GB2509917A - Mixed metal oxide materials of titanium and tungsten - Google Patents

Mixed metal oxide materials of titanium and tungsten Download PDF

Info

Publication number
GB2509917A
GB2509917A GB1300823.0A GB201300823A GB2509917A GB 2509917 A GB2509917 A GB 2509917A GB 201300823 A GB201300823 A GB 201300823A GB 2509917 A GB2509917 A GB 2509917A
Authority
GB
United Kingdom
Prior art keywords
particles
mixed metal
metal oxide
palladium
catalyst
Prior art date
Legal status (The legal status 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 status listed.)
Withdrawn
Application number
GB1300823.0A
Other versions
GB201300823D0 (en
Inventor
Brian Elliott Hayden
Jonathan Conrad Davies
Claire Mormiche
Xiaojuan Lu
Current Assignee (The listed assignees may be inaccurate. Google has not performed a legal analysis and makes no representation or warranty as to the accuracy of the list.)
Ilika Technologies Ltd
Original Assignee
Ilika Technologies Ltd
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 Ilika Technologies Ltd filed Critical Ilika Technologies Ltd
Priority to GB1300823.0A priority Critical patent/GB2509917A/en
Publication of GB201300823D0 publication Critical patent/GB201300823D0/en
Priority to PCT/GB2014/050121 priority patent/WO2014111717A1/en
Publication of GB2509917A publication Critical patent/GB2509917A/en
Withdrawn legal-status Critical Current

Links

Classifications

    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01GCOMPOUNDS CONTAINING METALS NOT COVERED BY SUBCLASSES C01D OR C01F
    • C01G23/00Compounds of titanium
    • C01G23/04Oxides; Hydroxides
    • BPERFORMING OPERATIONS; TRANSPORTING
    • B01PHYSICAL OR CHEMICAL PROCESSES OR APPARATUS IN GENERAL
    • B01JCHEMICAL OR PHYSICAL PROCESSES, e.g. CATALYSIS OR COLLOID CHEMISTRY; THEIR RELEVANT APPARATUS
    • B01J23/00Catalysts comprising metals or metal oxides or hydroxides, not provided for in group B01J21/00
    • B01J23/38Catalysts comprising metals or metal oxides or hydroxides, not provided for in group B01J21/00 of noble metals
    • B01J23/40Catalysts comprising metals or metal oxides or hydroxides, not provided for in group B01J21/00 of noble metals of the platinum group metals
    • B01J23/44Palladium
    • BPERFORMING OPERATIONS; TRANSPORTING
    • B01PHYSICAL OR CHEMICAL PROCESSES OR APPARATUS IN GENERAL
    • B01JCHEMICAL OR PHYSICAL PROCESSES, e.g. CATALYSIS OR COLLOID CHEMISTRY; THEIR RELEVANT APPARATUS
    • B01J35/00Catalysts, in general, characterised by their form or physical properties
    • B01J35/30Catalysts, in general, characterised by their form or physical properties characterised by their physical properties
    • B01J35/33Electric or magnetic properties
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01GCOMPOUNDS CONTAINING METALS NOT COVERED BY SUBCLASSES C01D OR C01F
    • C01G23/00Compounds of titanium
    • C01G23/04Oxides; Hydroxides
    • C01G23/047Titanium dioxide
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01GCOMPOUNDS CONTAINING METALS NOT COVERED BY SUBCLASSES C01D OR C01F
    • C01G41/00Compounds of tungsten
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M4/00Electrodes
    • H01M4/86Inert electrodes with catalytic activity, e.g. for fuel cells
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M4/00Electrodes
    • H01M4/86Inert electrodes with catalytic activity, e.g. for fuel cells
    • H01M4/90Selection of catalytic material
    • H01M4/9075Catalytic material supported on carriers, e.g. powder carriers
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M4/00Electrodes
    • H01M4/86Inert electrodes with catalytic activity, e.g. for fuel cells
    • H01M4/90Selection of catalytic material
    • H01M4/92Metals of platinum group
    • H01M4/925Metals of platinum group supported on carriers, e.g. powder carriers
    • BPERFORMING OPERATIONS; TRANSPORTING
    • B01PHYSICAL OR CHEMICAL PROCESSES OR APPARATUS IN GENERAL
    • B01JCHEMICAL OR PHYSICAL PROCESSES, e.g. CATALYSIS OR COLLOID CHEMISTRY; THEIR RELEVANT APPARATUS
    • B01J35/00Catalysts, in general, characterised by their form or physical properties
    • B01J35/30Catalysts, in general, characterised by their form or physical properties characterised by their physical properties
    • B01J35/391Physical properties of the active metal ingredient
    • B01J35/393Metal or metal oxide crystallite size
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01PINDEXING SCHEME RELATING TO STRUCTURAL AND PHYSICAL ASPECTS OF SOLID INORGANIC COMPOUNDS
    • C01P2002/00Crystal-structural characteristics
    • C01P2002/50Solid solutions
    • C01P2002/52Solid solutions containing elements as dopants
    • C01P2002/54Solid solutions containing elements as dopants one element only
    • 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
    • Y02E60/00Enabling technologies; Technologies with a potential or indirect contribution to GHG emissions mitigation
    • Y02E60/30Hydrogen technology
    • Y02E60/50Fuel cells

Landscapes

  • Chemical & Material Sciences (AREA)
  • Organic Chemistry (AREA)
  • Chemical Kinetics & Catalysis (AREA)
  • Engineering & Computer Science (AREA)
  • Materials Engineering (AREA)
  • General Chemical & Material Sciences (AREA)
  • Electrochemistry (AREA)
  • Inorganic Chemistry (AREA)
  • Life Sciences & Earth Sciences (AREA)
  • Environmental & Geological Engineering (AREA)
  • General Life Sciences & Earth Sciences (AREA)
  • Geology (AREA)
  • Inert Electrodes (AREA)
  • Catalysts (AREA)

Abstract

Mixed metal oxide materials of titanium and tungsten are disclosed for use in fuel cells. The mixed metal oxides are useful as catalyst support materials and as composite materials comprising a catalyst supported on or within support materials. The mixed metal oxides comprises up to 40at.% tungsten. The catalyst, supported by the mixed metal oxide may be palladium or a palladium alloy. An alloy of palladium and cobalt is preferred.

Description

Catalytic and Composite Materials The present invention relates to materials for use in a ifiel cell and particularly to catalytic materials for use in fuel cells, more particularly to catalyst support materials and composite materials comprising a catalyst supported on or within support materials.
A fuel cell comprises an anode where fuel is oxidised and a cathode where an oxidising species such as oxygen is reduced. An electrolyte allows for the transport of ions between the electrodes. Fuel and oxidiser are supplied separately to each electrode. As fuel is oxidised at the anode, electrons are released and pass through an external circuit to the cathode where they are consumed in the reduction of the oxidising species. In a polymer electrolyte membrane fuel cell (PEMFC) the fuel used is usually hydrogen and the oxidising species is usually oxygen; the polymer electrolyte is capable of allowing the flow of protons from the anode to the cathode.
Platinum and platinum alloys have been found to be the most efficient catalysts for facilitating the oxygen reduction reaction (ORR) which takes place at the cathode.
Platinum, however, has a high cost and there has been much interest in reducing the amount of precious metal. Usually, platinum is dispersed over a carbon support to increase the surface area relative to the mass of platinum used. Figure 1 shows a plot of the mass activity in ORR obtained at 0.85 V vc. RUE from the steady state measurement experiments versus Pt particle size, on a carbon support. The measurements were performed in 0.5 M HCIO4(aq) and at 25°C, following the procedure set out in the examples below. Previously, the inventors have studied the effects of particle size for a series of commercially available Pt nano-particlcs supported on carbon which were supplied by Johnson Matthey plc [Reference 1].
Both studies and others in the literature suggest that the particle size associated with optimum mass activity for carbon supported platinum is ca. 3.5 -4 nm [References 1- 4]. This is also in good agreement with predictions made by Kinoshita, that suggested a maximum at 3.5 nm in mass activity corresponding to an optimum in size for the p) exposure of (100) facets for cubo-octahedral particles [5]. This suggests that the high throughput methodologies and thin-film models used by the current inventors are a good approximation for powder catalysts as would be conventionally used in fuel cells.
It is also of interest to compare and contrast the particle size dependence of the ORR activity of carbon supported platinum to carbon supported palladium. Palladium is considered to be a potentially cheaper alternative to platinum in thel cell cathode catalysts. There are a number of published reports on electrochemical studies of palladium and palladium alloys in acidic electrolytes, in particular, Pd [7-13] PdAu [14], PdCo [15-22], PdCr [18], PdFc [23], PdTi [24], PdCoAu [24, 25], PdCoMo [26], PdV [27], PdMn [271, PdPt [28-33], PdRh [30], PdRhPt [30, 34], and PdNi [18].
Figure 2 shows a plot of the mass activity in ORR obtained using the same procedure at 0.65 V vs. RHE from the steady state measurement experiments versus Pd particle size, on carbon support. Note that the Pt nanoparticle data were shown at 0.85 V c.
RilE, but the Pd data shows considerably lower current at this potential, which means that Pd has intrinsically lower activity than Pt on carbon. There appears to be no clear maximum for Pd supported on carbon, however looking at Figure 3: The specific activity at 0.70 V w. RHE of a thin film of Pd (see dashed line) is lower than that obtained for particles with a diameter between 9 and 12 nm (at the same potential), which means that there must be an optimum particle size, but that this has not been reached by 12 nm.
Although pure Pd is less active than Pt, recent studies have suggested that alloying of Pd with a second element such as Co or Fe can lead to activities which are comparable to, or even better than Pt [20, 211. The main disadvantage of palladium based catalysts is the intrinsic instability of palladium under ORR conditions, although there are also suggestions that alloying Pd can considerably improve its stability [15, 35].
The ORR activity of PdCo alloys with approximately 30 at.% Co has been investigated by Savadogo eta! [20, 21]. They found good stability of sputtered PdCo films during extended potential cycling. It was proposed that this could be due to an electronic interaction between Pd and Co in the amorphous alloy. The onset potential measured for the ORR for the PdCo alloy was similar to that for Pt (0.90 V w. RHE) whilst the current observed at 0.80V vs. RHE for PdCo (0.58 mA cm2) was higher than that obtained with Pt (0.25 mA cm2). Similar conclusions were obtained by Fernandez et a!. [17] who prepared a series of PdCo particles with compositions ranging from 0-100 at.% Pd on glassy carbon, and screened them using a Scanning Electrochemical Microscopy (SECM) technique. They concluded that the addition of -30 % Co to Pd produced a material with a superior catalytic activity to pure Pd and close to that of pure Pt, although the activity was observed to degrade over time.
The inventors have previously studied PdCo thin films, the steady-state current (b) and surface area (a) were published in a previous patent application [15] and are shown in Figure 4. Combined they suggest an optimum activity (specific current density per unit area -(a)) for PdCo alloys with 40-50 at.% Co (50-60 at.% Pd). This is similar to the maximum found (in both specific and mass specific current density) for PdCo particles deposited on carbon as shown in Figure 5. The data was obtained from steady state potential step measurements at 0.75 V vs. RHE. There does not seem to be any influence of the particle size. The main trend seems to follow the composition with increased activity as the percentage of Co increases, until 60 at.% Co in PdCo at which the particles become unstable.
In this work, palladium and palladium cobalt alloy particles with a tungsten doped titanium oxide support have been produced. The effect of cobalt is studied in terms of improvcmcnts of the stability (durability) and activity of thc catalyst in low temperature fuel cells. Metal oxides have previously been investigated for use as fuel cell catalyst supports [36-48j. Carbon can be oxidised under fuel cell operating conditions leading to degradation of the catalyst and limiting the lifetime of the fuel cell [49]. Oxides may be less prone to oxidative corrosion and therefore more stable than carbon in a fuel cell environment [45].
Several patents have described the use of metal oxides as supports for fuel cell catalysts and methods of synthesising them [36, 37, 39, 42, 47].Traditionally, dopants have been added to metal oxides in order to increase conductivity by creating defects (such as oxygen vacancies). However, the inventors believe that in addition to increased conductivity, dopants can also help to create defects on the surface of the metal oxides and hence more sites for nucleation of the active catalyst. Moreover it was found that the propensity for large palladium particles formation is significantly enhanced by increasing the substrate temperature during deposition, which means that the number of particles is also reduced. All depositions of palladium and palladium cobalt particles were carried out at 350°C.
In general, published reports of Pd and Pd alloys studics focus on a limited number of compositions due to the time taken to prepare the materials using standard laboratory techniques. An inherent advantage of the inventors High Throughput Physical Vapour Deposition (HT-PVD) synthesis method is that a wide range of compositions can be prepared in a single experiment and the resulting materials can subsequently be screened in a parallel electrochemieal experiment [50]. This approach varies from others in that thin film materials of varying composition are prepared via simultaneous deposition of the component elements[51]. This method allows for the preparation of graded compositional thin films of alloy materials through the controlled and simultaneous deposition of the component elements on a substrate at room temperature. A significant advantage of this method is that the alloys formed arc not subjected to annealing steps which could alter their relative bulk and surface compositions and cause the formation of thermodynamically stable phases. Rather, it is possible to prepare thin film alloys with compositions determined by the deposition conditions. Previously, these have been demonstrated to be continuous polycrystalline solid solutions with good agreement between the surface and bulk compositions for the as-deposited films. The high-throughput methodologies also provide a means of obtaining good statistical data under identical electroehemical conditions to provide both particle size and support effects in eleetroeatalysis [44, 52-551. For example, the effects of particle size and substrate support interactions on the activity of Au nano-particles supported on carbon and on sub-stoichiomctric titanium dioxide (also referred to as titania) (TiO) has been successfully demonstrated using the high-throughput synthetic and screening methodology [52].
In the current work, high throughput methodologies have been used to oplimise the particle size andlor composition of Pd or PdCo particles on W-doped titanium oxide supports.
In its broadest sense, the present invention provides a mixed metal oxide material of titanium and tungsten.
Preferably, the mixed metal oxide material comprises tungsten in an amount of 40 atomic percent (at.%) or less, calculated on a metals only basis.
Preferably, the material comprises an oxide of titanium doped with tungsten, with an oxygen content near to stolehiometrie.
In one embodiment, the material is formed in an amorphous phase. Preferably, the atomic percentage of tungsten is 26 at.% or less.
In an alternative embodiment, the material is formed in a crystalline phase.
Suitably, the material is formed in a crystalline phase by annealing of an amorphous material as described above.
More preferably, in a crystalline phase, the atomic percentage of tungsten is 23 at.% or less, even more preferably, 18 at.% or less, most preferably about 7 at.%.
The present invention also provides a catalytic material comprising a catalyst support and a catalyst supported by the catalyst support, wherein the catalyst support is a mixed metal oxide material of titanium and tungsten Preferably, the catalyst is palladium or a palladium alloy.
Preferably, the catalyst is formed as particles on the catalyst support.
Preferably, the particle size is SOnm or less, preferably from mm to lOnm; more preferably from Snm to 8 rim.
Preferably, the catalyst is a palladium alloy of palladium and cobalt.
More preferably, the palladium alloy consists essentially of palladium and up to about 70 at.% cobalt; preferably up to about 40 at.% cobalt, more preferably up to about at.% cobalt; most preferably, 15-27 at.% cobalt.
The present invention also provides a fuel cell comprising a catalytic material.
The present invention also provides the use of a mixed metal oxide material as defined above orofa catalytic material as defined above in a fuel cell.
For convenience, mixed metal oxide materials of tungsten and titanium will also be referred to by the shorthand formulation TiWO. It will be appreciated that this is not used as a chemical formula to indicate any specific stoiehiometry and should not be taken to be so limiting in any way.
The term mixed metal oxide indicates an oxide of a mixture of metals; a mixture of metal oxides or a combination thereof The above and other aspect of the invention will now be described in further detail, by way of example only, with reference to the following examples and the accompanying figures, in which: Figure 1: Mass activity per g of platinum at 0.85 V vs. RHE (obtained from steady state step measurements at discrete potentials) versus Pt particle diameter for particles supported on carbon. Measurements performed in 0.5 M HCIO4(aq) and at 25°C; Figure 2: Mass activity pcr g of palladium, at 0.65 V vs. RHE (obtained from steady state step measurements at discrete potentials) versus Pd particle diameter for particles supported on carbon for 2 different arrays. The measurements were performed in similar conditions in 0.5 M HC1O4(aq) and at 25°C; Figure 3: Specific activity obtained at 0.70 V vs. RHE (from steady state step measurements at discrete potentials) versus Pd particle diameter for particles supported on carbon for 2 different arrays. The measurements were performed in similar conditions in 0.5 M HC1O4(aq) and at 25°C. The dashed line represents the limit of a thin film of Pd calculated by CO stripping area; Figure 4(a): Elcctrochcmically active surface area, (b) Steady state current and (c) Specific current density (mA cni2) of PdCo thin film for ORR at 0.8 V vs. RHE; IS Figure 5(a): Mass activity (A gpd1) and (b) Specific current density (A cm2) of PdCo particles on Carbon at 0.75 V vc. RHE. [151 Figure 6: Schematic view of an electrochemical array substrate, highlighting the x and y coordinates 1, 4, 7 and 10; Figure 7: XRD patterns of TiO on glass after annealing at 450 O(: for 6 hours in I atm of oxygen. The dashed black line corresponds to TiO2 anatase (00-021-1272) (Reference 6); Figure 8: XRD patterns of TiWO on glass after annealing at 450 °C for 6 hours in 1 atm of oxygen. The dashed black line corresponds to TiO2 anatase (00-021-1272); Figure 9: TEM images for TEM grids!TiO(amorphous)!Pd, obtained at an equivalent x-coordinate position of a) 1; b) 4; c) 7; d) 10; Figure 10: TEM images for TEM grids!TiWO (amorphous)!Pd, obtained at an equivalent x-eoordinate position of a) 1; b) 4; c) 7; d) 10 with 25 at.Vo W. Figure 11: Average particle diameter versus the grid position (x) for Pd particles on amorphous TiO and TiWO (25 at.% W); Figure 12: Selected fields: Slow cyclic voltammetry (5 my -1) in oxygenated 0.5 M HCIO4 obtained between -0.05 and 1.10 V uc. RHE for sample #662 1 -Pd particles on amorphous TiWO (5 at.% W). The mean particle sizes range from approximately 2-7 nm. The black dashed line shows the chosen current (0.7 uA) for the ignition potential analysis. The markers on some of the plots do not represent real data points; they arc only there to guide the eye; Figure 13: a) lOx 10 plot of oxygen reduction ignition potential for Pd particles on amorphous TiWO with 7 at.% W; 1) lOx 10 plot of oxygen reduction ignition potential for Pd particles on amorphous TiWO with 10 at.% W; c) data per column shown in a) ; and d) data per column shown in b). The black dashed lines in c) and d) show the limit of braking down of the Pd particles around 4 nm diameter. The current at which the ignition potential was taken is 0.7 jiA; Figure 14: Cyclic voltammetry (100 mY 1) in deoxygenated 0.5 M HC1O4 obtained between -0.05 and 1.00 V vs. RHE for Pd particles on amorphous TiW0 (7 at.% W) performed before and after slow cycling experiments in saturated oxygen electrolyte.
The black dashed line shows the Pd particle diameter at which Pd breakdown is believed to start occurring; Figure 15: a) Raw oxygen reduction current; b) Specific activity and c) Mass activity per g of palladium, all at 0.65 V (obtained from cathodic sweep of slow sweep voltammetry (5 mY in 02 saturated electrolyte) versus Pd particle diameter for particles supported on amorphous TiWOx Figure 16: TEM images for TEM window!/TiO(crystalline)!Pd, obtained at an equivalent x-coordinate position of a) 1; b) 4; e) 7; d) 10; Figure 17: TEM images for TEM window!TiWO (crystalline)!Pd, obtained at an equivalent x-coordinate position of a) 1; b) 4; c) 7; d) 10 with 25 at.Vo W; Figure 18: Average particle diameter versus the grid position (x) for Pd particles on crystalline TiO and TiWO (25 at.% W); Figure 19: Selected fields: Slow cyclic voltammetry (5 mV s') in oxygenated 0.5 M HCIO4 obtained between -0.05 and 1.00 V vs. RilE for Pd particles on crystalline TiWO (8 at.% W). The mean particle sizes range from approximately 2-8 nm. The black dashed line shows the chosen current (0.7 iA) for the ignition potential analysis. The markers on some of the plots do not represent real data points; they are only there to guide the eye; Figure 20: a) 10 x 10 plot of oxygen reduction ignition potential for Pd particles on crystalline TiWO with 8 at.% W; b) 10 x 10 plot of oxygen reduction ignition potential for Pd particles on amorphous TiWO with 29 at.% W; c) data per column shown in a); and ci) data per column shown in b). The black dashed line in d) show the limit of breaking down of the Pd particles around 5.5 nm diameter; Figure 21: a) Raw oxygen reduction current; b) Specific activity and c) Mass activity per g of palladium, all at 0.65 V vs. RUE (obtained from cathodic sweep of slow sweep voltammetry (5 mV s1 in 02 saturated electrolyte) versus Pd particle diameter for particles supported on crystalline TiW0.
Figure 22: Mass activity per g of palladium at 0.65 V vs. RHE (obtained from cathodic slow sweep voltammetry (5 mV in 02 saturated electrolyte) versus Pd particle diameter for particles supported on both amorphous and crystalline TiO supports; Figure 23: A (10 x 10) macro of slow cyclic voltammetry (5 mY 1) in oxygenated 0.5 M HCIO4 electrolyte obtained between -0.05 and 1.00 V vs. RHE for PdCo particles on amorphous TiWOX (20 at.% W). Average particle sizes range from approximately 2-9 nm; Figure 24: Selected fields: Slow cyclic voltammetry (5 mY -1) in oxygenated 0.5 M HCIO4 obtained between -0.05 and 1.00 V vc. RHE for a sample with PdCo particles on amorphous TiWOX (20 at.% W), with the atomic percentage of cobalt varying between 0 and 100 at.% in PdCo (corresponding to row S highlighted by the plane rectangle selection in Figure 23) and the particle size varies from 4.5 to 7.3 nm. The markers on some of the plots do not represent real data points; they are only there to guide the eye; Figure 25: Selected fields: Slow cyclic voltammetry (5 mV -1) in oxygenated 0.5 M HCIO4 obtained between -0.05 and 1.00 V vs. RHE for a sample with PdCo particles on amorphous TiWOX (20 at.% W), with the atomic percentage of cobalt constant at at.%, and the particles size varying from 2.7 to 8.0 nm (corresponding to column 4 highlighted by the dashed rectangle selection in Figure 23). The markers on some of the plots do not represent real data points; they are only there to guide the eye; Figure 26: a) Plot of ignition potential against field position for PdCo particles on amorphous TIWOX (20.0 at.% W); b) Plot of atomic percentage of Pd in PdCo; e) Plot of particle diameter (urn); Figure 27: Activity per unit mass of Pd, at 0.65 V vs. RHE (obtained from the cathodic sweep of slow sweep voltammetry 5 mY 1) as a function of atomic percentage of Co, for the 3 equivalent thickness of particles deposited (from 5.3 to 8.3 nm diameter) supported on amorphous TiWOX supports; Figure 28: A (10 x 10) macro of slow cyclic voltammetry (5 mY s') in oxygenated 0.5 M HCIO4 electrolyte obtained between -0.05 and 1.00 V uc. RHE for PdCo particles on crystalline TiWO (13 at.% W). Average particle sizes range from approximately 2-10 nm; Figure 29: Selected fields: Slow cyclic voltammetry (5 mV -1) in oxygenated 0.5 M HClO obtained between -0.05 and 1.00 V vs. RHE for PdCo particles on crystalline TiWO (13 at.% W), with the atomic percentage of cobalt varying between 0 and at.% in PdCo (corresponding to rows highlighted by the plane rectangle selection in Figure 28) and the particle size varies from 4.4 to 5.6 nm. The markers on some of the plots do not represent real data points; they are only there to guide the eye; Figure 30: Selected fields: Slow cyclic voltammetry (5 mV 1) in oxygenated 0.5 M I-1C104 obtained between -0.05 and 1.00 V vs. RIlE for PdCo particles on crystalline TiWO (13 at.% W), with the atomic percentage of cobalt constant at 35 at.%, and the particles size varying from less than 3.0 to 8.1 nm diameter (conesponding to column 4 highlighted by the dashed rectangle selection in Figure 28). The markers on some of the plots do not represent real data points; they are only there to guide the eye; Figure 31: a) Plot of ignition potential against field position for PdCo particles on crystalline TiWO (13.0 at.% W) and b) zoom ofplot a) in the range 0.6 to 0.67 V vs. RHE highlighting the optimum; and Figure 32: Activity per unit mass of Pd, at 0.65 V vs. RHE (obtained from the cathodic sweep of slow sweep voltammetry 5 mV 1) as a function of atomic percentage of Co, for the 3 equivalent thickness of particles deposited (from 5.3 to 8.3 nm diameter) supported on crystalline TiWO supports.
Experimental Synthesis of samples: Thin film samples were deposited using a high throughput physical vapour deposition system described in Reference 51. The depositions were carried out in a cryo-pumpcd ultrahigh vacuum system with a base pressure of 1 x l0b0 Ton, incorporating 6 off-axis PVD sources (3 electron beam sources and 3 Knudsen cells). A range of oxide samples were synthcsised onto different substrates including silicon wafers, glass wafers, SSTOP (Si'Si02/Ti02/Pt where the Ti02 layer is 10 nm thick and Pt layer is urn thick) wafers and electrochernical arrays (10 x 10 arrays of gold contact pads as represented schematically in Figure 6). Titanium (Alfa Aesar Puratronic 99.995%) S and tungsten (Alfa Aesar 99.97%) were both deposited from an electron beam (E-gun) source, in the presence of either molecular oxygen or an atomic oxygen source (operating at 400 -600 W power) at a pressure in the range of 5 x I o to 5 x I o-Torr to form TiWOX films. Amorphous thin films were prepared by depositing TiWOX at 25°C. Crystalline thin films were prepared by post-annealing the as-deposited oxide films for 6 hours at 450°C in an atmosphere of oxygen in order for crystallisation to take place. For comparison, carbon (Alfa Acsar 99.995%) films were deposited from an E-gun source.
The Pd and PdCo particles were deposited onto the pre-deposited oxide thin films IS from a Knudsen cell source (Pd (Allgemeine 99.95%)) and an E-gun source (Co (Alfa Acsar 99.95%)), during deposition the oxide substrates were heated to 350°C in order to dehydroxylate the surface and ensure a good distribution and size of particles. A shutter that was moved during deposition allowed different equivalent thicknesses of catalyst to be deposited onto different fields. These deposition procedures allowed for a varying PdCo alloy composition in one direction and a varying size of particles in the other (orthogonal) direction, while the oxide support composition was left uniform across the substrate. The amounts of Pd and Co deposited were calibrated by depositing thicker films onto silicon substrates and measuring the thickness of the films by optical profilomctry and producing a calibration curve against deposition time. For comparison, Pt (Algemein 99.9%) particles were deposited onto pre-deposited carbon films using an E-gun source using a similar procedure.
Oxide characterisation: The composition of the oxide films was determined using a Laser Ablation Inductively Coupled Plasma Mass Spectrorneter (LA-ICP-MS, New Wave 213 nrn laser & Perkin Elmer Elan 9000). This method gives the relative composition of the metallic elements, but is not capable of measuring the oxygen concentration.
Thcrcforc oxidc compositions shown arc thc relative atomic masses of thc two metallic elements only within the oxide (i.e. Ti and W). As the ICP-MS measurements are destructive, composition measurements were made on samples deposited onto silicon wafers. Thc same deposition conditions wcre thcn uscd to deposit onto equivalent eleetroehemical arrays. The samples used for stability test of the support were made by depositing the oxide thin films onto SSTOP substrates with a varying oxide composition across the substrate. The composition range on each substrate was measured by ICP-MS on the four comers of the film.
X-Ray diffraction (XRD) patterns were obtained using the Bruker D8 Discover diffractometer, a powerful XRD tool with a high-precision, two-circle goniometer with independent stepper motors and optical encoders for the Theta and 2 Theta circles. The D8 diffractometer system is equipped with a GADDS detector operating at 45 kV and 0.65 mA. A high intensity X-ray IRS Incoatec source (with Cu Ku radiation) is incorporated allowing high intensity and collimated X-rays to be localised on thin film materials providing an efficient high throughput structural analysis. This analysis was carried out on oxide films deposited onto glass substrates.
Eleetroehemieal stability tests have been carried out on the oxide films deposited on SSTOP substrates as continuous thin films. The samples were placed in an electrochemical cell and left at 1.2 V w. RHE the reversible hydrogen electrode (RHE for a period of 100 hours. Cyclic voltammograms were performed between 0.025-1.2 V vs. RHE at 25 hour intervals, at a sweep rate of 50 mY s' to determine the initial, intermediate, and final state of the oxide layers. The temperature was maintained at 25°C and a mercury/mercury sulphate (MMSE) reference electrode was used. A Pt mesh counter electrode was used, in a glass compartment separated from the working electrode compartment by a glass fit. The electrolyte used for all experiments was 0.5 M HCIO4 prepared from concentrated HCIO4 (double distilled, GFS) and uhrapure water (ELGA, 18 MQ cm).
The apparatus used for these measurements has been previously reported [1], and the high-throughput cell was simply adapted to provide data for a continuous film through a single channel. Photographs and ICP-MS measurements of the samples were obtained before and after this test.
Eleetroehemieal screening: A range of eleetroehemieal arrays have been prepared, amorphous and crystalline Ti0 and TiWO supports combined with either Pd or PdCo particles; and, for comparison, Pt and Pd particles on carbon supports.
High-throughput electroehemical screening equipment enables electrochemical experiments on 100 independently addressable electrodes arranged in a 10 10 array to be carried out in a parallel screening mode which has been described in detail elsewhere [11. The geometric areas of the individual working electrodes on the electrochemical array are 1.0 mm2.
The samples were screened using the procedure described in Table 1, in 0.5 M HCIO4 at 25 °C, prepared from concentrated HCIO4 (double-distilled, GFS) and ultrapure water (ELGA, 18 M 12 cm). The gases used (Ar, 02 and CO) were of the highest commercially available purity (Air Products). Unless stated otherwise, experiments were performed under an atmosphere of argon. Unless noted otherwise, the maximum potential applied to the electrodes was 1.0 V vs. RilE.
Table 1: Electrochemical screening procedure. ___________ Experiment Gas Potential limits / Sweep rate Vvs. RUE /mVs' Bubbling Ar 25 mm CVs in deoxygcnatcd solution Ar above solution -0.050 to 1.000 (start at 0.075) 100 ___________________________ Bubbling CO 25 mm At 0.075 ____________ _____________________________ Bubbling Ar 35 mm At 0.075 _____________ CO sfripping Ar above solution -0.050 to 1.000 (start at 0.075) 50 (1)2 saturation Bubbling Ar 60s At 0.400 _____________ __________________________ Bubbling 02 10 mm At 0.400 ___________ 02 reduction steps Bubbling °2 in solution Steps 0.400 to 0.800 and hack to 0.400 in 0.050 V increments every 90 s 3-5 CVs in 02 saturated 0 above solution -0.050 to 1.000 (start at 0.800) 5 solution ___________________________ Bubbling Ar 25 mm __________________________ ____________ 3-5 CVs in deoxygcnatcd Ar above solution -0.050 to 1.000 (start at 0.075) 100 solution ___________________________ Bubbling CO 25 mm At 0.075 ____________ ___________________________ Bubbling Ar 35 mm At 0.075 ____________ CC) stripping Ar above solution -0.050 to 1.000 (start at 0.075) 50 Results and Discussion XRD Characterisation: TiO and TiWO samples deposited at 25°C showed no XRD features, suggesting the direct formation of an amorphous oxide.
Figure 7 shows XRD patterns of TiO deposited on glass substrates at 25°C after annealing at 450°C for 6 hours in one atmosphere of oxygen. Crystallisation is much more effective for sub-stoichiometric, as-deposited, oxide, compared to the oxides deposited with a higher oxygen content (or using the atom source). Oniy the titanium oxide deposited using a pressure of oxygen of I x b-5 Torr with the molecular source show XRD peaks after annealing. The peaks seen are in good agreement with those of Ti02 anatase space group 141/amd (card 00-021-1272), which displays a single peak at 20=25.28° for the Miller indices (hkl) (101), a triplet around 20=37.80° for (004), and single peaks at 20=48.05° for (200) and at 20=53.89° for (105) (see dashed black line).
The experimental data was also compared to the diffraction patterns of Ti02 rutile (space group P42/mnm) (card 00-021 -1276) reported in the database used. However no assignment of diffraction peaks to the rutile phase could be made. Indeed none of the present features matched any other phase reported in this database.
Figure 8 shows the XRD patterns of T1WO, on glass substrates deposited at 25°C after annealing at 450°C for 6 hours in one atmosphere of oxygen. As observed for the pure titanium oxide, samples deposited using a lower pressure of oxygen lead to an oxide which gives better crystallisation after annealing. The dashed black line corresponds to Ti02 anatase(00-021-1272).
The data shown here have been scparated into groups of oxides deposited undcr the same conditions for the oxygen source, but with a varying compositional range of Ti vs. W. It is clear that with a higher titanium content the XRD features are in good agreement with that of the Ti02 anatase space group (see dashed black line). As the content of W increases (i.e. the content of Ti decreases) the formation of the Ti02 anatase phase is disrupted. No assignment of any peaks associated with any tungsten and/or tungsten oxide phases was possible. Hsu et aL [56] synthesized these oxides by sol-gel methods and further annealed their samples. They made similar observations and concluded that during the heat treatment of the mixed Ti02 and W03 precursor gels, nano-crystalline Ti02 outgrew the counterpart W03 phase.
Electrochenilcal stability tests: Tests have been carried out on oxide films deposited on SSTOP substrates. The samples were placed in an electrochemical cell and left at 1.2 V vs. RHE for a period of 100 hours as detailed in the experimental section.
TiO was found to be very stable both in amorphous and anatase phases. The amorphous binary TiWO showed that adding up to 30 at.% W only leads to a small loss of tungsten, with a final atomic percentage of W of 27 at.% after 100 hours. The crystalline samples showed no significant loss when adding up to 17 at.% W. Pd particles on amorphous TiO and TiWO: High-throughput methodologies have been used to investigate Pd particles on amorphous TiO and TiWO supports. The effect of Pd particle size and the W-content of the support have been obscrved.
TEM data has been obtained for a range of different palladium particle sizes on amorphous supports as shown in Figure 9 and Figure 10 for TiO and TiWO (25 at.% W), respectively. The images were obtained for positions equivalent to coordinates x1, 4, 7 and 10 on the eleetroehemieal array (Figure 6), therefore providing an insight into the variations observed across the range of electrochemical measurements performed.
Figure 11 shows that the average particle size of the Pd varies from approximately 1.6 to 6.8 nm. Very little difference is observed between Pd on TiO and TiWO (with at.% W).
Figure 12 shows cyclic voltammograms obtained from slow voltammetric cycling (5 mV s4) in oxygenated 0.5 M HCIO4 electrolyte obtained between -0.05 and 1.10 V vs. RilE for sample #662 1 for selected fields from a (10 x 10) array (Pd particles supported on amorphous TiWO (5.0 at.% W)), with varying particle size. The mean particle sizes range from approximately 2-7 nm. The black dashed line shows the chosen current (0.7 jiA) for the ignition potential analysis. An ignition or onset potential is defined as the potential at which the absolute value of the current starts to increase from the background (double layer) level in a cyclic voltammetry experiment indicating that an oxidation or reduction reaction is taking place at that potential. The markers on some of the plots have been added to help differentiate the different data sets. It will be appreciated that the data actually includes many more data points than indicated by the markers.
In the slow cycling experiment the main feature observed is a reduction current associated with the oxygen reduction reaction. This has an onset at approximately 0.68 -0.71 V vs. RHE for the largest Pd particles on all samples. The onset (ignition) potential appears to shift gradually to lower voltages with decreasing particle size indicating a decrease in activity for the oxygen reduction reaction. The ignition potential is also altered depending on the support. Figure 13a) and b) show maps of the ignition potential on 2 (10 x 10) arrays of Pd particles deposited on amorphous TiWO supports with 7 and 10 at.% W respectively measured at a fixed current of 0.7 jiA. Figure 13 c) and d) show the ignition potential versus Pd particle diameter on both supports. Below 4 nm particle diameter (as shown by the black dashed lines) a large distribution in the data is clearly apparent which is likely to represent the limit at which a thin film or larger islands of Pd begin to break down into particles of inhomogeneous size. In comparison, the Pd thin film on carbon studied showed an ignition potential of around 0.83 V vs. RHE at the same current of 0.7 jiA. The increase in the over-potential can be attributed to an increase in the reversibility of the Pd/PdO couple on oxide supports analogous to the previously observed Pt on Ti02 [44,55].
Figure 14 shows cyclic voltammograms (CV5) (100 my s') in dcoxygcnatcd 0.5 M FlClO4 obtained between -0.05 and 1.00 V vs. RHE for Pd particles on amorphous TiWO (7 at.% W) performed before and after slow cycling experiments in saturated oxygen electrolyte. The black dashed line shows the Pd particle diameter at which Pd breakdown is believed to start occurring. A comparison of the CVs before and after oxygen reduction gives an indication of the stability or instability of the Pd particles depending on their size and the substrate on which they were deposited. Figure 14 shows that at the larger particle sizes there is no or little difference in the voltammetry before and after oxygen reduction as shown for electrode (2,3). For small particle size the difference between the cyclic voltammograms is clear, in certain cases (which correspond to the lowest ignition potential observed) the PdO to Pd surface reduction peak is not even present anymore which suggests an increased irreversibility of the palladium oxide formation and reduction reaction, possibly due to break down of the Pd particles on the titanium tungsten oxide surface as shown for electrode (9,2). In comparison for electrode (9,7) which has Pd particles of the same size as electrode (9,2), the difference between the two cyclic voltammograms is small, suggesting that the Pd particles were morc stable. Electrode (6,3) highlights the size of particles at which the instability starts to occur. The conelation between change in the cyclic voltammograms and the ignition potential seen during the oxygen reduction reaction (an indicator of activity) is clear.
Figure 15 shows plots of the raw oxygen reduction current (a), specific activity (b) and mass activity (c) in ORR obtained at 0.65 V from the cathodic sweep of the slow cycling (5 mV in 02 saturated electrolyte) experiments versus Pd particle size, on the amorphous TiWO supports. A maximum in specific activity and mass activity is found around 5.5 -6 nm particle diameter, decreasing rapidly towards 3 nm diameter particles. This is similar to the results obtained for Pt particles on carbon shown in Figure 1, where a maximum in mass activity and specific activity is found at 3.5 - 4 nm particle diameter but is in contrast with results obtained for Pd particles on carbon shown in Figure 2 which do not shown any clear optimum. The increase seen at lower particles size is likely to be due to some uncertainty when calculating the surface area from TEM data. There is no obvious trend with the W doping of the support, however only a limited number of TiW0 compositions were studied.
Hence, a more detailed study of the effect of the support composition might highlight a complex trend in the activity ofPd.
Pd particles on crystalline TiO, and TiWO: High-throughput methodologies were also used to investigate Pd particles on crystalline Ti0 and TiW0 supports. The effect of Pd particle size and the W-content of the support have been observed.
Initially, TEM data has been obtained for a range of different palladium particle sizes on the crystalline supports as shown in Figures 16 for Ti0 and TiWO (25 at.% MO, respectively. The images were obtained for positions equivalent to coordinates x = 1, 4, 7 and 10 on the electrochemical array, therefore providing an insight into the variations observed across the range of electrochemical measurements performed.
Figure 18 shows thc average particle diameter vcrsus grid position x for Pd particles on crystalline TiO and TiWO (25 at.%w) and shows that the average particle size of the Pd varies from approximately 2.6 to 9.4 nm. Very little difference is observed between Pd on crystalline TiO and TiWO (with 25 at.% W).
However a slight difference between the particle sizes is observed between the amorphous and erystaHine substrates (see Figure 11 and Figure 18), with the particles smaller on the amorphous substrate which may be due to a higher number of nucleation sites on the amorphous substrate, due to a higher degree of disorder, leading to a higher number of smaller particles for the equivalent metal loading.
Figure 19 shows CVs from selected fields from the oxygen reduction slow cycling experiment (5 mV s') in oxygenated 0.5 M HCIO4 obtained between -0.05 and 1.00 V vs. RHE obtained from one sample (Pd particles supported on crystalline TiWO (8.0 at.% W)), demonstrating the change with varying particle size. The mean particle sizes range from approximately 2-8 nm. The black dashed line shows the chosen current (0.7 tA) for the ignition potential analysis. The markers on some of the plots have been added to differentiate different data sets. It will be appreciated that the data actually includes many more data points than are indicated by the markers.
The main feature observed is the current associated with the oxygen reduction reaction which has an onset at approximately 0.67 -0.70 V vs. RHE for the largest Pd particles on all samples. The onset (ignition) potential appears to shift to lower voltages with decreasing particle size as can be seen from two different samples on Figure 20, suggesting decreasing activity.
At particles with a diameter of below around 5.5 nm (as shown in Figure 20 by the black dashed line) a steep switching in ignition potential is seen (Figure 20 (d)) for a TiWO crystalline support with 29 at.% W (3.5 nm for 18 at.% W and 6.5 nm for 24 at.% W (not shown here). A gradual decrease in ignition potential is seen for a support with only 8 at.% W. In fact, looking at various supports (from 0 to 40 at.% W) the behaviour of the ignition potential varies significantly with varying W content, which is probably duc to a) the Pd particles wetting very differently, b) the distribution of particle size varying and/or c) the stability of the Pd particles on the substrate varying depending on the crystalline support. Hence by optimising the composition of the crystalline support it is possible to minimise the size of particles (hence the amount of material used) and still keep an ignition potential around 070 V vs. RilE.
Figure 21 shows plots of the raw oxygen reduction current (a), specific activity (b) and mass activity (c) in ORR obtained at 065 V vs. RilE from the cathodic sweep of the slow cycling experiments (5 mY s in 02 saturated electrolyte) versus Pd particle size, on the crystalline TiW0 supports. All variables are seen to increase as the particle diameter increases in the range studied, although a maximum at higher particle sizes cannot be ruled out. Below 3 nm particle diameter the oxygen reduction current is negligible. Contrary to the amorphous support, the W-doped crystalline materials show a higher activity than the undoped anatase titanium oxide, therefore the same specific or mass activity can be obtained for smaller particle size than on the pure anatase Ti0, by W-doping.
Comparing data from the amorphous and crystalline supports it is clear that the activity is higher on the amorphous supports than on the crystalline, as can be shown for undoped titanium oxide in Figure 22, which shows mass activity per g of palladium at 0.65 V vs. RHE (obtained from cathodic slow sweep voltammetry (5 mV in 02 saturated electrolyte) versus Pd particle diameter for particles supported on both amorphous and crystalline Ti0 supports.
The results obtained here contrast with previous data obtained for carbon supported Pt shown in Figure 1 but show similar behaviour to that obtained on carbon supported Pd particles shown in Figure 2 where no clear maximum in specific or mass activity were found for the range of size of particles studied.
PdCo particles on amorphous TiO and TiWO: Figure 23 shows a typical 10 x 10 data set obtained from slow voltammetric cycling (5 my -1) in oxygenated 0.5 M HCIO4 electrolyte obtained between -0.05 and 1.00 V vc. RHE for PdCo partides supported on amorphous TiWO (20 at.% W). Average particle sizes range from approximately 2-9 nm. The data demonstrates a clear trend in activity relating to both particle size and PdCo composition which is repeated across all samples.
Figure 24 shows selected fields from row y5 in Figure 23 (as highlighted by the solid black rectangle). The percentage of cobalt in the particles varies from 0 to at.%, while the particle size varies from 4.5 to 7.3 nm. In fact, although it has been dcmonstratcd (not shown here) a constant thin film thickness can be obtained during deposition across a row, the particles size may vary with composition, that is palladium and cobalt do not wet the support in the same way, leading to differences in the particle size and density. Figure 25 shows selected fields from Figure 23 (see dashed rectangle selection) where the percentage of cobalt has been kept constant at at.% in PdCo and the size of particles varies from 2.7 to 8.0 nm diameter.
Together these figures show that by adding a small amount of cobah to palladium (15 at.% Co) the current improves and that the optimum particles size appears around 7nm.
The ignition potential measured (at 0.7 iA) for the array (with 20 at.% W) is plotted against the field in Figure 26 a). The white area represents the region of highest ignition potentia' hence giving the conditions in particle afloy composition and size of particles for the most promising materials. Figure 26 b) and c) show the alloy composition and the particle diameter across the array, respectively. Electrodes with high Co content gave a very poor signal due to the unstable nature of particles very rich in cobalt (see column x = 10). Still it is clear that the area of interest for the highest ignition potential is represented by the white region which is for a particle diameter between 4.5 and 8.0 nm and for 0 to 35 at.% Co in PdCo.
An in-dcpth analysis of all data for PdCo particlcs with varying size and composition on amorphous TiWO supports with a varying amount of W was performed. The specific activity, the mass activity and the ignition potentials where extracted from these slow cyclic voltammetry data.
Comparison of PdCo particles on amorphous T1WO1 support (3 equivalent thickness): Figure 27: shows the Activity per unit mass of Pd, at 0.65 V (obtained from the cathodic sweep of slow sweep voltammetry 5 my -1) as a function of atomic percentage of Co, for the 3th equivalent thickness of particles deposited (from 5.3 to 8.3 nm diameter supported on amorphous TiWO supports.
Particles size diameter varies from: 5.3 to 8.3 depending on the ratio ofPdCo and the amount of W within the substrate (5.3-5.9 for TiO and 6.1 to 8.3 for TiWO), although the equivalent thickness of material deposited is identical.
This plot demonstrates the effect of the atomic percentage of Co on the mass activity (for mass of Pd) obtained at 0.65 V vs. RHE on amorphous TiWO. The behaviour observed suggests that there is a slight enhancement on the addition of some Co for some of these systems, and certainly, there is no considerable negative effect of adding up to approx. 30 at.% Co. At higher Co contents there is a sharp decrease in the activity, and by 60-70 at.% the activity is very low. This trend is observed for the range of samples investigated. For the amorphous materials, the TiO support appears to lead to the most active systems.
PdCo particles on crystalline TiO and TiWO: Figure 28 shows a typical 10 x 10 data set obtained from slow voltammetrie cycling (5 my 1) in oxygenated 0.5 M HCIO4 electrolyte obtained between -0.05 and 1.00 V vc. RHE for PdCo particles supported on TiWO (13 at.% W). Average particle sizes range from approximately 2-10 nm. This demonstrates a clear trend in activity relating to both particle size and PdCo composition which is repeated across all samples.
Figure 29 shows corresponding CYs from selected fields from the oxygen reduction slow voltammetric cycling experiment, obtained from PdCo particles supported on crystalline TiWO at 13.0 at.% W (see Figure 28 row 5 highlighted by the solid rectangle selection). The percentage of cobalt in the particles varies from 0 to at.%, while the particles size varies from 4.4 to 5.6 nm. As mentioned before, although it has been demonstrated (not shown here) that we could obtain a constant thin film thickness during deposition across a row, the particles size may vary with composition. Figure 30 shows selected fields from Figure 28 (see dashed rectangle selection) where the percentage of cobalt has been kept constant at 35 at.°/b in PdCo and the size of particles varies from 3.0 to 8.1 nm diameter. Together these figures show that by adding a small amount of cobalt to palladium (15 at.% Co) the current improves and that the optimum particle size appears around 7 nm, similarly to the amorphous support.
The ignition potential (measured at 0.7 pA) for the array (with 13 at.% W) is plotted against the field in Figure 3 Ia). The white area represents the region of highest ignition potential hence giving the conditions in particle alloy composition and size of particles for the most promising materials. Figure 31b) is an enlargement in the ignition potential scale of Figure 31a), in the range 0.6 to 0.67V vc. RHE which highlights the hotspot which corresponds to a peak activity at approximately 6-7 nm for PdCo with 15 to 27 at.% Co. An in-depth analysis of all data for PdCo particles with varying size and composition on crystalline TiWO supports with a varying amount of W was performed. The specific activity, the mass activity and the ignition potentials where extracted from these slow cyclic voltammetry data.
Comparison of PdCo particles on Crystalline TiWO support (3 equivalent thickness): Figurc 32 shows mass activity of Pd, at 065 V vs. RHE (obtaincd from thc cathodic swccp of slow swcep voltammetry 5 mV) as a function of atomic pcrccntagc of Co, for the 31 equivalent thickness of particles deposited (from 5.3 to 8.3 nrn diameter) supported on crystalline T1WO, supports.
Particles size diameter varies from: 5.3 to 8.3 depending on the ratio of PdCo and the amount of W within the substrate (5.3 to 5.9 for TiO and 6.1 to 8.3 for TiWO) although thc equivalent thickncss of material dcposited is identical. Thcsc plots dcmonstratc thc cffcct of thc atomic pcrccntagc of Co on the raw current (proportional to the geometric current density), specific activity and mass activity (for mass of Pd) obtained at 0.65 V vs. RHE on amorphous TiWO. The behaviour observed suggests that there is no loss of activity on the addition of up to approx. 30 at.% Co for some of these systems. At higher Co contents there is a sharp decrease in the activity, and by 70 at.% the activity is very low. This trend is observed for the range of samples investigated. For the crystalline materials, a binary oxide support with 7 at.% W appears to lead to the most active systems.
References: 1. Guerin, S., et al., ombinatoriai Electrochemical Screening qf Fuel Cell Electrocatalvsts. Journal of Combinatorial Chemistry, 2003. 6(1): p. 149-158.
2. Gasteiger, HA., et al., Activity henchmarlcc and requiretnentc for Pt, Pt-alloy, and non-Pt oxygen reduction catalysts for PEMFCs. Applied Catalysis B: Environmental, 2005. 56(1-2): p. 9-35.
3. Peuckert, M., et al., Oxygen Reduction on Small Supported Platinum Particles.
Journal of The Electrochcmical Society, 1986. 133(5): p. 944-947.
4. Xu, Z., et al., Effect of particle size on the activity and durability of the Pt/C electrocatalyst for proton exchange membrane fuel cells. Applied Catalysis B: Environmental, 2012. 111-112(0): p. 264-270.
5. Kinoshita, K., Particle-size effects JOr oxygen reduction on highly dispersed platinum in acid electrolytes. Journal of the Electrochemical Society, 1990.
137(3): p. 845-848.
6. International Centre for Diffraction Data database 2010 [00-021-1273].
7. Burke, L.D. and J.K. Casey, Journal of Electrochemical Society, 1993. 145: p. 1284-1291.
8. Hayakawa, Y., Journal of Applied Electrochemistry, 1992. 22: p. 1017-1024.
9. Hemandez-Creus, A., Journal of Physical Chemistry B, 2004. 108: p. 10785-10795.
10. Lin, Y., X. Cui, and X. Ye, Electrochemistry Communications, 2005. 7: p. 267-274.
11. Pattabiraman, R.,Applied Catalysis A: General, 1997. 153: p. 9-20.
12. Salyador-Pascual, J.J., S. Citalan-Cigarroa, and 0. Solorze-Feria, Journal of Power Sources, 2007. 172: p. 229-234.
13. Yang, W., 2007. 45: p. 397-401.
14. Schmidt, T.J., Journal of Electroanalytical Chemistry, 2001. 501: p. 132-140.
15. Brace, K.e.a., W02007!042841 Palladium Alloy Catatvsts hr Fuel Cell Gathodes, 2007.
16. Fernandez, J., Langmuir, 2006. 22: p. 10426-10431.
17. Fernandez, J., D.A. Walsh, and A.J. Bard, Journal of the American Chemical Society, 2005. 127: p. 357-365.
1.. Lee, K., Journal of The Electrochemical Society, 2006. 153: p. A20-A24.
19. Mustain, W.E., K. Kepler, and J. Prakash, Electrochemistry Communications, 2006. 8: p. 406-410.
20. Savadogo, 0., et al., Journal of New Materials /br Electrocheinical Systems.
2004. 7: p. 77-83.
21. Savadogo, 0., et al., Electrochemistry Communications, 2004. 6: p. 105-109.
22. Shao, M.H., Langmuir, 2006. 22: p. 10409-10415.
23. Shoa, M., K. Sasaki, and R.R. Adzic, Journal of the American Chemical Society, 2006. 128: p. 3526-3527.
24. Fernandez, J.L., et al., Journal of the American Chemical Society, 2005. 127: p. 13100-13101.
25. Raghuvccr, V., P.J. Fcrrcira, and A. Manthiram, Electrochemistry Communications, 2006. 8: p. 807.
26. Raghuveer, V., A. Manthiram, and A.J. Bard, Journal of Physical Chemistry B, 2005. 109: p. 22909-22912.
27. Walsh, D.A., J. Fernandez, and A.J. Bard, Journal of Electrochemical Society, 2006. 153: p. E99-E103.
28. Gucrin, S. and GA. Attard, Electrochemistry Communications, 2001. 3: p. 544-548.
29. Kadirgan, F., Journal of Electroanalytical Chemistry, 1981. 125: p. 89-103.
30. Lukaszewski, M. and A. Czerrwinski, Journal of Electroanalytical Chemistry, 2007. 606: p. 117-133.
31. Solla-Gullon, J., Electrochemistry Communications, 2002. 4: p. 716-721.
32. Watson, D.J. and GA. Attard, Surface Science, 2002. 515: p. 87-93.
33. Zoltowski, P. and E. Makowska, Physical Chemistry Chemical Physics, 2001. 3: p. 2935-2942.
34. Zejuilovic, R., Journal of Applied Electrochemistry, 1984. 14: p. 481-488.
35. Bard, A.J. and L.R. Faulkner, Elcctrochcmical methods: thndamcntals and applications. 2nd edition John Wiley & Sons, 2001.
36. Adzic, R., M. Vukmirovic, and K. Sasaki, Synthesis of metal-metal oxide catalysts and electrocatalysts using a metal cation adsorption/reduction and adatom replacement by more noble ones US 7 704918 B2, 2010: United States.
37. Adzie, R., J. Zhang, and M. Vukmirovic, Electrocataivst for oxygen reduction with reduced platinum oxidation and dtssolution rates US 2006/0263675 Al, 2006: United States.
38. Antolini, E. and E.R. Gonzalez, Ceramic materials as supports fhr low- temperature fuel cell catalysts. Solid State lonics, 2009. 180(9-10): p. 746-763.
39. Cai, M., et a!., Eiectrocatah'st Supports/br Fuel Cells US 2007/003 7041 Al, 2007: United States.
40. Chen, Y., et al., Atomic layer deposition assisted Pt-Sn02 hybrid catalysts on nitrogen-doped CNTc with enhanced electrocatalytic activities S/hr low temperature fhel celLc. International Journal of Hydrogen Energy, 2011.
36(17): p. 11085-11092.
41. Cui, X., et al., Graphitized mesoporous carbon supported Pt-Sn02 nanoparticles as a catalyst for methanol oxidation. Fuel, 2010. 89(2): p. 372-377.
42. Do, T.B., M. Cai, and MS. Ruthkosky, Mesoporous electrically conductive metal oxide catalyst supports WO 2009/152003 A2, 2009.
43. Guo, D.-J. and J.-M. You, Highly catalytic activin' of Pt electrocatalyst supported on sulphated Sn02/muiti-nulied carbon nanotube composites /br methanol electro-oxidation. Journal of Power Sources, 2012. 198(0): p. 127-131.
44. Hayden, B.E., et al., The influence of support and particle size on the platinum catalysed oxygen reduction reaction. Physical Chemistry Chemical Physics, 2009. 11(40): p. 9141-9148.
45. Huang, S.-Y., et al., Development of a Titanium Dioxide-Supported Platinum Catalyst with Ultrahigh Stability fhr Polymer Electrolyte Membrane Fuel Cell Applications. Journal of the American Chemical Society, 2009. 131(39): p. 13898-13899.
46. Huang, S.-Y., P. Ganesan, and B.N. Popov, Electrocatalytic activity and stability of niobium-doped titanium oxide supported platinum catalyst for polymer electrolyte membrane fuel cells. Applied Catalysis B: Environmental, 2010. 96(1-2): p. 224-231.
47. Weidner, J.W. and B.L. Garcia, Electrocatalyst support and catalyst supported thereon US 2009/0065 738 Al, 2009: United States.
48. Ye, J., et a!., Preparation of Pt supported on W03-C with enhanced catalytic activity by nzicrowave-pyrolysis method. Journal of Power Sources, 2010.
195(9): p. 2633-2637.
49. de Bruijn, F.A., V.A.T. Dam, and G.J.M. Janssen, Review: Durability and Degradation Issues of PEM Fuel Cell Components. Fuel Cells, 2008. 8(1): p. 3-22.
50. Guerin, S., et al., High-Throughput Synthesis and Screening of Ternary Metal ALoys for Electrocataivsis. The Journal of Physical Chemistry B, 2006.
110(29): p. 14355-14362.
51. Guerin, S. and BE. Hayden, Physical Vapor Deposition. Method Jbr the High-Throughput Synthesis of Solid-State Material Libraries. Journal of Combinatorial Chemistry, 2006. 8(1): p. 66-73.
52. Gucrin, S., et al., A combinatorial approach to the study of particle size tfèctc on supported electrocatalysts: oxygen reduction on gold. Journal of Combinatorial Chemistry, 2006. 8: p. 679-686.
53. Guerin, S., et a!., combinatorial approach to the study of particle size effects in electrocataivsis: synthesis of supported gold nanoparticles. Journal of Combinatorial Chemistry, 2006. 8: p. 791-798.
54. Hayden, B.E., et al., co oxidation on gold in acidic environments: particle size and substrate etjècts. journal of physical chemistry C, 2007. 111: p. 17044-17051.
55. Hayden, B.E., et a!,, The influence of Pt particle size on the suifbce oxidation of titania supported platinum. Physical Chemistry Chemical Physics, 2009.
11(10): p. 1564-1570.
56. Hsu, C.-S.L., C-K.; Chan, C.-C.; Chang, C-C.; Tsay, C.-Y., Thin Solid Films, 2006. 494: p. 228-233.

Claims (16)

  1. Claims 1 A mixed metal oxide material of titanium and tungsten comprising up to about atomic percent tungsten on a metals basis.
  2. 2 A mixed metal oxide material as claimed in Claim 1 comprising up to about atomic percent tungsten.
  3. 3 A mixed metal oxide material as claimed Claim I or Claim 2 wherein the material is formed in an amorphous phase.
  4. 4 A mixed metal oxide material as claimed in any preceding claim wherein the material is formed in a crystallinc phase.
  5. A mixed metal oxide material as claimed in Claim 4, wherein the material is formed in a crystalline phase by annealing of a material as claimed in Claim 3.
  6. 6 A mixed metal oxide material as claimed in any preceding claim wherein the atomic perccntagc of tungstcn is 26 at.% or less, prefcrably 23 at.% or css, more preferably 18 at.% or less.
  7. 7 A mixed metal oxide material as claimed in Claim 6 wherein the atomic pcrcentagc of tungstcn is about 7 at.%.
  8. 8 A catalytic material comprising a catalyst support and a catalyst supported by the catalyst support, wherein the catalyst support is a mixed metal oxide material of titanium and tungsten.
  9. 9 A catalytic medium as claimed in Claim 8 wherein the mixed metal oxide material is a material as claimed in any one of Claim I to 7.
  10. A catalytic material as claimcd in Claim 9 wherein the catalyst is formed as particles on the catalyst support.
  11. 11 A catalytic material as claimed in Claim 10 wherein the particle size is 50 nm or less; preferably from I nm to 10 nm; more preferably from 5 nm to S nm.
  12. 12 A catalytic material as claimed in any one of claims 9 to 11 whcrcin thc catalyst is palladium or a palladium alloy.
  13. 13 A catalytic material as claimed in Claim 12 wherein the catalyst is a palladium alloy of palladium and cobalt.
  14. 14 A catalytic material as claimed in Claim 13 wherein the palladium alloy consists essentially of palladium and up to about 70 at.% cobalt preferably up to about 40 at.% cobalt, more preferably up to about 30 at.% cobalt; most preferably 15- 27 at.% cobalt.
  15. 15 A fuel cell comprising a catalytic material as claimed in any one of claims 9 to 14.
  16. 16 Usc of a mixed metal oxide material as claimed in any one of claims Ito 7 or of a catalytic material as claimed in any one of claims 8 to 14 in a fuel celL
GB1300823.0A 2013-01-16 2013-01-16 Mixed metal oxide materials of titanium and tungsten Withdrawn GB2509917A (en)

Priority Applications (2)

Application Number Priority Date Filing Date Title
GB1300823.0A GB2509917A (en) 2013-01-16 2013-01-16 Mixed metal oxide materials of titanium and tungsten
PCT/GB2014/050121 WO2014111717A1 (en) 2013-01-16 2014-01-16 Catalytic and composite materials

Applications Claiming Priority (1)

Application Number Priority Date Filing Date Title
GB1300823.0A GB2509917A (en) 2013-01-16 2013-01-16 Mixed metal oxide materials of titanium and tungsten

Publications (2)

Publication Number Publication Date
GB201300823D0 GB201300823D0 (en) 2013-02-27
GB2509917A true GB2509917A (en) 2014-07-23

Family

ID=47758095

Family Applications (1)

Application Number Title Priority Date Filing Date
GB1300823.0A Withdrawn GB2509917A (en) 2013-01-16 2013-01-16 Mixed metal oxide materials of titanium and tungsten

Country Status (2)

Country Link
GB (1) GB2509917A (en)
WO (1) WO2014111717A1 (en)

Citations (4)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
JP2005013946A (en) * 2003-06-27 2005-01-20 Mitsubishi Chemicals Corp Catalyst for water gas reaction, method for reducing carbon monoxide and fuel cell power generation system
US20070082257A1 (en) * 2005-09-29 2007-04-12 Kabushiki Kaisha Toshiba Membrane electrode, membrane electrode assembly, fuel cell, method of manufacturing the membrane electrode
EP1859861A1 (en) * 2005-03-18 2007-11-28 Nippon Shokubai Co.,Ltd. Catalyst for removing oxygen and method for removing oxygen using the catalyst
WO2010135576A2 (en) * 2009-05-21 2010-11-25 Cornell Research Foundation, Inc. Conducting metal oxide and metal nitride nanoparticles

Patent Citations (4)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
JP2005013946A (en) * 2003-06-27 2005-01-20 Mitsubishi Chemicals Corp Catalyst for water gas reaction, method for reducing carbon monoxide and fuel cell power generation system
EP1859861A1 (en) * 2005-03-18 2007-11-28 Nippon Shokubai Co.,Ltd. Catalyst for removing oxygen and method for removing oxygen using the catalyst
US20070082257A1 (en) * 2005-09-29 2007-04-12 Kabushiki Kaisha Toshiba Membrane electrode, membrane electrode assembly, fuel cell, method of manufacturing the membrane electrode
WO2010135576A2 (en) * 2009-05-21 2010-11-25 Cornell Research Foundation, Inc. Conducting metal oxide and metal nitride nanoparticles

Non-Patent Citations (4)

* Cited by examiner, † Cited by third party
Title
K. Akurati et al. Flame-made WO3/TiO2 nanoparticles: Relation between surface acidity, structure and photocatalytic activity. Applied Catalysis B: Environmental 79, 53-62, 2008 *
M.N. Taylor et al. Synergy between tungsten and palladium supported on titania for the catalytic total oxidation of propane. *
M.N. Taylor et al. Synergy between tungsten and palladium supported on titania for the catalytic total oxidation of propane. Journal of Catalysis 285, 103-114, 2012 *
Q Lin et al. Hydrogenation of pyrene using Pd catalysts supported on tungstated metal oxides. Applied Catalysis A: General 387, 166-172, 2010 *

Also Published As

Publication number Publication date
GB201300823D0 (en) 2013-02-27
WO2014111717A1 (en) 2014-07-24

Similar Documents

Publication Publication Date Title
EP2945909B1 (en) A core-shell catalyst comprising a mixed metal oxide material of tin and titanium
US9837668B2 (en) Cost-effective core-shell catalyst with high electrochemical stability
US10041179B2 (en) Non-noble metal based electro-catalyst compositions for proton exchange membrane based water electrolysis and methods of making
US8158548B2 (en) Electrocatalysts based on mono/plurimetallic carbon nitrides for fuel cells fueled with hydrogen
JP6997167B2 (en) Method for Producing Electrode Catalyst from Perovskite Metal Oxide
US11552303B2 (en) Methods, catalysts, and supports for electrochemical devices
US9614228B2 (en) Composite materials
JP2007257888A (en) Oxygen pole catalyst for solid polymer fuel cell, oxygen reduction electrode using it, and manufacturing method of those
Ziegelbauer et al. Chalcogenide electrocatalysts for oxygen-depolarized aqueous hydrochloric acid electrolysis
Park et al. Direct fabrication of gas diffusion cathode by pulse electrodeposition for proton exchange membrane water electrolysis
JP5382114B2 (en) Catalyst electrode for fuel cell
US7695849B2 (en) Low Pt content direct methanol fuel cell anode catalyst: nanophase PtRuNiZr
US9634332B2 (en) Composite materials
US9040210B2 (en) Stable, durable carbon supported catalyst composition for fuel cell
Sun et al. Thin films fabricated by pulsed laser deposition for electrocatalysis
Davies et al. Stabilising oxide core—platinum shell catalysts for the oxygen reduction reaction
US11142837B2 (en) Electrocatalytic materials for oxygen evolution of formula A1-xBxO3-δ
GB2509917A (en) Mixed metal oxide materials of titanium and tungsten
JP5207461B2 (en) Electrocatalyst and electrode using the same
JP2015150480A (en) catalyst for fuel cell

Legal Events

Date Code Title Description
WAP Application withdrawn, taken to be withdrawn or refused ** after publication under section 16(1)