Electrocatalytic activity of AuCeO2/C towards ethylene glycol oxidation and oxygen reduction reactions

Virginija Kepenienė1*,

Raminta Stagniūnaitė1,

Daina Upskuvienė1,

Loreta Tamašauskaitė-Tamašiūnaitė1,

Vidas Pakštas1,

Audrius Drabavičius1,

Mindaugas Andrulevičius2,

Eugenijus Norkus1

1 Department of Catalysis, Center for Physical Sciences and Technology, 3 Saulėtekio Avenue, 10257 Vilnius, Lithuania

2 Institute of Materials Science of Kaunas University of Technology, 59 Baršausko Street, 50131 Kaunas, Lithuania

AuCeO2/C and Au/C catalysts were obtained by adsorption of AuNPs on the CeO2/C and pure carbon (C) substrates from an Au colloidal solution. It has been found that AuNPs of ca. 50 nm in size were adsorbed on the surfaces of CeO2/C and C; however, electrocatalytic activity of the  investigated AuCeO2/C and Au/C catalysts was different. Ethylene glycol oxidation current density values are ca. 3 times higher on the AuCeO2/C catalyst as compared to those of the bare Au/C catalyst. Moreover, the AuCeO2/C catalyst showed more positive onset potentials, as well as higher current in the mixed-kinetic-diffusion region towards the oxygen reduction reaction in an alkaline medium compared to that of the Au/C catalyst.

Keywords: gold nanoparticles, cerium(IV) oxide, ethylene glycol oxidation, oxygen reduction

INTRODUCTION

Low-temperature polymer electrolyte membrane fuel cells (PEMFCs) are undergoing rapid development for mobile applications and, in particular, for the transport sector [1]. Different fuels have been used for fuel cells, whereas hydrogen, methanol and ethanol have been mostly explored. The operation of PEMFCs is based on the oxidation of fuel at the  anode and the  oxygen reduction reaction (ORR) at the cathode. In the case of direct methanol, ethanol or ethylene glycol (EG) fuel cells (DMFC, DEFC or DEGFC), acidic or alkaline alcohols solutions can be used as fuel. However, oxidation of fuels like alcohols in an alkaline medium is higher than in an acidic one. Moreover, PEMFCs that used an alkaline alcohols solution as fuel are more attractive and less polluted in comparison with the ones that used an acidic solution as fuel [24]. Forasmuch, oxidation reactions of ethanol, methanol or EG in an alkaline medium as well as the ORR have drawn increasing interest with the aim to develop active catalysts for DMFC, DEFC or DEGFC in an alkaline medium.

Application of DMFC is widely investigated, but problems such as the low methanol electro-oxidation kinetics and methanol permeation across the proton exchange membrane obstruct its commercialization [1, 5]. Although ethanol has been recognized as the most suitable fuel for the alkaline fuel cell, however, usually the main products of the  oxidation of ethanol on Au at the  lower temperatures are acetaldehyde and acetic acid, which complicate to get the clean catalytic process [68]. In this case, as the  actual and theoretical electron transfer numbers per one ethanol molecule are 4 and 12 for partial oxidation of ethanol to acetic acid and complete oxidation of ethanol to carbon dioxide, respectively, the electron transfer rate (ETR) of ethanol in the alkaline DEFC is only 33% [8]. EG has been found as another choice for alkaline fuel cells. It was determined that the ETR of EG in the alkaline DEGFC is ca. 80%. This value is much higher than that in the alkaline DEFC. In addition, EG owns many remarkable advantages including low toxicity, favourable storage, ease of transport and production from biomass and industry [810].

In order to improve the  performance of fuel cells, it is very important to develop effective and selective catalysts for fuel oxidation and ORR. Au has been investigated as an excellent catalyst in an alkaline medium [5, 1114]. Comparing peculiarities of the application of Au and platinum (Pt) for the oxidation reaction of alcohols in an alkaline medium, a higher oxidation activity shows Au as compared with that for Pt. This can be explained by the resistance of the Au surface to the poisoning by intermediates, which are formed during the oxidation of alcohol [2, 14]. Ureta-Zanartu et al. compared activity of Au for methanol and EG oxidation reactions and explored that EG is oxidized at a higher rate [13]. Au has been found to be the best catalyst for the oxidation of EG owing to the fact that EG adsorbs strongly onto the surface of Au [1521]. The highly dispersed Au particles on the  metal oxides supports, including Co3O4, SnOx, MnO2, RuO2, etc., exhibit an extraordinarily high activity for the oxidation reaction of various materials at low temperatures [2226] as well as for ORR [27, 28]. Cerium(IV) oxide (CeO2) is an attractive co-catalyst for various noble metals due to its oxygen vacancy, i.e. oxygen storage capacity ability, functioning as an oxygen buffer by storing/releasing O2 due to the  Ce4+/Ce3+ redox couple. This results in supplying sufficient OHads at low potential values and the efficient eliminating of poisoning intermediates during the  reaction. Other positive properties of CeO2, such as low price, synergistic effect between CeO2 and metal as well as tolerance to the  CO poisoning, determine the choice of this oxide as an additive in the preparation of various catalysts [22, 29, 30]. CeO2 is widely used for combining it with Pt, Pd or other oxides [3133]. However, nanosized Au particles supported on metal oxides have been extensively studied because of their high reactivity in a  variety of important catalytic reactions. Although, the interaction between Au and CeO2 is not widely described, the  combination of CeO2 and Au may lead to the  formation of materials that have an enhanced electrocatalytic activity for the ORR and EG oxidation [13, 22].

In the  paper presented herein, we investigated the  electrocatalytic activity of AuCeO2/C and Au/C catalysts prepared by the adsorption of AuNPs on the CeO2/C and C substrates from an Au colloidal solution. The composition, morphology and structure of the prepared catalysts have been characterized using inductively coupled plasma optical emission spectroscopy (ICP-OES), field emission scanning electron microscopy (FESEM), transmission electron microscopy (TEM), X-ray photoelectron spectroscopy (XPS) and X-ray diffraction (XRD). Electrocatalytical properties of the prepared AuCeO2/C and Au/C catalysts have been investigated towards the  oxidation of EG in an alkaline medium using cyclic voltammetry (CV), whereas the reduction of oxygen has been investigated by means of linear sweep voltammetry (LSV) using the rotating disk electrode (RDE) method.

It has been found that the AuCeO2/C catalyst showed an enhanced electrocatalytic activity towards the oxidation of EG and ORR in an alkaline medium compared to the bare Au/C catalyst.

EXPERIMENTAL

Chemicals

HAuCl4 (99.995%), CeO2 powder (99.9%) and carbon powder (99.999%) were purchased from Sigma-Aldrich and Alfa-Aesar Supplies. Polyvinylidenefluoride (PVDF), N-methyl-2-pyrrolydone (NMP), ethylene glycol (CH2OH)2, ethanol (96%), H2SO4 (96%), glucose (99.5%), ascorbic acid (99%) and NaOH (98.8%) were purchased from Chempur Company. Oxygen gas (99.999%) was used for the saturation of the NaOH solution. Pt/C wt. 46.4% Pt (TEC10E50E) was purchased from Tanaka Kikinzoku Kogyo K. K. Supplier. All chemicals were of analytical grade. Ultra-pure water with a resistivity of 18.2 MΩ cm–1 was used for preparing the solutions.

Preparation of catalysts

For the  formation of the  AuCeO2/C catalyst, at first, CeO2/C substrate was prepared and further AuNPs were deposited on the  CeO2/C surface from the Au colloidal solution. The CeO2/C substrate was prepared by the  following procedure: dry CeO2 powder was mixed with carbon powder (mass ratio being 1:1) in a 2-propanol solution by ultrasonication for 30 min with further drying of the  obtained mixture. The  Au colloidal solution was prepared by adding 1 ml of the solution containing 0.1 M glucose and 0.05 M ascorbic acid to 30 ml of the 0.3 mM HAuCl4 solution. The solution quickly got cherry red colour indicating reducing of Au3+ to Au0. Then, for the preparation of the AuCeO2/C catalyst, 100 mg of the prepared CeO2/C powder was added to the  obtained Au colloidal solution. The mixture was stirred for 1 h until the  solution became transparent. The  obtained AuCeO2/C catalyst powder was filtered and dried in a vacuum oven at 80°C temperature for 2 h. The Au/C catalyst was prepared in the same manner by using 100 mg of pure carbon powder instead of CeO2/C.

Characterization of catalysts

XRD patterns of the  studied AuCeO2/C and Au/C powders were measured using an X-ray diffractometer SmartLab (Rigaku) equipped with an X-ray tube with a  9  kW rotating Cu anode. The measurements were performed using Bragg–Brentano geometry with a  graphite monochromator on a diffracted beam and a step scan mode with a step size of 0.02° (in 2θ scale) and counting time of 1s per step. The measurements were conducted in a 2θ range of 20–70°. Phase identification was performed using the software package PDXL (Rigaku) and the ICDD powder diffraction data base PDF4+ (2018 release). The size of crystallites was calculated from XRD peaks broadening using the  graphical Halder–Wagner method in the PDXL software.

A shape and size of AuNPs were examined using a transmission electron microscope Tecnai G2 F20 X-TWIN equipped with an EDAX spectrometer with an r-TEM detector. For microscopic examinations, 10 mg of the sample was first sonicated in 1 ml of ethanol for 1 h and then the obtained mixture was deposited on the Cu grid covered with a continuous carbon film.

The morphology and composition of the fabricated catalysts were characterized using a SEM/FIB workstation Helios Nanolab 650 with an energy dispersive X-ray (EDX) spectrometer INCA Energy 350 X-Max 20.

The chemical composition of the prepared catalysts was determined by means of X-ray photoelectron spectroscopy (XPS) using an ESCALAB 250Xi Avantage spectrometer (VG Scientific, UK) equipped with an Al Kα X-ray radiation source (1486.6 eV) and operated at a fixed pass energy of 40 eV. The base pressure in the analytical chamber was better than 8 × 10–7 Pa.

The Au loading in the  prepared catalysts was estimated by means of ICP-OES using an ICP optical emission spectrometer Optima 7000DV (Perkin Elmer).

Electrochemical measurements

Electrochemical measurements were assessed using the RDE and the AUTOLAB electrochemical workstation. A  three-electrode conventional cell was used, where the glassy carbon RDE with a geometric area of 0.07 cm2 was used as the working electrode, a  Pt sheet was used as a  counter electrode and an Ag/AgCl/KClsat electrode was used as a reference. The  catalysts suspensions were obtained according to the following steps: at first, 10 mg of the AuCeO2/C and Au/C catalysts were dispersed ultrasonically in 100 µl of 2% PVDF in the NMP solution for 1 h. With the aim to keep a similar amount of Au during the measurement, a required amount of the prepared catalyst suspension was pipetted onto the polished surface of a glassy carbon electrode and dried in an oven at 80°C temperature for 4 h.

In order to determine electrochemically active surface areas (ESAs) of AuNPs in the  prepared catalysts, the  cyclic voltammograms (CVs) were recorded in a 0.5 M H2SO4 solution. The electrode potential was cycled in a range of –0.2–1.4 V with a scan rate of 50 mV s–1. The oxidation of EG in an alkaline medium was investigated by recording CVs in a 1 M (CH2OH)2 + 0.5 M NaOH solution at a scan rate of 50 mV s–1 from –0.3 to 1.1 V. Chronoamperometric measurements (CA) were carried out in the same solution at a constant potential value of 0.5  V for 30  min. The  electrode potential values are quoted versus the silver/silver chloride electrode (Ag/AgCl/KClsat).

For ORR measurements, an O2-saturated 0.1 M NaOH solution was used. Linear sweep voltammetry curves (LSVs) were recorded in the  electrode potential range from 1 to 0.2 V in the cathodic direction at a  scan rate of 5  mV  s–1 and varying the  rotation speed from 0 to 2000  rpm. CA for ORR were recorded at a potential value of 0.5 V in the O2-saturated 0.1 M NaOH solution for 30 min. The electrode potential values for ORR are quoted versus the  reversible hydrogen electrode (RHE).

Koutecky–Levich (K–L) equations (13) were used for calculation of the electron transfer numbers per oxygen molecule (n) in ORR:

j 1 = j k 1 + j d 1 ,                                        (1)
j d =0 .62nFD 2/3  Co 2 ν 1/6 ω 1/2 ,                  (2)
j 1 = j k 1 + ω 1/2 ×A.                            (3)

Here j, jk and jd are the measured current density, kinetic current density and the diffusion-limiting current density (mA cm–2), respectively; n is the number of electrons transferred in the reaction; F is the Faraday constant (F = 96.485 C mol–1); D is the diffusion coefficient of the reactant (1.9 × 10–5 cm2 s–1); Co2 is the concentration of the reactant in the bulk electrolyte (1.2 × 10–3 mol L–1); ν is the kinetic viscosity of the electrolyte (1.13 × 10–2 cm2 s–1); ω is the rotation rate; A is the slope of the linear plot of j−1 versus ω−1/2 (K–L plot) [34].

All measurements were carried out at 25°C temperature. All solutions were dearated with Ar before each measurement, except ORR measurements. The presented current densities are normalized with respect to the geometric area of the working electrode.

RESULTS AND DISCUSSION

Figure 1 presents the XRD profiles of the Au/C (pattern a) and AuCeO2/C (pattern b) catalysts. As evident, the XRD pattern of the Au/C catalyst shows diffraction peaks at 2θ  =  38.18, 44.39 and 64.58° which were assigned to the (111), (200), and (220) crystallographic planes of cubic Au phase (ICDD # 00-004-0784). The received Au diffraction peaks are rather broad, which indicates that the Au crystallites are quite small with an average size of 9.8 ± 0.5 nm (Fig. 1, pattern a). In the case of AuCeO2/C (Fig. 1, pattern b), the obtained diffraction peaks were assigned to the same crystallographic planes of cubic Au phase at the same 2θ values. The characteristic diffraction peaks obtained at 2θ = 28.58, 33.12, 47.54 and 56.41° were assigned to the (111), (200), (220) and (311) crystallographic planes of cubic CeO2 phase (ICDD  #  00-034-0394) (Fig.  1, pattern  b). The Au and CeO2 crystallites ca. 4.5 ± 0.4 nm and ca. 16.2 ± 0.6 nm in size, respectively, were estimated in the AuCeO2/C catalyst. The theoretical cubic Au lattice constant A is 4.0786 Å, while that of CeO2 is 5.4113 Å. For both Au/C and AuCeO2/C catalysts, no changes in the  lattice parameters for Au and CeO2 were observed. Therefore, it can be assumed that solid solutions did not form in the synthesized AuCeO2/C catalyst. However, the size of Au crystallites in the  AuCeO2/C catalyst was approximately 2-fold smaller than those in the pure Au/C catalyst.

X-ray photoelectron spectroscopy was used for the  determination of surface elemental composition and chemical state of the  prepared catalysts. The XPS spectra of the Au 4f, O 1s, Ce 3d and C 1s regions for AuCeO2/C and Au/C are given in Fig. 2. The  binding energy values (BEs) of spin-orbital splitting peaks are listed in Table 1. The Au4f XPS spectra of the Au/C and AuCeO2/C catalysts showed doublets assigned to the 4f7/2 and 4f5/2 phases for metallic Au with binding energy differences of 4.03 and 3.6 eV, respectively (Fig. 2a) [35, 36]. For the O 1s XPS spectra, the BE peak at 532.22 eV was attributed to the presence of the surface of adsorbed O2, H2O or CO2 for both Au/C and AuCeO2/C catalysts (Fig. 2b) [37]. The BE peak at 529.62 eV in the O 1s spectrum for the AuCeO2/C catalyst was assigned to lattice oxygen species (Fig. 2b, a dashed line) [37]. The Ce 3d spin-orbital splitting peaks were attributed to the 3d5/2 and 3d3/2 transitions of Ce, which correspond to Ce4+ (Fig. 2c, Table 1) [38, 39]. Figure 2 d shows XPS spectra of the C 1s region with a spin-orbital splitting peak at 284.49 eV, which is assigned to carbon [40, 41].

Fig. 1. XRD patterns of the Au/C (a) and AuCeO2/C (b) catalysts
Fig. 2. XPS spectra of the Au 4f (a), O 1s (b), Ce 3d (c) and C 1s (d) regions for the Au/C (a solid line) and AuCeO2/C (a dashed line) catalysts
Table 1. XPS analysis of the elemental composition of catalysts
Sample Au 4f Ce 3d O 1s C 1s
Eb, eV at.% Eb, eV at.% Eb, eV at.% Eb, eV at.%
AuCeO2/C 87.36 0.60 916.40 9.96 532.22 27.38 284.49 62.06
83.76 907.20 529.62
898.41
882.41
Au/C 87.99 0.19 532.22 3.07 284.49 96.73
83.96

In order to evaluate the size and shape of AuNPs in the prepared catalysts, TEM analysis was carried out (Fig. 3). It is clearly seen that spherical AuNPs were deposited on the surfaces of CeO2/C (Fig. 3a) and carbon (Fig. 3b). The size of AuNPs is ca. 40–50 nm in both prepared catalysts. SEM images of the AuCeO2/C and Au/C catalysts (Fig. 4) confirmed the data obtained by the TEM analysis. It is evident that spherical AuNPs of ca. 50 nm in size were deposited on the CeO2/C (Fig. 4a) and carbon (Fig. 4b) substrates.

Electrochemically active surface area values of AuNPs in the  AuCeO2/C and Au/C catalysts were determined from CVs recorded in a 0.5 M H2SO4 solution (Fig. 5) and by integrating the area of the peak associated with the reduction of AuO at about 1.0  V with a  charge density of 400 μC cm–2 [42]. The summarized data are presented in Table 2. The Au loading in the prepared catalysts determined by ICP-OES are also listed in Table 2.

Fig. 3. TEM images of the AuCeO2/C (a) and Au/C (b) catalysts
Fig. 4. SEM images of the AuCeO2/C (a) and Au/C (b) catalysts
Fig. 5. CVs of AuCeO2/C (a solid line) and Au/C (a dashed line) recorded in a 0.5 M H2SO4 solution at 50 mV s–1

The electrocatalytic activity of the  AuCeO2/C and Au/C catalysts were examined for the oxidation of EG in an alkaline medium by cyclic voltammetry. Long-term CVs for the  investigated catalysts recorded in a  1 M (CH2OH)2  +  0.5  M NaOH solution at a sweep rate of 50 mV s–1 are shown in Fig. 6. As evidence, two anodic current peaks I and II in the forward and reverse scans, respectively, were recorded at both AuCeO2/C and Au/C catalysts (Fig. 6a, b). These anodic peaks are related to the oxidation reaction of hydrocarbons in the  forward scan and with incompletely oxidized carbonaceous residues on the catalyst surface during the reverse scan [43]. The maximum current density of anodic peaks I was reached at ca. 0.5 V at both AuCeO2/C and Au/C catalysts, whereas anodic peaks II appear at ca. 0.25 V in the  reverse scan (Fig.  6a, b). It should be noted that during long-term scanning, EG oxidation current density values corresponded to the peak I on both catalysts increase, indicating a high activity and stability of catalysts.

Fig. 6. CVs of the Au/C (a) and AuCeO2/C (b) catalysts recorded in a 1 M (CH2OH)2 + 0.5 M NaOH solution at 50 mV s–1; 25°C. (c) Stabilized positive potential-going scans of the same catalysts at the same conditions

Particular current density values of the  anodic peaks of forward (jF) and reverse (jR) scans are given in Table  2. The  ratio (jF/jR) of forward anodic peak current density to the reverse anodic peak expresses an index of the catalysts tolerance to the poisoning species accumulated on the surface of the  electrode. For AuCeO2/C and Au/C jF/jR ratios were calculated and they are equal to 3.42 and 2.13, respectively. Higher jF/jR ratio indicates a more efficient oxidation of EG with a little accumulation of carbonaceous residues during the forward scan [43, 44]. According to the jF/jR data, the  CeO2 supported AuNPs catalyst shows its better tolerance to the poisoning of carbonaceous residues as compared with that of the carbon supported AuNPs catalyst.

The stabilized positive potential-going scans (10th cycle of scan) of AuCeO2/C and Au/C are presented in Fig. 6c. It was found that the onset potential for the oxidation of EG on the AuCeO2/C catalyst starts at ca. –0.2 V. Moreover, this value is more negative compared to that of the  Au/C catalyst, which is ca. –0.1  V. EG oxidation current density that corresponds to the peak I is 2.5 times higher at the AuCeO2/C catalyst, compared with that for the bare Au/C catalyst. These results greatly highlighted the enhanced catalytic activity and better stability of the  AuCeO2/C catalyst, which may be attributed to the  synergistic catalytic effects of AuNPs and CeO2.

Figure  7 illustrates the  EG oxidation current densities, which were normalized by the electrochemically active surface areas of AuNPs in the catalysts and the Au loading for each catalyst. The values obtained represent the  mass (mA  mgAu–1) (Fig. 7a) and specific (mA cm–2) (Fig. 7b) activities of the AuCeO2/C and Au/C catalysts towards EG oxidation. The mass and specific activities for EG oxidation were ca. 2 and 8 times higher, respectively, at the AuCeO2/C catalyst compared with those for the Au/C catalyst (Fig. 7a, b).

Table 2. Electrochemical parameters of the Au/C and AuCeO2/C catalysts
Catalyst ESA, cm–2 Au loading, μg cm–2 jF, mA cm–2 jR, mA cm–2
Au/C 0.018 56.00 13.81 6.47
AuCeO2/C 0.006 81.00 34.99 10.24
Fig. 7. Comparison of the mass (a) and specific (b) activities towards the oxidation of EG under peaks I for AuCeO2/C and Au/C recorded in 1 M (CH2OH)2 + 0.5 M NaOH at 50 mV s–1

The  electrochemical stability of as-prepared catalysts for the oxidation of EG was investigated using CA (Fig. 8). The potential was firstly held at open circuit for 10 s, then set to 0.5 V for 1900 s in a 1  M (CH2OH)2  +  0.5  M NaOH solution at 25°C. As seen in Fig. 8a, a current density decay was observed at first 5 min. Then it gradually stabilizes and reaches the current density values ca. 5 and 13 mA cm–2 for Au/C and AuCeO2/C, respectively. As evident, the  AuCeO2/C catalyst shows a higher current density at the end of experiment as compared to that at Au/C (Fig. 8a). The mass and specific activity values were calculated from the current densities obtained at the end of the experimental period (t  =  1900  s) and are given in Fig. 8b and c. It is evident that mass (Fig. 8b) and specific (Fig. 8c) activity values of the AuCeO2/C catalyst are ca. 2 and 8 times higher as compared to those for Au/C, respectively. These data confirm the results obtained by CV.

The as-prepared Au/C and AuCeO2/C catalysts were also investigated towards the oxygen reduction reaction. Figures 9a and b show the ORR polarization curves recorded on the investigated catalysts in the O2-saturated 0.1 M NaOH solution. The  electrode potential was scanned in the  cathodic direction from 1.0 to 0.2 V at a scan rate of 5 mV s−1. The rotation speed was varied from 400 to 2000 rpm (Fig. 9a, b). From the data obtained, it is seen that AuCeO2/C and Au/C demonstrate a typical view of ORR curves with the onset potential of ca. 0.80 V. According to the literature data, at potential values between 0.7 and 0.9 V (vs RHE) the ORR response is dominated by the kinetics of electrocatalysts [45]. The  Koutecky–Levich plots were obtained from the RDE data on oxygen reduction at potentials from 0.2 up to 0.5 V for both catalysts and are shown in Fig. 9c and d. In general, the K–L curves display a linear relationship between j–1 and w–1/2 indicating the first-order dependence of the ORR kinetics at different potentials (0.2–0.5 V). According to the K–L equations (1–3), the  calculated number of electrons transferred (n) on both Au/C and AuCeO2/C catalysts was obtained ca. 2.5 (Fig. 9c, d, the insets). This indicates that the both AuCeO2/C and Au/C catalysts favour a  2-electron transfer reaction path and O2 is reduced to HO2and OH [46, 47].

Fig. 8. CA data for AuCeO2/C and Au/C studied in 1 M (CH2OH)2 + 0.5 M NaOH at 0.5 V vs Ag/AgCl for 1900 s. Bar columns of EG oxidation current densities obtained at the end of experimental period and normalized by the Au loading (b) and electrochemically active surface areas of AuNPs (c) for each catalyst

Figure  10a compares the  LSV scans of AuCeO2/C, Au/C and the commercial Pt/C catalyst that has the Pt loading of 0.071 mg Pt cm–2

at a rotation speed of 1600 rpm. The determined onset potential for Pt/C was equal to 0.95  V. It is clearly seen that the  AuCeO2/C onset potential should be more positive compared to that of the Pt/C catalyst. Nevertheless, LSVs display that AuCeO2/C demonstrated higher current density in the mixed-kinetic-diffusion region as compared with that at the Au/C catalyst. It can be assumed that the  CeO2 supported AuNPs catalyst shows an enhanced electrocatalytic activity towards the oxygen reduction reaction. The ORR current densities were also normalized by the  electrochemically active surface areas of AuNPs and Au loadings for both catalysts at a potential value of 0.75 V at a rotation speed of 1600 rpm (Fig. 10b). As seen from the  data obtained, the  mass and specific activity is 1.5 and 6 times higher, respectively, at the  AuCeO2/C catalyst compared with that at Au/C (Fig.  10b). The  data received confirms the  synergetic action between present metal and metal oxide in the  catalyst composition. Although the  mixed-kinetic-diffusion current densities are not low, the  onset potential should be at least of 0.1 V more positive for the efficient oxygen reduction reaction. The durability of the  prepared Au/C and AuCeO2/C catalysts was evaluated by CA measurements at a potential value of 0.5  V (vs RHE) for 1900s (Fig.  10c). It is seen that after approximately 200 s the current density settled down and only slightly increased with time. Figure  10c’ presents the  normalized ORR current densities (%) for each catalyst after 200  s. It is seen that the  Au/C, AuCeO2/C and commercial Pt/C catalysts maintained 82.2, 78.5 and 80.1%, respectively, of their initial signals. This manifests that the as-synthesized catalysts demonstrated a similar stability to commercial Pt/C during the ORR process in an alkaline medium.

We compared the  electrocatalytic activity towards the oxidation of EG and ORR of our prepared catalysts with other Au or CeO2 supported catalysts recently reported in the  literature (Table  3). From the  data in Table  3, it is seen that the as-prepared AuCeO2/C catalyst exhibits higher electrocatalytic activity towards the  ethylene glycol oxidation reaction.

Fig. 9. LSVs of the Au/C (a) and AuCeO2/C (b) catalysts recorded in an O2-saturated 0.1 M NaOH solution at 5 mV s–1 and varying the rotation speed from 0 to 2000 rpm. Koutecky–Levich plots of ORR collected for Au/C (c) and AuCeO2/C (d). The insets show the calculated electron transferred number for each catalyst
Fig. 10. (a) Comparison of LSVs of the Au/C, AuCeO2/C and commercial Pt/C catalysts recorded in an O2-saturated 0.1 M NaOH solution at 1600 rpm. (b) Bar columns of the ORR mass and specific activities for Au/C and AuCeO2/C calculated from LSVs at 0.75 V. (c) CA data of Au/C, AuCeO2/C and Pt/C recorded in a O2-saturated 0.1 M NaOH solution at 0.5 V vs RHE. Inset c’ represents chronoamperometric responses (percentage of current density retained vs operation time) of all the catalysts
Table 3. The comparison of onset potentials for the ORR and EG oxidation current densities of various catalysts presented in the literature and this study
Catalyst ORR EG oxidation Ref.
Eonset, V vs RHE Electrolyte jpeak I, mA cm–2 Electrolyte
Au/C 0.78 0.1 M NaOH 13.81 0.5 M NaOH + 1 M (CH2OH)2 This study
AuCeO2/C 0.82 34.99
Pd/Au 0.86 0.1 M NaOH [32]
CeO2/MnO2 0.92 0.1 M KOH [33]
Au-RGO 0.93 1 M KOH [48]
Au nanodendrites 0.88
Au/RGO 16.6 0.5 M NaOH + 0.1 M (CH2OH)2 [49]
Au 6.7 0.5 M NaOH + 0.1 M (CH2OH)2 [50]
PtCeO2 (4:1)/C 9.8 1 M KOH +1 M (CH2OH)2 [31]
PtCeO2 (2:1)/C 15.1
PtCeO2 (1.3:1)/C 19.2
PtCeO2 (1:1)/C 14.7

CONCLUSIONS

In summary, AuCeO2/C and Au/C catalysts have been prepared by the  adsorption of AuNPs on the CeO2/C and pure C substrates from an Au colloidal solution. The  oxidation of ethylene glycol and the  reduction of oxygen have been investigated in an alkaline medium.

The obtained results suggest that the AuCeO2/C catalyst showed ca. 3 times enhanced electrocatalytic activity towards the  oxidation of ethylene glycol compared to that of Au/C. Moreover, AuCeO2/C showed a more positive onset potential, as well as higher current density in the mixed-kinetic-diffusion region towards ORR in an alkaline medium compared to that at the bare Au/C catalyst. The enhanced electrocatalytic activity of AuCeO2/C catalyst should be attributed to the synergistic catalytic effects of AuNPs and CeO2.

Received 27 December 2019

Accepted 19 February 2020

References

  1. E. Antolini, E. R. Gonzalez, J. Power Sources, 195, 3431 (2010).
  2. P. Quaino, N. B. Luque, R. Nazmutdinov, E. Santos, W. Schmickler, Angew. Chem. Int., 51, 12997 (2012).
  3. K.  Matsuoka, Y.  Iriyama, T.  Abe, M.  Matsuoka, Z. Ogumi, J. Power Sources, 150, 27 (2005).
  4. M. L. Perry, T. F. Fuller, J. Electrochem. Soc., 149, S59 (2002).
  5. C. Bianchini, P. K. Shen, Chem. Rev., 109, 4183 (2008).
  6. L. An, T. S. Zhao, J. B. Xu, Int. J. Hydrogen Energy, 36, 13089 (2011).
  7. L. An, Z. H. Chai, L. Zeng, P. Tan, T. S. Zhao, Int. J. Hydrogen Energy, 38, 14067 (2013).
  8. L. An, R. Chen, J. Power Sources, 329, 484 (2016).
  9. K. Matsouka, Y.  Iriyama, T.  Abe, M.  Matsouka, Z. Ogumi, J. Electrochem. Soc., 152, A727 (2005).
  10. T. M. Brueckner, E. Wheeler, B. Chen, E. N. E. Sawy, P. G. Pickup, J. Electrochem. Soc., 166, F942 (2019).
  11. B. Beden, I. Cetin, A. Kahyaoglu, D. Takky, C. Lamy, J. Catal., 104, 37 (1987).
  12. Y. Kwon, S. C. S. Lai, P. Rodriguez, M. T. M. Koper, J. Am. Chem. Soc., 133, 6914 (2011).
  13. M. S. Ureta-Zanartu, A. Alarcon, G. Munoz, C. Gutiérrez, Electochim. Acta, 52, 7857 (2007).
  14. L. Xin, Z. Zhang, J. Qi, D. Chadderdon, W. Li, Appl. Catal. B, 125, 85 (2012).
  15. D.  Huang, Y.  Luo, S.  Li, M.  Wang, Y.  Shen, Electrochim Acta, 174, 933 (2015).
  16. X. Zhang, X. Wang, L. Le, A. Ma, S. Lin, J. Mater. Chem. A, 3, 19273 (2015).
  17. F.  Papiya, A.  Nandy, S.  Mondal, P.  P.  Kundu, Electrochim. Acta, 254, 1 (2017).
  18. E. N. Alvar, B. Zhou, S. H. Eichhorn, J. Mater. Chem. A, 17, 6540 (2016).
  19. J.  Liu, L.  Jiang, T.  Zhang, J.  Jin, L.  Yuan, G.  Sun, Electrochim. Acta, 205, 38 (2016).
  20. X. Lv, W. Lv, X. Zheng, C. Zhang, L. Zhi, Q. H. Yang, Chem. Commun., 51, 3911 (2015).
  21. H.  Y.  Y.  Ko, M.  Mizuhata, A.  Kajinami, S.  Deki, J. Electroanal. Chem., 559, 91 (2003).
  22. S. Yongprapat, A. Therdthianwong, S. Therdthianwong, J. Electroanal. Chem., 697, 46 (2013).
  23. D. Van Dao, G. Adilbish, T. Duc Le, T. T. D. Nguyen, I. H. Lee, Y. T. Yu, J. Catal., 377, 589 (2019).
  24. H. Gao, Y. Cao, Y. Chen, et al., J. Alloys Compd., 732, 460 (2018).
  25. L. Karuppasamy, C. Y. Chen, S. Anandan, J. J. Wu, Electrochim. Acta, 246, 75 (2017).
  26. I. Kaskow, P. Decyk, I. Sobczak, Appl. Surf. Sci., 444, 197 (2018).
  27. V.  Kepenienė, R.  Stagniūnaitė, L.  Tamašauskaitė-Tamašiūnaitė, et al., Mat. Chem. Phys., 241 (2020).
  28. N. Ostojic, Z. Duan, A. Galyamova, G. Henkelman, R. M. Crooks, J. Am. Chem. Soc., 140, 13775 (2018).
  29. V.  Kepenienė, L.  Tamašauskaitė-Tamašiūnaitė, J. Vaičiūnienė, R. Kondrotas, V. Pakštas, E. Norkus, Mater. Sci., 22, 243 (2016).
  30. M. M. Yusuf, A. B. Radwan, R. A. Shakoor, et al., J. Appl. Electrochem., 48, 391 (2018).
  31. J. Yang, X. Tan, Y. Qian, et al., J. Appl. Electrochem., 46, 779 (2016).
  32. S.  Štrbac, I.  Srejic, Z.  Rakočevic, J. Electroanal. Chem., 789, 76 (2017).
  33. J. Yang, J. Wang, L. Zhu, W. Zeng, J. Wang, Mater. Lett., 234, 331 (2019).
  34. K. M. Shi, X. Cheng, Z. Y. Jia, J. W. Guo, C. Wang, J. Wang, Int. J. Hydrogen Energy, 41, 16903 (2016).
  35. J. C. Fuggle, E. Kallne, L. M. Watson, D. J. Fabian, Phys. Rev. B, 16, 750 (1977).
  36. C. Battistoni, G. Mattogno, R. Zanoni, L. J. Naldini, Electron. Spectrosc. Relat. Phenom., 28, 23 (1982).
  37. P. T. Hsieh, Y. C. Chen, K. S. Kao, C. M. Wang, Appl. Phys. A, 90, 317 (2008).
  38. L.  Wang, L.  Zhuang, H.  Xin, Y.  Huang, D.  Wan, Open J. Inorg. Chem., 5, 12 (2015).
  39. E.  Beche, P.  Charvin, D.  Perarnau, S.  Abanades, G. Flamant, Surf. Interf. Anal., 40, 264 (2008).
  40. H. K. Jeong, E. Echeverria, P. Chakraborti, H. T. Le, P. A. Dowben, RSC Adv., 7, 10968 (2017).
  41. Y.  Zhu, M.  Yue, V.  Natarajan, et  al., RSC Adv., 8, 14879 (2018).
  42. H.  Angerstein-Kozlowska, B.  E.  Conway, A.  Hamelin, L.  Stoicoviciu, Electrochim. Acta, 31, 1051 (1986).
  43. Y. Zhao, L. Zhan, J. Tian, S. Nie, Z. Ning, Electrochim. Acta, 56, 1967 (2011).
  44. H. Qin, H. H. Yang, X. S. Zhang, P. Li, C. Mab, Int. J. Hydrogen Energy, 35, 7667 (2010).
  45. U.  Paulus, T.  Schmidt, H.  Gasteiger, R.  Behm, J. Electroanal. Chem., 495, 134 (2001).
  46. C. Song, J. Zhang, in: PEM Fuel Cell Electrocatalysts and Catalyst Layers, Springer, London (2008).
  47. X. Ge, A. Sumboja, D. Wuu, et al., ACS Catal., 5, 4643 (2015).
  48. L. Karuppasamya, C. Chena, S. Anandanc, J. Wub, Catal. Today, 307, 308 (2018).
  49. C. Jin, J. Zhu, R. Dong, Q. Huo, Electrochim. Acta, 190, 829 (2016).
  50. C. Jin, C. Sun, R. Dong, Z. Chen, Electrochim. Acta, 56, 321 (2010).

Virginija Kepenienė, Raminta Stagniūnaitė, Daina Upskuvienė, Loreta Tamašauskaitė-Tamašiūnaitė, Vidas Pakštas, Audrius Drabavičius, Mindaugas Andrulevičius, Eugenijus Norkus

AuCeO2/C ELEKTROKATALIZINIŲ SAVYBIŲ ĮTAKA ETILENGLIKOLIO OKSIDACIJOS IR DEGUONIES REDUKCIJOS REAKCIJOMS

Santrauka

Tirtas elektrokatalizinis AuCeO2/C ir Au/C katalizatorių aktyvumas etilenglikolio oksidacijos ir deguonies redukcijos reakcijoms. Minėti katalizatoriai buvo gauti adsorbcijos būdu, nusodinant aukso nanodaleles (AuNPs) ant cerio (IV) oksido / anglies (CeO2/C) ir grynos anglies (C) pagrindų iš koloidinio aukso tirpalo. Ant abiejų naudotų pagrindų nusodintos apie 50 nm dydžio AuNPs.

Nustatyta, kad AuCeO2/C katalizatorius pasižymi didesniu elektrokataliziniu aktyvumu tirtose etilenglikolio oksidacijos ir deguonies redukcijos reakcijose, palyginti su Au/C katalizatoriumi. Etilenglikolio oksidacijos metu AuCeO2/C katalizatorius rodė apie tris kartus didesnes srovės tankio vertes, palyginti su Au/C katalizatoriumi. Deguonies redukcija naudojant AuCeO2/C katalizatorių prasideda esant teigiamesnėms elektrodo potencialo vertėms nei su Au/C katalizatoriumi. Be to, išmatuotos kinetinės-difuzinės srities srovės tankio vertės taip pat yra didesnės naudojant pastarąjį elektrodą.

AuCeO2/C katalizatoriaus aktyvumas siejamas su sinergetine sąveika tarp Au nanodalelių ir CeO2. Gauti rezultatai leidžia manyti, kad šiomis sąlygomis gautas AuCeO2/C katalizatorius yra tinkamesnis atliekant etilenglikolio oksidacijos tyrimus.

* Corresponding author. Email: virginija.kepeniene@ftmc.lt