The present disclosure concerns catalyst compositions and methods of making and using such compositions. More specifically, the catalyst compositions described herein are formed by selective metal leaching from initial ABO3 perovskite compounds having an initial M1M2M3O3 formula wherein the catalysts are useful, for example, for the electrolysis of water under acidic conditions.
Increased energy demands and global warming issues are forcing society to rely more on renewable energy sources, which are often intermittently available. To mitigate this issue, converting electrical energy (provided by renewable energy sources) into chemical bonds through electrocatalysis, such as water electrolysis for hydrogen fuel, is a viable choice for energy storage.
In water electrolysis, the oxygen evolution reaction (OER), which occurs via a four-electron-transfer process, plays a pivotal role in determining the energy conversion efficiency. Its sluggish reaction kinetic greatly constrains the efficiency of the whole reaction, making the development of highly efficient OER catalysts one of the major challenges for implementing water electrolysis. To date, a wide variety of transition-metal-based materials have been explored for catalyzing the OER, both in acid and alkaline environments, and substantial improvements have been achieved. More recently, the existence of OER-induced surface reconstruction, mainly ion leaching and/or structural reorganization, has been widely detected with various catalysts, which range from metal alloys, metal sulfides/selenides/nitrides/phosphides, and metal oxides. Among them, the perovskite-type complex oxides, such as (Ba0.5Sr0.5) (Co0.8Fe0.2)O3-δ and strontium iridate (SrIrO3), have demonstrated superior OER activity due to the presence of unique surface reconstructions. However, although the observations of perovskite surface reconstruction have been intensively reported, designing advanced perovskite pre-catalysts to generate highly active reconstructed surfaces for OER is still a challenge. This, to a great extent, is due to the complexity of the entire reconstruction process, which has not been fully understood.
The present disclosure provides a method for designing advanced perovskite pre-catalysts and using such advanced perovskite pre-catalysts to make catalysts having highly active reconstructed surfaces that are useful for, for example, performing the OER. Catalysts made by the method are also disclosed.
One disclosed embodiment for making a catalyst comprises providing an initial compound having a perovskite lattice structure according to formula I
M1M2M3O3 FORMULA I.
With reference to Formula I, M1 is about 1 relative elemental ratio strontium (Sr); M2 is from greater than 0 to 0.7 relative elemental ratio, preferably 0.5, and is selected from cobalt (Co), scandium (Sc), iron (Fe), nickel (Ni), and titanium (Ti); and M3 is 0.3 to 0.6 elemental ratio, preferably 0.5 elemental ratio, iridium (Ir). Certain initial compounds, comprising 0.5 Sc or Co and 0.5 iridium, included SrSc0.5Ir0.5O3 (SSI) and SrCo0.5Ir0.5O3 (SCI). M1 and/or M2 cations are then selectively leached from the initial compound, such as by electrochemical cycling, to produce a catalyst having substantially increased catalytic performance compared to the initial compound. Exemplary catalysts include SSI-H, SCI-H, SSI-OH and SCI-OH.
Leaching metal atoms, such as strontium atoms, from an initial crystalline perovskite lattice forms an amorphous surface having reduced metal concentration. Electrochemically cycling involves cycling the initial compound plural times in an acid, such as perchloric acid (HClO4), or a base, such as a metal hydroxide, exemplified by KOH. The initial compounds may be electrochemically cycled until a compositional steady state is reached, a structural steady state is reached, and/or a catalytic activity steady state is reached. Cycling SSI or SCI in an acid produces SSI-H or SCI-H, whereas cycling SSI or SCI in a base produces SSI-OH or SCI-OH. Without being limited to a particular theory of operation, cycling is believed to: reconstruct the perovskite surface from a crystalline structure to an amorphous structure with A-site cation (Sr) leaching, which induces an activity improvement of approximately one order of magnitude; leach B-site cations, which induces further activity improvement of approximately one order of magnitude; increases surface area available for catalytic activity; and any and all combinations thereof.
Metal leaching from the initial compound forms a highly active amorphous IrOxHy surface phase where X and Y fulfill an equation 4+Y=2X. For certain embodiments, X is 3, Y is 2, and the catalyst had an amorphous H2IrO3-honeycomb structure and an electrochemical surface area substantially higher than that of the initial compound. The catalyst typically has an amorphous surface structure having a depth of greater than 0 nanometers to at least 50 nanometers, more particularly 10 nanometers to at least 50 nanometers. For certain embodiments, subsequent to cycling in acid, the compound was SSI-H or SCI-H and the strontium surface concentration was reduced to substantially 0 relative elemental ratio to 0.2 relative elemental ratio. For other embodiments, subsequent to cycling in base, the compound was SCI-OH and the strontium surface concentration was reduced to 0.6 relative elemental ratio to 0.7 relative elemental ratio.
Cycling substantially increases catalytic activity. SCI-H, for example, has a geometry-surface-area-normalized OER current (jgeo) at 1.5 V versus a reversible hydrogen electrode (RHE) that is 150 times greater than that of SSI-OH. For particular embodiments, the catalyst is SCI-H having a Brunauer-Emmett-Teller (BET) normalized activity (jgeo) of 7.5±1.0 mA cm2; the catalyst is SSI-H having a BET-normalized activity (jgeo) of 3.5±0.5 mA cm−2; the catalyst is SCI-OH having a BET-normalized activity (jgeo) of 0.4±0.1 mA cm−2; or the catalyst is SSI-OH having a BET-normalized activity (jgeo) of 0.05±0.01 mA cm−2. The amorphous IrOxHy surface phase has an intrinsic activity (jECSA), which refers to the current density normalized to electrochemical surface area (ECSA) at 1.5 V versus RHE as shown by
The catalysts of the present disclosure may be particularly formulated for use in an acid environment. For example, certain disclosed embodiments are designed for use in an acidic environment having a pH of 3.0 or less, preferably a pH of 2.3 or less.
A particular disclosed method embodiment for making catalysts according to the present invention comprises providing an initial compound having a perovskite lattice structure according to formula I
M1M2M3O3 FORMULA I.
With reference to Formula I, M1 is about 1 relative elemental ratio strontium (Sr); M2 is from greater than 0 to 0.7 relative elemental ratio and is selected from cobalt (Co) and scandium (Sc); and M3 is 0.3 to 0.6 relative elemental ratio iridium (Ir). Two representative initial compounds are SrSc0.5Ir0.5O3 (SSI) and SrCo0.5Ir0.5O3 (SCI). M1 and/or M2 cations are then selectively leached from the initial compound, such as by electrochemically cycling the initial compound plural times in a base or an acid, thereby producing a catalyst having substantially increased catalytic performance compared to the initial compound. Electrochemical cycling continues until the initial compound reaches a compositional steady state, a structural steady state, and/or a catalytic activity steady state. For certain disclosed embodiments, the initial compound was cycled approximately 50 times to achieve a suitable steady state. The initial compound may be electrochemically cycled in an acid, which produces SSI-H or SCI-H compounds. For SSI-H and SCI-H, the strontium surface concentration was substantially reduced to be within the range of substantially 0 elemental ratio to 0.2 elemental ratio. The initial compound also can be electrochemically cycled in a base, which produces SSI-OH or SCI-OH. For SSI-OH and SCI-OH, the strontium surface concentration was reduced to be within the range of between 0.6 elemental ratio to 0.7 elemental ratio. Electrochemically cycling the initial compound leaches cations from a surface portion of the initial compound, thereby forming a highly active amorphous surface phase having a depth of greater than 0 nanometers to at least 50 nanometers. For certain disclosed embodiments, the highly active amorphous surface phase was an H2IrO3-honeycomb phase.
A more particular embodiment of the present invention comprises calcining appropriate stochiometric amounts of reagents selected from SrCO3, IrO2, Co3O4, and Sc2O3 (Sigma Aldrich, 99.9%) at a temperature of 1,100° C. or greater to form an initial compound selected from SrSc0.5Ir0.5O3 (SSI) or SrCo0.5Ir0.5O3 (SCI). Again, the initial compound is electrochemically cycled in an acid to produce SSI-H or SCI-H, or electrochemically cycled in a base to produce SSI-OH or SCI-OH, thereby forming a catalyst having a highly active amorphous H2IrO3-honeycomb surface phase having a depth of greater than 0 nanometers to at least 50 nanometers.
The present invention also concerns catalysts produced according to disclosed method embodiments. Certain disclosed catalyst embodiments comprise a core portion having a formula I
M1M2M3O3 FORMULA I.
With reference to Formula I, M1 is about 1 relative elemental ratio strontium (Sr); M2 is from greater than 0 to 0.7 relative elemental ratio and is selected from cobalt (Co) and scandium (Sc); and M3 is 0.3 to 0.6 relative elemental ratio iridium (Ir). Disclosed catalysts also comprise an outer surface portion from which M1 and/or M2 cations have been selectively leached in an acid, such as perchloric acid, thereby reducing the strontium concentration in the outer surface portion to a range between 0 relative elemental ratio to 0.2 elemental ratio. Alternatively, M1 and/or M2 cations may be selectively leached from the initial compound in a base, thereby reducing the strontium concentration in the outer surface portion to a range between 0.6 elemental ratio to 0.7 elemental ratio relative to the core portion concentration. The outer surface portion extends from the surface of the catalyst to a depth of at least 50 nanometers. Metal leaching forms a highly active amorphous IrOxHy surface phase where X and Y fulfill an equation 4+Y=2X. For certain embodiments, X is 3, Y is 2, and the catalyst had an amorphous H2IrO3-honeycomb structure and an electrochemical surface area substantially higher than that of the initial compound.
Catalysts produced according to the present invention can be used to perform catalytic reactions. For example, disclosed catalysts can be used to perform a water oxidation reaction.
The foregoing and other objects, features, and advantages of the invention will become more apparent from the following detailed description, which proceeds with reference to the accompanying figures.
where Q is the integral charge from redox peak (Ir4+ to Ir3+), e is the charge of a single electron. The calculated surface area is 1.03 cm2 and 0.71 cm2 for the (100) facet and (001) facet, respectively. These two surface areas are close to the measured ECSA of 1.25±0.32 cm2 from impedance analysis (
The following explanations of terms and abbreviations are provided to better describe the present disclosure and to guide those of ordinary skill in the art in the practice of the present disclosure. As used herein, “comprising” means “including” and the singular forms “a” or “an” or “the” include plural references unless the context clearly dictates otherwise. The term “or” refers to a single element of stated alternative elements or a combination of two or more elements, unless the context clearly indicates otherwise.
Unless explained otherwise, all technical and scientific terms used herein have the same meaning as commonly understood to a person of ordinary skill in the art to which this disclosure pertains. Although methods and materials similar or equivalent to those described herein can be used in the practice or testing of the present disclosure, suitable methods and materials are described below. The materials, methods, and examples are illustrative only and not intended to be limiting. Other features of the disclosure are apparent from the following detailed description and the claims.
The disclosure of numerical ranges should be understood as referring to each discrete point within the range, inclusive of endpoints, unless otherwise noted. Unless otherwise indicated, all numbers expressing quantities of components, molecular weights, percentages, temperatures, times, and so forth, as used in the specification or claims are to be understood as being modified by the term “about.” Accordingly, unless otherwise implicitly or explicitly indicated, or unless the context is properly understood by a person of ordinary skill in the art to have a more definitive construction, the numerical parameters set forth are approximations that may depend on the desired properties sought and/or limits of detection under standard test conditions/methods as known to those of ordinary skill in the art. When directly and explicitly distinguishing embodiments from prior art, the embodiment numbers are not approximates unless the word “about” is recited.
The present disclosure concerns novel catalyst compounds and methods for making and using such compounds. The catalysts are produced from initial compounds by electrically cycling such compounds in a base or acid electrolyte, which leaches cations, such as strontium, from the initial compound. Cation leaching continues with continued cycling and results in reconstructing the surface of the initial compound to have both a new composition and a new 3-dimensional structure. This produces a catalyst having substantially enhanced activity relative to the initial compound.
Initial compounds according to the present invention generally are ABO3 perovskites of Formula I
M1M2M3O3 FORMULA I.
For most inorganic ABO3 perovskites, the A-site is occupied by alkaline-earth-metals and lanthanides, which, with their relatively large ionic size (>1 Å), are indispensable for supporting the framework of corner-shared B-site octahedra. Some A-site cations are soluble in water, even in alkaline conditions. As a result, a heavy dissolution of A-site cations is widely observed during the surface reconstruction in benchmark perovskite catalysts. The B-site can be occupied by various transition metals, which are active toward catalyzing OER. Normally, transition metals with good thermodynamic stability are used as B-site cations. In alkaline conditions, Fe, Co, and Ni are generally used, while Ir is used in acidic conditions because of its high corrosion resistance. The ABO3 perovskite may have, for example, strontium as an A site element, and a transition metal or metal pair B site, such as an Sc/Co B site.
As a result, with reference to Formula I, specific compounds of the present invention include those where M1 is 1 elemental ratio strontium (Sr); M2 is from greater than 0 elemental ratio to 0.7 elemental ratio, typically 0.5 elemental ratio, and is selected from cobalt (Co), scandium (Sc), iron (Fe), nickel (Ni) and titanium (Ti), preferably cobalt and scandium; and M3 is 0.3 elemental ratio to 0.6 elemental ratio, typically 0.5 elemental ratio, Iridium (Ir).
Two particular exemplary initial compounds exemplifying the present disclosure are SrCo0.5Ir0.5O3 (also referred to as SCO) and SrSc0.5Ir0.5O3 (also referred to as SCI). These initial compounds are prepared using SrCO3 (Sigma Aldrich, 99.9%), IrO2 (Sigma Aldrich, 99.9%), Co3O4 (Sigma Aldrich), and Sc2O3 (Sigma Aldrich, 99.9%). Appropriate stochiometric amounts of these reagents were thoroughly ground and calcined for 12 hours under ambient air at 1,150° C. to produce SrCo0.5Ir0.5O3, and at 1,350° C. to produce SrSc0.5Ir0.5O3.
The catalytic activity of initial compounds according to Formula I are substantially improved by cycling in either a base, such as 0.1 KOH, or in an acid, such as HClO4. Cycling can be accomplished, for example, by forming an electrode comprising the initial compound, and cycling the electrode for 50 cyclic voltammetry cycles (between 0.3 V and 1.8 V vs. RHE without iR correction) to ensure that the reconstructed surface reaches a steady status. Cycling SrCo0.5Ir0.5O3 in an acid, such as 0.1 M HClO4, produces a new structure, referred to herein as SCI-H, whereas cycling SrSc0.5Ir0.5O3 in 0.1 M HClO4 produces a new structure referred to herein as SSI-H. Similarly, cycling SrCo0.5Ir0.5Oin a base, such as 0.1 M KOH, produces SCI-OH, and cycling SrSc0.5Ir0.5O3 in 0.1 M KOH produces SSI-OH.
Cycling also changes the physical structure of the initial compound to form an amorphous surface structure having a depth of greater than 0 nanometers to at least 50 nanometers, and more typically a depth of 10 nanometers to 50 nanometers. The amorphous structure is an IrOxHy structure that likely is IrO2-rutile, H2IrO3-honeycomb, H2IrO3-F, or IrOOH-brucite, most likely H2IrO3-honeycomb. X and Y fulfill an equation 4+Y=2X, and for certain exemplary embodiments, X=3 and Y=2.
Cation leaching forms the highly active amorphous IrOxHy surface phase. For example, the amorphous IrOxHy surface phase may have an intrinsic activity that is more than two orders of magnitude higher than the activity of a rutile IrO2. The activities of SSI-OH, SCI-OH, SSI-H and SCI-H are provided by
In most inorganic ABO3 perovskites, the A-site is occupied by alkaline-earth-metals and lanthanides, which, with their relatively large ionic size (>1 Å), are indispensable for supporting the framework of corner-shared B-site octahedra. Some A-site cations are soluble in water, even in alkaline conditions. As a result, a heavy dissolution of A-site cations is widely observed during the surface reconstruction in benchmark perovskite catalysts. The B-site can be occupied by various transition metals (TMs), which are active towards catalyzing OER. Normally, TMs with good thermodynamic stability are employed as B-site cations. For example, in alkaline conditions, Fe, Co, and Ni are generally utilized, while Ir is used in acidic conditions due to its high corrosion resistance. Thus, the leaching of B-site cations is rather weak as compared to A-site cations. Cation leaching is generally accompanied by a considerable increase in electrochemical surface area for OER. Nonetheless, deliberate cation leaching (as sacrificial agent) from the initial bulk can induce the formation of unique local structural environments, such as reactive surface hydroxyls and activated oxygen ligands, which can promote activity. Given that a typical perovskite contains two types of metal cations (A-site and B-site) in totally different structural environments and both of them can leach out during surface reconstruction, identifying the role of metal cation leaching at each site is pivotal for understanding the relationship between the activity evolution and the surface reconstruction. On the other hand, a consensus is that the reconstructed surface, in contact with an electrolyte, is the final active phase towards OER. Therefore, a better understanding of the formed surfaces is essential too.
Prior to the present disclosure, the formed active surface phases after metal cation leaching have been elucidated case-by-case. A resemblance of the reconstructed surface phases to the (oxy)hydroxides has been suggested based on the detection of edge-shared octahedra with X-ray absorption analysis. The formation of phases, akin to the initial perovskite structure, have also been predicted by density functional theory (DFT) calculations in a SrIrO3 perovskite. Nevertheless, probing the exact structural motif of the surface phase is still challenging as the perovskite surface is generally amorphized after the reconstruction.
The present disclosure presents a step-by-step strategy to control the metal cation leaching from each geometric site in perovskites, such as two Ir-based exemplary perovskites, to understand their activity evolution and the role of the metal leaching at each site. The perovskites of SrSc0.5Ir0.5O3 (SSI) and SrCo0.5Ir0.5O3 (SCI) are employed as model catalysts. Metal cation leaching and the accompanied surface reconstruction can be controlled by tailoring the thermodynamic stability of B-site cations. A thorough reconstruction, including the metal cation leaching and structural rearrangement, induces a remarkable activity improvement by approximately 150 times (1.5 V vs. reversible hydrogen electrode (RHE)), which makes SCI among the best catalysts for OER in acid. A-site cation leaching creates more electrochemical area available for catalyzing OER and the additional B-site cation leaching induces the formation of a highly active amorphous IrOxHy surface phase, which has an intrinsic activity more than two orders of magnitude higher than the activity of a rutile IrO2. The surface-sensitive X-ray absorption analysis and DFT simulations indicate a honeycomb-like structure of the reconstructed amorphous IrOxHy.
To evaluate activity evolution during surface reconstruction and to identify the effects of metal leaching, the ability to precisely manipulate surface metal leaching was a prerequisite. The present disclosure provides a surface reconstruction mechanism, starting from the perovskite structure, which mechanism allows precise control of the perovskite surface metal leaching and accompanied reconstruction.
Strontium dissolution includes two stages (
As shown by
GIS4 and Gfinal are the free energies of the surfaces before and after the desorption of surface Sr, respectively. μSr is the chemical potential of the Sr atom. Such chemical potential is estimated by calculating the chemical potential of a Sr metal model, in which a face-centered cubic Sr crystal is constructed for calculation.
The second sub-step is the dissolution of atomic Sr into the electrolyte (Sr=Sr2++2e−). The free-energy change for dissolution of atomic Sr (ΔG2) can be expressed as:
Where μSr
Substitution of the above two equations into equation (2) gives
where ΔG602 SHE and aSr
At the ideal surface of SrSc0.5Ir0.5O3 without any B-site (Sc) defects, the migration of a lattice Sr from sub-surface to the outer-surface requires an activation energy (Ea) of 3.03 eV (
A similar decrease in activation energy (from 2.45 eV to 1.95 eV) is obtained in SrCo0.5Ir0.5O3 by creating a surface B-site (Co) vacancy (
Two model SSI and SCI perovskites were synthesized via a solid-state method as discussed below in Example 1. Their double perovskite structures, having a chemical formula A2(BB′)O6 as compared to perovskite's ABO3 formula, with high phase purity, are confirmed by Rietveld refinement of the XRD patterns (
The local crystal structure of model perovskites was determined using high-resolution transmission electron microscopy (HR-TEM). As shown in
The initial states of Ir in the lattice of both perovskites were studied with X-ray absorption spectroscopy (XAS). As shown by
For a better understanding of the electronic structure of Ir in perovskites, the second derivative of Ir Lm-edges spectra (white lines) of SSI and SCI are plotted in
The local structure environment around Ir in the lattice was studied based on the corresponding extended X-ray absorption fine structure (EXAFS) at the Ir Lm-edge (
Surface reconstruction of catalysts during electrochemical cycling usually is accompanied with gradually increased double-layer capacitance (DLC), intensity of redox peaks, and catalytic activity. The activity improvement is generally related to increased electrochemical area and/or the formation of a more active surface phase(s) after reconstruction. Cyclic voltammetry (CV) measurements were first used to understand the surface reconstructions of model perovskites.
Additional redox peaks above ˜1.2 V versus RHE are also observed. These peaks can be related to the redox transition of Ir4+/Ir5+, which occurs at high overpotentials. Interestingly, during the cycling tests, redox peak intensities and the OER current gradually decrease. In SSI, Sr2+ and Sc3+ have fixed oxidation states and therefore do not contribute to the redox peaks. Thus, the gradual disappearance of redox peaks in the CV indicates the loss of surface-active sites (Ir) in SSI. Although Ir can be one of the most stable elements, the highly oxidized Ir6+ (IrO42−) can be soluble. Thus, Ir in the outermost surface, which contributes to the measured OER catalytic activity, can dissolve during the cycling. Meanwhile, the DLC remains unchanged during the cycling, hinting that the surface of SSI can still be highly stable. According to the HR-TEM image of the SSI after the cycling (
IrO2-related nanocrystallites form in five anodic cycles but disappear when the number of anodic cycles increases to 130. However, with similar electrochemical tests, the reconstructed surface of a polycrystalline monoclinic SrIrO3 (with mixed edge-/corner-shared octahedrons) is strictly amorphous and no nanocrystallites can be observed. Given their identical composition, the substantial difference in crystal structure may account for the formation of reconstructed surfaces with different properties. This deduction is also supported by considering that the reconstructed Ir-based surface is more active if the IrO6 octahedrons are corner-shared in the initial perovskite structure. IrO6 octahedrons in the initial SCI and SSI compounds are also corner-shared. However, the IrO2-related nanocrystallites cannot be found over the reconstructed surfaces perhaps because the presence of foreign B-site metals (Co and Sc) prohibits the evolution of rutile IrO2.
The stable surface of SSI-OH is further supported by XPS results (
In low pH-value solutions, Sc is thermodynamically unstable (
The CV profiles of SrCO0.5Ir0.5O3 cycled in acid (SCI-H) are shown in
The surface composition changes of all four samples are summarized in
To highlight the promoting effect of surface reconstruction for OER, the geometry-surface-area-normalized (GEO-normalized) activities of SSI-OH, SCI-OH, SSI-H, and SCI-H are shown in
The ECSA was estimated using advanced impedance spectrum analysis. More information is provided with reference to
Representative impedance spectra from SSI-OH, SCI-OH, SSI-H, and SCI-H are provided by
Where T is a frequency-independent constant with F(P−1)m−2 units, I is the square root (−I),
w is the angular frequency of the AC signal, and P is a parameter ranging from 0 to 1. The electrochemical double-layer capacitance can be obtained with the equation of
The electrochemical surface area can be calculated with
Where Cs is the specific capacitance of the sample. In
Interestingly, all Tafel plots for the three surface-reconstructed catalysts SCI-OH, SSI-H, and SCI-H exhibited prominent “bending” behavior with a change of Tafel slope from ˜40 mV/dec at low potentials to ˜100 mV/dec at high potentials. Specifically, the Tafel plot of SCI-OH shows a bending at a potential of ˜1.53 V, while the bending of Tafel plots from SSI-H and SCI-H starts at a lower potential of ˜1.49 V. More recently, a universal linear relationship between log (current) and pseudocapacitive charge storage (catalyst deprotonation with hole formation) has been demonstrated for rutile IrO2. The observed bending of Tafel plot at ˜1.58 V is related to the change in the response of IrO2 surface hole coverages to the potential applied. With pulse voltammetry tests, the logarithm of OER current from SCI-OH and SCI-H was also proportional to the corresponding charge stored during OER (
To further confirm that the high intrinsic activity is related to the formation of superior Ir sites, the turnover frequencies (TOF) of Ir from different surfaces were calculated by taking structural effects (the arrangement of surface Ir atoms) into account. Based on well-defined surfaces, the overpotentials required to reach a TOF of 0.03 s−1 are compared in
To better understand the structure of the reconstructed surface, especially the formed highly active Ir, the local structure and oxidation state of Ir in the reconstructed surface region were characterized by XAS using total-electron-yield (TEY) detection. SCI-H, which has a fully reconstructed surface, was used as a model sample.
The local structural environment of Ir in the reconstructed SCI-H surface was then studied by EXAFS.
On the other hand, the typical peak reflecting corner-shared IrO6 octahedra in rutile IrO2 does not appear in the spectrum of SCI-H. The additional fitting of the first peak revealed that the Ir center in the reconstructed SCI surface is fully coordinated with six oxygen atoms (Table 3). However, due to the reduction of Ir5+after surface reconstruction, the average Ir—O bond length increases to 1.973 Å, which is higher than 1.950 Å in pristine SCI but comparable to that in rutile IrO2 (1.983 Å). In addition to the bond length, the Debye-Waller (DW) factor, which corresponds to the mean-square-displacement of the Ir—O bond length due to the vibration and/or the static disorder, can be obtained from the fitting. For the vibration, a longer Ir—O bond length with stronger thermal vibration should induce a larger DW factor. As a result, a positive correlation between Ir—O bond length and DW factor has been found in Ir-based perovskites (the dashed line in
Soft XAS (in TEY mode) characterization at the O K-edge was performed to better assess the effect of surface reconstruction on the local electronic state of Ir. As a result of the low energy of soft X-ray, the probing depth of soft X-ray in TEY mode is around a couple of nanometers. This makes the O K-edge spectrum highly sensitive to the surface. Because the unoccupied oxygen 2p band hybridizes with the unoccupied metal bands, the O K-edge spectrum can reflect the surface electronic structure changes before and after reconstruction.
Similar changes can be also observed in the O K-edge spectra of the STEM-EELS analysis (
The observed evolution of O K-edge spectra corresponds well with the detected surface reconstructions in SSI and SCI (
To explore the most likely structure of the IrOxHy phase in the reconstructed perovskite surface, the O K-edge spectrum of rutile IrO2 was measured and compared to that of SCI-H with a reconstructed surface. As shown by
As displayed in
Note that, if the honeycomb character is ignored, the structure of H2IrO3 is similar to those of transition-metal-(oxy) hydroxides, which are also popular OER catalysts. Additionally, the formation of certain (oxy) hydroxide(s) that feature edge-sharing octahedra has also been considered as the real active phase(s) of some highly active complex oxides with surface reconstruction. On the other hand, a layered IrOOH has also been synthesized for catalyzing OER, but the activity of this IrOOH is reported to be inferior to the rutile IrO2. Considering that the intrinsic activity of the Ir (oxy) hydroxide(s) in the amorphous perovskite surface is superior to both rutile IrO2 and perovskite SSI (
Considering the unusual Tafel plot bending (
Additionally, the OER free energy diagrams were also computed to investigate the thermodynamic features of H2IrO3. As shown by
Reconstruction-induced activity improvement for compounds disclosed herein is likely due to two factors. First, the surface reconstruction with A-site metal cation leaching makes more electrochemical area available for OER. The second is the formation of a highly active IrOxHy phase in thoroughly reconstructed surfaces with mixed A-site and B-site metal cation leaching, and the B-site cation leaching is pivotal to the formation of such an active phase. Subsequently, with surface-sensitive O K-edge spectra as fingerprints, the active phase possesses a honeycomb-like structure, which is responsible for the high activity. The activity of SCI-H after surface reconstruction is among the best toward water oxidation in acid. Given that surface reconstruction with ion leaching has been intensively observed in the catalysts for electrocatalysis, the step-by-step leaching strategy disclosed herein can be extended to other complex catalysts to investigate the roles of element leaching in surface reconstruction processes for better catalyst design.
The following example is provided to illustrate certain features of the present invention. A person of ordinary skill in the art will appreciate that the scope of the invention is not limited to the features of this particular example.
This example concerns the synthesis of model perovskites. Both SrCo0.5Ir0.5O3 and SrSc0.5Ir0.5O3 were synthesized using a solid-state reaction. Stoichiometric amounts of SrCO3 (Sigma Aldrich, 99.9%), IrO2 (Sigma Aldrich, 99.9%), Co3O4 (Sigma Aldrich), and Sc2O3 (Sigma Aldrich, 99.9%) were thoroughly ground and calcined at 1150° C. (for SrCo0.5Ir0.5O3) or 1350° C. (for SrSc0.5Ir0.5O3) for 12 hours under ambient air.
The electrodes were prepared by drop-casting as-prepared catalyst ink on a glassy carbon rotating electrode with a diameter of 5 mm (Pine Research Instrumentation). The catalyst loading was fixed at 0.05 mg. Specifically, the ink was prepared by mixing 2.5 mg catalyst powder with 1 mg acetylene black carbon, which was ultrasonically dispersed in a solution containing 375 μL H2O, 112.5 μL isopropanol, and 12.5 μL Nafion solution (5 wt. %, Sigma Aldrich). 10 μL of well-dispersed ink was drop-casted onto the polished glassy carbon electrode, which was dried in ambient air until a robust catalyst layer formed. Note that, for evaluating the intrinsic activity (normalized to ECSA) of different samples and for estimating charge storage with pulse voltammetry, the catalyst inks were prepared without acetylene black carbon. The electrochemical tests were conducted in either 0.1 M KOH or 0.1 M HClO4. OER measurements were performed with a biologic SP-150 potentiostat coupled with the modulated speed rotator (Pine Research Instrumentation). The glassy carbon electrode was used as the working electrode, a Pt wire was used as the counter electrode, and a saturated calomel electrode (SCE) was used as the reference electrode. The rotation speed for all tests was fixed at 1600 rpm. At least three measurements were performed when evaluating the OER activities of different catalysts. The pulse voltammetry was performed in the same electrochemical setup used for activity evaluation. Before pulse voltammetry tests, the electrodes were pre-treated for 50 cyclic voltammetry cycles (between 0.3 V and 1.8 V vs. RHE without iR correction) to ensure the reconstructed surfaces reaching the steady status. In pulse voltammetry, a low potential of 1.35 V, below the onset of OER, was selected, and the high potential changed from 1.42 V to 1.8 V with a step of 20 mV. For both anodic and cathodic sections, the duration was fixed at 10 seconds and the current was recorded every 0.001 second.
X-ray powder diffraction (XRD) measurements were performed with a BRUKER D8 Advance diffractometer in Bragg-Brentano geometry with Cu Kα radiation. A GSAS program and EXPGUI interface were used for the Rietveld refinement. TEM was performed on JEOL 2100F with UHR configuration. The EELS were collected with a Gatan 963 Quantum GIF SE, and the spectrum was processed with GMS3 software. The X-ray photoelectron spectroscopy (XPS) tests were performed using PHI-5400 equipment with Al Ka beam source (250 W) and position-sensitive detector. An XPSpeak41 software is applied for peak fitting. The BET surface areas were measured with nitrogen adsorption-desorption tests (ASAP Tri-star II 3020).
The samples for ex-situ XAS measurements were collected by performing the tests on a large working electrode with increased catalyst loading. Specifically, 50 mg of catalyst were loaded onto a large carbon paper (3*3 cm2). The cycling tests were performed in a three-electrode system (single cell) without electrode rotating. The hard XAS measurements at Ir L-edge and Co K-edge were performed at beamline 9-BM of the Advanced Photon Source (APS) at Argonne National Laboratory. The Athena and Artemis software packages were used for the data analysis. Soft XAS measurements (O K-edge) were carried out at 4-ID-C at APS. Calculations of the O K-edge XANES were performed using the finite difference method as implemented within the Finite Difference Method Near Edge Scattering (FDMNES) package using a free form SCF potential of radius 6.0 Å around the absorbing atom. Broadening contributions due to the finite mean-free path of the photoelectron and to the core-hole lifetime were accounted for using an arctangent convolution.
The overpotential required to reach a TOF of 0.03 s−1 is obtained from
where e is the electric charge carried by a single electron, and ρIr is the surface density of Ir atoms. While calculating the ρIr, different surface atom arrangements are considered. For IrO2 (110) film with a stable surface, the (110) facet is considered and the lattice parameters are from the reference. For SSI-OH with a stable surface, the B-site Sc and Ir were assumed to be fully ordered to simplify the calculation. Two cases of the (100) facet (with the lowest Ir density) and (001) facet (with the highest Ir density) were considered. The refined lattice parameters from Table 2 were used for calculations. For SCI-OH, SSI-H, and SCI-H with reconstructed surfaces, two optimized structures of H2IrO3 (honeycomb) and IrOOH (brucite) were considered. In both structures, the (001) facet (with the highest Ir density) was used. The lattice parameters are obtained from DFT calculations.
The spin-polarized DFT calculations were performed using the Vienna Ab Initio Simulation Package, employing the projected augmented wave (PAW) model. The exchange and correlation effect was described by Perdew-Burke-Ernzerhof (PBE) functional. J. P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865 (1996). The GGA+U calculations were performed using the model proposed by Dudarev et al. [S. Dudarev, G. Botton, S. Savrasov, C. Humphreys, A. Sutton, Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+U study. Phys. Rev. B. 57, 1505 (1998)], with the Ueff(Ueff=Coulomb U−exchange J) values of 1 eV, 3.3 eV, and 3 eV for Ir, Co, and Sc, respectively. In all the calculations, the cutoff energy was set to be 450 eV. The Monkhorst-Pack k-point meshes were set to be 6×6×5 and 2×2×1 for performing the bulk and surface calculations of the perovskite structure, respectively. H. J. Monkhorst, J. D. Pack, Special points for Brillouin-zone integrations. Phys. Rev. B. 13, 5188 (1976). The force and energy convergence tolerance were set to be 0.05 eV Å−1 and 10−5 eV, respectively.
The OER free energies were calculated based on the following four elementary steps:
OH-+*→*OH+e−
*OH+OH-→*O+H2O+e−
*O+OH-→*OOH+e−
*OOH+OH-→+O2+H2O+e−
where * denotes the cation sites on the catalyst surface. Based on the above mechanism, the free energies of the three intermediate states, *OH, *O, and *OOH, are crucial in determining the OER activity of a given material.
The computational hydrogen electrode (CHE) model [J. K. Nørskov, J. Rossmeisl, A. Logadottir, L. Lindqvist, J. R. Kitchin, T. Bligaard, H. Jonsson, Origin of the overpotential for oxygen reduction at a fuel-cell cathode. J. Phys. Chem. B. 108, 17886-17892 (2004)] was used to evaluate the energy state of the OER intermediates, where the free energy of an adsorbed species is defined as
With reference to this formula, ΔEads is the electronic adsorption energy, ΔEZPE is the zero-point energy difference between adsorbed and gaseous species, and TΔSads is the corresponding entropy difference between these two states. The electronic binding energy is referenced as ½ H2 for each H atom, and (H2O—H2) for each O atom, plus the energy of the clean slab. The corrections of zero-point energy and entropy of the OER intermediates are provided by Table 6.
The surface Pourbaix diagram was calculated based on the method proposed by Hansen et al. [H. A. Hansen, J. Rossmeisl, J. K. Nørskov, Surface Pourbaix diagrams and oxygen reduction activity of Pt, Ag and Ni (111) surfaces studied by DFT. Phys. Chem. Chem. Phys. 10, 3722-3730 (2008)], where the free energy of oxygen and hydroxyl exchange at a given surface at any pH and potential is calculated as
where ΔGfield is the change in the adsorption energy due to the electric field in the electrochemical double layer at the cathode. According to the work by Rossmeisl et al., [J. Rossmeisl, J. K. Nørskov, C. D. Taylor, M. J. Janik, M. Neurock, Calculated phase diagrams for the electrochemical oxidation and reduction of water over Pt (111). The J. Phys. Chem. B. 110, 21833-21839 (2006)], the relative stability change of O* and OH* under electric field is more than one-order magnitude lower than the change of free energy. Therefore, the trend in adsorption energies can be well described by neglecting ΔGfield in the construction of the surface Pourbaix diagram.
In view of the many possible embodiments to which the principles of the disclosed invention may be applied, it should be recognized that the illustrated embodiments are only preferred examples of the invention and should not be taken as limiting the scope of the invention. Rather, the scope of the invention is defined by the following claims. We therefore claim as our invention all that comes within the scope and spirit of these claims.
This application is a continuation of PCT/US2022/051987, filed Dec. 6, 2022, which claims the benefit under 35 U.S.C. § 119 (e) of U.S. Provisional Application No. 63/286,967, filed on Dec. 7, 2021, and U.S. Provisional Application No. 63/323,188, filed on Mar. 24, 2022, the entire contents of each of which are incorporated herein by reference.
This invention was made with government support under the Singapore National Research Foundation under its Campus for Research Excellence and Technological Enterprise (CREATE) program, through the Cambridge Center for Carbon Reduction in Chemical Technology (C4T), Grant No. EC8040 awarded by the Oregon State University Faculty Startup fund and Grant No. ECCS2025489 awarded by the National Science Foundation. The United States government has certain rights in the invention.
Number | Date | Country | |
---|---|---|---|
63323188 | Mar 2022 | US | |
63286967 | Dec 2021 | US |
Number | Date | Country | |
---|---|---|---|
Parent | PCT/US2022/051987 | Dec 2022 | WO |
Child | 18733669 | US |