GET THE APP

Effect of Ag-doping of Nanosized FeAlO System on its Structural,
Journal of Thermodynamics & Catalysis

Journal of Thermodynamics & Catalysis
Open Access

ISSN: 2157-7544

+44 1300 500008

Research Article - (2011) Volume 2, Issue 2

Effect of Ag-doping of Nanosized FeAlO System on its Structural, Surface and Catalytic Properties

Laila I. Ali*, Sahar A. El-Molla, Nabil H. Amin, Anwer A. Ebrahim and Hala R. Mahmoud
Chemistry Department, Faculty of Education, Ain Shams University, Roxy 11757, Cairo, Egypt, E-mail: saharel@als.edu.ac.com
*Corresponding Author: Laila I. Ali, Chemistry Department, Faculty of Education, Ain Shams University, Roxy 11757, Cairo, Egypt, Tel: (02012) 23184682, Fax: (02) 22581243 Email:

Abstract

The effects of Ag2O-doping on the physicochemical, surface and catalytic properties FeAlO system with various extent of Fe2O3 loading have been investigated. The dopant concentration was changed between 1.5 and 4.0 mol % Ag2O. Pure and variously doped solids were subjected to heat treatment at 400-800?C. The techniques employed for characterization of catalysts were TG/DTG, XRD, N2-adsorption at ?196?C and the catalytic decomposition of H2O2 at 25-40?C. The obtained results revealed that, the investigated catalysts consisted of nanosized ?-Al2O3 phase. The textural properties including SBET, porosity and St were modified by Ag2O-doping. The doping process with Ag-species improved the catalytic activity of FeAlO system. Increasing the precalcination temperature from 400 to 800?C increased the catalytic activity (k30°C) of 3.5 % AgFeAlO with 1.9 fold towards H2O2 decomposition. Furthermore, the maximum increase in the k30°C value due to doping with 3.5 mol% Ag2O attained about 15.1 fold for the solids calcined at 800°C.

Keywords: Fe2O3/Al2O3 catalyst; Ag2O-doping; Surface; Amorphous materials; H2O2 decomposition

Introduction

Supported transition metal oxides are interesting solids due to their surface acid-base properties [1] and oxidation-reduction potentials [2]. Iron oxide-based catalysts are very important catalysts in oxidation processes [3,4]. Inspite of its catalytic potential and its availability α-Fe2O3 has low thermal stability against sintering which is accompanied by deactivation [5]. Therefore, supported metal oxide [6] usually exhibit modification in its textural, structural, and catalytic properties [7]. It is known that the activity and selectivity of a large variety of catalysts can be modified by loading on a finely divided support and doping with certain foreign oxides [8]. γ-Al2O3 is one of the supports commonly used in the petroleum refining, petrochemical industries [9], the oxidation–reduction reactions [10] and catalytic oxidation of methane [11] owing to its high specific surface area, porous structure, high thermal stability, perfect mechanical strength and acidity [12]. The surface acidity of γ-Al2O3 connected closely with the dispersion of the supported active metal oxides since the acid–base interaction between the metal oxides and the acid sites of γ-Al2O3 support promoted the dispersion of the metal oxides, which was favorable for increasing the catalytic activity [13,14]. γ-Al2O3, copper oxide, and ferric oxide are known as important components in automotive exhaust emission control and industrial catalysts [15].

Doping system containing transition metal oxides with certain foreign oxides is accompanied by significant modifications in their thermal stability, electrical, optical, magnetic, surface, and catalytic properties [16,17]. Many authors studied the effect of Ag+-doping on physicochemical properties of different supported transition metal oxides [17]. Doping of NiO catalyst with Ag2O resulted in a progressive increase in its surface electric conductivity [18]. Doping La0.6Ce0.4CoO3 catalyst with Ag+ alter its physical and chemical properties, such as the oxidation state of cobalt, the density of oxygen vacancies and the mobility of lattice oxygen. All these factors play important roles for NO decomposition [19]. The well-dispersed Ag+ ions in (C,S)- doped TiO2 significantly promote the electron–hole separation and subsequently enhance its photoactivity [20,21]. Silver supported on γ-alumina catalyst is used for epoxidation of ethylene to ethylene oxide below 300°C [22], and exhibited relatively good activity and selectivity for NO reduction to N2 [23]. Doping Co3O4/Al2O3 and Co3O4/MgO with Ag2O increases their catalytic activities in CO oxidation by O2 [16,17]. Doping V2O5/Al2O3, NiO and CuO catalysts with silver oxide increased their activities towards decomposition of H2O2 [18,24,25]. Catalytic decomposition of H2O2 is an oxidation–reduction reaction, used as a green fuel/propellant instead of carcinogenic hydrazine in spaceflight at certain conditions [26]. Hydrogen peroxide can be used as an oxidizer on the fuel cells [27] instead of as liquid oxygen. Furthermore, H2O2 can be used as a suitable alternative fuel [28]. It has been used as a source of the hydroxyl radical (•OH) in the presence of UV irradiation for destruction of organic wastes [29]. Decomposition of H2O2 over metal oxides and their mixtures has been investigated by several investigators [18,24,25,30,31] to measure their catalytic activities towards the oxidation-reduction process [27].

In this paper, we aimed to investigate the influence of extent of Fe2O3 loading, Ag-doping and precalcination temperature of the Fe2O3/ Al2O3 system on its physicochemical, surface and catalytic properties. The techniques employed were TG/DTG, XRD, nitrogen adsorption at −196°C and the catalytic decomposition of H2O2 at 25-40°C.

Materials And Methods

Aluminum hydroxide sample was prepared by precipitating Al(NO3)3·9H2O (1M) solution using 0.2M ammonia solution at 70°C and pH = 8. The precipitate was carefully washed with bi-distilled water till free from ammonium and nitrate ions, then filtered and dried at 110°C till constant weight. Al(OH)3 sample was calcined in air at 400, 500 and 800°C for 4h.

Five specimens of Fe2O3/Al2O3 solids were prepared by impregnation known mass of Al(OH)3 sample with solutions containing different amounts of iron nitrate dissolved in the least amount of bi-distilled water. The obtained pastes were dried at 110°C till constant weight then calcined in air at temperatures ranged between 400 and 800°C for 4 h. The nominal compositions of the prepared solids were 0.025, 0.035, 0.045, 0.055, 0.065 Fe(NO3)2.9H2O: Al(OH)3. The iron oxide content in these specimens was 3.76, 5.19, 6.58, 7.92 and 9.23 wt %, respectively. The formula of prepared solids was abbreviated as xFeAlO.

The Ag2O-doped 0.045 Fe2O3/Al2O3 system was prepared using known mass of Al(OH)3 impregnated with solutions containing a fixed amount of iron nitrate and different proportions of silver nitrate. The obtained pastes were dried at 110°C till constant weight then calcined in air at temperatures ranged between 400 and 800°C for 4 h. The concentration of Fe2O3 (6.58 wt %) and of Ag2O added were 1.5, 2.0, 3.0, 3.5 and 4.0 mol % which corresponding 3.08, 4.06, 5.97, 6.89 and 7.80 wt %, respectively. The formula of the prepared samples was abbreviated as yAgFeAlO. All the chemicals employed were of analytical grade supplied by BDH Company.

Characterization techniques

Thermogravimetric analysis (TGA) and differential thermogravimetric analysis (DTG) of the catalysts were carried out using Shimadzu TGA-50H thermo-gravimetric analyzer; the rate of heating was kept at 10°C/min. X-ray diffractograms of various prepared solids were determined by using a Brucker diffractometer (Brucker Axs D8 Advance Germany). The Patterns were run with CuKα1 with secondly monochromator, (λ = 0.15404 Å) at 40 kV and 40 m A at scanning rate of 2° in 2θ /min.The surface characteristics, namely specific surface areas (SBET), total pore volume (Vp) and average pore radius (r) of the various catalysts were determined form nitrogen adsorption isotherms measured at −196°C using a Quantachrome NOVA 2000 automated gas-sorption apparatus model 7.11. All catalysts were degassed at 200°C for 2h under a reduced pressure of 10-5 Torr before undertaking such measurements.

The catalytic activities of the various catalysts were measured by studying the decomposition of H2O2 at 25–40°C using 100 mg of a given catalyst sample with 0.5 ml of H2O2 of known concentration diluted to 20 ml with distilled water. The reaction kinetics was monitored by measuring the volume of O2 liberated at different time intervals until no further oxygen was liberated.

Results And Discussion

Thermal properties

Thermogravimetric analysis (TGA) and differential thermogravimetric analysis (DTG) of uncalcined Al(OH)3, 0.045FeAlO and 3.5% AgFeAlO solids were determined and shown in (Figure 1(AC)), respectively. Inspection of Figure 1: (i) The TG curve of Al(OH)3 solid consists of three stages. The first and second thermal processes are indicative to desorption of physisorbed water and starting the decomposition of aluminum hydroxide to aluminum oxyhydroxide Al(OOH) [32]. The last step corresponds to the complete decomposition of aluminum hydroxide into the corresponding oxide Al2O3. (ii) TG curve of uncalcined 0.045FeAlO solid consists of three stages. The first step represents desorption of physisorbed water and water of crystallization of iron nitrate. The second step is indicative to start the decomposition of aluminum hydroxide to aluminum oxyhydroxide compound Al(OOH) [32]. The last step corresponds to the complete decomposition of iron nitrate and aluminum oxyhydroxide into Fe2O3 and Al2O3, respectively. (iii) TG curve of uncalcined 3.5%AgFeAlO solid consists of three stages. The first step represents desorption of physisorbed water and water of crystallization of iron nitrate. The second step represents the complete decomposition of silver nitrate and aluminum hydroxide into Ag2O and Al2O3, respectively. The last step corresponds to the complete decomposition of iron nitrate yielding the corresponding oxide Fe2O3. It can be concluded that doping Fe2O3/ Al2O3 system with Ag2O enhanced the thermal decomposition of aluminum hydroxide and ferric nitrate to Al2O3 and Fe2O3. This result confirmed by the previous published data [24], which indicated the presence of Ag2O enhanced the thermal decomposition of aluminum hydroxide and ammonium vanadate to Al2O3 and V2O5, respectively.

thermodynamics-catalysis-curves-uncalcined

Figure 1: TG and DTG curves of uncalcined (A) Al(OH)3, (B) 0.045FeAlO and (C) 3.5 %AgFeAlO solids.

XRD investigation of the prepared solids

XRD of Fe2O3/Al2O3 system: The X-ray diffractograms of γ-Al2O3, 0.045FeAlO and 0.065FeAlO solids being calcined at 500°C were determined and illustrated in Figure 2. The effect of Fe2O3 loading on the degree of ordering of γ-Al2O3 phase was investigated as shown in Table 1. Inspection of Figure 2 and Table 1: (i) the diffractograms contain diffraction lines at d-spacing = 2.78, 2.39, 1.98 and 1.396 ? due to γ-Al2O3 phase which is amorphous in nature (JCPDS 10-425). There is no diffraction lines related to Fe2O3 phase in the investigated solids calcined at 500°C. The absence of any XRD peaks attributable to iron oxide as separate phase in Fe2O3/Al2O3 system confirms their high dispersion and its small size to be detected by XRD tool [14,24]. So, γ-Al2O3 acted as a convenient support for hematite. On the other hand, the crystallization temperature of pure α-Fe2O3 is usually lower than 500°C, while in our samples the nucleation and growth of α-Fe2O3 grains were restrained by the amorphous structure of alumina matrix [33]. (ii) Increasing the Fe2O3 content from 4.5 to 6.5 mol% decreases the degree of ordering of γ-Al2O3 phase which has nano size ≤ 6 nm.

thermodynamics-catalysis-curves-uncalcined

Figure 2: XRD diffractograms of pure γ-Al2O3 and FeAlO solids with various Fe2O3 loading calcined at 500°C. Lines (1) refer to γ-Al2O3 phase.

Intensity count (a.u) Calcination
temperture
°C
Solid
γ-Al2O3
17.1 500 Al2O3
10.5 500 0.045FeAlO
8.4 500 0.065FeAlO
9.0 500 Ag-0.045FeAlO % 3.5
20.1 800 Al2O3
14.9 800 0.045FeAlO
13.5 800 Ag-0.045FeAlO % 3.5
12.5 800 Ag-0.045FeAlO % 4.0

Table 1: Intensity counts of the main diffraction lines of XRD of γ-Al2O3 phase for FeAlO samples calcined at 500 and 800°C.

XRD of Ag2O-doped Fe2O3/Al2O3 system: The X-ray diffractograms of γ-Al2O3, 0.045FeAlO and 0.06 system: The X-ray diffractograms of pure and Ag2O-doped 0.045FeAlO solids being calcined at 500 and 800°C were determined and illustrated in (Figure3A and 3B), respectively. Inspection of (Figure 3A and 3B) and (Table 1):

thermodynamics-catalysis-diffractograms-pure

Figure 3a: XRD diffractograms of pure and Ag2O treated 0.045FeAlO solids calcined at 500°C. Lines (1) refer to γ-Al2O3 phase.

thermodynamics-catalysis-solids-various

Figure 3b: XRD diffractograms of pure and treated 0.045FeAlO solids with various amounts of Ag2O calcined at 800°C. Lines (1) refer to γ- Al2O3 phase.

(i) the diffractograms of pure and doped solids calcined at 500°C consist of diffraction peaks due to γ-Al2O3 phase which is amorphous in nature and there are no diffraction lines due to Fe2O3 phase or silver species (Ag2O or Ag metal). (ii) The diffractograms of pure and doped solids precalcined at 800°C consist of diffraction lines related to poorly crystalline γ-Al2O3 phase and absence any lines related to Fe2O3 phase, iron aluminate spinel or silver species. (iii) Increasing the calcination temperature of pure and doped solids from 500 to 800°C increases the degree of ordering of γ-Al2O3 phase which has nano size (5-10 nm). (iv) Doping of 0.045FeAlO sample with Ag2O followed by calcination at 500 and 800°C didn’t much affect on the degree of ordering of γ-Al2O3 phase.

The above results can be explained on the light of the following: (i) absence the diffraction lines due to Ag-species (Ag2O or Ag metal) in doped solids calcined at 500°C or 800°C was expected because of the small amounts of silver oxide added were below the detection limits of the employed X-ray diffractometer [16]. It has been reported that heating of Ag2O at a temperature above 500°C gives metallic Agspecies [18,34] which are not detected by XRD technique because of their minute amounts. (ii) The absence of any solid-solid interaction between Fe2O3 and γ-Al2O3 yielding aluminate spinel can be attributed to the reaction between the transition metal oxides and Al2O3 to produce metal aluminate is strongly dependent upon the nature of the transition metal element. The rate of reaction between the metal oxide and Al2O3 deceases in the following order: Cu>Co>Ni>>Fe [35,36]. (iii) The slight decrease in the degree of ordering of γ-Al2O3 phase precalcined at 500 and 800°C due to Ag2O-doping could be attributed to a possible coating of the γ-Al2O3 crystallites with Ag2O film which hinders the particle adhesion process, thus limiting their grain growth during the course of heat treatment [37]. It has been reported that the presence of Ag effected the dispersion of Fe2O3 particles in the catalyst surface [38]. (iv) The increase in the degree of ordering of γ-Al2O3 phase for pure and doped solids due to increasing the calcination temperature from 500 to 800°C could be explained in the light of the grain growth mechanism or sintering processes [39].

The observed changes in both of degree of ordering and crystallite sizes of γ-Al2O3 present in the investigated catalysts as a result of Ag2O doping are expected to induce changes in the specific surface areas of FeAlO system.

Surface properties

The nitrogen adsorption isotherms were measured at -196°C for pure and Ag2O doped samples preheated at 500 and 800°C and illustrated in (Figure 4). The isotherms obtained are of type II of Brunauer’s classification [40] showing closed hysteresis loops. The specific surface areas were calculated from these adsorption isotherms by applying the BET equation [40] the data obtained are given in (Table 2 and Figure 5A and 5B), respectively. The total pore volumes (Vp) were taken at P/Po = 0.95 and are expressed in ml/gm. The average pore radius, r (Å), was calculated from the above-mentioned textural properties, applying the relationship: r = (2Vp/SBET) × 104 Å. Another series of specific surface areas St were computed from the VL-t plots of the various investigated adsorbents. These plots were constructed using the de Boer-t plot [41]. The computed St values are given also in (Table 2). Representative VL-t plot curves of investigated samples calcined at 500 and 800°C are shown in (Figure 6A and 6B), respectively.

thermodynamics-catalysis-nitrogen-adsorption

Figure 4: The representative nitrogen adsorption –desorption isotherms on pure and Ag2O-doped FeAlO samples calcined at 500 and 800°C.

thermodynamics-catalysis-Linear-BET

Figure 5a: Linear BET plot of Fe2O3, Al2O3, pure FeAlO and Ag2O-doped samples calcined at 500°C.

thermodynamics-catalysis-doped-samples

Figure 5b: Linear BET plot of Al2O3, pure0.045FeAlO and Ag2O-doped samples calcined at 800°C.

thermodynamics-catalysis-samples-calcined

Figure 6a: VL-t plots of Fe2O3, Al2O3, pure FeAlO and Ag2O-doped samples calcined at 500°C.

thermodynamics-catalysis-doped-samples

Figure 6b: VL-t plots of Al2O3, pure 0.045FeAlO and Ag2O-doped samples calcined at 800°C.

ŕ (Å) Vp
(ml/g)
St
(m2/g)
SBET (m2/g) BET-C
constants
Vm
(cc/g)
Calcination
Temperature °C
Solid
68.38 0.2383 74.9 69.7 25.00 16.00 500 Fe2O3
20.69 0.2239 221.2 216.4 46.35 49.72 500 Al2O3
19.43 0.2057 213.5 211.7 18.19 48.64 500 FeAlO
23.21 0.2263 199.5 195.0 31.40 44.80 500 0.065FeAlO
21.81 0.2367 217.2 217.1 23.30 49.88 500  Ag-0.045FeAlO % 3.5
36.68 0.2346 131.7 127.9 51.43 29.38 800 Al2O3
33.18 0.2162 130.2 130.3 17.49 29.94 800  FeAlO
35.63 0.2252 125.4 126.4 18.62 29.03 800 Ag-0.045FeAlO % 3.5
30.66 0.1852 117.5 120.8 13.92 27.75 800 Ag-0.045FeAlO % 4.0

Table 2: The surface characteristics of investigated pure and doped FeAlO samples calcined at 500 and 800°C.

Inspection of (Figure 5A and 5B) and (Table 2) show the following: (i) addition of different amounts of Fe2O3 (4.5& 6.5 mol %) to γ-Al2O3 support followed by calcination at 500°C resulted in a limited decrease in SBET of γ-Al2O3 attained about 2 and 10 %, respectively. (ii) The SBET of hematite precalcined at 500°C increased by its loading on γ-Al2O3 support with small amount (4.5 mol %), the increase in the SBET was about 204 %. Increasing Fe2O3 content to 6.5 mol % is accompanied by a decrease in SBET of FeAlO system with about 8 %. (iii) Doping of 0.045FeAlO sample with 3.5 mol % Ag2O followed by calcination at 500°C led to a slight increase in its SBET with about 3 %. (iv) The average pore radius r of the investigated samples precalcined at 500°C decreased by loading Fe2O3 on γ-Al2O3. (v) The rise in the calcination temperature of pure γ-Al2O3, 0.045FeAlO and 3.5% Ag-0.045FeAlO samples from 500 to 800°C brought about a significant decrease in their specific surface areas and an increase in the average pore radius r. The decrease in SBET values, due to increasing the calcination temperature, attained about 41, 39 and 42 %, respectively. (vi) Doping of 0.045FeAlO sample with 3.5 and 4.0 mol% Ag2O followed by calcination at 800°C led to a slight decrease in its SBET which attained about 3 and 7 %, respectively. (vii) Doping of 0.045FeAlO sample with Ag2O followed by calcination at 500 and 800°C didn’t much affect on the values of average pore radius r of the investigated system.

According to VL-t plot curves as shown in (Figure 6A and 6B) of investigated samples calcined at 500 and 800°C show the following: (i) Fe2O3 is mesoporous material. This behavior is indicated by the upward deviation following an initial linear region by demonstrating the existence of mesopores. (ii) The VL-t plots of γ-Al2O3, 0.045FeAlO, 0.065FeAlO, 3.5 % Ag-0.045FeAlO samples calcined at 500°C and 4.0 % Ag-0.045FeAlO sample calcined at 800°C reveal the microporosity character, as indicated by downward deviation from the initial straight line which passes through the origin. (iii) The VL-t plot of γ-Al2O3, 0.045FeAlO and 3.5 % Ag-0.045FeAlO samples calcined at 800°C, the initial linear region is followed by an upward deviation which is limited and a decrease in its slope is noted. This indicates the filling of some of the pores present by both multilayer formation and capillary condensation and the rest solely by multilayer formation. This indicates that these samples actually constitute of mixture of meso and micropores.

It is seen from (Table 2) that the values of SBET and St are close to each other which justifies the correct choice of standard t-curves used in the analysis.

The changes in the specific surface areas of the prepared and calcined samples can be explained as follow: (a) the observed increase in the SBET value of hematite due to loading on γ-Al2O3 support sample precalcined at 500°C can be discussed in the light of fine dispersion Fe2O3 particles on the surface of γ-Al2O3 [14]. Indeed, XRD peaks due to iron oxide were not detected in the investigated samples in the present work indicating that the iron oxide phase exists in a highly divided or amorphous state in these specimens. The change from the mesoporous structure to microporous structure as a result of supporting iron oxide on γ-Al2O3 as shown in (Figure 6A) is another factor. (b) The induced decrease in the surface area of 0.045FeAlO sample due to increasing the amount of Fe2O3 loading from 4.5 to 6.5 mol % may be ascribed to the aggregation of small iron species into larger bulk particles of iron in preparation process [42]. (c) The slight increase in SBET value due to Ag2O-doping precalcined at 500°C can be attributed to creation of new pores due to liberation of nitrogen oxides gases during the thermal decomposition of AgNO3 dopant added [43]. (d) The observed significant decrease in SBET of the investigated samples as result of increasing the calcination temperature from 500 to 800°C could be attributed to the sintering process. The sintering process might take place according to the collapse of the pore structure, pore widening [44] and/or the particle adhesion (grain growth) process together with possible phase transformation [45-47].

The observed changes in textural properties of the investigated solids as a result of increasing the extent of iron oxide loading, Agdoping and increasing the calcination temperature should modify the concentration of catalytically active constituents taking part in the catalyzed reaction.

Catalytic properties of the prepared solids

Effect of extent of Fe2O3 loading on the catalytic activity of FeAlO system: Catalytic decomposition of H2O2 is a model reaction chosen to study the redox properties of the prepared catalysts. (Figure 7) shows the First-order plots of H2O2 decomposition conducted at 30°C using FeAlO catalysts at different Fe2O3 loading calcined at 400°C. (Figure 7) shows that γ-Al2O3 support solid exhibits no catalytic activity towards H2O2 decomposition reaction. The catalytic activity of FeAlO catalyst is higher than that of pure Fe2O3 catalyst. The increase in amount of Fe2O3 content from 2.5 to 6.5 mol% is accompanied by increasing the catalytic activity of FeAlO system precalcined at 400°C. The maximum increase in the catalytic activity attained about 34 % for 0.065FeAlO catalyst at k30°C with respect to pure Fe2O3.

thermodynamics-catalysis-catalysts-calcined

Figure 7: First-order plots of H2O2 decomposition conducted at 30°C over pure and various extent of Fe2O3/Al2O3 catalysts calcined at 400°C.

XRD and SBET measurements showed that the high dispersion of Fe2O3 on γ-Al2O3 and the significant increase in the SBET may be responsible for the higher catalytic activity of FeAlO than the pure Fe2O3. The absence of any XRD patterns detected for Fe2O3 as separate phase reflected the decrease in the crystallite size of detected phase which becomes so small to be detected by the employed XRD technique and hence increasing the surface area of investigated solids. Other factor, we cannot overlook, is the creation of bivalent catalytic centers [32,48] such as Fe3+–Fe2+ ion pairs that are involved in H2O2- decomposition reaction. It has been reported that a favorable redox couple of Fe2+–Fe3+ is essential for catalytic decomposition of H2O2 through electron exchange [49].

Effect of calcination temperature on the catalytic activity of 0.045FeAlO system: Variation of the catalytic activity expressed as reaction rate constant (k min-1) as a function of precalcination temperature of 0.045FeAlO system in the range of 400-800°C towards H2O2 decomposition conducted at 25-40°C was investigated and determined as shown in (Figure 8). Inspection of (Figure 8) (i) Increasing the calcination temperature from 400 to 500°C increases the catalytic activity of 0.045FeAlO system; the increase in the k30°C value attained about 14 %. (ii) Increasing the calcination temperature from 500 to 800°C was accompanied by a progressive decrease in the catalytic activity of 0.045FeAlO system; the decrease in the k30°CC value attained about 45 %. (iii) The catalytic activity of 0.045FeAlO system increased with increasing the reaction temperature from 25 to 40°C.

thermodynamics-catalysis-calcination-temperatures

Figure 8: Variation of reaction rate constant (k) as a function of calcination temperatures for the catalytic decomposition of H2O2 conducted at 25-40°C over 0.045FeAlO catalysts.

Increasing the catalytic activity of 0.045FeAlO system as a result of increasing the calcination temperature from 400 to 500°C may be attributed to increase in the concentration of catalytically active constituents of Fe3+–Fe2+ ion pairs taking part in the catalysis of H2O2-decomposition reaction. The progressive decrease in the catalytic activity of 0.045FeAlO system as a result of increasing the calcination temperature from 500 to 800°C may be due to (i) increasing the degree of ordering of γ-Al2O3 phase (Table 1). (ii) The sintering process of catalytically active sites with subsequent decrease in its specific surface area from 211.7 to 130.3 m2 g-1 (Table 2) [50,51]. (iii) The effective removal of surface OH groups which act as active sites for the H2O2 decomposition reaction [37]

Effect of Ag2O-doping on the catalytic activity of 0.045FeAlO system calcined at different calcination temperatures: Variation of the catalytic activity expressed as reaction rate constant (k min-1) of H2O2 decomposition conducted at 25-40°C as a function of wt % of Ag2O for the solids precalcined at 500 and 800°C was investigated and graphically represented in (Figure 9A and 9B), respectively. The variation of k (min-1) of H2O2 decomposition over 3.5 % AgFeAlO conducted at 30°C as a function of calcination temperature is graphically represented in (Figure 10). It is seen from these (Figures 9 &10) that: (i) the catalytic activity of 0.045FeAlO catalysts increases progressively by increasing the amounts of dopant up to certain extent reaching to a maximum at 3.5 mol % Ag2O, the increase in the k30°C value of 3.5 % AgFeAlO calcined at 500°C attained about 502 %. (ii) Increasing Ag2O content to 4.0 mol % is accompanied by a sharp decrease in the catalytic activity falling to values greater than that of pure catalysts precalcined at the same temperature. The maximum decrease in the k30°C value on increasing the amount of dopant from 3.5 to 4.0 mol % attained about 40 and 30 % for the catalysts precalcined at 500 and 800°C, respectively. (iii) The catalytic activity of pure and doped 0.045FeAlO system increased with increasing the reaction temperature from 25 to 40°C. (iv) The catalytic activity of doped 0.045FeAlO solid with 3.5 mol % Ag2O increases progressively by increasing the calcination temperature from 400 to 800°C. The increase in the k30°C value due to doping with 3.5 mol %Ag2O attained about 383, 502, 635 and 1511 % for the catalysts precalcined at 400, 500, 600 and 800°C, respectively.

thermodynamics-catalysis-reaction-rate

Figure 9a: Variation of reaction rate constant (k) as a function of wt % of Ag2O for the catalytic decomposition of H2O2 conducted at 25-40°C over pure 0.045FeAlO catalysts calcined at 500°C.

thermodynamics-catalysis-decomposition

Figure 9b: Variation of reaction rate constant (k) as a function of wt % of Ag2O for the catalytic decomposition of H2O2 conducted at 25-40°C over pure 0.045FeAlO catalysts calcined at 800°C.

The significant enhancement in the catalytic activity of 0.045FeAlO solids in the investigated redox reaction as a result of Ag2O-doping could be discussed in terms of: (i) the slight decrease in the degree of ordering of γ-Al2O3 phase in 3.5 % AgFeAlO sample calcined at 500°C (as shown in XRD section). This effect could result from a possible coating of the γ-Al2O3 support with an Ag2O film which acts as an energy barrier opposing their particles adhesion. This reflected the role of Ag2O treatment in increasing the degree of dispersion of Fe2O3 and consequently increasing the catalytic activity of H2O2 decomposition. (ii) The possible changes in the concentration of ion pairs acting as active sites for the catalyzed reaction present in the outermost surface layers of the treated solids [24,52,53]. The created ion pairs due to Ag2O-doping may be Ag+–Fe2+, Ag–Fe3+ and Ag–Ag+ [25].

However, the observed significant decrease in the k30°C value on increasing the amount of dopant from 3.5 to 4.0 mol% can be attributed to decrease in the concentration of catalytically active species involved in the catalyzed reaction. This could result from the location of dopant species on the surface layers of the treated catalysts thus blocking some of the active constituents by dopant cation and/or metallic silver [54].

The observed increase in the catalytic activity of Ag-doped FeAlO solids by increasing the calcination temperature from 400 to 800°C (Figure 10) could be attributed to the possible presence of metallic Agspecies and increase in the concentration of surface excess oxygen as a result of heating at a temperature above 450°C [18,34] according to the following reaction [55].

thermodynamics-catalysis-calcination-temperatures

Figure 10: Variation of reaction rate constant (k) as a function of calcination temperatures for the catalytic decomposition of H2O2 conducted at 30°C over 3.5%AgFeAlO catalysts.

2 AgNO3 2Ag+2NO2+O2

In spite of increasing the calcination temperature decreases the surface areas of the Ag-doped solid catalysts. The observed increase in catalytic activity clearly reflects the minor role played by surface area in determining the catalytic activity of doped solids.

The formation of radicals from H2O2 and iron oxides has been proposed in the literature [56]. The mechanism of catalytic decomposition of H2O2 by radicals generation could be explained by the possible initiation the reaction with partially reduced surface specie, for example, Fe2+, according to the Haber–Weiss mechanism:

Fesurf2++ H2O2→ Fesurf3++ •OH + −OH (1)

The formation of O2 in a radical reaction can be very complex but a simple pathway can be proposed via the hydroperoxide radical:

H2O2 + •OH → H2O + •OOH (2)

Fesurf3++•OOH → Fesurf2++ H+ +O2 (3)

In these reactions a hydroperoxide radical intermediate is generated, which can then react with a surface species to produce O2 and H+. The H+ is then neutralized by the OH- generated in (Equation 1).

The present Fe3+ for the solids calcined at 800oC is reduced during the decomposition of hydrogen peroxide (as shown in Equation 3) there is the possibility to regenerate the Fe2+, which can catalyze the decomposition process [57].

Conclusions

Doping with Ag2O enhanced the thermal decomposition of aluminum hydroxide and ferric nitrate to Al2O3 and Fe2O3. Ag2Odoping led to a slight decrease in the degree of ordering of γ-Al2O3 phase precalcined at 500 and 800°C. Doping of 0.045FeAlO catalyst with 3.5 or 4.0 mol % Ag2O preheated at 800°C decreased its BET-surface area. The catalytic activity of FeAlO system towards H2O2 decomposition increased by increasing the Fe2O3 content and Ag2O amount up to 3.5 mol%. The maximum increase in the catalytic activity attained about 34 % for 0.065FeAlO catalyst at k30°C with respect to un-supported Fe2O3. The maximum increase in the catalytic activity measured at k30°C due to 3.5 mol % Ag2O-doping attained about 15.1 fold for the solids calcined at 800°C.

References

  1. Khaleel A, Shehadi I, Shamisi M (2010) Nanostructured chromium-iron mixed oxides: Physicochemical properties and catalytic activity. Colloids Surf A 355: 75-82.
  2. Cao JL, Wang Y, Yu XL, Wang SR, Wu SH et al. (2008) Mesoporous CuO- Fe2O3 composite catalysts for low-temperature carbon monoxide oxidation. Appl Catal B 79: 26-34.
  3. Al-Sayari S, Carley AF, Taylor SH, Hutchings GJ (2007) Au/ZnO and Au/Fe2O3 catalysts for CO oxidation at ambient temperature: comments on the effect of synthesis conditions on the preparation of high activity catalyst prepared by coprecipitation. Top Catal 44: 123-128.
  4. Wang HC, Chang SH, Hung PC, Hwang JF, Chang MB (2008) Catalytic oxidation of gaseous PCDD/Fs with ozone over iron oxide catalysts. Chemosphere 71: 388-397.
  5. Liu XH, Shen K, Wang YG, Wang YQ, Guo YL et al. (2008) Preparation and catalytic properties of Pt supported Fe-Cr mixed oxide catalysts in the aqueous-phase reforming of ethylene glycol. Catal Commun 9: 2316-2318.
  6. Klabunde KJ, Khaleel A, Park D (1995) Overlayer of iron oxide on nanoscale magnesium oxide crystallites: chlorocarbon destruction by catalytic solid-state ion/ion exchange.High Temp Mater Sci 33: 99-106.
  7. Wachs IE (2005) Recent conceptual advances in the catalysis science of mixed metal oxide catalytic Materials. Catal Today 100: 79-94.
  8. El-Molla SA (2006) Dehydrogenation and condensation in catalytic conversion of iso-propanol over CuO/MgO system doped with Li2O and ZrO2. Appl Catal A 298: 103-108.
  9. Jun Cheng L, Lan X, Feng X, Wen WZ, Fei W (2006) Effect of hydrothermal treatment on the acidity distribution of ?-Al2O3 support. Appl Surf Sci 253: 766- 770.
  10. El-Shobaky GA, El-Khouly SM, Ghozza AM, Mohamed GM (2006) Surface and catalytic investigations of CuO-Cr2O3/Al2O3 system. Appl Catal A 302: 296-304.
  11. Lippits MJ, Boer Iwema RRH, Nieuwenhuys BE (2009) A comparative study of oxidation of methanol on ?-Al2O3 supported group IB metal catalysts. Catal Today 145: 27-33.
  12. Stiles AB (1987) "Catalyst Supports and Supported Catalysts: Theoretical and Applied Concepts", Butterworths, Boston.
  13. Uguina MA, Delgado JA, Rodríguez A, Carretero J, Gómez-Díaz D (2006) Alumina as heterogeneous catalyst for the regioselective epoxidation of terpenic diolefins with hydrogenperoxide. J Mol Catal A 256: 208-215.
  14. Wu Y, Gao F, Liu B, Dai Y, Zhu H, et al. (2010) Influence of ferric oxide modification on the properties of copper oxide supported on ?- alumina. J Colloid Interface Sci 343: 522-528.
  15. Ozawa M, Suzuki S, Loong CK, Richardson JW, Thomas RR (1997) Structural phase transitions and lean NO removal activity of copper-modified Alumina. Appl Surf Sci 121-122: 441-444.
  16. El-Shobaky GA, Shouman MA, El-Khouly SM (2003) Effect of silver oxide doping on surface and catalytic properties of Co3O4/Al2O3 system. Mater Lett 58: 184-190.
  17. Deraz NM (2001) Effect of Ag2O doping on surface and catalytic properties of cobalt-magnesia Catalysts. Mater Lett 51: 470-477.
  18. Turky AM (2003) Electrical surface and catalytic properties of NiO as influenced by doping with CuO and Ag2O. Appl Catal A 247: 83-93.
  19. Liu Z, Hao J, Fu L, Zhu T (2003) Study of Ag/La0.6Ce0.4CoO3 catalysts for direct decomposition and reduction of nitrogen oxides with propene in the presence of oxygen. Appl Catal B 44: 355-370.
  20. Hamal DB, Klabunde KJ (2007) Synthesis, characterization, and visible light activity of new nanoparticle photocatalysts based on silver, carbon, and sulfur doped TiO2. J Colloid Interface Sci 311: 514-522.
  21. Zielinska-Jurek A, Kowalskab E, Sobczak JW, Lisowski W, Ohtani B, et al. (2011) Preparation and characterization of monometallic (Au) and bimetallic (Ag/Au) modified-titania photocatalysts activated by visible light. Appl Catal B 101: 504-514.
  22. Wu Y, Yan A, He Y, Wu B, Wu T (2010) Ni-Ag-O as catalyst for a novel onestep reaction to convert ethane to ethylene oxide. Catal Today 158:258-262.
  23. She X, Flytzani-Stephanopoulos M (2006) The role of Ag-O-Al species in silver-alumina catalysts for the selective catalytic reduction of NOx with methane. J Catal 237: 79-93.
  24. Shaheen WM (2006) Thermal solid-solid interactions and catalytic properties of V2O5/Al2O3 system treated with Li2O and Ag2O. Mater Sci Eng B 135: 30-37.
  25. TurkyA M, Radwan N R E, El-Shobaky G A (2001) Surface and catalytic properties of CuO doped with MgO and Ag2O. Colloids Surf A 181: 57-68.
  26. Teshimal N, Genfa Z, Dasgupta P K (2004) Catalytic decomposition of hydrogen peroxide by a flow-through self-regulating platinum black heater. Anal Chim Acta 510: 9-13.
  27. Sanli AE, Ayta A (2011) Response to Disselkamp: Direct peroxide/peroxide fuel cell as a novel type fuel cell. Int Hydrogen Energy 36: 869-875.
  28. Wee JH (2006) Which type of fuel cell is more competitive for portable application: direct methanol fuel cells or direct borohydride fuel cells?. J Power Sources 161: 1-10.
  29. Bandara J, Mielczarski J A, Lopez, Kiwi J (2001) Sensitized degradation of chlorophenols on iron oxides induced by visible light. Comparison with titanium oxide. Appl Catal B 34: 321-333.
  30. Rao G R, Sahu H R, Mishra B G (2003) Surface and catalytic properties of Cu-Ce-O composite oxides prepared by combustion method. Colloids Surf A 220: 261-269.
  31. Renuka NK (2010) A green approach for phenol synthesis over Fe3+/MgO catalysts using hydrogen peroxide. J Mol Catal A 316: 126-130.
  32. Shaheen WM, Hong KS (2002) Thermal characterization and physicochemical properties of Fe2O3-Mn2O3/Al2O3 system. Thermochim Acta 381: 153-164.
  33. Liu M, Lib H, Xiao L, Yu W, Lu Y, Zhao Z (2005) XRD and Mössbauer spectroscopy investigation of Fe2O3-Al2O3 nano-composite. J Magn Magn Mater 294: 294-297.
  34. Imamura S, Yamada H, Utani K (2000) Combustion activity of Ag/CeO2 composite catalyst. Appl Catal A 192: 221-226.
  35. Gardener SD, Hoflund GB, Upchurch BT, Schryer DR, Kielin EJ, Schryer J (1991) Comparison of the performance characteristics of Pt/SnOx and Au/MnOx catalysts for low-temperature CO oxidation. J Catal 129: 114-120.
  36. Fagal GA, El-Shobaky GA, El-Khouly SM (2001) Surface and catalytic properties of Fe2O3-Cr2O3/Al2O3 solids as being influenced by Li2O and K2Odoping. Colloids Surf A 178: 287-296.
  37. El-Shobaky GA, Amin NH, Deraz NM, El-Molla, SA (2001) Decomposition of H2O2 on pure and ZnO-treated Co3O4/Al2O3 solids. Adsorp Sci Technol 19(1): 45-58.
  38. Brinen JS, Schmitt JL, Doughman WR, Achorn PJ, Siegel LA et al.(1975) X-ray photoelectron spectroscopy studies of the rhodium on charcoal catalyst: II. Dispersion as a function of reduction. J Catal 40: 295-300.
  39. Shaheen WM (2007) Thermal behaviour of pure and binary Fe(NO3)3·9H2O and (NH4)6Mo7O24·4H2O systems. Mater Sci Eng A 445-446:113-121.
  40. Brunauer S, Emmett PH, Teller E (1938) Adsorption of Gases in Multimolecular Layers. J Am Chem Soc 60: 309-319.
  41. Lippens BC, deBoer JH (1965) Studies on pore systems in catalysts: V. The t method. J Catal 4: 319-323.
  42. Wang JB, Lin SC, Huang TJ (2002) Selective CO oxidation in rich hydrogen over CuO/samaria-doped ceria. Appl Catal A 232: 107-120.
  43. El-Molla SA, Hammed MN, El-Shobaky GA (2004) Catalytic conversion of isopropanol over NiO/MgO system doped with Li2O. Mater Lett 58: 1003-1011.
  44. El-Shobaky GA, Deraz NM (2001) Surface and catalytic properties of cobaltic oxide supported on an active magnesia. Mater Lett 47: 231-240.
  45. Deraz NM (2008) Production and characterization of pure and doped copper ferrite nanoparticles. J Anal Appl Pyrolysis 82: 212-222.
  46. Deraz NM (2003) The formation and physicochemical characterization of Al2O3- doped manganese ferrites. Thermochim Acta 401: 175-185.
  47. Radwan NRE, El-Sharkawy EA, Youssef AM (2005) Influence of gold and manganese as promoters on surface and catalytic performance of Fe2O3/Al2O3 system. Appl Catal A 28: 93-106.
  48. Mucka V, Tabacik S (1991) Catalytic properties of nickel-cobalt mixed oxides and the influence of ionizing radiation. Radiat Phys Chem 38: 285-290.
  49. Parida KM, Dash SS, Mallik S, Das J (2005) Effect of heat treatment on the physico-chemical properties and catalytic activity of manganese nodules leached residue towards decomposition of hydrogen peroxide. J Colloid Interface Sci 290: 431-436.
  50. Shaheen WM (2002) Thermal solid-solid interaction and catalytic properties of CuO/Al2O3 system treated with ZnO and MoO3. Thermochim Acta 385: 105-116.
  51. Shaheen WM (2007) Effects of thermal treatment and doping with cobalt and manganese oxides on surface and catalytic properties of ferric oxide. Mater Chem Phys 101: 182-190.
  52. Shaheen WM, Ali AA (2001) Thermal solid-solid interaction and physicochemical properties of CuO-Fe2O3 system. Int J Inorg Mater 3: 1073-1081.
  53. Shaheen WM, Zahran AA, El-Shobaky GA (2003) Surface and catalytic properties of NiO/MgO system doped with Fe2O3. Colloids Surf A 231: 51-65.
  54. Radwan NRE, Turky AM, El-Shobaky GA (2002) Surface and catalytic properties of CuO doped with Li2O and Al2O3. Colloids Surf A 203: 205-215.
  55. Anand S, Srivastava ON (2002) Effect of AgNO3 additive/doping on microstructure and uperconductivity of the Pb doped Hg: 1223 thin film prepared through spray pyrolysis. J Phys Chem Solids 63: 1647-1653.
  56. Kwan WP, Voelker BM (2003) Rates of hydroxyl radical generation and organic compound oxidation in mineral-catalyzed Fenton-like systems. Environ Sci Technol 37: 1150-1158.
  57. Costa RCC, Lelis MFF, Oliveira LCA, Fabris JD, Ardisson JD et al. (2006) Novel active heterogeneous Fenton system based on Fe3-xMxO4(Fe, Co, Mn, Ni): The role of M2+ species on the reactivity towards H2O2 reactions. J Hazard Mater B 129: 171-17.
Citation: Ali LI, El-Molla SA, Amin NH, Ebrahim AA, Mahmoud HR (2011) Effect of Ag-doping of Nanosized FeAlO System on its Structural, Surface and Catalytic Properties. J Thermodyn Catal 2: 108.

Copyright: © 2011 Ali LI, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top