You are on page 1of 5

Powder Technology 253 (2014) 809–813

Contents lists available at ScienceDirect

Powder Technology
journal homepage: www.elsevier.com/locate/powtec

The effect of sulfate contents on the surface properties of iron–manganese


doped sulfated zirconia catalysts
Fatah H. Alhassan a,b,⁎, Umer Rashid c, Mothanna Sadiq Al-Qubaisi d,
Abdullah Rasedee d,e, Yun H. Taufiq-Yap a,b,⁎
a
Catalysis Science and Technology Research Centre, Faculty of Science, Universiti Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia
b
Department of Chemistry, Faculty of Science, Universiti Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia
c
Institute of Advanced Technology, Universiti Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia
d
Institute of Bioscience, Universiti Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia
e
Department of Veterinary Pathology and Microbiology, Faculty of Veterinary Medicine, University Putra Malaysia, Malaysia

a r t i c l e i n f o a b s t r a c t

Article history: The iron–manganese doped sulfated zirconia catalysts were prepared via precipitation method; the sulfation was
Received 25 June 2013 carried out by impregnation with different amounts of sulfate (4%, 10% and 16% SO−2 4 by weight) with the addi-
Received in revised form 15 December 2013 tion of Fe–Mn doped and calcined at 600 °C for 3 h. The prepared catalysts were characterized by TGA-DTA, XRD,
Accepted 27 December 2013
BET, FT-IR, TEM, TPD-NH3 and XPS. XRD and BET results revealed that the addition of sulfate imparts special sta-
Available online 5 January 2014
bilization to the catalytically active tetragonal phase of zirconia. All the iron–manganese doped sulfated zirconia
Keywords:
catalysts were found to have strong acid sites, high surface area and small crystallite size.
Iron–manganese doped sulfated zirconia 4% © 2014 Elsevier B.V. All rights reserved.
SO−4
2
by weight (FMSZ-4)
Sulfation
Dopants
Crystallite size

1. Introduction for the isomerization of n-butane at room temperature [4,5]. The iron–
manganese doped sulfated zirconia is reported to be about three orders
Zirconium oxide is an active catalyst for many reactions, whose sur- in magnitude more active than sulfated zirconia for isomerization of re-
face possesses four chemical properties: acidic, basic, reducing and oxi- action [5], with large amount of superacid sites, which shows that it has
dizing. It has become well established that the performance of a the highest acid strength among solid superacids containing sulfate ion
heterogeneous catalyst depends not only on the intrinsic catalytic activ- [6]. Findings from our literature review showed that no work has been
ity of its components, but also on its texture and stability. One of the done so far on the influence of sulfate amounts on the surface properties
most important factors in controlling the surface properties of a catalyst of iron–manganese doped sulfated zirconia catalyst.
involves the correct choice of additives. The adsorption of various an- The main purposes of this study therefore are to prepare zirconium
ions [1], particularly sulfate or phosphate anions, onto oxide, has been oxide catalyst acidified with sulfate anions, doped with iron and manga-
attempted as a means of improving their catalytic activity. The increase nese and to study the effect of sulfate amounts on the surface properties
in activity is believed to arise from an increase in the surface acidity of of it. ZrO2 was prepared from zirconyl oxychloride ZrOCl 2 and has
the modified oxide [2]. Modification of metal oxides with sulfate anion been sulfated using ammonium sulfate with different loading levels
can generate a strong acidity, even stronger than 100% sulfuric acid of support (4%, 10% and 16% SO − 2
4 ); Fe (1.5%) and Mn (0.5%) ions
and hence they tend to become superacid catalysts that are useful in re- were supported on SO− 2
4 /ZrO(OH)2 by stepwise equilibrium adsorp-
actions like isomerization, low temperature esterification, alkylation tion of Fe(NO3)3·9H2O and Mn(NO3)2·4H2O aqueous solutions. The
and cracking [3]. catalysts were characterized using different tools to determine the
The acid strength of iron–manganese doped sulfated zirconia cannot effect of sulfate levels on their surface properties.
be measured by the color-change indicators because of its dark color.
However, it is regarded as superacid, judging from its catalytic activities 2. Experimental procedure

2.1. Materials
⁎ Corresponding authors at: Catalysis Science and Technology Research Centre, Faculty
of Science, Universiti Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia. Tel.: + 60 3
89466809; fax: + 60 3 89466758.
Iron nitrate nonahydrate, manganese nitrate tetra hydrate, ammoni-
E-mail addresses: abuohamid9090@gmail.com (F.H. Alhassan), taufiq@upm.edu.my um sulfate, zirconium oxy chloride, and ammonium hydroxide (28–30%)
(Y.H. Taufiq-Yap). were obtained from (Sigma-Aldrich—Malaysia).

0032-5910/$ – see front matter © 2014 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.powtec.2013.12.045
810 F.H. Alhassan et al. / Powder Technology 253 (2014) 809–813

Fig. 1. TGA and DTG profile for zirconyl hydroxide gel.

2.2. Catalyst preparation

5 g of hydrate zirconyl oxychloride ZrOCl2 was dissolved in about


100 ml of deionised water, dropwise addition of 2.63 g of a 25% aqueous
solution of ammonia led to precipitation; as much of the ammonia solu-
tion necessary to reach a pH of ≈9–10 was added. The precipitate was Fig. 3. XRD patterns of sample Z, (FMSZ-4), (FMSZ-10), and (FMSZ-16) calcined at
600–3 h: (♦) tetragonal phase; (O) monoclinic phase.
allowed to age overnight, filtered by vacuum, washed several times
until it tested negative to chloride ion, and then dried for about 24 h
at 120 °C.
Fe(NO3)3·9H2O, Mn(NO3)2·4H2O and ammonium sulfate were dis- Cu-Kα radiation to generate diffraction patterns from powder crystalline
solved in about 100 ml of deionised water in amounts corresponding to samples at ambient temperature. The Cu-Kα radiation was generated by
the desired stoichiometry, solution was added dropwise to zirconium Philips glass diffraction X-ray tube broad focus 2.7 kW type. The crystal-
oxyhydroxide solution and the mixture was vigorously stirred at room lite size D of the samples was calculated using the Debye–Scherrer's rela-
temperature for 4 h. The mixture dried at 120 °C overnight and calcined tionship [8–11].
at 600 °C −3 h [7].
D ¼ 0:9λ=ðB cosθÞ
2.3. Catalyst characterization
where D is the crystallite size, λ is the incident X-ray wavelength, β is the
full width at half-maximum, and θ is the diffraction angle.
The TGA analysis was carried out on a Mettler Toledo TG-SDTA
The Fourier transform infrared (FT-IR) analysis was done with
apparatus (Pt crucibles, Pt / Pt − Rh thermocouple) with the purge
PerkinElmer spectrometer model 100 series (sample preparation
gas (nitrogen) flow rate of 30 ml min − 1 and the heating rate of
UATR). The total surface area of the catalysts was obtained using
10 °Cmin− 1 from 25 to 1000 °C.
Brunauer–Emmer–Teller (BET) method with nitrogen adsorption at
The powder X-ray diffraction analysis was performed using a
− 196 °C. Analysis was conducted using a Thermo Fisher Scientific S.p.A
Shimadzu diffractometer model XRD 6000. The diffractometer employed
(model: SURFER ANALYZER) nitrogen adsorption–desorption analyzer.
The transmission electron microscopy (TEM, Model Hitachi H-7100,
Japan) was used to determine the crystal shape and homogeneity of the
zirconia and iron–manganese doped sulfated zirconia nanoparticles.
Briefly, the powder was dispersed in deionized water, dropped onto
carbon-cover copper grids placed on a filter paper and dried at room
temperature.
The acidity of the catalyst was evaluated by temperature pro-
grammed desorption using NH3 as probe molecule. The TPD-NH3 exper-
iment was conducted by Thermo Finnigan TPDRO 1100 apparatus
equipped with a thermal conductivity detector. About 0.1 g of catalyst
was placed in the reactor, treated under 150 °C for 15 min in N2
(20 ml/min). 10% NH3 in helium gas was ramped at 1 °C/min for

Table 1
Influence of sulfate loading on the surface area, pore volume and crystallite size.

Sample Specific surface area (m2 g−1), pore volume (cm3 g−1),
and crystallite size (nm)

Z 34.8 0.00056 25.5


FMSZ-4 72.0 0.01887 14.0
FMSZ-10 139.0 0.05336 9.5
FMSZ-16 149.0 0.05966 8.0
Fig. 2. TGA and DTG profile for FMSZ-4.
F.H. Alhassan et al. / Powder Technology 253 (2014) 809–813 811

Table 2
The distribution of acid site strength of the catalysts under
investigation.

Catalysts NH3 desorbed (μmol/g)

Z 630.4
FMSZ-4 4318.4
FMSZ-10 5458.0
FMSZ-16 7544.0

step approximately between 413 °C and 437 °C with about 0.315%


mass reduction was due to dehydroxylation of zirconyl hydroxide
while the third stage between 447 °C and 743 °C of about 0.7% weight
reduction was assigned to the dehydration of hydrated zirconium
oxide. Correspondingly, the differential thermogravimetric profile
showed two peaks at around 105 °C and 432 °C [12]. Therefore, the ex-
hibited thermal events involved in general, elimination of water of dif-
ferent strengths with a final formation of ZrO2 agree well with the data
Fig. 4. FT-IR of Z, (FMSZ-4), (FMSZ-10), (FMSZ-16).
of XRD [13].
The iron–manganese doped samples exhibited a higher thermal sta-
bility. The commencement of decomposition in these cases occurred
60 min. The purging with N2 was done at room temperature for 45 min only above 700 °C. Thus, it can be inferred that, besides delaying the
to remove NH3 in the gas phase. The analysis of NH3description was crystallization process, the addition of iron–manganese also helps in
then carried out between 50 °C and 900 °C under helium flow stabilizing the surface sulfate species.
(15 °C min−1, 20 ml min−1) and detected by a thermal conductivity The TGA profile of iron–manganese doped sulfated zirconia 4% SO−2 4
detector. by weight (FMSZ-4) sample depicted five mass loss steps (Fig. 2). The
XPS spectra were obtained using a modified AXIS ULTRA DLD first step, which ends at 236 °C, was attributed to the removal of phys-
(KRATOS) photoelectron spectrometer, equipped with a hemispherical ically adsorbed water, whereas the other three stages, which began at
analyzer. For excitation, AlKα (1486.6 eV) radiation was used. The ana- around 238 °C and finished at about 518 °C, were assigned to the
lyzer was operated in the fixed analyzer transmission mode at 40 eV decomposition of iron and manganese nitrate and dehydroxylation
pass energy. The X-ray gun was operated at 4 mA emission current process of ZrO(OH)2 [14]. The mass loss at higher temperature,which
and 15 kV acceleration voltage. Samples were pretreated under vacuum began at about 673 °C and ended at 926 °C, was referred to as the
(1 × 10−9 Torr) at 22 ± 2 for 2 h before measurements. The base pres- decomposition of sulfate group [15–17].
sure of the analyzer chamber was 1 × 10−9 Torr; during data collection, The delay in the transition from amorphous to crystalline material
the pressure was still always b 1 × 10−9 Torr. Binding energy (BE) cor- by sulfate doping was supported by thermal analysis results. Consistent
responding to C 1 s, Zr 3d, Fe 2p, Mn 2p, S 2p, and O 1s were measured. with earlier reports [18–21], an increase in the sulfate loading shifted
All binding energies are relative to C 1s peak taken as 284.5 eV. the peak maxima towards the high temperature region.

3. Results and discussion


3.2. X-ray diffraction
3.1. Thermal analysis
The XRD patterns of pure ZrO2 and iron–manganese doped sulfated
The thermal behavior of zirconyl hydroxide (Fig. 1) showed three zirconia samples were illustrated in Fig. 3. The high temperature calci-
stages of weight loss between 38 °C and 600 °C. The first weight loss nation seems essential for the formation of the catalytically active te-
step between 38 °C and 105 °C was attributed to the removal of sur- tragonal phase. Pure zirconia was crystallized even at 500 °C with
face physisorbed water molecules with 14.6%, the second mass loss only the tetragonal phase.

Fig. 5. TEM image of zirconia (a) and FMSZ-4 (b).


812 F.H. Alhassan et al. / Powder Technology 253 (2014) 809–813

Fig. 6. XPS spectra of zirconia (a–b) and FMSZ-4 (c–g).

The XRD patterns of catalysts calcined at 600 °C of zirconia consist of dispersed Fe2 O3 and MnO particles along with bulky sulfate species
a mixture of monoclinic and tetragonal phases. The sharp diffraction on the surface of zirconia particles prevent their agglomeration during
peaks at around 2θ = 30.1, 35.1, 50.3 and 60.1° are due to the tetrago- calcination [23].
nal form of zirconia (JSPDS file no.: 01-081-1545). While the diffraction
lines at about 2θ = 17.2, 24.1, 27.8, 31.4, 34.2, 38.7, 41.2, 45.5, 49.3,
3.3. FT-IR spectroscopy
54.2, 55.7, 57.5, 58.0, 63.1, 66.0, 69.2, 71.1 and 73.2° are attributed to
the monoclinic phase of zirconia ((JSPDS file no.: 00-065-2357). XRD
Fig. 4 displays the FT-IR spectra of FMSZ (4-6-16). From these spec-
patterns of iron–manganese doped catalysts showed metastable phase
tra, the presence of sulfate groups was confirmed by the peak at a range
of tetragonal ZrO2 which appeared at 2θ = 30.1, 35.1, 50.4 and 60.1°
of 1225 to 1070 cm−1, which were due to the asymmetric and symmet-
(JSPDS file no.: 01-081-1545). Whereas, all the XRD patterns showed
ric stretching frequency of the O_S_O and O\S\O group. The band
only peaks of ZrO2 phases.
which is around 1380 cm −1 is characteristic for the surface sulfate
The intensity of the peak corresponding to the monoclinic phase was
species having S_O covalent bonds. The only effect of sulfate loading
not shown in the iron–manganese-doped samples signifying the stabi-
resulted in increasing the band intensities of the sulfate group [2,24].
lization of the tetragonal phase. The absence of characteristic peaks
matching to Fe2O3and MnO implies that the metal oxides are present
in the form of solid solution or it is well dispersed on the zirconia sur- 3.4. Influence of sulfate loading on the surface area and pore volume
face. The dispersion of Fe2O3 and MnO particles causes the stabilization
of the tetragonal phase and also imparts a higher surface area [11]. The variation in the surface area and pore volume with sulfate load-
To know the effect of sulfate content on the iron–manganese-doped ing of iron–manganese doped sulfated zirconia is presented in Table 1.
catalysts quantitatively, their mean crystallite sizes were obtained from Consistent with earlier reports, the iron–manganese doped samples
the broadening of the strongest peak of the samples based on Debye– showed a higher surface area as compared to the zirconia sample. The
Scherrer's equation (Table 1). The addition of sulfate was associated dispersed Fe2O3 and MnO particles along with the sulfate species
with a reduction in crystallite size to become 14.0, 9.5 and 8.0 nm prevent the agglomeration of zirconia particles leading to an enhanced
for iron–manganese doped sulfated zirconia 4% SO− 4
2
by weight surface area [19,25–27]. In the case of iron–manganese doped samples,
(FMSZ-4), iron–manganese doped sulfated zirconia 10% SO− 4
2
by an increase in sulfate content resulted in an increase of surface area.
weight (FMSZ-10) and iron–manganese doped sulfated zirconia 16% Furthermore, the pore volume of the samples also showed an increased
SO−2
4 by weight (FMSZ-16), respectively. This may be attributed to the trend with increased sulfate loading. Thus it can be assumed that the
sulfate groups that remain bounded on the surface of the samples and incorporation of iron and manganese prevents surface area loss during
inhibit the growth of zirconia crystallites, agreeing thus with the other high temperature calcinations resulting in a gradual increase in the
transition metal oxides, i.e. titania, stannia and ferria [22]. The decrease surface area of the samples FMSZ-4, FMSZ-10 and FMSZ-16 to become
in the crystallite size can be explained by the hypothesis that the 72, 139 and 149 m2/g, respectively.
F.H. Alhassan et al. / Powder Technology 253 (2014) 809–813 813

3.5. Transmission electron microscope (TEM) Acknowledgments

Fig. 5 confirms nanosized crystal shape in zirconia and iron– The authors acknowledge the financial support of the International
manganese doped sulfated zirconia-4 catalysts as characterized by TEM. Graduate Research Fellowship (IGRF) from the Universiti Putra
The images of the crystals depicted in Fig. 5(a, b), were uniformly distrib- Malaysia.
uted at the nanometer scale having a clear shape and different size, being
approximately 50 nm with individual particles measuring 25–20.5 and References
11–18 nm for zirconia and FMSZ-4, respectively, which was very well
[1] G.A.H. Mekhemer, Surface characterization of zirconia, holmium oxide/zirconia and
in line with XRD data. The reproducibility of the method was established sulfated zirconia catalysts, Colloids Surf. A Physicochem. Eng. Asp. 274 (2006)
through the triplicate preparations and triplicate analyses. 211–218.
[2] A. Clearfield, G. Serrette, A. Khazi-Syed, Nature of hydrous zirconia and sulfated hy-
drous zirconia, Catal. Today 20 (1994) 295–312.
3.6. TPD-NH3 [3] T. Jyothi, K. Sreekumar, M. Talawar, S. Mirajkar, B. Rao, S. Sugunan, Physico-chemical
characteristic of sulfated mixed oxides of Sn with some rare earth elements, Pol. J.
Total number of acid sites and their relative strength for the catalysts Chem. 74 (2000) 801–812.
[4] M. Hino, K. Arata, Reaction of butane to isobutane catalyzed by iron oxide treated
under investigation can be measured by means of a total amount of NH3 with sulfate ion. Solid superacid catalyst, Chem. Lett. 8 (1979) 1259–1260.
desorbed as listed in Table 2. From these results one can conclude that [5] C.-Y. Hsu, C. Heimbuch, C. Armes, B. Gates, A highly active solid superacid catalyst
the Z sample has a low acid site density of 630.4 μmol/g. The addition for n-butane isomerization: a sulfated oxide containing iron, manganese and zirco-
nium, J. Chem. Soc. Chem. Commun. (1992) 1645–1646.
of Fe–Mn dopants and sulfate to ZrO2 can increase the acidic strength [6] C.-H. Lin, C.-Y. Hsu, Detection of superacidity on solid superacids; a new approach, J.
and create very strong acid sites on the surface of ZrO2 to become Chem. Soc. Chem. Commun. (1992) 1479–1480.
4318.4 μmol/g, 5458.0 μmol/g and 7544.0 μmol/g for the samples [7] F.C. Jentoft, A. Hahn, J. Kröhnert, G. Lorenz, R.E. Jentoft, T. Ressler, U. Wild, R. Schlögl,
C. Häßner, K. Köhler, Incorporation of manganese and iron into the zirconia lattice in
FMSZ-4, FMSZ-10 and FMSZ-16, respectively. doped sulfated zirconia catalysts, J. Catal. 224 (2004) 124–137.
[8] H.P. Klug, L.E. Alexander, X-ray diffraction procedures: for polycrystalline and amor-
3.7. X-ray photoelectron spectra phous materials, in: Harold P. Klug, Leroy E. Alexander (Eds.), 2nd edition,
Wiley-VCH, ISBN: 0-471-49369-4, May 1 1974, p. 992.
[9] A. Patterson, The Scherrer formula for X-ray particle size determination, Phys. Rev.
Both zirconia (Z) and FMSZ-4 samples showed a photoelectron 56 (1939) 978.
emission at 181.5 eV characteristic of Zr 3d5/2 and one at 530.0 eV [10] Z. Cheng, H. Yang, L. Yu, Y. Cui, S. Feng, Preparation and magnetic properties of
Y3Fe5O12 nanoparticles doped with the gadolinium oxide, J. Magn. Magn. Mater.
characteristic of O1s. These photoelectron peaks taken together, were 302 (2006) 259–262.
assigned to ZrO2 [28]. The peaks in Fig. 6(a, e) are assuming the presence [11] R.C. Garvie, The occurrence of metastable tetragonal zirconia as a crystallite size ef-
of two Zr(IV) components. The lower binding energy component is in fect, J. Phys. Chem. 69 (1965) 1238–1243.
[12] T. Sato, The thermal decomposition of zirconium oxyhydroxide, J. Therm. Anal.
the range of Zr(IV) in ZrO2, whereas the higher binding energy compo- Calorim. 69 (2002) 255–265.
nent (183.3, 185.5EV) corresponds to the formation of a Zr(IV) species [13] R. Srinivasan, D. Taulbee, B.H. Davis, The effect of sulfate on the crystal structure of
bound to a more electron-attractive species. The Zr 3d peak (Fig. 6(e)) zirconia, Catal. Lett. 9 (1991) 1–7.
[14] K. Wieczorek-Ciurowa, A. Kozak, The thermal decomposition of Fe (NO3)3·9H2O, J.
confirms no appreciable modifications with respect to the peak of the
Therm. Anal. Calorim. 58 (1999) 647–651.
undoped oxide (Fig. 6(a)). The total absence of effects on the shape of [15] H.A. Khalaf, Textural properties of sulfated iron hydroxide promoted with alumi-
the Zr 3d peak is very interesting and indicative of the different strength num, Monatsh. Chem. 140 (2009) 669–674.
and nature of the interactions involved between sulfates and zirconia [16] B.M. Reddy, P.M. Sreekanth, P. Lakshmanan, A. Khan, Synthesis, characterization and
activity study of SO2− 4 /CexZr1 − xO2 solid superacid catalyst, J. Mol. Catal. A Chem.
for (NH4)2SO4 impregnation at these concentrations [29]. The S2p emis- 244 (2006) 1–7.
sion was located at about 168.6 eV (Fig. 6(c)) and suggests the presence [17] M. Bi, H. Li, W.-P. Pan, W.G. Lloyd, B.H. Davis, Prepr. Pap. Am. Chem. Soc. Div Fuel
of S+6 in agreement with that relative to sulfur in sulfates and with pre- Chem. 41 (1996) 77–81.
[18] B.H. Davis, R.A. Keogh, S. Alerasool, D.J. Zalewski, D.E. Day, P.K. Doolin, Infrared study
vious results relative to sulfated zirconia samples [29]. of pyridine adsorbed on unpromoted and promoted sulfated zirconia, J. Catal. 183
Fig. 6(f) shows the XP spectrum of FMSZ-4 in the Fe 2p region. The (1999) 45–52.
Fe 2p3/2 and Fe 2p1/2 emissions, located at about 711.3 and 724.0 eV, re- [19] D.A. Ward, E.I. Ko, Sol–gel synthesis of zirconia supports: important properties for
generating n-butane isomerization activity upon sulfate promotion, J. Catal. 157
spectively, support the existence of iron as an oxide at the surface of the (1995) 321–333.
catalyst. This spectrum is a clear evidence of the presence of Fe+3and is [20] T. Mishra, K. Parida, Effect of sulfate on the surface and catalytic properties of
practically identical to the spectra that have been reported for Fe2O3 iron–chromium mixed oxide pillared clay, J. Colloid Interface Sci. 301 (2006)
554–559.
[30]. Mn peaks could be detected only with a rather poor noisy signal. [21] A. Corma, V. Fornes, M. Juan-Rajadell, J. Nieto, Influence of preparation conditions on
Fig. 6(g) displays the XPS spectrum of FMSZ-4 in the Mn 2p region. The the structure and catalytic properties of SO2− 4 /ZrO2 superacid catalysts, Appl. Catal.
Mn 2p3/2 and Mn 2p1/2 photoelectron lines were observed at approxi- A Gen. 116 (1994) 151–163.
[22] A. Khder, E. El-Sharkawy, S. El-Hakam, A. Ahmed, Surface characterization and cat-
mately 642.5 and 653.4 eV, respectively, which confirms the presence
alytic activity of sulfated tin oxide catalyst, Catal. Commun. 9 (2008) 769–777.
of manganese oxide at the surface of the catalyst. [23] A. Jogalekar, R. Jaiswal, R. Jayaram, Activity of modified SnO2 catalysts for acid‐
catalysed reactions, J. Chem. Technol. Biotechnol. 71 (1998) 234–240.
[24] G.A. Mekhemer, H.A. Khalaf, S.A. Mansour, A.K. Nohman, Sulfated alumina cata-
4. Conclusions lysts: consequences of sulfate content and source, Monatsh. Chem. 136 (2005)
2007–2016.
The obtained results show that the ZrO2 gel needs at least one mole- [25] A.F. Bedilo, K.J. Klabunde, Synthesis of catalytically active sulfated zirconia aerogels,
J. Catal. 176 (1998) 448–458.
cule of water to give the formula ZrO2·H2O and that the addition of ferric [26] K. Arata, Solid superacids, Adv. Catal. 37 (1990) 165–211.
and manganese as promoter and sulfate imparts a special stabilization to [27] M. Scurrell, Conversion of methane–ethylene mixtures over sulphate-treated zirco-
the catalytically active tetragonal phase of zirconia, which decreases the nia catalysts, Appl. Catal. 34 (1987) 109–117.
[28] M. Scheithauer, E. Bosch, U.A. Schubert, H. Knözinger, T.-K. Cheung, F.C. Jentoft, B.C.
crystallite size and consequently, increases the specific surface area. The
Gates, B. Tesche, Spectroscopic and microscopic characterization of iron-and/or
specific surface area is increased by increasing the loading of sulfate. manganese-promoted sulfated zirconia, J. Catal. 177 (1998) 137–146.
Moreover, the incorporation of promoters and sulfate onto ZrO2 can in- [29] S. Ardizzone, C. Bianchi, XPS characterization of sulphated zirconia catalysts: the role
crease its acidity and create a strong acidic site. The XPS results indicate of iron, Surf. Interface Anal. 30 (2000) 77–80.
[30] B.M. Weckhuysen, D. Wang, M.P. Rosynek, J.H. Lunsford, Conversion of methane to
the presence of the sulfate, Fe2O3, and MnO particles on the surface of benzene over transition metal ion ZSM-5 zeolites: II. Catalyst characterization by
zirconia catalyst. X-ray photoelectron spectroscopy, J. Catal. 175 (1998) 347–351.

You might also like