Next Article in Journal
Evaluation of the Zero Shear Viscosity, the D-Content and Processing Conditions as Foam Relevant Parameters for Autoclave Foaming of Standard Polylactide (PLA)
Next Article in Special Issue
Material Discovery and High Throughput Exploration of Ru Based Catalysts for Low Temperature Ammonia Decomposition
Previous Article in Journal
The Radial Piezoelectric Response from Three-Dimensional Electrospun PVDF Micro Wall Structure
Previous Article in Special Issue
Development of A Novel High Throughput Photo-catalyst Screening Procedure: UV-A Degradation of 17α-Ethinylestradiol with Doped TiO2-Based Photo-catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of Bismuth-Cerium Oxides for the Catalytic Oxidation of Diesel Soot

Faculty of Natural Science, Professorship of Chemical Technology, Chemnitz University of Technology, 09111 Chemnitz, Germany
*
Author to whom correspondence should be addressed.
Submission received: 28 February 2020 / Revised: 13 March 2020 / Accepted: 16 March 2020 / Published: 18 March 2020

Abstract

:
In this paper, the syntheses of a set of cerium-bismuth mixed oxides with the formula Ce1−xBixO2−x/2, where the range of x is 0.0 to 1.0 in 10 mol% steps, via co-precipitation methods is described. Two synthesis routes are tested: The “normal” and the so called “reverse strike” (RS) co-precipitation route. The syntheses are performed with an automated synthesis robot. The activity for Diesel soot oxidation is measured by temperature programmed oxidation with an automated, serial thermogravimetric and differential scanning calorimetry system (TGA/DSC). P90 is used as a model soot. An automated and reproducible tight contact between soot and catalyst is used. The synthesized catalysts are characterized in terms of the specific surface area according to Brunauer, Emmett and Teller (SBET), as well as the dynamic oxygen storage capacity (OSCdyn). The crystalline phases of the catalysts are analysed by powder X-ray diffraction (PXRD) and Raman spectroscopy. The elemental mass fraction of the synthesized catalysts is verified by X-ray fluorescence (XRF) analysis. A correlation between the T50 values, OSCdyn and SBET has been discovered. The best catalytic performance is exhibited by the catalyst with the formula RS-Ce0.8Bi0.2Ox which is synthesized by the reverse strike co-precipitation route. Here, a correlation between activity, OSCdyn, and SBET can be confirmed based on structural properties.

Graphical Abstract

1. Introduction

Diesel soot has proven to be carcinogenic [1] and furthermore, it can influence the environment, the vegetation, or the climate [2]. Diesel soot may cause lung or cardiovascular diseases [3,4]. Therefore, different technologies to control the emission of Diesel soot have been developed. Modern Diesel vehicles possess a Diesel particulate filter (DPF) system to filtrate Diesel soot from the exhaust stream. Over time, Diesel soot accumulates on the surface or in the channels of the DPF. The accumulated Diesel soot causes an increase in pressure drop, and this influences the fuel consumption or leads to filter failure [2]. In the absence of a filter regeneration process, the car engine will stop due to DPF plugging [5]. Today, active and passive regeneration processes are established. Another possibility is the use of fuel born catalysts based on Ce or Fe [6]. During the active regeneration process, the accumulated soot is periodically combusted in an oxidising atmosphere when the pressure drop reaches a preset limit. The heat for this process is generated by an electric heater, a flame-based burner, or microwave cavity [7]. At the active regeneration process, temperatures above 600 °C are reached. This causes on the one hand problems with materials of the DPF because of their melting point [8]. On the other hand, large amounts of energy are needed [9]. Therefore, the filter wall is coated with a catalyst [2]. During the passive regeneration, the accumulated soot is combusted continuously by chemical reaction. Here, no additional amount of fuel is needed and the oxidation temperature is equal to the exhaust temperature. The problem of the passive regeneration is the suitability and costs of catalysts. Common catalysts are based on precious metals [2].
Many researchers investigate soot oxidation with ceria as catalyst applied in three-way catalytic converters, in solid oxide fuel cells fed with hydrocarbons, or in the water-gas-shift reaction [10]. Ceria has the ability to incorporate oxygen vacancies into the lattice resulting in a substoichiometric phase CeO2−x through a reduction process [11]. The Ce1−xBixO2−x/2-system is a promising candidate for the passive regeneration of the DPF [12]. This mixed oxide system is free of precious metals. The investigation of this system is directly related to the introduction of vacancies in oxide ion sublattice. Oxide ion vacancies are inherent to ceria. Additionally, doping with ions with a lower formal charge than in Ce4+ and an asymmetric coordination sphere like Bi3+ ions increases the amount of oxide ion vacancies and reducibility of the oxide. Via s-p-hybridisation in the formal electron configuration of Bi3+ [Xe] 4f145d106s26p0 the coordination sphere of this ion is typically distorted (so-called lone pair) [10]. Dikmen et al. observed a high oxide ion conductivity for Bi-doped CeO2. They synthesized Bi-Ce mixed oxides via co-precipitation and subsequent hydrothermal aging at T = 900 – 1300 °C [13]. Zhao and Feng doped CeO2 with Bi3+ and M2+ (M = Ca2+, Sr2+, Ba2+). All of their synthesized oxides possess a high oxide ion conductivity. The highest oxide ion conductivity showed the following oxide: Ce0.95Ca0.05Bi0.4O2.55 [14]. Hund et al. synthesized Bi3+-doped CeO2 via solid state reaction at T = 800 °C. They found up to 40 mol% Bi3+ exclusively as fluorite-type CeO2 crystalline phase but with increased lattice constant of the cubic mixed oxide phase because of Bi3+ insertion [15]. A more detailed investigation of the crystal structure of Bi3+-doped CeO2 has not yet been made. This is because all the prepared Bi-Ce mixed oxides have nanocrystalline structures [10] and these reveal very broad reflections in PXRD pattern. First attempts to clarify the structure are made with the combination of extended X-ray absorption fine structure (EXAFS) and PXRD by Frolova et al. They discovered that the introduction of bismuth oxide in CeO2 leads to the formation of a single-phase solid solution system. The resulting structure is close to the fluorite-type structure of CeO2. With increasing Bi content, the unit cell parameter and the disorder of the structure increases [16]. Sardar et al. synthesized mixed oxides via hydrothermal synthesis starting from cerium chloride and sodium bismutate at T = 240 °C. They investigated the structure through XRD and X-ray absorption near-edge fine structure (XANES) measurements. They conclude that the mixed oxides have the formula Ce 1 x 4 + Bi x 3 + O 2 0 , 5 x (x ≤ 0.6) and reveal a locally distorted fluorite-type structure because of the unsymmetrical coordination of Bi3+ [10].
In this paper, a series of Bi-Ce mixed oxides with varying molar ratio between Ce and Bi are synthesized automatically via the synthesis robot Chemspeed Accelerator SLT 106. Two routes of co-precipitation are used: The normal co-precipitation route and the reverse strike co-precipitation route according to Lee at al. [17]. In the normal co-precipitation route, the precipitation agent is added to the metal salt precursor solution. The reaction runs from low to high pH values. Due to hydrolysis reactions, we have to deal with different beginnings of precipitation for different precursors. For the reverse strike co-precipitation route, the premixed precursor solution is added to the precipitation agent. Thus, the reaction starts in the basic pH range and ends at pH = 6. This method guarantees a homogenous start of the precipitation and prevents the oligomerisations of hydrated bismuth oxides in the acidic pH range [17,18]. The catalysts are characterized via PXRD, Raman spectroscopy, X-ray fluorescence (XRF), and specific surface measurements according to Brunauer, Emmett, and Teller (SBET). The catalytic activity for Diesel soot oxidation and dynamic oxygen storage capacity (OSCdyn) are measured by automated serial thermogravimetric analysis methods.

2. Materials and Methods

2.1. Catalyst Preparation

The chemicals specified below were used without any further purification.
A set of cerium-bismuth mixed oxides with the formula Ce1−xBixO2−x/2 in the range of 0.0 ≤ x ≤ 1.0 in 10 mol% steps were prepared via co-precipitation methods. The syntheses were performed with the synthesis robot Accelerator SLT106 from Chemspeed Technologies AG, Switzerland. Two co-precipitation synthesis routes were used: The normal co-precipitation with continuous addition of the precipitation agent via a 4-needle head (4NH) of the synthesis robot for liquid dosing to a given solution mixture of both metal salts and the reverse strike co-precipitation route according to Lee et al. [17] with reversal of this sequence of solution combination. The syntheses were performed in lined beakers with magnetic stirring bars, which were tempered via a thermostat from Huber Corp. to a temperature of T = 25 °C. The control of the thermostat was implemented in the software “ApplicationExecutor” for the synthesis robot. During the normal co-precipitation route, 0.5 M Ce(NO3)3·6H2O and 0.5 M Bi(NO3)3·5H2O (99.5%, Alfa Aesar) were dissolved in 1 M HNO3 and transferred in beakers as reservoirs. Via the 4NH of the synthesis robot, the corresponding molar amounts of the solutions were pipetted into lined beakers, mixed by magnetic stirring and kept at T = 25 °C. In a different lined beaker the precipitant, in our case a 1 M (NH4)2CO3 (food grade, BASF) solution, which was dissolved in ultra-pure water, was also brought to a temperature T = 25 °C. Via 4NH, the precipitant was transferred to the premixed nitrate solutions. The white to pale yellow precipitations were stirred for a further t = 0.5 h at T = 25 °C. After stirring, the precipitations were separated from the supernatants via a 3-fold parallel pressure filtration station at an overpressure of p = 1 bar. Polyethersulfone membranes (Sartorius Stedim Biotech Corp.) with 0.1 µm mean pore size were used for filtration. The filtrated samples were dried over night at room temperature, crushed and ground in an agate mortar until fine powders were received and subsequently calcined at T = 800 °C for t = 5 h. During the reverse strike co-precipitation route the corresponding molar amounts of the nitrates were premixed. The premixed nitrate solution was continuously dosed via the 4NH to the tempered precipitant until a preset pH was achieved (pH = 6). The subsequent downstream processes of stirring, filtration, and calcination were equal to the normal co-precipitation route. The dispensing and aspirating speeds are summarized in Table 1.

2.2. Activity Measurements—Soot Oxidation and OSCdyn

The catalytic performance for soot oxidation of the synthesized catalysts was investigated by automated serial thermogravimetric analyses coupled with heat flow differential scanning calorimetry (TGA-DSC 1 1600; Mettler Toledo Corp.). The model soot used was the carbon black P90 supplied by Evonik Degussa. In this work, the so-called tight contact in an automated mode was used. The tight contact was realized with an asymmetric dual centrifuge in a highly reproducible mode of operation. For this purpose, the SpeedmixerTM from Hauschild Corp. was used. The soot and catalyst with a weight ratio of 1:4 were mixed for t = 300 s at a rotation speed of rs = 3000 rpm. Ca. m = 10 mg of the mixed soot and catalyst was placed in a corundum crucible of volume V = 70 µl and heated to T = 700 °C at a heating rate of r = 5 °C·min−1 with a synthetic air flow of V ˙ = 25 mL·min−1. The activity of the catalyst was determined by the characteristic combustion temperature at a specific weight loss. In this work, the T50 temperatures were used. These were the temperatures where 50% of the soot was oxidized. These values were determined with the help of software Stare from Mettler Toledo Corp.
Another significant property for soot oxidation catalysts is the dynamic oxygen storage capacity (OSCdyn). In this work, OSCdyn was determined with the help of the TGA-DSC11600 from Mettler Toledo Corp., described above. To calculate the OSCdyn, two heating and cooling cycles were performed. First, m = 30 mg of the catalyst which was calcined at T = 800 °C was placed in a V = 70 µl corundum crucible and heated to T = 700 °C under a nitrogen flow of V ˙ = 25 mL·min−1 to release adsorbed oxygen, water, or other compounds. The system was stabilized for t = 10 min. Next, the system was cooled to T = 150 °C under a synthetic air flow of V ˙ = 25 mL·min−1 for oxygen uptake and stabilized for t = 10 min. This procedure was repeated and the weight loss of the second heating step was used for calculating the OSCdyn as mass difference of m O 2 150   ° C m N 2 700   ° C M O 2 · m c a t , where m O 2 150   ° C is the mass of catalyst at a temperature T = 150 °C after oxygen uptake,   m N 2 700   ° C is the mass of catalyst at T = 700 °C after oxygen release,   M O 2 is the molar mass of oxygen, and m c a t is the buoyancy-corrected mass of catalyst weighed in a V = 70 µl corundum crucible after desorption of compounds like water. OSCdyn is represented in µ m o l O 2 · g c a t 1 . For measuring the OSCdyn a heating/cooling rate of r = 5 °C·min−1 was used.

2.3. Catalyst Characterisation

Powder X-Ray diffraction (PXRD) was performed on a Bruker D8 diffractometer with Co fine focus X-ray source (Ni filter, λKα = 1.79021 Å, Θ-Θ geometry, VDS, Lynxeye detector) for samples of the normal synthesis route and on a STOE-Stadi P with Cu fine focus X-ray source (monochromator, λKα1 = 1.54056 Å, Θ-Θ geometry, MYTHEN2 R1K detector) for samples of the reverse strike synthesis route. The selected intensity data were collected in the 2θ range from 20 to 85°. Qualitative phase identification was achieved by diffraction pattern assignment according to ICDD data (International Centre for Diffraction Data). Following ICDD data are used: CeO2 ICDD#75-76, α-Bi2O3 ICDD#71-2274, and β-Bi2O3 ICDD#78-1793. Crystalline phases were identified through Rietveld refinement with the program TOPAS version 4.2 according to [19,20,21] adapted for the cerium-bismuth mixed oxides. Instrumental parameters for the fundamental parameter TOPAS refinement were determined by refinement of a LaB6 reference sample diffraction pattern.
Raman spectra were recorded with an inVia Raman Microscope from Renishaw Corp. A frequency-doubled Nd:Yag laser of the model RL473C from Renishaw with a wavelength of λ = 532 nm and a UVDD-CCD-Array detector with a grid consisting of 1800 lines/mm was used. The laser power depends on the measuring time. Usually laser powers of 0.5% of the maximal laser power P = 50 mW and a measuring time of t = 60 s were employed. The calcined samples were placed on a 96-well microplate with V-shaped bottom from Greiner Corp. The microscope was used in line focus. The sample were measured at three different points.
The determination of the specific surface according to Brunauer, Emmett and Teller was performed on a NOVA touch LX4 surface area and pore size analyzer from Quantachrome Corp. The evaluation was performed with the program TouchWinTM version 1.11 from Quantachrome Corp. The specific surface was measured by N2 adsorption-desorption isotherms at the temperature of liquid N2 (T = −196 °C). The samples were degassed under high vacuum for t = 4 h at T = 250 °C prior to the measurement. To determine the specific surface SBET, the seven-point BET multi-point method in a relative pressure range of 0.02–0.25 was used.
Energy-dispersive X-ray fluorescence (XRF) analyses to determine the mass fractions of Bi and Ce were performed on a FISCHERSCOPE® X-RAY XAN® instrument and the evaluations were conducted with the program WinFTM® from Fischer Corp. The excitation voltage was U = 30 keV and the measuring time t = 60 s. The catalysts which were calcined at T = 800 °C were placed on top of a cellulose foil. Three repeated measurements at different points of the sample were employed and averaged in composition. Calibration of the measurement data was done by a standardless fundamental parameter method supplied by Fischer Corp (for more details see [22,23,24,25]).

3. Results

3.1. Raman and PXRD Analysis

As depicted in the Raman spectra in Figure 1, the synthesized Ce1.0Bi0.0Ox (pure ceria) samples for both co-precipitation routes show one band at ν = 462 cm−1. Note that with “x” we denote in the following that the amount of oxygen is variable and depending on preparation conditions, whereas the other subscripts in the all formulae represent the molar composition. The band at ν = 462 cm−1 was attributed to the F2g mode of CeO2 [26]. This confirms that the pure CeO2 was synthesized in both cases without any defect structure. For the Ce0.0Bi1.0Ox (pure bismuth(III) oxide) all bands typical for α-Bi2O3 were observed for both synthesis methods. These bands were the following one: ν = 120 cm−1, 140 cm−1, 152 cm−1, 184 cm−1, 212 cm−1, 281 cm−1, 314 cm−1, 411 cm−1, 449 cm−1, and 532 cm−1 [26,27,28,29]. For both synthesis routes, with increasing Ce content the F2g band of CeO2 dominates the Raman spectra but with a higher F2g band FWHM (full width at half maximum, see supplementary information Table S1) compared to the intense F2g band of the Ce1.0Bi0.0Ox spectra.
Figure 1a shows the Raman spectra of the normal co-precipitation route. In the range from Ce0.1Bi0.9Ox to Ce0.4Bi0.6Ox bands for β- and α-Bi2O3 can be identified. Especially, the spectra of the compositions Ce0.1Bi0.9Ox, Ce0.2Bi0.8Ox, and Ce0.4Bi0.6Ox show the band profile of β-Bi2O3 with the following typical bands: ν = 125 cm−1, 232 cm−1, 314 cm−1, and 462 cm−1 [29]. Through position-resolved Raman micro-spectroscopy at different locations of the samples, we found that the samples were inhomogeneous with regards to phase composition. Areas with predominantly α-Bi2O3 and areas with β-Bi2O3 mixed with CeO2 were observed. The F2g band of CeO2 at ν = 464 cm−1 and the band of β-Bi2O3 at ν = 466 cm−1 are overlapping [26,29]. Therefore, it is hardly possible to distinguish between CeO2 and β-Bi2O3 in the range of high Bi content. With increasing Ce content, the typical bands for the bismuth oxides are disappearing and the F2g band of CeO2 dominates the spectra except the spectrum for Ce0.7Bi0.3Ox. This spectrum contains bands for β-Bi2O3 and CeO2. The spectra for Ce0.8Bi0.2Ox and Ce0.9Bi0.1Ox show two broad bands at ν = 510 cm−1 and 570 cm−1. According to Schilling et al. these bands can be assigned to oxygen vacancy formation C e 3 + O 7 V ö and C e 4 + O 7 V ö [26]. Figure 1b shows the Raman spectra of the reverse strike co-precipitation samples. The oxides RS-Ce0.1Bi0.9Ox, RS-Ce0.2Bi0.8Ox, and RS-Ce0.3Bi0.7Ox show clearly the typical bands originating from α-Bi2O3. The two oxides RS-Ce0.4Bi0.6Ox and RS-Ce0.5Bi0.5Ox have the typical F2g band of CeO2 and the bands of β-Bi2O3. From RS-Ce0.6Bi0.4Ox to RS-Ce0.9Bi0.1Ox only the F2g band and the two bands at ν = 510 cm−1 and 570 cm−1 resulting from the defects of the fluorite-type structure are to be recognized.
Figure 2 presents the PXRD pattern of the synthesized samples for both synthesis routes. In general, the PXRD data confirm the results from Raman analyses. For both synthesis routes, the PXRD pattern shows that the oxides Ce0.0Bi1.0Ox adopt the monoclinic α-Bi2O3 crystal structure with space group P21/c. The synthesized Ce1.0Bi0.0Ox crystallises cubic in a fluorite-type structure with the space group Fm 3 ¯ m . The PXRD pattern of the mixed oxides with increasing Bi content shows that, additionally to the reflections of ceria lattice structure, diffraction signals of the monoclinic α-Bi2O3 structure and of the tetragonal β-Bi2O3 with space group P 4 ¯ 21c are appearing. For the normal synthesis route, only for the oxides Ce0.8Bi0.2Ox and Ce0.9Bi0.1Ox a single phase of Bi3+-doped CeO2 was observed. The lattice parameter a and the mass fraction of Bi3+ in the CeO2 lattice have been determined by Rietveld refinement. Doping CeO2 with Bi3+ increases the lattice parameter a from 5.4110(1) Å for CeO2 to 5.4174(1) Å for Ce0.9Bi0.1Ox and further on 5.4235(1) Å for Ce0.8Bi0.2Ox. In contrast, with the reverse strike synthesis route in the range of pure ceria to RS-Ce0.6Bi0.4Ox a single Bi3+-doped CeO2 crystalline phase was observed, and the range of existence was thus broader than for the normal precipitation process. Figure 3 shows the lattice parameter a from Rietveld refinement of the Bi3+-doped CeO2 phases, and in Tables S2 and S3 all refined parameters are specified in detail. The insertion of Bi3+ ions into the CeO2 lattice increased the lattice parameter. For RS-Ce0.6Bi0.4Ox the lattice parameter a was 5.4273(3) Å. The lattice parameter a for the synthesized pure RS-CeO2 was 5.4071(1) Å.

3.2. XRF and BET Results

To verify the nominal Bi and Ce metal mass percentages of the oxides, i.e. without inclusion of oxygen content, XRF measurements were performed. Table 2 shows the acquired mass percentages. Note that there were no larger deviations between nominal and the expected mass percentages of the samples. Furthermore, the mass percentages of Bi and Ce differed not more than 2% between the two synthesis routes.
Table 3 shows the total surface area SBET of the synthesized samples determined by N2 physisorption experiments. The SBET was determined by the multi-point BET method. For the normal precipitation route, the pure Bi2O3 had with SBET = 0.4 m2∙g−1 the lowest specific surface. With increasing Ce content, the SBET increases as well. The highest surface area was observed for the pure CeO2 with SBET = 26.6 m2·g−1. For the compositions Ce0.6Bi0.4Ox with SBET = 18.7 m2·g−1 and Ce0.8Bi0.2Ox with SBET = 13.1 m2·g−1 also relatively high surface areas within this test series were observed. Surprisingly, the composition Ce0.7Bi0.3Ox has only a SBET of 6.5 m2·g−1. The differences in the observed Raman and PXRD data of this composition compared to others also reflected this behavior.
For the reverse strike co-precipitation route, as Ce content increased the SBET also increased. The maximum of the surface area was reached at a Ce content of 80 mol% with SBET = 21.3 m2·g−1. A further increase of the Ce content led to a decrease in SBET. For pure CeO2 prepared by the reverse-strike precipitation method only SBET = 6.2 m2·g−1 was observed.

3.3. OSCdyn and TGA Results

The dynamic oxygen storage capacities OSCdyn were measured via thermogravimetric analyses in defined gas atmospheres and are shown in Table 4. First, each sample was heated with a specific heating rate under nitrogen for desorption of compounds like water, and to reduce the sample to substoichiometric oxides. Second, the sample was cooled under synthetic air. At this period, the sample was reoxidized. These two steps were repeated. The second reoxidation step was used for calculating the OSCdyn as difference in masses under the two atmospheres at specific temperatures (see chapter 2.2. for details). The OSCdyn is an important factor for the Diesel soot oxidation because the total available surface lattice oxygen is related to the OSC [31]. For the normal synthesis route, Ce0.8Bi0.2Ox had the highest OSCdyn with 34.0 µmolO2∙gKat−1 and Ce0.0Bi1.0Ox revealed the lowest OSCdyn with 1.6 µmolO2∙gcat−1. For this synthesis route, there was no obvious correlation between the composition of oxide and the determined OSCdyn. In contrast to that, the OSCdyn for the reverse strike synthesis route was lower. Nevertheless, a correlation between the composition of oxide and the determined OSCdyn seems to exist. Starting from pure bismuth oxide the OSCdyn increased with an increase in Ce content. The composition RS-Ce0.8Bi0.2Ox with a value of OSCdyn = 26.4 µmolO2∙gcat−1 seems to represent the maximum in this synthesis series. After that, the OSCdyn decreased to a value of 1.9 µmolO2∙gcat−1 for RS-Ce1.0Bi0.0Ox. Furthermore, it is to be noted that the OSCdyn in the intermediate range between the pure cerium oxides and bismuth oxides were different for both synthesis routes. The OSCdyn for the reverse strike co-precipitation of pure bismuth oxide was eight times higher than for the normal synthesis route and the reverse strike co-precipitation of pure cerium oxide was 13 times smaller than the normal synthesis route. The SBET reflected the same trend. There, the SBET for the reverse strike route was five times higher than the normal synthesis route. This shows clearly that the OSCdyn really represents a surface parameter, in this case the surface oxygen lattice ions to be utilized for oxidation processes.
In thermogravimetric (TGA) measurement data, the T50 value is used to characterise the activity for the Diesel soot oxidation. The T50 value describes the temperature where 50% of the model soot is oxidized in a model gas mixture of defined flow rate and heating rate. This value is commonly used in the literature. Table 5 presents the determined T50 values. Only the T50 values of tight soot-catalyst contact conditions were presented, but also other contact modes were prepared and screened.
The lowest T50 value was reached for the compositions Ce0.8Bi0.2Ox and RS-Ce0.8Bi0.2Ox. Like for the SBET and OSCdyn, there was a less distinct correlation between T50 and the composition of the oxide for the normal synthesis route indicating nuisance factors in the syntheses. In contrast to the normal synthesis route, there was a clear dependency in T50 values as a function of oxide composition for the reverse strike synthesis route. Starting from pure bismuth oxide the T50 decreased with higher Ce content to a minimum for composition RS-Ce0.8Bi0.2Ox with a T50 value of 444 °C. After that, the T50 value increased again to 586 °C for RS-Ce1.0Bi0.0Ox. In general, and for both synthesis routes, the T50 values of pure CeO2 and Bi2O3 were higher than that of the mixed oxides.

4. Discussion

4.1. Normal Synthesis Route

PXRD data refinements reveal that in compositions with high Bi content two crystalline Bi2O3 phases can be detected, i.e., monoclinic α-Bi2O3 and tetragonal β-Bi2O3. Additionally reflections representing about 5m% bismutite Bi2(CO3)O2 can also be detected in PXRD determined by Rietveld refinement. Confirmation of these results is achieved by Raman spectroscopy. With increasing Ce content up to the composition Ce0.7Bi0.3Ox, a β-Bi2O3 phase is present. One reason for this observation could be the initial stages of precipitation of the oxides. From a visual inspection of the precipitation process, we find the beginning of precipitation for bismuth oxide at Ph < 1 and for cerium oxide near to pH = 4. After calcination, the α-Bi2O3 polymorph should be formed as this is the thermodynamic stable form [32]. Additionally, it is known that hydrated bismuth oxides oligomerise in aqueous acidic media [18]. In our case, also the β-Bi2O3 form is present. The oligomerisation of the hydrated bismuth oxides at low pH values and the incorporation of cerium oxide into the crystal lattice, may both serve as an explanation for the observation of the tetragonal β-Bi2O3 form. Another reason is that the formation of β-Bi2O3 may result from decomposition of bismutite [33]. With increasing Ce content, the F2g mode in the Raman spectra gets the dominant band. Above a Ce content of 50 mol% there are two more bands to be assigned to CeO2. The band at ν ~ 510 cm−1 is caused by the formation of Ce3+ ions in combination with oxygen vacancies. The band at ν ~ 570 cm−1 originates from oxygen vacancies in combination with Ce4+ ions (see above and ref. [19]). From Figure 1 it can be seen that these two defect bands are most noticeable for the compositions Ce0.8Bi0.2Ox and Ce0.9Bi0.1Ox and that in the samples with these compositions only the Bi3+-doped CeO2 phase is verifiable. Interestingly, these two compositions have, for this synthesis route, the lowest T50 values and highest OSCdyn values compared to the other compositions. This is illustrated in Figure 4.

4.2. Reverse Strike Synthesis Route

To avoid the oligomerisation of the hydrated bismuth oxides in aqueous acidic media and the sequential precipitation of oxides with different compositions, i.e., cerium and bismuth oxides separately, the reverse strike co-precipitation method is applied. During this synthesis route, the premixed acidic precursor solutions are added to the basic precipitation agent solution. The precipitation is stopped at pH = 6. Therefore, an acidic pH range is avoided. According to the PXRD data, see Figure 2b, below a molar fraction of 60 mol% Ce only a Bi3+-doped CeO2 phase is observed. These results are in accordance to the results of Hund et al. [15]. Compared to the normal synthesis route, with the reverse strike route the mixed oxide phase extends to even lower Ce fractional amounts. These results are supported by Raman analyses (see Figure 1b). Additionally, only bismutite is present in the Raman spectra of Ce0.2Bi0.8Ox to Ce0.8Bi0.2Ox as an additional phase in a minor amount. Furthermore, the two defect bands at ν = 510 cm−1 and ν = 570 cm−1 are very prominent within the composition range of Ce0.6Bi0.4Ox to Ce0.8Bi0.2Ox. At higher Ce content the intensity of the two defect bands decreases again, and are not to be detected in the Raman spectrum of Ce1.0Bi0.0Ox. In Section 3.1. we concluded that the formation of oxygen sublattice defects in ceria of fluorite-structure type may influence the activity of the catalyst for the Diesel soot oxidation. The more pronounced the defect structure is, the higher the activity will be for Diesel soot oxidation. In Figure 5, the correlation between T50, OSCdyn and SBET is presented for the samples of this synthesis route. For this synthesis series, a clear correlation between activity and characterization data can be seen. Starting from pure ceria with increasing Bi content, T50 values decrease revealing that the soot oxidation activity increases. At the same time, both OSCdyn and SBET increase. This trend reaches its maximum – low T50 value and high OSCdyn and SBET – at 80 mol% Ce content. This is in accordance with the described effect of the Raman spectra with the two defect bands at ν = 510 cm−1 and ν = 570 cm−1. At still lower Ce content, the T50 values increase, and the OSCdyn and SBET values decrease again. In the same sequence, the intensities of the two defect bands in Raman spectra decrease. The two defect bands are not detected in the Raman spectrum of pure ceria.

5. Conclusions

In this paper, two automated routes of co-precipitation methods to synthesize a complete series of Bi-Ce mixed oxides Ce1−xBixO2−x/2 (0 ≤ x ≤ 1) for catalytic oxidation of Diesel soot are described: The normal synthesis and the reverse strike synthesis. The synthesized oxides are characterized by PXRD, Raman spectroscopy, and nitrogen adsorption to measure the specific surface. Furthermore, the dynamic oxygen storage capacities OSCdyn are determined by TGA methods. The catalytic activities are also measured by TGA under continuous gas flow conditions and specified by T50 values. Raman spectroscopy reveals that, with an increasing Ce content, an oxygen defect structure of the fluorite-type ceria lattice as a result of Bi3+ insertion into the cation sublattice occurs (in Kröger-Vink notation, see equation 1):
B i 2 O 3 ( C e O 2 ) 2 B i C e + V ö + 3 O O x
At least two bands resulting from oxygen defects of the ceria lattice are observed in the Raman spectra. Additionally, a relation between OSCdyn, SBET and T50 values is found indicating that catalytic activity for Diesel soot oxidation and defect structure are highly correlated. This relation is very striking for the mixed oxides, which are synthesized through the reverse strike method, but less distinct for samples prepared by the normal precipitation route. Here, the maximal catalytic activity with the maximal OSCdyn and SBET is reached at the maximal intensity of the two defect bands in the Raman spectra at the optimal composition of Ce0.8Bi0.2O1.9. Compared to the pure reverse strike co-precipitated CeO2 RS-Ce1.0Bi0.0Ox, the T50 decreases from 586 °C to 444 °C. This is an improvement of ΔT50 ≈ 140 °C. Simultaneously, the OSCdyn increases from 1.9 µmolO2∙gcat−1 for the RS-Ce1.0Bi0.0Ox to 26.4 µmolO2∙gcat−1 for the RS-Ce0.8Bi0.2Ox. This is an increase of ΔOSCdyn = 24.5 µmolO2∙gcat−1. Additionally, the SBET also increases from 6.2 m2∙g−1 for the RS-Ce1.0Bi0.0Ox to 21.3 m2∙g−1 for the RS-Ce0.8Bi0.2Ox. This is an increase of ΔSBET = 15,1 m2∙g−1. PXRD data reveal that with this composition we are still in the phase field of a cubic defect structure of the fluorite-type structure and that the field of stability for this mixed oxide phase is limited to the compositional range 0 ≤ x ≤ 0.4. At higher x also pure bismuth oxide phases are formed. Due to ionic radii differences (r(Bi3+) = 1.03 Å, r(Ce3+) = 1.01 Å, r(Ce4+) = 0.87 Å, all for CN = 6 [34]) the incorporation of Bi3+ into the ceria lattice Ce4+(O2−)2 “widens” the crystal lattice as can be seen from the increase in the lattice parameter a. Additionally, due to the lone pair configuration of Bi3+, a local distortion of the cation coordination environment by oxygen anions occurs. Both factors facilitate the incorporation of larger Ce3+ ions into the cationic Ce4+ sublattice, thus, increasing the oxygen storage capacity, as for each O2− removed for charge neutrality two Ce4+ ions have to be reduced to Ce3+ (so-called Schottky defect set in square brackets, see equation 2):
2 C e O 2 ( C e O 2 ) [ 2 C e C e + V ö ] + 3 O O x + 1 2 O 2
As on the other side, too much Bi3+ is destabilising the lattice, as seen from the crystallisation of additional bismuth oxide phases. There is an optimum of oxygen storage capacity and thus minimum of T50 value, in this case at x = 0.2.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/1996-1944/13/6/1369/s1. Table S1: FWHM (full width at half maximum) of the F2g band from the CeO2 lattice of Raman spectra. Given is the average of three repeated measurements and the standard deviation. FWHM is determined with the software Origin Pro 2019b; Table S2: Results of Rietveld-Refinement of the normal precipitation route. Samples are calcined by T = 800 °C; Table S3: Results of Rietveld-Refinement of the reverse strike precipitation route. Samples are calcined by T = 800 °C.

Author Contributions

S.C.H. build the experimental setups, conducted experiments, analyzed data and wrote this paper. K.S. supervised the work and reviewed the article. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

We thank the professorship Coordination Chemistry (Prof. M. Mehring, M.Sc. I. Köwitsch) at TUC for measuring PXRD pattern and the professorship of Material and Surface Engineering (Prof. T. Lampke, M.Sc. Th. Mehner, M.Sc. Pügner, Dipl.-Ing. E. Benedix) at TUC for the possibility to measure XRF as well as measuring PXRD pattern.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Janssen, N.A.; Gerlofs-Nijland, M.E.; Lanki, T.; Salonen, R.O.; Cassee, F.; Hoek, G.; Fischer, P.; Brunekreef, B.; Krzyzanowski, M. Health Effects of Black Carbon; World Health Organization: Geneva, Switzerland, 2012; pp. 1–86. [Google Scholar]
  2. Prasad, R.; Bella, V.R. A review on diesel soot emission, its effect and control. Bull. Chem. React. Eng. Catal. 2010, 5, 69–86. [Google Scholar] [CrossRef]
  3. Channell, M.M.; Paffett, M.L.; Devlin, R.B.; Madden, M.C.; Campen, M.J. Circulating factors induce coronary endothelial cell activation following exposure to inhaled diesel exhaust and nitrogen dioxide in humans: Evidence from a novel translational in vitro model. Toxicol. Sci. 2012, 127, 179–186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Schwarze, P.E.; Totlandsdal, A.I.; Låg, M.; Refsnes, M.; Holme, J.A.; Øvrevik, J. Inflammation-related effects of diesel engine exhaust particles: Studies on lung cells in vitro. BioMed Res. Int. 2013, 2013, 685142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Johansen, K. Multi-catalytic soot filtration in automotive and marine applications. Catal. Today 2015, 258, 2–10. [Google Scholar] [CrossRef]
  6. Zhang, Z.H.; Balasubramanian, R. Effects of Cerium Oxide and Ferrocene Nanoparticles Addition As Fuel-Borne Catalysts on Diesel Engine Particulate Emissions: Environmental and Health Implications. Environ. Sci. Technol. 2017, 51, 4248–4258. [Google Scholar] [CrossRef]
  7. Zhang, B.; Jiaqiang, E.; Gong, J.; Yuan, W.; Zuo, W.; Li, Y.; Fu, J. Multidisciplinary design optimization of the diesel particulate filter in the composite regeneration process. Appl. Energy 2016, 181, 14–28. [Google Scholar] [CrossRef]
  8. Adler, J. Ceramic diesel particulate filters. Int. J. Appl. Ceram. Technol. 2005, 2, 429–439. [Google Scholar] [CrossRef]
  9. Krishna, K.; Bueno-López, A.; Makkee, M.; Moulijn, J.A. Potential rare earth modified CeO2 catalysts for soot oxidation. I. Characterisation and catalytic activity with O2. Appl. Catal. B Environ. 2007, 75, 189–200. [Google Scholar] [CrossRef] [Green Version]
  10. Sardar, K.; Playford, H.Y.; Darton, R.J.; Barney, E.R.; Hannon, A.C.; Tompsett, D.; Fisher, J.; Kashtiban, R.J.; Sloan, J.; Ramos, S.; et al. Nanocrystalline cerium-bismuth oxides: Synthesis, structural characterization, and redox properties. Chem. Mater. 2010, 22, 6191–6201. [Google Scholar] [CrossRef]
  11. Aneggi, E.; Boaro, M.; De Leitenburg, C.; Dolcetti, G.; Trovarelli, A. Insights into the redox properties of ceria-based oxides and their implications in catalysis. J. Alloys Compd. 2006, 408–412, 1096–1102. [Google Scholar] [CrossRef]
  12. Wolf, V. Neuartige Katalysatoren für die Oxidation von Dieselruß in der Abgasnachbehandlung von Fahrzeugen. Ph.D. Thesis, Universität des Saarlandes, Saarbrücken, Germany, 2015. [Google Scholar]
  13. Dikmen, S.; Shuk, P.; Greenblatt, M. Hydrothermal synthesis and properties of Ce1-xBixO2-delta solid solutions. Solid State Ionics 1998, 112, 299–307. [Google Scholar] [CrossRef]
  14. Zhao, H.; Feng, S. Hydrothermal Synthesis and Oxygen Ionic Conductivity of Codoped Nanocrystalline Ce1-xMxBi0.4O2.6-x, M = Ca, Sr, and Ba. Chem. Mater. 1999, 11, 958–964. [Google Scholar] [CrossRef]
  15. Hund, F. Fluoritmischphasen der Dioxide von Uran, Thorium, Cer und Zirkonium mit Wismutoxide. Zeitschrift für Anorg. und Allg. Chemie 1964, 333, 248–255. [Google Scholar] [CrossRef]
  16. Frolova, Y.V.; Kochubey, D.I.; Kriventsov, V.V.; Moroz, E.M.; Neofitides, S.; Sadykov, V.A.; Zyuzin, D.A. The influence of bismuth addition on the local structure of CeO2. Nucl. Instrum. Methods Phys. Res. Sect. A 2005, 543, 127–130. [Google Scholar] [CrossRef]
  17. Lee, K.T.; Lidie, A.A.; Jeon, S.Y.; Hitz, G.T.; Song, S.J.; Wachsman, E.D. Highly functional nano-scale stabilized bismuth oxides via reverse strike co-precipitation for solid oxide fuel cells. J. Mater. Chem. A 2013, 1, 6199–6207. [Google Scholar] [CrossRef]
  18. Hollemann, N.W.A.F. Lehrbuch der Anorganischen Chemie; 102. Edtion; Walter de Gruyter: Berlin, Germany; New York, NY, USA, 2007; pp. 822–860. [Google Scholar]
  19. Cheary, R.W.; Coelho, A.A. A fundamental parameters approach to X-ray line-profile fitting. J. Appl. Crystallogr. 1992, 25, 109–121. [Google Scholar] [CrossRef]
  20. Cheary, R.W.; Coelho, A.A.; Cline, J.P. Fundamental parameters line profile fitting in laboratory diffractometers. J. Res. Natl. Inst. Stand. Technol. 2004, 109, 1–25. [Google Scholar] [CrossRef]
  21. Kern, A.; Coelho, A.A.; Cheary, R.W. Convolution Based Profile Fitting. In Diffraction Analysis of the Microstructure of Materials; Mittemeijer, E.J., Scardi, P., Eds.; Springer: Berlin/Heidelberg, Germany, 2004; pp. 17–50. [Google Scholar]
  22. Lagultton, D.; Parrish, W. Simultaneous Determination of Composition and Mass Thickness of Thin Films by Quantitative X-ray Fluorescence Analysis. Anal. Chem. 1977, 49, 1152–1156. [Google Scholar] [CrossRef]
  23. Criss, J.W.; Birks, L.S. Calculation Methods for Fluorescent X-Ray Spectrometry: Empirical Coefficients vs. Fundamental Parameters. Anal. Chem. 1968, 40, 1080–1086. [Google Scholar] [CrossRef]
  24. Shiraiwa, T.; Fujino, N. Theoretical Calculation of Fluorescent X-Ray Intensities in Fluorescent X-Ray Spectrochemical Analysis. Jpn. J. Appl. Phys. 1966, 5, 886–899. [Google Scholar] [CrossRef]
  25. Ohno, K.; Fujiwara, J.; Morimoto, I. X-ray fluorescence analysis without standards of small particles extracted from super-alloys. X-Ray Spectrom. 1980, 9, 138–142. [Google Scholar] [CrossRef]
  26. Schilling, C.; Hofmann, A.; Hess, C.; Ganduglia-Pirovano, M.V. Raman Spectra of Polycrystalline CeO2: A Density Functional Theory Study. J. Phys. Chem. C 2017, 121, 20834–20849. [Google Scholar] [CrossRef]
  27. Taylor, P.; Sunder, S.; Lopata, V.J. Structure, spectra, and stability of solid bismuth carbonates. Can. J. Chem. 1984, 62, 2863–2873. [Google Scholar] [CrossRef]
  28. Selvamani, T.; Anandan, S.; Granone, L.; Bahnemann, D.W.; Ashokkumar, M. Phase-controlled synthesis of bismuth oxide polymorphs for photocatalytic applications. Mater. Chem. Front. 2018, 2, 1664–1673. [Google Scholar] [CrossRef]
  29. Salazar-Pérez, A.J.; Camacho-López, M.A.; Morales-Luckie, R.A.; Sánchez-Mendieta, V. Structural evolution of Bi2O3 prepared by thermal oxidation of bismuth nano-particles. Superf. y vacío 2018, 18, 4–8. [Google Scholar]
  30. Diffrac. Suite TOPAS; Bruker AXS: Karlsruhe, Germany, 2015. [Google Scholar]
  31. Li, P.; Chen, X.; Li, Y.; Schwank, J.W. A review on oxygen storage capacity of CeO2-based materials: Influence factors, measurement techniques, and applications in reactions related to catalytic automotive emissions control. Catal. Today 2019, 327, 90–115. [Google Scholar] [CrossRef]
  32. Mehring, M. From molecules to bismuth oxide-based materials: Potential homo- and heterometallic precursors and model compounds. Coord. Chem. Rev. 2007, 251, 974–1006. [Google Scholar] [CrossRef]
  33. Iyyapushpam, S.; Nishanthi, S.T.; Padiyan, D.P. Synthesis of β-Bi2O3 towards the application of photocatalytic degradation of methyl orange and its instability. J. Phys. Chem. Solids 2015, 81, 74–78. [Google Scholar] [CrossRef]
  34. Shannon, R.D.; Prewitt, C.T. Effective ionic radii in oxides and fluorides. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1969, 25, 925–946. [Google Scholar] [CrossRef]
Figure 1. Raman spectra of the synthesized Bi-Ce-mixed oxides with varying Bi-Ce-molar ratios in 10 mol% steps. Each sample is measured three times. Shown are representative spectra of these measurements. All spectra are normalized for comparison. Oxides have been calcined at T = 800 °C. Laser wavelength is λ = 532 nm. (a) Normal synthesis route; (b) Reverse strike synthesis route.
Figure 1. Raman spectra of the synthesized Bi-Ce-mixed oxides with varying Bi-Ce-molar ratios in 10 mol% steps. Each sample is measured three times. Shown are representative spectra of these measurements. All spectra are normalized for comparison. Oxides have been calcined at T = 800 °C. Laser wavelength is λ = 532 nm. (a) Normal synthesis route; (b) Reverse strike synthesis route.
Materials 13 01369 g001
Figure 2. Powder X-ray diffraction (PXRD) pattern of the synthesized Bi-Ce mixed oxides with varying Bi-Ce molar ratios in 10 mol% steps. Oxides calcined at T = 800 °C. (a) Normal synthesis route. Measured with Co-Kα radiation, λ = 178.92 nm; (b) Reverse strike synthesis route. Measured with Cu-Kα radiation, λ = 154.06 nm. Reference files for CeO2: ICDD#75-76; α-Bi2O3: ICDD#71-2274; β-Bi2O3: ICDD#78-1793.
Figure 2. Powder X-ray diffraction (PXRD) pattern of the synthesized Bi-Ce mixed oxides with varying Bi-Ce molar ratios in 10 mol% steps. Oxides calcined at T = 800 °C. (a) Normal synthesis route. Measured with Co-Kα radiation, λ = 178.92 nm; (b) Reverse strike synthesis route. Measured with Cu-Kα radiation, λ = 154.06 nm. Reference files for CeO2: ICDD#75-76; α-Bi2O3: ICDD#71-2274; β-Bi2O3: ICDD#78-1793.
Materials 13 01369 g002
Figure 3. Lattice parameter a with standard deviation (STD) of a of cubic fluorite-type Bi3+-doped CeO2 phase in dependence of the nominal molar Ce content determined by Rietveld refinement with the program TOPAS [30].
Figure 3. Lattice parameter a with standard deviation (STD) of a of cubic fluorite-type Bi3+-doped CeO2 phase in dependence of the nominal molar Ce content determined by Rietveld refinement with the program TOPAS [30].
Materials 13 01369 g003
Figure 4. T50 (tight contact), SBET and OSCdyn values in dependence of the nominal molar Ce content for samples prepared by normal precipitation. Shown are duplicate measurements together with their standard deviation.
Figure 4. T50 (tight contact), SBET and OSCdyn values in dependence of the nominal molar Ce content for samples prepared by normal precipitation. Shown are duplicate measurements together with their standard deviation.
Materials 13 01369 g004
Figure 5. T50 (tight contact), SBET and OSCdyn values in dependence of the nominal molar Ce content for samples prepared by reverse strike precipitation. Shown are duplicate measurements with their standard deviation.
Figure 5. T50 (tight contact), SBET and OSCdyn values in dependence of the nominal molar Ce content for samples prepared by reverse strike precipitation. Shown are duplicate measurements with their standard deviation.
Materials 13 01369 g005
Table 1. Dispensing and aspirating speeds of the pipetting steps. The 4-needle head (4NH) of the synthesis robot Chemspeed Accelerator SLT106 is used for liquid dosing.
Table 1. Dispensing and aspirating speeds of the pipetting steps. The 4-needle head (4NH) of the synthesis robot Chemspeed Accelerator SLT106 is used for liquid dosing.
SolutionDispensing Speed [mL·min−1]Aspirating Speed [mL·min−1]
0.5 M Ce(NO3)3·6H2O4040
0.5 M Bi(NO3)3·5H2O4040
premixed nitrate solution4040
1 M (NH4)2CO34080
Table 2. Mass percentages m% of the synthesized mixed oxides determined by X-ray fluorescence (XRF) analysis. Shown are the averaged values of three repeated measurements of one sample at different areas with standard deviations.
Table 2. Mass percentages m% of the synthesized mixed oxides determined by X-ray fluorescence (XRF) analysis. Shown are the averaged values of three repeated measurements of one sample at different areas with standard deviations.
Composition of OxideNominal Mass PercentagesMeasured Mass Percentage
Formulae Normal SynthesisReverse Strike Synthesis
m%Cem%Bim%Cem%Bim%Cem%Bi
Ce0.0Bi1.0Ox0.00100.000.00 ± 0.00100.0 ± 0.000,00 ± 0.00100.0 ± 0.00
Ce0.1Bi0.9Ox7.5992.416.89 ± 0.1792.85 ± 0.157.52 ± 0.1592.48 ± 0.15
Ce0.2Bi0.8Ox15.5984.4114.36 ± 0.1685.26 ± 0.1615.15 ± 0.4484.85 ± 0.44
Ce0.3Bi0.7Ox24.0575.9522.86 ± 0.2676.51 ± 0.3023.52 ± 0.0476.48 ± 0.04
Ce0.4Bi0.6Ox33.0067.0031.71 ± 0.2167.64 ± 0.2630.07 ± 0.5569.31 ± 0.45
Ce0.5Bi0.5Ox42.4957.5145.52 ± 1.3354.18 ± 1.6444.09 ± 0.1955.91 ± 0.19
Ce0.6Bi0.4Ox52.5647.4457.13 ± 0.4442.27 ± 0.1955.58 ± 0.2044.42 ± 0.20
Ce0.7Bi0.3Ox63.2936.7160.64 ± 0.7438.50 ± 0.7958.25 ± 0.3740.71 ± 0.36
Ce0.8Bi0.2Ox74.7225.2875.42 ± 0.7523.85 ± 0.2778.66 ± 0.1021.34 ± 0.10
Ce0.9Bi0.1Ox86.9313.0785.64 ± 0.4013.96 ± 0.1689.17 ± 0.0910.83 ± 0.09
Ce1.0Bi0.0Ox100.000.00100.0±0.000.00 ± 0.00100.0 ± 0.000.00 ± 0.00
Table 3. Calculated specific surface areas according to Brunauer, Emmett, and Teller (SBET) of the synthesized mixed Bi-Ce oxides determined by nitrogen adsorption measurements. Oxides are calcined at T = 800 °C.
Table 3. Calculated specific surface areas according to Brunauer, Emmett, and Teller (SBET) of the synthesized mixed Bi-Ce oxides determined by nitrogen adsorption measurements. Oxides are calcined at T = 800 °C.
Composition of OxideSBET,normal synthesis [m2·g−1]SBET,reverse strike synthesis [m2·g−1]
Ce0.0Bi1.0Ox0.42.0
Ce0.1Bi0.9Ox1.51.0
Ce0.2Bi0.8Ox2.22.1
Ce0.3Bi0.7Ox5.41.0
Ce0.4Bi0.6Ox6.41.5
Ce0.5Bi0.5Ox7.82.2
Ce0.6Bi0.4Ox18.76.7
Ce0.7Bi0.3Ox6.515.1
Ce0.8Bi0.2Ox13.121.3
Ce0.9Bi0.1Ox20.211.6
Ce1.0Bi0.0Ox26.66.2
Table 4. Calculated dynamic oxygen storage capacity (OSCdyn) of the synthesized mixed oxides determined by thermogravimetric analysis (TGA). Measurement temperature in the range of T = 150–700 °C with heating rate of r = 5 °C∙min−1 in synthetic air and nitrogen. The average of double determination with the standard deviation is shown.
Table 4. Calculated dynamic oxygen storage capacity (OSCdyn) of the synthesized mixed oxides determined by thermogravimetric analysis (TGA). Measurement temperature in the range of T = 150–700 °C with heating rate of r = 5 °C∙min−1 in synthetic air and nitrogen. The average of double determination with the standard deviation is shown.
Composition of OxideOSCdyn,normal synthesis [µmolO2·gcat−1]OSCdyn,reverse strike synthesis [µmolO2·gcat−1]
Ce0.0Bi1.0Ox1.6 ± 1.513.4 ± 1.0
Ce0.1Bi0.9Ox12.6 ± 7.91.9 ± 0.0
Ce0.2Bi0.8Ox10.7 ± 0.91.4 ± 0.0
Ce0.3Bi0.7Ox11.3 ± 0.32.9 ± 2.4
Ce0.4Bi0.6Ox13.9 ± 1.55.4 ± 1.1
Ce0.5Bi0.5Ox24.7 ± 4.69.6 ± 0.2
Ce0.6Bi0.4Ox20.3 ± 1.818.5 ± 0.9
Ce0.7Bi0.3Ox13.3 ± 3.526.0 ± 1.6
Ce0.8Bi0.2Ox34.0 ± 0.426.4 ± 5.0
Ce0.9Bi0.1Ox23.4 ± 5.717.3 ± 2.3
Ce1.0Bi0.0Ox24.7 ± 2.01.9 ± 0.7
Table 5. T50 values for Diesel soot oxidation in tight contact mode of the synthesized mixed oxides determined by TGA. Measurement temperature in the range of T = 25–700 °C with heating rate of r = 5 °C∙min−1 in synthetic air. The averages of double determination with standard deviations are given.
Table 5. T50 values for Diesel soot oxidation in tight contact mode of the synthesized mixed oxides determined by TGA. Measurement temperature in the range of T = 25–700 °C with heating rate of r = 5 °C∙min−1 in synthetic air. The averages of double determination with standard deviations are given.
Composition of OxideT50.normal synthesis [°C]T50.reverse strike synthesis [°C]
Ce0.0Bi1.0Ox537.7 ± 0.5549.1 ± 1.1
Ce0.1Bi0.9Ox518.4 ± 2.1549.1 ± 1.8
Ce0.2Bi0.8Ox507.6 ± 1.4554.8 ± 1.3
Ce0.3Bi0.7Ox483.9 ± 0.5558.0 ± 1.0
Ce0.4Bi0.6Ox519.8 ± 1.4555.4 ± 2.3
Ce0.5Bi0.5Ox494.0 ± 0.5543.9 ± 1.0
Ce0.6Bi0.4Ox480.2 ± 0.2534.9 ± 2.0
Ce0.7Bi0.3Ox489.3 ± 1.5498.7 ± 0.2
Ce0.8Bi0.2Ox450.3 ± 0.8443.7 ± 2.7
Ce0.9Bi0.1Ox457.5 ± 0.1525.9 ± 0.8
Ce1.0Bi0.0Ox509.2 ± 1.6585.8 ± 1.7

Share and Cite

MDPI and ACS Style

Hebert, S.C.; Stöwe, K. Synthesis and Characterization of Bismuth-Cerium Oxides for the Catalytic Oxidation of Diesel Soot. Materials 2020, 13, 1369. https://0-doi-org.brum.beds.ac.uk/10.3390/ma13061369

AMA Style

Hebert SC, Stöwe K. Synthesis and Characterization of Bismuth-Cerium Oxides for the Catalytic Oxidation of Diesel Soot. Materials. 2020; 13(6):1369. https://0-doi-org.brum.beds.ac.uk/10.3390/ma13061369

Chicago/Turabian Style

Hebert, Sabrina C., and Klaus Stöwe. 2020. "Synthesis and Characterization of Bismuth-Cerium Oxides for the Catalytic Oxidation of Diesel Soot" Materials 13, no. 6: 1369. https://0-doi-org.brum.beds.ac.uk/10.3390/ma13061369

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop