cathodic photocurrent shift came from the action of solvated cobalt as a redox mediator, a set of control experiments involving deliberate addition of solvated Co2+ was performed by adding Co(OH)4 2" to the 1 M NaOH electrolyte of the PEC cell. Co(OH)4 2" was prepared by dissolving cobalt nitrate in a 50 wt% concentrated NaOH aqueous solution to make a -0.005M Co(OH)42~ solution, which was then added to distilled water to reach pH - 13. The final solution was added dropwise to the electrolyte of the PEC cell under operating conditions, where its influence on dark and photocurrent densities of various photoanodes could be monitored. Although sparingly soluble at pH 13.6, the precipitation of solid Co(OH)2 from the electrolyte solution was likely slow, as indicated by observation of the characteristic Co(OH)4 2" 4Tj(P) ligand field band in the absorption spectrum of the pH - 13 stock solution even
-30 min after preparation (FIGURE 4), which is roughly three times longer than required to collect the PEC data. Although solvated Co(OH)4 2- was difficult to detect spectroscopically in the PEC cell at very low concentrations, its presence was readily detected electrochemically by an increase in dark current for both FTO and CC-Fe203-modified FTO anodes. When Co(OH)42~ was added to the PEC cell during CC-Fe203 film measurement, an increase in dark current was observed but no significant change in photocurrent resulted, arguing against interpretation of the cathodic shift described in the manuscript as arising from participation of a soluble Co2+ redox mediator. Independent control experiments in which I-V scans of unmodified FTO anodes were measured before and after repeated photocurrent measurements on Co-Pi/cc-Fe203 composite photoanodes without replacing the electrolyte also failed to detect solvated Co2+, again arguing against interpretation of the cathodic photocurrent shift described in the manuscript as arising from participation of a soluble Co2+ redox mediator.
As the control experiments showed, the possibility of dissolved cobalt acting as redox mediator, or of an unidentified sacrificial reagent contributing to photocurrent, was eliminated by the following observations: (i) Addition of solvated Co2+ to the electrolyte had no noticeable effect on photocurrent densities; (ii) Replacement of the PEC electrolyte solution with new stock solution caused no change in photocurrent and did not lead to a photocurrent induction period; (iii) Continuous photocatalysis at 1 V vs RHE for about 10 hrs or longer showed no change in performance. Therefore, the cathodic shift in FIGURE 2 A reflected the ability to drive Equation 1 (of the Background section) at much smaller overpotentials using the composite photoanodes than with cc-Fe203 alone.
Most cc-Fe203 PEC cells operating under similar conditions show negligible photocurrent densities below 1 V vs RHE. Modification of the cc-Fe203 surface by adsorption of Co2+ from aqueous 10 mM Co(N03)2 was previously shown to cause an
-17% increase in current density at 1.23 V vs RHE and an 80 mV cathodic shift of the onset potential. Similarly, growth of Ru02 onto cc-Fe203 surfaces led to a 120 mV cathodic shift of the onset potential with about 80 μΑ/cm2 or lower at 1 V vs RHE. Interestingly, cc-Fe203 nanorods have shown greater relative photocurrent densities at low bias than typical mesostructured cc-Fe203 photoanodes, but with photocurrent densities of ~2 μΑ/cm2. Without being bound to any theory, it is possible that the conformal catalyst deposition facilitates interfacial hole transfer from cc-Fe203 to Co-Pi, allowing photon absorption and redox catalysis to be effectively decoupled while retaining photocurrent densities. Efficient hole transfer from cc-Fe203 to Co-Pi should enhance the electron gradient in the cc-Fe203 mesostructure under irradiation, also contributing to the driving force for electron diffusion to the FTO and reducing deleterious carrier recombination processes. Catalyst electrochemical deposition onto a-Fe203 may also passivate surface defects.
The experimental results for the Co-Pi/cc-Fe203 composite photoanodes may be summarized in FIGURE 5 (and this model can be used to describe the behavior of all composite photoanodes provided in the disclosed aspects and embodiments herein). Referring to FIGURE 5, photoexcitation of cc-Fe203 generates an electron-hole pair.
Photogenerated holes are trapped by the Co-Pi catalyst, which excels at water oxidation. Photogenerated electrons migrate to the FTO back contact and pass through the circuit to the Pt counter electrode, where water reduction occurs in the 3-electrode configuration.
The present results demonstrate that partitioning photoabsorption, charge separation, and redox catalysis in composite photoanodes offers promising opportunities for improving solar water- splitting PECs.
Si doped cc-Fe203 photoanodes were fabricated on fluorine doped tin oxide (FTO) glass (50 x 13 x 2.3 mm TEC 15 Hartford Glass Co.) at 470°C for 5 min by atmospheric pressure chemical vapor deposition (APCVD) following procedures known in the art. The a-Fe203 films investigated here were typically -400-500 nm thick. For Co-Pi deposition onto cc-Fe203 photoanodes for the following data, electrical tape with an aperture that matched the irradiated area during photoelectrochemical (PEC) experiments (0 = 6 mm diameter) was applied onto the cc-Fe203. As the working electrode, a-Fe203 was submerged into a solution of 0.5 mM cobalt nitrate in 0.1 M pH 7 potassium phosphate (KPi) buffer. A Pt mesh was used as the counter electrode and saturated Ag/AgCl was used as the reference electrode. Co-Pi was electrodeposited at +1.1 V vs Ag/AgCl for 15 (FIGURE 6) or 30 (FIGURE 7) min. Typical current densities during deposition were -20-200 μΑ/cm2 (FIGURES 6 and 7) For electrolysis studies involving Co-Pi on FTO, Co-Pi was electrodeposited for 15 min using the above conditions and a mask of 1 cm x 1 cm.
Photoelectrochemical experiments. Current-voltage characteristics were measured using an Eco Chemie μ-Autolab II potentiostat in a home-built three-electrode optical cell using Ag/AgCl as the reference electrode and a Pt wire as the counter electrode. Contact to the photoanodes was made by a titanium clasp attached to the exposed FTO surface at the top of the anode, while the lower portion containing the sample was submerged in the electrolyte. Measurements were performed in 1 M NaOH(aq) at pH 13.6, 0.1 M KPi buffered at pH 8, and 0.1 M NaCl(aq) buffered at pH 8 with 0.1 M KPi. Potentials are reported vs Ag/AgCl (measured) or RHE (obtained using the relationship ERHE = EAg/AgQ + 0.0591*pH + 0.1976 V). Photocurrent densities were measured under 1 sun, AM 1.5 simulated sunlight using an Oriel 96000 solar simulator equipped with a 150 W Xe arc lamp and an Oriel 81094 filter. The photoanodes were masked to illuminate a circular area of 6 mm diameter. Power dependence measurements were performed using an Ag variable neutral density filter, Thorlabs NDC-50C-2M. Unless otherwise stated, all films in this example were illuminated from the front side of the photoanode. Unless otherwise specified, all experiments in this example were performed at room temperature in air atmosphere.
Oxygen detection. The detection of 02 was performed using a YSI 5000 dissolved oxygen meter equipped with a YSI 5010 self-stirring Clark-type probe in a three-neck flask with an optical window. Before use, the electrolyte (0.1 M KPi buffered at pH 8) was degassed and purged with argon gas. Measurements were conducted in argon in the same three-electrode configuration described for PEC experiments using the same light source. Again, the photoanodes were masked to illuminate a circular area of 6 mm in diameter. Consecutive measurements were taken at +1.0, 1.1, and 1.23 V vs RHE for two hours at each potential. While the light was off between voltages (-160 seconds), there was no increase and sometimes even a decrease in the 02 level due to consumption by the Clark electrode.
Co-Pi/cc-Fe203 photoanode performed under front-side illumination and mild pH conditions.
Optimization of the composite photoanodes for front-side illumination was carried out at mild pH conditions. To reduce photon absorption by the catalyst, Co-Pi deposition times were decreased from the original one-hour duration.
The basic electrolyte (pH 13.6) used previously herein is generally undesirable for practical applications. A gradual decrease in photocurrent density from cc-Fe203 anodes alone was observed under continuous illumination in 0.1 M KPi electrolyte at pH 7 and
+1.3 V vs RHE (FIGURE 8). Therefore PEC measurements were carried out in electrolytes at pH 8, which is around the pH of natural seawater. Two approaches were used to achieve this pH. One involved use of 0.1 M potassium phosphate (KPi), buffered to pH 8. The second involved 0.1 M NaCl buffered to pH 8 with 0.1 M KPi.
FIGURES 9A and 9B show current-voltage (J-V) curves collected for Co-Pi/cc-Fe203 composite photoanodes prepared with 30 min deposition of Co-Pi and measured in various electrolytes. Each data set was compared to analogous data collected for the same Fe203 film measured before Co-Pi deposition. FIGURE 9A shows the J-V curves collected using 1 M NaOH at pH 13.6 and FIGURE 9B shows data collected using 0.1 M KPi at pH 8. As described above, in 1 M NaOH, Co-Pi deposition yields a cathodic shift of about 350 mV or higher in the photocurrent onset potential relative to CC-Fe203 (FIGURE 9 A). PEC measurements in 0.1 M KPi electrolyte at pH 8 also showed similar shifts. With front- side illumination there was a slight decrease in the photocurrent density at high applied potentials (+1.3-1.6 V) compared to cc-Fe203 alone, attributed to partial photon absorption by the catalyst layer. The data in FIGURES 9A and 9B demonstrated that Co-Pi/cc-Fe203 composite photoanodes can operate under reduced pH conditions and with front-side illumination.
Kinetic bottleneck in Co-Pi/cc-Fe203 composite photoanodes. In the course of efforts to optimize the Co-Pi/cc-Fe203 composite photoanodes, it was recognized that improvements in efficiency were often accompanied by increasingly apparent symptoms of kinetic limitations. For example, FIGURES 9A and 9B also showed J-V curves of the same Co-Pi/cc-Fe203 composite photoanodes measured at the slower scan rate of 10 mV/s (dashed line). The open circles in FIGURES 9 A and 9B were the quasi- steady state photocurrent densities measured after 200 s of simulated solar irradiation at each applied potential. The cathodic shift and photocurrent densities both decreased as the scan rate was slowed, converging on those of the underlying cc-Fe203 at slowest scan rates. Upon increasing the scan rate again, the J-V curves recovered their original shape, even after 10+ hours of continuous illumination. Illuminating with chopped light also recovers the photocurrent enhancement. This behavior is largely independent of electrolyte or pH
(FIGURES 9 A and 9B), and suggests the existence of a kinetic bottleneck in the performance of these Co-Pi/cc-Fe203 composite photoanodes.
To detail this kinetic bottleneck, its symptoms were explored in various complementary measurements on a single Co-Pi/cc-Fe203 composite photoanode in
0.1 M KPi electrolyte at pH 8. The resulting data are summarized in FIGURES 10A and 10B. Expanding on FIGURES 9 A and 9B, photocurrents were measured at a greater variety of scan rates to show the evolving characteristics of the J-V curves. With faster scan rates, the first maximum shifted to higher potentials. The inset showed the photocurrent response vs time upon unblocking the light path, measured at +1.1 V vs RHE. A large initial spike in photocurrent upon illumination was followed by multi-exponential decay to a lower steady-state current density with an effective time constant on the order of 10 sec, i.e., comparable to the data collection timescale (10s of seconds). Without being bound to any theory, it is believed that some of the initial current density in this trace is attributed to cobalt oxidation, which is an essential step in the water oxidation mechanism. Oxygen detection experiments in 0.1 M KPi at 1 V vs RHE showed that this initial high current density was accompanied by a spike in the oxygen evolution rate, however, which then also decreased as the current density decays (FIGURE 11). The spike and subsequent current density decay therefore cannot be ascribed solely to an initial charging current for the redox active Co-Pi layer. FIGURE 10B plots the steady-state photocurrent density for a Co-Pi/cc-Fe203 composite photoanode measured as a function of illumination power density between 0 and 1 sun. There was a marked saturation in the photocurrent as the light intensity was increased.
Overall, four major symptoms of this kinetic bottleneck can be identified: (i) a scan rate dependence, (ii) a kinetic decay in the photocurrent density, (iii) photocurrent saturation upon increased illumination, and (iv) a sweep-rate-dependent maximum at the beginning of the J-V curve. Without being bound by any theory, the sweep-rate dependence of this maximum is a consequence of the superposition of an increasing current density from increasing bias with a current decay.
Parallel measurements were performed for cc-Fe203 photoanodes alone under the same experimental conditions. Under the conditions represented in FIGURES 9A and 9B, the J-V curves of cc-Fe203 photoanodes did not change significantly with scan rate
(data collected at 10 mV/s are plotted in FIGURES 9A and 9B). In time-dependence measurements, irradiation typically induced a small initial current spike followed by relaxation to a similar steady-state value within a few seconds. Finally, the cc-Fe203 photocurrents increased much more linearly with increasing light intensity under these conditions (FIGURE 10B). Without being bound by any theory, it is believed that the kinetic bottleneck is associated with the Co-Pi modification, perhaps as an intrinsic limitation of the catalyst under these conditions, or perhaps because of slow interfacial electron transfer.
To test the catalyst alone, Co-Pi was electrodeposited on FTO and electrochemical experiments were conducted in 0.1 M KPi electrolyte buffered to pH 7 with stirring. FIGURE 12 showed the J-V characteristics of Co-Pi at various scan rates, and the current density time dependence under typical electrolysis conditions of +1.1 V vs Ag/AgCl. Like in FIGURES 10A and 10B, the bulk electrolysis by Co-Pi on FTO also showed a scan rate dependence and a decay in the current density in the region where water oxidation was normally observed, +1.3 V vs NHE, or +1.7 V vs RHE. These observations support that the kinetic bottleneck observed in the Co-Pi/cc-Fe203 photoanodes was associated with the Co-Pi catalyst itself rather than with the Co-Pi/a-Fe203 interface.
Alleviating the kinetic problem in Co-Pi/cc-Fe203. Without being bound to any theory, if there is a kinetic bottleneck in Co-Pi/cc-Fe203 photoanodes, it is possible that even further reduction of the Co-Pi deposition time may remediate the problem. For example, thick layers of Co-Pi may inhibit rapid charge or proton transport from electrolyte through the catalyst, thus restricting current flow and allowing other non-productive recombination pathways to become competitive. To test this possibility, Co-Pi was electrodeposited on cc-Fe203 photoanodes for 15 min. FIGURES 13A-13D showed SEM images of the resulting Co-Pi/cc-Fe203 composite photoanode. Unlike the dense coverage of Co-Pi on cc-Fe203 after 1 hr of electrochemical deposition described above (FIGURE 1C), 15 min deposition resulted in sparse coverage of the cc-Fe203 photoanode by Co-Pi, which displayed ring-like patterns on the surface (FIGURE 13A). These ring patterns were formed from smaller patches of Co-Pi (FIGURES 13B-13D). These patches were estimated to be thinner than 100 nm, compared to -200 nm thick film that was deposited during 1 hour deposition (FIGURE 1C), and they showed cracking due to drying. The Co-Pi layer conformed to the topology of the cc-Fe203 well and the micro structure of the cc-Fe203 surface could be seen through the catalyst in places.
Without being bound by any theory, it is hypothesized that these patches were somehow associated with scratches or pinholes in the cc-Fe203 that allowed current to flow more readily during electrochemical deposition. Preliminary energy dispersive X-ray analysis (ED AX) experiments demonstrated the existence of a very thin catalyst layer over the entire cc-Fe203 surface, and without being bound by any theory, it is possible that catalysis was also distributed over the entire surface.
PEC measurements were performed on these thinly covered Co-Pi/cc-Fe203 photoanodes and the results were shown in FIGURES 14A and 14B. Cathodic shifts of -180 mV were observed, as were enhanced photocurrents across the entire potential range. The J-V curves measured in NaOH electrolytes no longer exhibited the marked scan rate dependence that was observed for the parallel set of photoanodes with greater
Co-Pi coverage (FIGURES 9A and 9B). With a thinner layer of Co-Pi, the enhanced current density is maintained even after 200 sec (FIGURES 14A and 14B), and the cathodic shift is stable. For PEC measurements conducted in 0.1 M KPi at pH 8, a gradual decay in the photocurrent over time was still evident, and the cathodic shift decreased from -200 mV to 150 mV after 200 sec of continuous illumination
(FIGURE 14B). PEC measurements were also performed in 0.1 M NaCl buffered to pH 8 with KPi. The resulting J-V curves with NaCl added are essentially indistinguishable from those without NaCl (FIGURE 14B), demonstrating that the present of chloride did not interfere with PEC water oxidation with Co-Pi/cc-Fe203 composite photoanodes.
FIGURES 15A and 15B compared the kinetic responses of Co-Pi/cc-Fe203 photoanodes with thicker (FIGURE 15 A, 30 min deposition, see FIGURES 16A-16D) and thinner (FIGURE 15B, 15 min deposition, see FIGURES 13A-13D) Co-Pi coverage. The photocurrent decay curves of FIGURE 15A showed a large initial spike in current density, followed by a multiexponential decrease with τ - 10 sec to a small steady- state photocurrent density close to that of the underlying cc-Fe203 photoanode. In contrast, the photoanodes with thinner Co-Pi coverage showed substantially more stable performance. The steady-state photocurrent densities in FIGURE 15B were enhanced relative to those of the parent cc-Fe203 photoanodes. The results showed a sustainable photocurrent density that was enhanced relative to cc-Fe203 by more than an order of magnitude at 0.83 V, where cc-Fe203 alone did not exhibit significant photocurrent (FIGURE 15B).
Gains in photocurrent were less substantial at higher applied potentials, without being bound by any theory, it was likely due to contributions directly from cc-Fe203.
FIGURES 14A, 14B, 15A, and 15B demonstrated that reduced Co-Pi deposition onto CC-Fe203 photoanodes circumvented the major kinetic limitations identified above, while still shifting the onset potential of cc-Fe203 by -180 mV, and simultaneously facilitated front-side illumination for maximum photocurrent densities.
Decreased deposition of Co-Pi onto cc-Fe203 largely overcame the kinetic limitations described in FIGURES 10A and 10B, but there was still some evidence of such kinetic effects in KPi electrolyte (FIGURES 14B, 15B) that were not observed in 1 M NaOH. For 1 M NaOH, there was a small initial spike in the photocurrent followed by a small gradual increase to steady state. Without being bound by any theory, it is possible that limited mobility of protons through the amorphous catalyst may contribute to the kinetic bottleneck described by FIGURES 9 A, 9B, 10A, 10B, 12A, and 12B, and that OH" is better able to overcome this limitation. Overall, FIGURES 14A, 14B, 15A, and 15B showed that this bottleneck was lessened by changing the electrolyte from pH 8
KPi to pH 13.6 NaOH, and was effectively circumvented by reducing the density of catalyst on the cc-Fe203 surface.
Oxygen evolution. In addition to current density measurements, PEC 02 evolution by the Co-Pi/cc-Fe203 composite photoanodes was also examined. Oxygen evolution was measured at various applied potentials before and after 15 min of Co-Pi electrochemical deposition onto an cc-Fe203 photoanode. Measurements were performed in 0.1 M KPi electrolyte at pH 8. FIGURE 17 A showed the J-V characteristics of the CC-Fe203 photoanode used for these measurements, before and after Co-Pi deposition, and for both front and backside illumination. Photocurrent densities increased substantially with front-side illumination, particularly at low potentials. FIGURE 17B plotted the photocurrent density vs time along with the 02 concentrations measured simultaneously using the Clark-type electrode. Sustained photocurrent was observed for the Co-Pi/cc-Fe203 composite photoanode over the course of this ~6 hour experiment. This steady- state photocurrent was enhanced over that of the parent cc-Fe203 film, even after several hours of illumination, and was accompanied by a correspondingly large enhancement in the 02 evolution rate. The photocurrent density and 02 evolution enhancement factors
J (Co-Pi/ ar- Fe203 ) 402l/ di(Co-Pi/ ar- Fe203 )
( and — , respectively) measured at
J (ar- Fe203 ) d [02] / dt(a- Fe203 )
each applied potential were indicated in FIGURE 17B. The amount of dissolved 02 detected by the Clark-type electrode was lower than the theoretical maximum for the measured current densities, but without being bound by any theory, this difference may be attributable to the adherence of bubbles on the rough surface of the Co-Pi/cc-Fe203 photoanode. Occasional jumps in the photocurrent density were observed for the composite photoanodes and may be related to release of these bubbles.
FIGURE 17B showed that PEC 02 evolution by the Co-Pi/cc-Fe203 composite photoanode was enhanced over that of the same cc-Fe203 photoanode without Co-Pi.
Despite the gradual decline in photocurrent density for these photoanodes when measured in 0.1 M KPi electrolyte at pH 8 (FIGURES 14A, 14B, 15A, and 15B), after 2 hours of continuous irradiation at +1.0 V vs RHE, ~5 times more oxygen was produced for the Co-Pi modified cc-Fe203 photoanode with no detectable degradation in performance.
Example 3. Co-Pi/ 0C-Fe2O3 photoanodes prepared by photoelectrochemical deposition.
Co-Pi catalyst was photoelectrochemical deposited onto cc-Fe203 photoanodes by using light and an external applied bias to deposit Co-Pi. Without being bound by theory, in principle, photogenerated holes can be used to oxidize Co2+ from the electrolyte to form Co-Pi on the cc-Fe203 photoanode and the electron can be removed by water reduction. Because photogenerated electrons in the conduction band of cc-Fe203 are below the energy needed to reduce protons to hydrogen, a very low bias was applied to assist in photoelectrochemical deposition of Co-Pi. Any light source with sufficient energy to excite the band gap of cc-Fe203 can be used in a photoelectrochemical deposition on cc-Fe203. In this embodiment, photoelectrochemical deposition on CC-Fe203 was conducted in a three-electrode configuration from a solution of Co2+ in potassium phosphate (KPi) buffer under 1 sun AM 1.5 simulated solar irradiation. A Pt mesh was used as the counter electrode and saturated Ag/AgCl was used as the reference electrode. Typical current densities during deposition were -1-100 μΑ/cm2.
FIGURE 18 shows a -120 mV cathodic shift in the J-V curve of an a-Fe203 photoanode after Co-Pi deposition measured in 0.1 M KPi at pH 8. Photoelectrochemical deposition and electrochemical deposition, of Co-Pi had similar effect of shifting the onset potential for water oxidation of a-Fe203.
In another exemplary embodiment, a photo-assisted electrochemical deposition approach (i.e., photoelectrodeposition) was used to deposit a cobalt-phosphate water oxidation catalyst ("Co-Pi") onto dendritic mesostructures of cc-Fe203. A comparison between this approach, electrochemical deposition of Co-Pi, and Co2+ wet impregnation showed that photo-assisted electrochemical deposition of Co-Pi yields superior cc-Fe203 photoanodes for photoelectrochemical water oxidation. Stable photocurrent densities of 1.0 mA/cm2 at 1.0 V and 2.8 mA/cm2 at 1.23 V vs RHE measured under standard illumination and basic conditions were achieved. By allowing deposition only where visible light generates oxidizing equivalents, photo-assisted electrochemical deposition provides a more uniform distribution of Co-Pi onto a-Fe203 than obtained by electrochemical deposition. This approach of fabricating catalyst-modified metal-oxide photoelectrodes may be attractive for optimization in conjunction with tandem or hybrid photoelectrochemical cells.
By way of background, the maturation of photoelectrochemical (PEC) water splitting as a viable solar fuels technology has been hindered by the need to identify photoelectrode materials that are simultaneously efficient at solar energy conversion, stable under reaction conditions, and inexpensive. Whereas high solar-to-hydrogen conversion efficiencies of 12.4 % have been demonstrated using semiconductor multilayer devices, these efficiencies are not sustainable even on the one-day timescale because of rapid electrode decomposition. Metal oxides have been widely studied as chemically robust alternatives, beginning with TiC^, but have been limited by various factors including low carrier mobilities, low absorption coefficients, or poor catalytic proficiencies. Hematite (a-Fe203) has emerged as a prototype photoanode for PEC water oxidation because of its balance of visible light absorption (bandgap of 2.1 eV), chemical stability, low cost, and large positive valence band edge potential. Low mobilities
(10"2-10_1 cm2 V"1 s"1) and short hole diffusion lengths (2-4 nm or 20 nm) have generally led to low PEC water oxidation efficiencies in bulk cc-Fe203, but doping and nano structuring have been used to sidestep these shortcomings, by increasing carrier density, decreasing the distance minority carriers have to travel to reach the reactive surface, and increasing semiconductor-electrolyte interfaces. Doping with silicon has been suggested to increase photocurrent densities by several orders of magnitude in mesostructure cc-Fe203 films. Nanowires and nanotubes of cc-Fe203 have also shown increased photocurrent densities relative to bulk, although such structures have so far been limited to absolute one-sun current densities on the order of μΑ/cm2.
Interfacing such mesostructured metal-oxide photoanodes with competent water oxidation catalysts offers one approach to improving their performance. Similar to Nature's photosynthesis, the separation of photon absorption, charge separation, and water oxidation tasks in composite photoelectrodes allows components performing each task to be optimized independently and thereby enables a greater flexibility in the selection of component materials.
Electrochemical deposition of Co-Pi, as described herein (e.g., Example 2), forms an adequate junction between the catalyst and semiconductor for interfacial charge transfer, and the resulting Co-Pi/cc-Fe203 composite photoanodes are stable under photolysis conditions. A kinetic bottleneck was observed with thick layers of Co-Pi that hindered the steady-state turnover of the composite photoanodes, especially at low applied potentials. This kinetic limitation was remediated by reducing the Co-Pi coverage, but at the expense of overpotential. With short electrochemical deposition times, however, Co-Pi was found to deposit preferentially at pinholes, scratches, or other imperfections in the cc-Fe203 film, where more current can flow from the underlying conductive FTO substrate. This inhomogeneity affects the performance of Co-Pi/cc-Fe203 photoanodes by creating areas where the catalyst layer is too thick (kinetic bottleneck), and it influences the reproducibility of the Co-Pi deposition itself.
Ultimately, a stable and efficient water oxidation photoanode is desired, and methods to apply a uniform thin catalyst layer onto highly mesostructured metal- oxide photoanodes, such as cc-Fe203 are therefore needed.
In this Example, we describe photo-assisted electrochemical deposition ("photoelectrodeposition") of Co-Pi onto mesostructured cc-Fe203 photoanodes, and present a comparison between this approach, electrochemical deposition of Co-Pi and Co2+ adsorption. These three approaches are summarized in FIGURE 21A- FIGURE 21C. Among these, photo-assisted electrochemical deposition of Co-Pi is found to yield superior PEC performance by all metrics, including absolute onset potential, cathodic shift of the onset potential, and maximum current density. In combination with recently improved dendritic cc-Fe203 photoanodes, the photo-assisted electrochemical deposition of Co-Pi yields arguably the best overall cc-Fe203 photoanodes for PEC solar water splitting reported to date, with stable current densities of 1.0 mA/cm2 at 1.0 V and 2.8 mA/cm2 at 1.23 V vs RHE measured under standard 1 sun, AM 1.5 illumination conditions at pH 13.6.
Mesostructured cc-Fe203 photoanodes were fabricated on FTO glass by the
APCVD method described in Example 2. Masks with apertures of 6 mm in diameter were applied to define the active surface areas. Co-Pi was electrodeposited onto cc-Fe203 photoanodes by modification of published procedures. A three-electrode cell was used with a-Fe203 as the working electrode, Ag/AgCl as the reference electrode, and Pt mesh as the counter electrode. 0.9 V vs Ag/AgCl was applied in a solution of 0.5 mM cobalt nitrate in 0.1 M potassium phosphate buffer at pH 7. The amount of Co-Pi deposited was controlled by the deposition time, which ranged between 200-500s. Current densities were typically -2-10 μΑ/cm2 during deposition.
Photo-assisted electrochemical deposition of Co-Pi onto mesostructured cc-Fe203 was performed from the same electrolyte composition used for electrochemical deposition, 0.5 mM cobalt nitrate in 0.1 M potassium phosphate buffer at pH 7, but with 1 sun AM 1.5 simulated sunlight illumination. Because conduction-band electrons in CC-Fe203 do not have sufficient potential to reduce water, an external bias (-0.1-0.4 V) was applied. The amount of Co-Pi was again controlled by the deposition time, which ranged between 500-750 s. Current densities were typically -2-5 μΑ/cm2 during deposition.
Following Example 4, Co2+ adsorption onto mesostructured cc-Fe203 photoanodes was achieved by dipping the photoanode in a solution of 0.1 M cobalt nitrate for 5 minutes. The amount of Co2+ adsorbed was optimized by repetition of this dipping process. Typically, PEC enhancement reached its maximum after about three cycles. Subsequent cycles resulted in either no change or a decrease in the PEC performance.
PEC measurements were conducted in 1M NaOH (pH 13.6) using a three-electrode configuration, with the photoanode as the working electrode, Ag/AgCl as the reference electrode, and Pt as the counter electrode. Photocurrent densities were measured with front-side illumination under 1 sun AM 1.5 simulated sunlight using an Oriel 96000 solar simulator equipped with a 150 W Xenon arc lamp and an Oriel AM 1.5 filter. Potentials vs. RHE are calculated using the Nernst equation ERHE = EAg/AgQ + 0.0591(pH) +0.1976 V. Very similar a-Fe203 photoanodes were used for all PEC measurements. The amount of catalyst applied was optimized to give the largest sustainable cathodic shift and overall current density by controlling the amount of catalyst loading, either by adjusting the time of deposition for Co-Pi or the number of cobalt dipping cycles for Co2+ adsorption. Cathodic shifts were calculated as the average voltage shifts in the window where current densities range from 0.5-1.5 μΑ/cm2. For uniformity, reported photocurrent increases with catalyst deposition refer specifically to the difference in photocurrent at 1.1 V vs RHE. Photocurrent onset potentials were calculated by extrapolation to zero current from the linear portion of the J-V curve where current densities range from 0.5-1.5 mA/cm2.
Scanning electron microscopy (SEM) and energy dispersive X-ray (EDX) analyses were performed using a FEI Sirion SEM equipped with an energy dispersive spectrometer. No conductive coating was deposited onto samples for these measurements.
SEM images of a representative mesostructured cc-Fe203 photoanode are shown in FIGURE 22A and 22B. The photoanode possesses the dendritic features typical of CC-Fe203 grown by APCVD. All catalyst-modified photoanodes show similar dendritic features but the images are slightly blurred (FIGURE 22 C-H), suggesting that the catalysts make the surfaces more insulating and hence more susceptible to charging effects from the electron beam. All photoanode surfaces appear uniform except for the one involving electrodeposited Co-Pi (FIGURE 22 C,D), which shows patches of Co-Pi. On the other films, the catalyst itself is not resolved by the SEM measurement, but it can be detected by EDX. EDX measurements on large and small areas of the films from FIGURE 22 E-H yield similar results, indicating uniform cobalt coverage on these length scales (TABLE 1).
TABLE 1
Material Probe area Fe ( ) Co ( )
0C-Fe2C>3 1700 μιη2 38.11
Photodeposited 430 μιη· 33.35 1.00 Co-Pi/a-Fe203 Material Probe area Fe ( ) Co ( )
Photodeposited 0.25 μιη2 35.52 1.06
Co-Pi/a-Fe203 (single nodule)
Electrodeposited 430 μιη2 33.71 1.23
Co-Pi/a-Fe203
Co2+ adsorbed/a-Fe203 1700 μιη2 34.02 0.94
The amount of catalyst on cc-Fe203 photoanodes that yields the largest sustainable
PEC enhancement can be roughly estimated using the EDX results. As expected for surface deposition, increasing the probe depth by increasing the electron acceleration voltage from 10 to 15 keV results in a substantial decrease in the relative cobalt peak intensity. Approximating the probe depth of a 10 keV electron beam to be ~ 200 nm the assumption of a uniform flat surface would yield a Co-Pi thickness of -30 nm, but this value represents an upper limit because of the very high surface roughness of the a-Fe203 mesostructure (roughness ~ 20). The active Co-Pi cluster is believed to possess seven cobalt ions, with a volume of -700 A, from which an upper limit of 34 clusters thickness is obtained. In all likelihood, the actual thickness is substantially smaller. For example, it is interesting to note that the amount of cobalt detected by EDX is about the same for optimized Co-Pi/cc-Fe203 as for Co2+-impregnated cc-Fe203. Co2+ adsorption has previously been suggested to yield only monolayer coverage, implying closer to one monolayer of Co-Pi cluster as well. Overall, these results clearly indicate that Co-Pi/cc-Fe203 composite photoelectrodes optimized for steady-state photocurrents possess far thinner Co-Pi layers than the analogous Co-Pi-coated electrodes used in electrocatalysis. This difference relates to the kinetic bottleneck described previously, which likely reflects the important role of surface electron-hole recombination under PEC conditions.
FIGURE 23A-C compares current-voltage (J-V) characteristics of representative
Co-Pi/cc-Fe203 and Co2+-modified cc-Fe203 photoelectrodes. All photoelectrodes have been optimized to give the largest steady-state cathodic shift and PEC enhancement compared to their parent cc-Fe203 photoanodes. Photo-assisted electrochemical deposition of Co-Pi onto cc-Fe203 (FIGURE 23A) yields the greatest cathodic shift of the onset potential for PEC water oxidation, -170 mV. Similar results were described previously for electrochemical deposition of Co-Pi onto cc-Fe203 following optimization. The best electrodeposited Co-Pi/cc-Fe203 photoanode in this set showed a -100 mV cathodic shift (FIGURE 23B), and the best Co2+-impregnated cc-Fe203 photoanode showed an -80 mV shift (FIGURE 23C).
The current densities for each of these films are stable and reproducible after multiple J-V scans and under illumination for over 72 hours, even after weeks of storage at room temperature in air. FIGURE 28 illustrates the time dependence of the photocurrent density of a Co-Pi/a-Fe203 photoanode prepared by photo-assisted electrochemical deposition, measured at 1.0 V vs RHE in 1 M NaOH under continuous 1 sun, AM 1.5 simulated solar irradiation. The electrolyte was not stirred. The electrolyte was replaced after 75 hrs (dashed line), resulting in recovery of photocurrent density.
Interesting variations in performance are observed from film to film, much of which derives from variations in the underlying cc-Fe203 photoanode performance. Such differences are illustrated in FIGURE 24A-24C, which shows the photocurrent responses of two quite different composite photoanodes in comparison with those of is their parent a-Fe203 photoanodes. The photoanode in FIGURE 24A shows large, stable photocurrent densities at high bias, whereas the one in FIGURE 24B excels at low bias. These differences are due to a small variation in the deposition temperature. These data also emphasize that Co-Pi surface deposition has a similar effect on each parent cc-Fe203 photoanode, despite their absolute performance differences. Both films show comparable cathodic shifts of their photocurrent onset potentials and small enhancements of their maximum photocurrent densities upon deposition of Co-Pi.
Incident-photon-to-current conversion efficiency (IPCE) measurements on a Co-Pi/cc-Fe203 photoanode prepared by photo-assisted electrochemical deposition (FIGURE 24C) show a value of 40% at 400 nm and 1.23 V vs RHE, with large visible-light conversion efficiencies even at lower bias. The photocurrent response spectrum of the Co-Pi/cc-Fe203 photoanode exhibits the same features as cc-Fe203, indicating the primary photoresponse is from the cc-Fe203 mesostructure and not in the
Co-Pi itself. We note the particularly strong response from the indirect bandgap feature at -550 nm relative to many other cc-Fe203 PEC cells. The prominence of this band here is attributed to the very high surface areas of these dendritic cc-Fe203 photoanodes, which allow hole harvesting even following excitation of such a localized transition, and to the role of Co-Pi in facilitating productive use of those holes for water oxidation.
To put the above comparisons on a more quantitative footing, FIGURES 25A-25C summarizes the PEC results obtained from the investigation of a total of 12 catalyst-modified photoanodes, with particular care given to ensuring that they all involved very similar parent a-Fe203 photoanodes as their starting points. The maximum cathodic shift (FIGURE 25A), maximum photocurrent density increase
(FIGURE 25B), and absolute photocurrent onset potentials (FIGURE 25C) of the best photoanodes in each category are plotted as a bar graph. The average performance in each category is indicated by a horizontal line in the top two graphs and by an empty bar in the bottom graph. Plotting one metric vs another confirms the linear relationship between cathodic shift and reduced photocurrent onset potential (FIGURE 26).
Similarly, greater absolute photocurrent densities are strongly correlated with larger increases in photocurrent density upon Co-Pi deposition (FIGURE 27). From these data, it is concluded that photo-assisted electrochemical deposition of Co-Pi onto cc-Fe203 photoanodes yields both a lower onset potential and a greater increase in photocurrent density than either Co-Pi electrochemical deposition or Co2+ surface adsorption.
Specifically, FIGURE 26 illustrates the average cathodic shifts plotted vs average onset potentials for Co-Pi/a-Fe203 photoanodes prepared by photo-assisted electrochemical deposition (P-Dep) and electrochemical deposition (E-dep) of Co-Pi, and for Co2+/a-Fe203 photoanodes prepared by surface adsorption of Co2+ (Co-dip) for the films used to generate the data of FIGURES 25A and 25C. The open symbols represent the parent a-Fe203 photoanodes. These data show a strong correlation between the two performance metrics, with photo-assisted electrochemical deposition of Co-Pi leading to the lowest onset potentials and the greatest cathodic shifts.
FIGURE 27 illustrates the Average photocurrent density increase vs photocurrent at 1.1 V vs RHE (one-sun) for the films used to generate the data of FIGURES 25B and 25C. Photo-assisted electrochemical deposition of Co-Pi yields the largest photocurrent density increases and the highest absolute photocurrent densities. The open symbols (grouped at the base of the dashed line) represent the parent a-Fe203 photoanodes.
Overall, these data clearly reveal the superiority of photo-assisted electrochemical deposition over simple electrochemical deposition for the preparation of Co-Pi/a-Fe203 composite photoanodes. They also illustrate the improvements in cc-Fe203 PEC performance obtained using Co-Pi rather than surface-adsorbed Co2+ as the electrocatalyst.
In addition to their ease of preparation, Earth- abundant composition, and highly stable photocurrent densities, the absolute performances of Co-Pi/cc-Fe203 photoanodes are comparable with those of Ir02/oc-Fe203 photoanodes prepared by attachment of nanocrystals of the well-known water oxidation catalyst, IrC^, onto similar a-Fe203 photoanodes. Compared to the Co-Pi/cc-Fe203 photoanode in FIGURE 24A, the best Ir02/oc-Fe203 photoanode showed a 50 mV greater cathodic shift, a 60 mV lower onset potential, and a -13% larger photocurrent density at 1.23 V vs RHE. An important difference between Co-Pi/a-Fe203 and Ir02/oc-Fe203 photoanodes, however, is that the photocurrent responses of the Ir02/oc-Fe203 photoanodes appear to diminish on short (200 s) timescales because of detachment of the IrC^ particles from the a-Fe203 surface. The Co-Pi/cc-Fe203 composite photoanodes show no similar instability (see supporting information).
Despite the reduced onset potential, a positive voltage must still be applied in order to drive PEC water oxidation using a-Fe203. Ideally, this voltage would be supplied by a photovoltaic (PV) device in a tandem configuration. For the photoanode in
FIGURE 24A, the increase in photocurrent density from 2.1 to 2.8 mA/cm2 at 1.23 V (the thermodynamic potential for electrolysis) following Co-Pi deposition corresponds to a 33% improvement and yields a solar to hydrogen conversion efficiency of /zstll = 3.4%, based on the Gibbs free energy of the reaction and assuming a faradaic efficiency of unity. At 1.43 V, the photocurrent density of 3.3 mA/cm2 corresponds to /zstll = 4.1%.
Unfortunately, most low cost PV devices such as DSSCs provide less than 1.0 V at open circuit, and multiple PV devices connected in series would thus be required to achieve the, above efficiencies. To minimize cost, PEC photocurrent densities at low applied potentials should be optimized, and the cathodic shifts provided by Co-Pi modification are therefore of interest. The Co-Pi/a-Fe203 photoanode of FIGURE 24B shows a relatively high photocurrent density of 1.0 mA/cm2 at 1.0 V vs RHE, which constitutes a 500% improvement over a-Fe203 alone at the same voltage (0.2 mA/cm2).
In summary, photo-assisted electrochemical deposition of Co-Pi onto mesostructured cc-Fe203 yields better performing photoanodes than either electrochemical deposition of Co-Pi or simple Co2+ wet impregnation. A stable -170 mV cathodic shift was observed with photoelectrochemical deposition of Co-Pi, while the electrochemical deposition of Co-Pi gave cathodic shifts of -100 mV, and Co2+ impregnation gave -80 mV cathodic shifts. Photo-assisted electrochemical deposition provides a more uniform distribution of Co-Pi on cc-Fe203 than obtained by electrochemical deposition by allowing deposition only where visible light generates oxidizing equivalents. Optimization of the photo-assisted electrochemical deposition conditions allowed elimination of all nodules and islands to yield thin uniform films of
Co-Pi over the entire photoanode surface. The resulting catalyst-modified metal-oxide photoelectrodes are attractive for solar water oxidation in tandem or hybrid PEC cells.
Example 4. Deposition of cobalt oxide catalysts on oc-Fe2Q3 by deposition from an aqueous solution of Co2+, such as from cobalt nitrate, cobalt acetate or cobalt sulfate.
Electrochemical deposition and photoelectrochemical deposition of a cobalt oxide catalyst, referred to here as "CoOx," on cc-Fe203 were produced by deposition from an aqueous solution of Co2+, such as from cobalt nitrate, cobalt acetate or cobalt sulfate. X-ray diffraction experiments showed that CoOx did not match the typical diffraction patterns of known cobalt oxides, CoO, C02O3, or C03O4. In one embodiment, CoOx was electrodeposited from an aqueous solution of lOmM cobalt nitrate (pH -4) at 0.7-1.4 V vs Ag/AgCl. A Pt mesh was used as the counter electrode and saturated Ag/AgCl was used as the reference electrode. In another embodiment, CoOx was photodeposited on a-Fe203 from the same Co2+ electrolyte under a light bias, 1 sun AM 1.5 simulated solar irradiation, and at 0.1-0.4 V vs Ag/AgCl. FIGURE 19 shows the J-V characteristics of a-Fe203 and the composite CoOx/cc-Fe203 photoanode after electrochemical deposition in 1 M NaOH. Dark current (dotted) and photocurrent (solid) are illustrated. A -100 mV cathodic shift of the onset potential for water oxidation was observed in CoOx-modified
CC-Fe203. A similar effect was also observed in CoOx/cc-Fe203 composite photoanodes after photoelectrochemical deposition of CoOx (FIGURE 20). These data demonstrated that electrochemical deposition and photoelectrochemical deposition of catalysts on CC-Fe203 are not limited to Co-Pi, but are applicable to other Co-based oxygen evolving catalysts as well.
Example 5. Composite Co-Pi/TiOq-NW Photoanode
A photo-assisted electrochemical deposition (photoelectrochemical) approach was employed to achieve selective deposition of Co-Pi onto Ti02 nanowires (NWs).
FIGURE 29 illustrates current density-voltage curves of a Ti02 nanowire photoanode before and after Co-Pi photoelectrochemical deposition, measured under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dashed) in 0.1M potassium phosphate buffer at pH 7. Co-Pi modification results in a -190 mV cathodic shift in the photocurrent.
These data demonstrate that the cobalt-containing catalyst Co-Pi can be photoelectrochemically deposited onto other semiconductor materials of different shapes, such as Ti02 nanowires, as well as onto dendritic a-Fe203 photoanodes. Simple electrodeposition of Co-Pi onto the same Ti02 nanowire structures grown on conductive FTO substrates results in preferential catalyst deposition onto the exposed more- conductive FTO instead and does not improve PEC water oxidation performance. Direct photodeposition of Co-Pi did not result in successful application of the catalyst. By photo-assisted electrodeposition using a wavelength at which only the Ti02 absorbs, Co- Pi was successfully applied specifically to the Ti02 nanowires, yielding the significant cathodic shift in the PEC water oxidation potential. This result also shows that Co-Pi can be used to improve the PEC water oxidation of a semiconductor such as Ti02 with an already low onset potential towards PEC water oxidation. The successful Co-Pi modification of Ti02 nanowires demonstrates the versatility of this photoelectrochemical deposition method to apply cobalt-containing water oxidation catalysts onto semiconductor materials of various shapes and sizes.
Example 6. Composite Co-Pi/ Amorphous Ti02/CdS/Ti02 NW Photoanodes. FIGURE 30 illustrates current density-voltage curves of a Ti02 nanowire photoanode sensitized with CdS nanoparticles coated with a thin amorphous Ti02 protective layer, before and after Co-Pi photoelectrochemical deposition, measured under 1 sun, AM 1.5 simulated solar irradiation in 0.5M sodium thiosulfate.
These data demonstrate that the catalyst Co-Pi can be deposited by photoelectrochemical deposition onto complex electrodes involving visible-light- absorbing sensitizers, such as CdS, integrated with UV light absorbing wide-bandgap semiconductors, such as Ti02, via photoexcitation of the sensitizer and an applied potential. By modifying the Ti02 nanowire surfaces with CdS (bandgap 2.4 eV), the PEC water oxidation electrode is made more sensitive to visible light (i.e., sunlight), as seen by the large photocurrent enhancement. A cathodic shift is also observed after catalyst modification, demonstrating the compatibility of this catalyst deposition method with sensitizers such as CdS and with complex electrodes involving both sensitizers and wide- gap oxides. Example 7. Composite Co-Pi/Co :ZnO Photoanodes.
FIGURE 31 illustrates current density- voltage curves of a Co2+:ZnO photoanode before and after Co-Pi photoelectrochemical deposition, measured under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dotted) in 0.1M potassium phosphate buffer at pH 11.
These data demonstrate that photoelectrochemical deposition can also be applied to deposite Co-Pi onto wide-gap semiconductors doped with cationic impurities (Co2+) introduced to extend PEC water oxidation into the visible region and increase the solar photocurrent densities relative to undoped ZnO. Photoelectrochemical deposition of catalysts onto such doped semiconductors can also be achieved via excitation of mid-gap electronic transitions arising from the dopants, demonstrating that the photoelectrochemical deposition method is not limited to bandgap excitation of semiconductors. Regardless of the electronic transition used for photoelectrochemical deposition, the result is an increase in the overall PEC water oxidation efficiency.
Example 8. Composite Co-Pi/W iBiVCy Photoanodes.
FIGURE 32 illustrates current density- voltage curves of a W-doped BiV04 photoanode before and after Co-Pi photoelectrochemical deposition, measured under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dashed).
These data demonstrate that photoelectrochemical deposition of Co-Pi can also be used to deposition water-oxidation catalysts onto doped ternary semiconductors, such as W:BiV04. The result is a substantial >300 mV cathodic shift in the onset potential for PEC water oxidation and a significantly lower onset potential (<400mV vs RHE) than can be achieved with Co-Pi/cc-Fe203 composite photoanodes. As with a-Fe203, a thin uniform layer of catalyst is desired for large stable photocurrent improvements. Increased deposition of Co-Pi onto W:BiV04 results in decreased PEC performance associated with thick catalyst layers. Photoelectrochemical deposition is a useful approach for applying thin well-dispersed catalyst layers onto various types of semiconductor photoanodes.
Example 9. Composite Cobalt Methyl-Phosphonate/a-FeqO Photoanodes.
FIGURE 33 illustrates current density-voltage curves of an -Fe203 photoanode before and after cobalt methyl-phosphonate (Co-MePi) photoelectrodeposition, measured under 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dotted) in 1 M NaOH, pH 13.6. These data demonstrated that other cobalt-containing water-oxidation electrocatalysts besides Co-Pi, such as Co-MePi, can be successfully photoelectrochemically deposited onto a semiconductor, such as -Fe203 to improve electrode performance. The resulting cathodic shift in the onset potential for PEC water oxidation is very similar to that achieved by Co-Pi deposition, indicating that the photoelectrodeposition method can be expanded to include other oxygen evolving electrocatalysts as well.
Example 10. Composite Nickel Borate/q-Fe2Q3 Photoanodes.
FIGURE 34 illustrates current density- voltage curves of an -Fe203 photoanode before and after nickel borate (Ni-Bi) electrodeposition, measured under backside illumination with 1 sun, AM 1.5 simulated solar irradiation (solid) and in the dark (dotted) in 1 M NaOH, pH 13.6.
These data demonstrate that other electrocatalysts, such as the nickel-based Ni-Bi catalyst, can be applied onto semiconductors, such as a-Fe203, by photoelectrochemical deposition to yield a favorable cathodic shift of the onset potential for PEC water oxidation. This illustration indicates that photoelectrochemical deposition is a general approach for the application of various oxygen evolving electrocatalysts onto a variety of semiconductor photoanodes. While illustrative embodiments have been illustrated and described, it will be appreciated that various changes can be made therein without departing from the spirit and scope of the invention.

Claims

The embodiments of the invention in which an exclusive property or privilege is claimed are defined as follows:
1. A method of forming a composite photoanode, comprising photoelectrodepositing a solid conformal layer of an electrocatalyst from an electrolyte solution on a surface of a semiconductor submerged in the electrolyte solution by simultaneously:
(1) impinging the surface of the semiconductor with electromagnetic radiation having a first wavelength and a first irradiance, to provide a first photoenergy that is sufficient to excite an electronic transition of the semiconductor; and
(2) applying a first electric bias to the semiconductor, wherein the first electric bias is less than an electrochemical deposition bias, said electrochemical deposition bias being the minimum voltage required to electrodeposit the electrocatalyst onto the surface of the semiconductor without impinging the surface of the semiconductor with electromagnetic radiation having the first photoenergy;
wherein the combination of the first photoenergy and the first electric bias are sufficient to deposit catalyst components from the electrolyte to form the solid conformal layer of the electrocatalyst.
2. The method of Claim 1, wherein the electronic transition is a bandgap transition.
3. The method of Claim 1, wherein the semiconductor has a physical shape selected from the group consisting of dendrites, wires, belts, rods, mesostructures, nanotubes, and thin films.
4. The method of Claim 3, wherein said physical shape has nanoscopic dimensions.
5. The method of Claim 1, wherein the semiconductor comprises hematite iron oxide dendrites.
6. The method of Claim 1, wherein the electrocatalyst is selected from the group consisting of a cobalt-containing catalyst, an iridium-containing catalyst, a manganese-containing catalyst, a ruthenium-containing catalyst, a nickel-containing catalyst, a cobalt-containing oxygen evolving catalyst, a cobalt oxide/hydroxide catalyst, and a cobalt oxide catalyst.
7. The method of Claim 1, wherein the electrocatalyst is cobalt phosphate.
8. The method of Claim 1, wherein the layer of the electrocatalyst has a thickness of from 0.5 nm to 30 nm.
9. The method of Claim 1, wherein the semiconductor is an n-type semiconductor.
10. The method of Claim 1, wherein the formed photoanode, when incorporated into a photoelectrochemical cell for electrolysis of water into oxygen, reduces a water electrolysis onset voltage compared to a second photoanode comprising the semiconductor without the electrocatalyst.
11. The method of Claim 10, wherein the water electrolysis onset voltage is reduced by 50 mV to 400 mV.
12. The method of Claim 1, wherein the first wavelength of the electromagnetic radiation is from 300 nm to 800 nm.
13. The method of Claim 1, wherein the first irradiance of the electromagnetic radiation is from 0.1 W/m2 to 1100 W/m2, or the equivalent in pulsed irradiation.
14. The method of Claim 1, wherein the electromagnetic radiation is selected from the group consisting of continuous radiation and pulsed radiation.
15. The method of Claim 1, wherein the first electric bias is applied to the semiconductor as part of an electrochemical deposition system comprising a power source in electrical communication with the semiconductor and a counter electrode.
16. The method of Claim 1, wherein the electrolyte solution comprises cations selected from the group consisting of cobalt, iridium, manganese, nickel, and ruthenium.
17. The method of Claim 1, wherein the electrolyte solution comprises anions selected from the group consisting of phosphate, methyl phosphonate, borate, acetate, sulfate and hydroxide. 18. The method of Claim 1, wherein the semiconductor comprises a sensitizer having a sensitizer absorbance wavelength, said sensitizer absorbance being different from a semiconductor absorbance wavelength.
20. The method of Claim 1, wherein the semiconductor comprises a group IV semiconductor having a formula selected from the group consisting of binary, ternary, and quaternary.
21. The method of Claim 20, wherein the group IV semiconductor further comprises ions selected from the group consisting of cations and anions.
22. The method of Claim 1, wherein the semiconductor comprises a material selected from the group consisting of an iron oxide, a zinc oxide, a titanium oxide, a tungsten-bismuth-vanadium oxide, a tungsten oxide, a gallium-zinc-oxide-nitride, or these materials also containing additional cations or anions.
23. The method of Claim 1, the wherein the combination of the first photoenergy and the first electric bias are sufficient to oxidize cations to deposit catalyst components from the electrolyte to form the solid conformal layer of the electrocatalyst.
PCT/US2011/027603 2010-03-08 2011-03-08 Composite photoanodes WO2011112620A2 (en)

Priority Applications (1)

Application Number Priority Date Filing Date Title
US13/606,439 US20130240364A1 (en) 2010-03-08 2012-09-07 Composite photoanodes

Applications Claiming Priority (2)

Application Number Priority Date Filing Date Title
US31172410P 2010-03-08 2010-03-08
US61/311,724 2010-03-08

Related Child Applications (1)

Application Number Title Priority Date Filing Date
US13/606,439 Continuation US20130240364A1 (en) 2010-03-08 2012-09-07 Composite photoanodes

Publications (2)

Publication Number Publication Date
WO2011112620A2 true WO2011112620A2 (en) 2011-09-15
WO2011112620A3 WO2011112620A3 (en) 2012-01-05

Family

ID=44564086

Family Applications (1)

Application Number Title Priority Date Filing Date
PCT/US2011/027603 WO2011112620A2 (en) 2010-03-08 2011-03-08 Composite photoanodes

Country Status (2)

Country Link
US (1) US20130240364A1 (en)
WO (1) WO2011112620A2 (en)

Cited By (2)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
CN105032467A (en) * 2015-07-22 2015-11-11 宁波工程学院 Photoelectrical catalyst of high purity WO<3>/C<3>N<4> composite mesoporous nano-belt
CN106906488A (en) * 2017-01-18 2017-06-30 常州大学 A kind of method for preparing cobalt hydroxide modified titanic oxide light anode

Families Citing this family (9)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
US9513328B2 (en) * 2012-04-23 2016-12-06 Arizona Board Of Regents On Behalf Of Arizona State University Systems and methods for eliminating measurement artifacts of external quantum efficiency of multi-junction solar cells
EP3476979A1 (en) * 2017-10-31 2019-05-01 Universität Zürich A hybrid catalyst system for a water photooxidation process
CN110302785A (en) * 2019-06-10 2019-10-08 广东工业大学 A kind of unformed cobalt oxide/di-iron trioxide composite photo-catalyst and its preparation method and application
CN111389430B (en) * 2020-05-15 2022-10-25 郑州大学 Catalyst CoP for hydrogen production by water electrolysis x S y MWCNTs and preparation method thereof
CN111593353A (en) * 2020-05-29 2020-08-28 深圳大学 Photoelectrochemistry anti-corrosion protection composite photo-anode and preparation method and application thereof
CN111646536A (en) * 2020-06-18 2020-09-11 闽江学院 Method for directly and photoelectrically degrading basic dye based on FTO conductive glass
CN115074746A (en) * 2021-03-10 2022-09-20 中国科学院大连化学物理研究所 Organic-inorganic semiconductor hybrid double-photoelectrode unbiased photoelectrocatalysis full-decomposition water hydrogen production method
CN113003555B (en) * 2021-03-12 2023-05-26 江南大学 Mesoporous carbon-nitrogen co-doped cobalt-based phosphate material, and preparation method and application thereof
CN115074739B (en) * 2022-07-19 2023-10-13 中国科学院海洋研究所 CdS@TiO for metal photogenerated cathode protection 2 Preparation method and application of NTAs composite material

Citations (3)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
EP1207572A1 (en) * 2000-11-15 2002-05-22 Dr. Sugnaux Consulting Mesoporous electrodes for electrochemical cells and their production method
US20040262154A1 (en) * 2003-06-27 2004-12-30 Gibson Thomas L. Photoelectrochemical device and electrode
US20060243587A1 (en) * 2004-05-05 2006-11-02 Sustainable Technologies International Pty Ltd Photoelectrochemical device

Family Cites Families (5)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
US4626322A (en) * 1983-08-01 1986-12-02 Union Oil Company Of California Photoelectrochemical preparation of a solid-state semiconductor photonic device
US7368045B2 (en) * 2005-01-27 2008-05-06 International Business Machines Corporation Gate stack engineering by electrochemical processing utilizing through-gate-dielectric current flow
US20080035489A1 (en) * 2006-06-05 2008-02-14 Rohm And Haas Electronic Materials Llc Plating process
CN101595535A (en) * 2006-09-24 2009-12-02 肖克科技有限公司 Utilize the auxiliary technology of carrying out plating substrate devices of voltage switchable dielectric material and light
US20100133110A1 (en) * 2008-10-08 2010-06-03 Massachusetts Institute Of Technology Catalytic materials, photoanodes, and photoelectrochemical cells for water electrolysis and other, electrochemical techniques

Patent Citations (3)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
EP1207572A1 (en) * 2000-11-15 2002-05-22 Dr. Sugnaux Consulting Mesoporous electrodes for electrochemical cells and their production method
US20040262154A1 (en) * 2003-06-27 2004-12-30 Gibson Thomas L. Photoelectrochemical device and electrode
US20060243587A1 (en) * 2004-05-05 2006-11-02 Sustainable Technologies International Pty Ltd Photoelectrochemical device

Cited By (2)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
CN105032467A (en) * 2015-07-22 2015-11-11 宁波工程学院 Photoelectrical catalyst of high purity WO<3>/C<3>N<4> composite mesoporous nano-belt
CN106906488A (en) * 2017-01-18 2017-06-30 常州大学 A kind of method for preparing cobalt hydroxide modified titanic oxide light anode

Also Published As

Publication number Publication date
US20130240364A1 (en) 2013-09-19
WO2011112620A3 (en) 2012-01-05

Similar Documents

Publication Publication Date Title
US20130240364A1 (en) Composite photoanodes
Zhang et al. Near-complete suppression of oxygen evolution for photoelectrochemical H2O oxidative H2O2 synthesis
Zhong et al. Photo-assisted electrodeposition of cobalt–phosphate (Co–Pi) catalyst on hematite photoanodes for solar water oxidation
Ma et al. Hierarchical Cu2O foam/g-C3N4 photocathode for photoelectrochemical hydrogen production
Liu et al. Synergistic enhancement of charge management and surface reaction kinetics by spatially separated cocatalysts and pn heterojunctions in Pt/CuWO4/Co3O4 photoanode
Li et al. Surfactant-assisted controlled synthesis of a metal-organic framework on Fe2O3 nanorod for boosted photoelectrochemical water oxidation
Morales-Guio et al. Hydrogen evolution from a copper (I) oxide photocathode coated with an amorphous molybdenum sulphide catalyst
Walter et al. Solar water splitting cells
Kumbhar et al. Interfacial growth of the optimal BiVO4 nanoparticles onto self-assembled WO3 nanoplates for efficient photoelectrochemical water splitting
US8809843B2 (en) Nickel-based electrocatalytic photoelectrodes
Bledowski et al. Improving the performance of hybrid photoanodes for water splitting by photodeposition of iridium oxide nanoparticles
Lee et al. Electrodeposited heterogeneous nickel-based catalysts on silicon for efficient sunlight-assisted water splitting
Zahran et al. Concisely Synthesized FeNiWO x Film as a Highly Efficient and Robust Catalyst for Electrochemical Water Oxidation
Siavash Moakhar et al. AuPd bimetallic nanoparticle decorated TiO 2 rutile nanorod arrays for enhanced photoelectrochemical water splitting
Kim et al. Toward high-performance hematite nanotube photoanodes: charge-transfer engineering at heterointerfaces
Li et al. Electrodeposition of a cobalt phosphide film for the enhanced photoelectrochemical water oxidation with α-Fe2O3 photoanode
Hoogeveen et al. Photo-electrocatalytic hydrogen generation at dye-sensitised electrodes functionalised with a heterogeneous metal catalyst
Chen et al. Surface-and interface-engineered heterostructures for solar hydrogen generation
Wei et al. 2D/3D WO3/BiVO4 heterostructures for efficient photoelectrocatalytic water splitting
Li et al. Assembly of copper phthalocyanine on TiO2 nanorod arrays as co-catalyst for enhanced photoelectrochemical water splitting
Tang et al. Facile synthesis of polypyrrole functionalized nickel foam with catalytic activity comparable to Pt for the poly-generation of hydrogen and electricity
Zahran et al. Nickel sulfate as an influential precursor of amorphous high-valent Ni (III) oxides for efficient water oxidation in preparation via a mixed metal-imidazole casting method
Burungale et al. Surface modification of p/n heterojunction based TiO2-Cu2O photoanode with a cobalt-phosphate (CoPi) co-catalyst for effective oxygen evolution reaction
Fabre et al. Silicon Photoelectrodes Prepared by Low-Cost Wet Methods for Solar Photoelectrocatalysis
Tao et al. Near-infrared-driven photoelectrocatalytic oxidation of urea on La-Ni-based perovskites

Legal Events

Date Code Title Description
121 Ep: the epo has been informed by wipo that ep was designated in this application

Ref document number: 11753952

Country of ref document: EP

Kind code of ref document: A2

NENP Non-entry into the national phase

Ref country code: DE

122 Ep: pct application non-entry in european phase

Ref document number: 11753952

Country of ref document: EP

Kind code of ref document: A2