Acessibilidade / Reportar erro

Thermal behaviour of titanium in host matrix clay

Abstract

Titanium in hosted by a clay matrix has important applications in a variety of catalysts. However, there is insufficient information about of textural and structural characteristics of the titanium in matrix clay at room temperature and after thermal treatment. In this study, titanium in host matrix clay was obtained by treating montmorillonite clay minerals with different OH-Ti solutions. The titanium-solutions containing partially hydrolyzed titanium were obtained using a TiCl4 solution with and without the addition of glycerol. The amount of Ti added to the solid was 20 mmol Ti per gram of clay. Thermal treatments were carried out at 500 and 1,000 °C in an air atmosphere. The solids were characterized by X ray diffraction, thermal analyses, infrared spectroscopy, N2 adsorption-desorption isotherms. The products presented different structural and textural characteristics and, after calcination at 1,000 °C, anatase and rutile in different proportions were present in the samples.

titanium; clay; thermal treatment; montmorillonite


REGULAR ARTICLES

Thermal behaviour of titanium in host matrix clay

Cristina Volzone * * e-mail: volzcris@netverk.com.ar ; Ana María Cesio

Centro de Tecnología de Recursos Minerales y Cerámica (CETMIC)-CIC/ CONICET, La Plata, C.C. 49, Cno. Centenario y 506, (1897) M.B. Gonnet, Prov. Buenos Aires, Argentina

ABSTRACT

Titanium in hosted by a clay matrix has important applications in a variety of catalysts. However, there is insufficient information about of textural and structural characteristics of the titanium in matrix clay at room temperature and after thermal treatment. In this study, titanium in host matrix clay was obtained by treating montmorillonite clay minerals with different OH-Ti solutions. The titanium-solutions containing partially hydrolyzed titanium were obtained using a TiCl4 solution with and without the addition of glycerol. The amount of Ti added to the solid was 20 mmol Ti per gram of clay. Thermal treatments were carried out at 500 and 1,000 °C in an air atmosphere. The solids were characterized by X ray diffraction, thermal analyses, infrared spectroscopy, N2 adsorption-desorption isotherms. The products presented different structural and textural characteristics and, after calcination at 1,000 °C, anatase and rutile in different proportions were present in the samples.

Keywords: titanium, clay, thermal treatment, montmorillonite

1. Introduction

Titanium, as dioxide, presents interesting properties because it is an inert and nontoxic solid. It may be used as white pigment, for water and air purification1, remediation, in catalysts, as an adsorbent, as compounds in bioceramic materials, etc. Different methods are applied to improve the surface area, size particles, porosity, grain size, and morphology properties.

The titanium in a host matrix solid can be used in appropriate applications: for example, photocatalytic properties can be improved by using TiO2[2] as a pillar in clay3, and visible light photocatalytic activity4 could be improved by preparing mesoporous carbon-doped TiO2.

Studies related to titanium-clay were started around 20 year ago by Sterte5, thereafter solids were analyzed for catalytic uses: butanol hydroxylation6, hydroxylation of phenol7, dehydration of pentanol8; decomposition of propanol9, selective catalytic reduction of NO by Ti-pillared clays10,11.

More recently, clay with a titanium compound can be use as a catalyst to produce polyester (for example, polyethylene terephthalate, PET), where the catalyst is used to catalyze the esterification, transesterificaton and/or polycondensation12.

The titanium-clays may be obtained by first using cation exchange of polynuclear hydroxy-titanium species to introduce the titanium ions between the interlayers of the clay. Then, upon heating, a stable metal oxide cluster of molecular dimensions is formed5,13,14.

The clays are fine grained phyllosilicate minerals constituted by sheets of aluminum in octahedral coordination and sheets of silica in tetrahedral coordination. The montmorillonite is one of the smectite clay classification species constituted by 2:1 layer type (one sheet of aluminum in octahedral coordination is between two sheets of silica in tetrahedral coordination, triple-sheet-layer). Isomorphic substitution results in negative charges on the three-sheet-layer. These charges are balanced by adsorbed cations (Na+, Ca2+, K+, Mg2+) in interlayer position, which are easily exchangeable15. Clays are solids which structure may undergo important changes after thermal treatment, mainly from 500 °C, due to removing of the OH- constituent of the clay from structure16.

The aim of the manuscript was to analyze the structural and textural changes of the titanium in host matrix clay at room temperature and after thermal treatment.

2. Materials and Methods

A montmorillonite supplied by Georgia Kaolin Co. was used in this study and referred to as E. The E sample is a clay mineral with high montmorillonite as clay composition. Quartz and α-cristobalite impurities in low proportions were detected in the sample. The montmorillonite has a cation-exchange capacity (CEC) of 110 meq/100g and the following structural formula: (Na0.32 K0.12 Ca0.25) (Si7.95 Al0.05)(Al2.95 Fe0.18 Ti0.03 Mg0.80) O20 (OH)4 was obtained by using a <2 μm size fraction17.

2.1. Titanium-solutions

The titanium-solutions (containing partially hydrolyzed titanium cations) were obtained by different methods: a) using a TiCl4 solution and 6.0 M HCl5, where the final Ti concentration was 0.80 M, and the HCl/Ti ratio was 1.22. The solution was aged for 3 hours at 20 °C before use (TiT-solution); b) a TiCl4 solution was slowly added to absolute alcohol (ethanol), this solution was then added to glycerol (50% v.v-1 in water) following Lin et al.18 and Bovey et al.8. This was named as TiG solution.

2.2. Preparation of titanium-clays

Natural montmorillonite clay (E) and acid treated montmorillonite clay (AE) were used as clays to prepare OH-Ti-clay. The acid activated montmorillonite (AE) was prepared by treating the montmorillonite with a 6 NH2SO4 solution at boiling temperature for 30 minutes with a solid/acid ratio of 5%w.v-1.

The OH-Ti-clays were obtained by adding OH-Ti-solution addition drop wise to a stirred suspension of montmorillonite clay in water. After 4 hours in continuous stirring, the solid was separated by centrifugation, and washed several times with distilled water until being chloride free. The amount of Ti added to the solid was 20 mmol Ti per gram of clay. The solids were named TiT-E, TiG-E and TiG-AE.

The thermal treatment at 500 and 1000 °C of the Ti-clays were obtained in an electrically heated furnace by using a Pt crucible in air for one hour (TiT-E5, TiG-E5 and TiG-AE5; and TiT-E10, TiG-E10 and TiG-AE10).

2.3. Characterization methods

Energy Dispersive Analysis of X ray (EDAX) was obtained by EDAX DX PRIME and Philips® SEM505 microscopy. X ray diffractions (XRD) of the samples were obtained with a Philips® 3020 Goniometer with PW 3710 Controller using Cu Kα radiation at 40 kV and 20 mA and Ni filter. The solids were conditioned at 55% HR.

The infrared spectra (IR) were obtained by using Spectrum One Perkin-Elmer® equipment. The samples were diluted in KBr and compacted in the form of thin pellets.

Diagrams of differential thermal analyses (DTA) were carried out with a Netzsch® 409 apparatus using α-Al2O3 as a reference at a heating rate of 10 °C/min. All samples were carried to similar relative humidity previous to the analysis.

The adsorption isotherms were obtained at liquid nitrogen temperature with N2 in Micromeritics® ASAP 2020 equipment. Samples were outgassed at 100 °C and 10-3 torr before nitrogen treatment for 12 hours. The BET surface area (SBET) was calculated from the first part of the isotherms (P/Po < 0.3) using 0.162 nm2 as the molecular area of adsorbed nitrogen molecule. The t-plot was made according to Lippens and De Boer19.

3. Results and Discussion

The chemical analyses of the samples by EDAX are shown in Table 1. The chemical analysis of the E sample corresponds to a typical montmorillonite clay in which, in this type of analysis, the water content was not taken into account. The Si4+, Al3+, Fe2/3+, Ti4+ are elements situated in tetrahedral (Si4+, Al3+) and octahedral (Al3+, Fe2/3+, Ti4+) sheets of the clay. The Mg2+ can be situated in octahedral and/or interlamellar position, whereas Na+, K+ and Ca2+ are typical cations in interlamellar position.

After treatment of the montmorilllonite with OH-Ti solutions, the interlayer Ca2+, Na+ and part of K+ were exchanged by OH-Ti cation species. Einaga20 proposed the existence of the polynuclear cations species ((TiO)8(OH)12)+4 in solutions prepared in similar conditions. The amount of titanium in montmorillonites after OH-Ti-solution treatments expressed as oxide were 15.88, 13.01 and 9.22% for TIT-E, TIG-E and TIG-AE samples, respectively. The reduction of aluminum, iron, and magnesium of the montmorillonite structure after OH-Ti treatment can be attributed to pH conditions mainly due to treatment with TiCl4 solutions (pH: 4). The acid treatment on the clay reduces the cation exchange capacity of montmorillonite21 and also may release the elements from the octahedral sheet of the clay, depending on acid concentration, treatment time, etc.22. A reduced amount of Ti was observed in TiG-AE (9.22% TiO2) compared to TiG-E (13.01%) where TiG-AE was prepared from acid clay (AE) and TiG-E from natural clay (E). The increase of the Si/(Al + Fe) ratio of the clays from 2.50 to 4.02-4.19, Table 1, is indicative of the released cations from octahedral sheets due to treatment with OH-Ti solutions, attributed to the low pH of the solution used for treating montmorillonites. In the same Table the amount of Al and Fe released due to OH-Ti treatment were in the range of 42-51 and 3.3-15 mol, respectively.

Figure 1 shows the differential thermal analysis diagrams for montmorillonite before and after titanium solution treatments. The temperature of the first endothermic peak in the range of 20-200 °C corresponds to the loss of interlamellar water (free and hydrated cations) and the positions of these peaks are related to the type of exchangeable cation, Na+, K+ and Ca2+, being present in this case17. The temperature of the first peak of the original sample (E), 154 °C, shifted to lower after titanium intercalation: 103-133; 152 and -114-148 °C for TiT-E; TiG-E and TiG-AE treated motmorillonites, respectively, showing a broad peak with doublet peaks in TiT-E and TiG-AE samples. The minor intensities of the peaks for montmorillonite after treatment with a titanium TiCl4 - solution (TiT-E) and for acid clays treated with TiCl4 - glycerol (TiG-AE) may be attributed to a slight modification of the structural sample due to the acid condition of the treatment. The second endothermic peak at 679 °C, corresponding to a loss of hydroxyl groups, confirmed the presence of montmorillonite as the smectite specie of clay in sample E23. The presence of titanium in interlamellar spacing shifted the peak at a lower temperature (628, 673 and 646 °C for TiT-E; TiG-E and TiG-AE, respectively). This kind of peak may depend on the perfection of the stacking of the layers influenced by the presence of titanium.


The montmorillonite treated with Ti, using glycerol in the preparation (TiG-E), showed an important exothermic peak which appeared at around 294 °C with a shoulder at 455 °C indicating the loss of organic substance due to the presence of glycerol. The sample -TiG-AE presented similar behavior with less pronounced peaks at minor temperature attributed to an altered structure due to the acid activated montmorillonite (AE) used for preparing TiG-AE.

The endothermic-exothermic peaks of the montmorillonites at around 850-930 °C correspond to the destruction of the lattice and its recrystallization into new phases16. The third endothermic peak, situated at 854 °C of sample E, shifted at higher temperature after Ti treatments, Figure 1, showing that TiT-E, TiG-E and TiG-AE samples have a higher thermal resistance to structural changes for to originate new phases.

Figure 2 shows the thermal weight loss of the samples. The total loss for natural montmorillonite, E, was 13.87% (Table 2), corresponding to two marked zones of weight loss, up to 200 °C (10%) and the second zone from 200 to 680 °C (3.87%), corresponding to water intercalated and the OH- structural loss, respectively, typical of montmorillonite clays16. The total weight losses of OH-Ti-montmorillonites were different depending on the Ti treatment, nevertheless, in all treated samples the loss was continuous in all temperature ranges up to around 700 °C, Figure 2. The total loss of TiT-E was 15.20%. TiG-E and TiG-AE showed three differentiate losses with higher amount for the first (17.35 and 13.95%, respectively). The different values are related to the presence of organic or previous acid-treated montmorillonite (AE). In Table 2 the values of the loss at different temperature ranges are shown. The higher loss of TiT-E regarding E up to 200 °C may be due to the loss of the OH- water of the Ti species intercalated in interlayer spacing. The similar loss of weight for TiG-E and TiG-AE (8.6 and 8.8%, respectively) in respect of TiT-E (11.2 %) suggests different Ti species formation with less OH-contribution. Assuming that weight loss up to 200 °C corresponds to hydroxyl of OH-Ti species situated in interlayer position, and taking into account the amount of titanium retained (Table 1), it was possible to calculate the OH/Ti for each sample. The OH/Ti values: 3.08, 3.00 and 4.24 of TiT-E; TiG-E and TiG-AE, respectively, are indicative that the species in interlayer for the two first samples were similar and higher for TiG-AE, indicating that the last sample has more a higher rate of OH in titanium species. The TiG-E sample needs a slightly higher temperature (152 °C) in respect of the TIG-AE (148 °C) for to lose the same value at low temperature (Figure 2). Finally, according to Table 2, the different total loss of each sample is related to different loss contributions.


The total loss for E sample was 13.87% and corresponds to the losses of: the free/hydration-cations water, and OH- of the structure. In TiT-E the increase in loss weight (15,20%) is due to the loss of OH- of Ti species. Nevertheless, the total loss for TiG-E is higher due to the contribution of organic compounds that contribute with 6.48%. The lower value for TiG-AE may be due to a less organic retention because the content of structural montmorillonite was reduced for a previous acid treatment. The less total weight for TiG-AE may be also attributed to the loss of structural montmorillonite. The loss of weight up to about 500 °C is continuous for the three Ti-montmorillonites, probably due to the dehydroxylation of OH- groups associated with the interlayer titanium complex (Figure 2).

The clays treated with Ti solutions showed a higher basal spacing (19-22 Å) than the one in the original sample (15 Å), Table 3, due to the replacement of Na+, K+, Ca2+ cations in interlayer positions by OH-Ti species. The different values may be attributed to the arrangement of OH-Ti-species present in each hydrolyzed Ti-solution as a different OH/Ti ratio was calculated in the above paragraph.

The heating of montmorillonite, E, at 500 °C originated an interlayer spacing of 9.3Å, corresponding to a collapsed distance between 2:1 layer. Nevertheless, the interlayer spacing after calcinations of the Ti-montmorillonite reached values from 16 to 19Å (Table 3) showing that, at 500 °C, the 2:1 layer of the clay presents separated interlayer, in a similar form to the pilarization processed6.

Figure 3a and 3b show N2 adsorption-desorption isotherms of the Ti-samples before and after treatment at 500 °C, respectively. The shape of the adsorption-desorption isotherm of the raw sample, E, was H3 type and corresponded to the system by slit-shaped pore according to Gregg and Sing24, Figure 3a. The adsorption-desorption isotherms of intercalated titanium montmorillonites maintain a form similar to the one of the raw sample, indicating that the pore preserves a similar form, nevertheless, the higher volume adsorbed at low P/Po of the Ti-samples (TiT-E, TiG-E and TiG-AE) indicated a contribution of higher micropore volume. Figure 3b shows the prepared Ti-montmorillonites after 500 °C. The TiT-E5, TiG-E5 and TiG-AE5 isotherms showed a higher total volume adsorbed than without thermal treatment (Figure 3a, Table 4). The isotherms of samples were similar, although the TiG-AE5 showed a higher volume adsorbed in the intermediate zone of the P/Po, indicating a contribution with different mesopore size (Figure 3b). The textural parameters resulting from the analysis of the isotherms are given in Table 4. The BET surface increased after Ti treatment from 46 m2.g-1, original E, to 122, 63 and 102 m2.g-1 for TiT-E, TiG-A and TiG-AE, respectively. The thermal treatment at 500 °C preserved the BET surface for TiT-E (120 m2.g-1) and an increase was observed for montmorillonites treated with Ti and glycerol, TiG-E5 and TiG-AE5 samples that reached 139 and 193 m2.g-1, respectively.



The total volume of the montmorillonite, 0.117 cm3.g-1 (at P/Po = 0.983), in general, also increased after treatment (Table 4). After thermal treatment at 500 °C of the Ti-montmorillonites values duplicated from 0.083 to 0.159 cm3.g-1 for TiG-E and from 0.122 to 0.237 cm3.g-1 for TiG-AE, respectively, except for TiT-E that only increased a third, from 0.150 to 0.194 cm3.g-1. The micropore volumes (Vmic) obtained by the high-pressure branch extrapolated to the adsorption axis in t-plots24 are shown in Table 4. These values increased between 3-6 times after titanium treatments from original montmorillonite. The percentage in micropore volume to total volume contributions were the same (15%) for Ti-montmorillonites. This percentage was maintained after 500 °C thermal treatment for TiT-E. Nevertheless, it was twofold and threefold in samples treated with glycerol, reaching 29 and 36%. Then, the increase in total volume was due to micropore contribution.

The infrared absorption band of the montmorillonite at 3,624 cm-1 is attributed to the stretching vibrations of the OH- groups of Mg-Al-OH and, at 3,454 cm-1, it is related to the presence of H-O-H. The amount of adsorbed water in clays is related to the deformation vibrations of the H-O-H group (1661 cm-1). The bands at 1042 and 798 cm-1 are attributed to Si-O stretching vibrations. The bands at 518 and 465 cm-1 correspond to deformation vibrations of Si-O-AlIV and Si-O-Si, respectively. The main modification in the IR spectra after titanium treatments can be seen at two different zones as shown in Figure 4. The band at 3,624 cm-1 was reduced in intensity and may be attributed to the release of octahedral cations from the structure and to the replacement of interlayers cations by titanium in interlayer spacing, as it was confirmed by chemical analyses (Table 1). The band at 3,454 cm-1 attributed H-O-H (Figure 4) was shifted to less wavelength, 3,418, 3,429 cm-1, indicating a high bond of OH- of the titanium species in interlayer spacing.


The slight reduction of the intensity band at 518 cm-1 in respect of the intensity band at 465 cm-1 (Figure 4) after titanium treatment was indicative that the bond of Si-O-AlIV was modified after Ti treatment, which is attributed to the effect of the pH treatment.

The samples heated at 1,000 °C were analyzed by X ray diffraction and infrared.

The montmorillonite clay is destroyed at 800-900 °C and, generally at 1,000 °C, new phases are starting to form depending on the initial chemical composition16. In Table 5 the phases found at 1,000 °C of all samples are shown. The original montmorillonite, E, developed μ-cordierite (2MgO.2Al2O3.5SiO2 ), β-quartz (SiO2), β-cristobalite (SiO2) and anorthite (Ca.Al2Si2O8) after heating at 1,000 °C, similar as it was found in a previous paper26. The β-quartz (SiO2), β-cristobalite (SiO2) were preserved in all Ti-samples. Nevertheless, due to the release of interlayer and octahedral cations from the structure after titanium treatment (Table 1), other phases were found. The presence of titanium as anatase in different proportion in the samples and rutile, mainly in TiT-E, are shown in Figure 5, and Table 5. Anatase, rutile and brookite are three phase structures of titanium dioxide and, as the anatase phase shows a higher photocatalytic activity than rutile, the sample TiT-E10 may be considered as better than the other samples obtained.


The IR spectra of samples heated at 1,000 °C are shown in Figure 6. Infrared spectra may be used to identify an oxide and to distinguish between different polymorphism forms25.


The presence of anatase in TiT-E10, TIG-E10 and TiG-AE10 could be detected at near 695 cm-1[25] and a smaller band at 400 cm-1, attributed to rutile in TiT-E10 sample. The presence of chlorine in the TIT-E sample transforms to rutile at high temperature27,28. At 1,000 °C the Si-O stretching and Si-O bending bands shifted to 1,090 cm-1 and 470 cm-1, respectively, implying changes in the structure of the tetrahedral sheets. The band at 790 cm-1, attributed to quartz (β-quartz (SiO2), β-cristobalite (SiO2)), remained. The presence of a developed intensity broad band centered at 3,450 cm-1 of the Ti-samples at a high temperature treatment is attributed to silicate species, M2+(SiO4=), Farmer25, which is indicative of a new substance, in concordance with an X ray diffraction (Table 5).

4. Conclusions

Treatment of montmorillonite clays with two different solutions containing the same amount of titanium addition resulted in a different uptake of titanium, which can be attributed mainly to the type of hydrolyzed OH-Ti solutions used for the treatments and the way OH-Ti species interact with clay.

The treatment of the clay with hydrolyzed OH-Ti solutions resulted in an increase in the interlayer spacing of the clay, and a higher surface area and the micropore volume on the solids. This may be due to the removal of exchangeable interlayer cations and part of structural cation of the clay. The thermal treatment at 500 °C improved the textural properties (surface area, pore volume) and the interlayer spacings slowly decreased but did not collapse (as it is typical in original montmorillonite clay). At 1,000 °C the structure of the clay was transformed to phases of calcined clay and the anatase was the principal compound found in the solid. Nevertheless, the rutile phase was also present in the product obtained without glycerol in the preparation of OH-Ti solutions.

Received: January 14, 2009;

Revised: April 24, 2009

  • 1. Mills A, Hunte SL. An overview of semiconductor photocatalysis. Journal of Photochemistry and Photobiology A: Chemistry 1997; 108(1):1-35.
  • 2. Jiaguo Yu, Guohong W, Bei C, Minghua Z. Effects of hydrothermal temperature and time on the photocatalytic activity and microstructures of bimodal mesoporous TiO2 powders. Applied Catalysis B: Environmental 2007; 69(3-4):171-180.
  • 3. Yoneyama H, Haga S, Yamanaka S. Photocatalytic activities of microcrystalline titania incorporated in sheet silicates of clay. Journal of Physical Chemistry 1989; 93(12):4833-4837.
  • 4. Ren W, Ai Z, Jia F, Zhang L, Fan X, Zou Z. Low temperature preparation and visible light photocatalytic activity of mesoporous carbon-doped crystalline TiO2 Applied Catalysis B: Environmental 2007; 69(3-4):138-144.
  • 5. Sterte J. Synthesis and properties of titanium oxide cross-linked montmorillonite. Clays and Clay Minerals 1986; 34(6):658-664.
  • 6. Del Castillo HL, Grange P. Preparation and catalytic activity of titanium pillared montmorillonite. Appied Catalysis A: General 1993; 103(11):23-34.
  • 7. Del Castillo HL, Gil A, Grange P. Hydroxilation of phenol on titanium pillared montmorillonite. Clays and Clay Minerals 1996; 44(5):706-709.
  • 8. Bovey J, Kooli F, Jones W. Preparation and characterization of Ti-Pillared acid-activated clay catalysts. Clay Minerals 1996; 31(4):501-506.
  • 9. Kooli F, Bovey J, Jones W. Dependence of the properties of Titanium-pillared clays on the host matrix: a comparison of montmorillonite, saponite and rectorite pillared materials. Journal of Materials Chemistry 1997; 7(1):153-158.
  • 10. Yang RT, Baksh MSA. Pillared clays as a new class of sorbents for gas separation. AIChE Journal 1991; 37(5):679-686.
  • 11. Bahranowski K, Jonas J, Machej T, Serwicka EM, Vartikian LA. Vanadium doped titania-pillared montmorillonite clays as a catalyst for selective catalytic reduction of NO by ammonia. Clay Minerals 1997; 32(4):665-672.
  • 12. Putzig DE. Composition comprising titanium and clays and process therewith. Patent WO/2006/017568. Disponível em: <http://www.wipo.int>
  • 13. Bernier A, Admaiai LF, Grange P. Synthesis and characterization of titanium pillared clays. Influence of the temperature of preparation. Applied Catalysis 1991; 77(2):269-281.
  • 14. Vicente MA, Bañares-Muñoz MA, Toranzo R, Gandia LM, Gil A. Influence of the Ti precursor on the properties of Ti-pillared smectites. Clay Minerals 2001; 36(1):125-138.
  • 15. MacEwan DMC. Montmorillonite minerals. In: Brown G, editor. The X-ray identification and crystal structures of clay minerals London: Mineralogical Society; 1961. p. 143-207.
  • 16. Mackenzie RC. Differential thermal analysis. In: Mackenzie RE, editor. Fundamental aspects London, New York: Academic Press; 1970.
  • 17. Volzone C. Hydroxy-Chromim smectite: influence of Cr(III) added. Clays and Clay Minerals 1995; 43(3):377-382.
  • 18. Lin JT, Jong SJ, Cheng S. A new method for preparing microporous titanium pillared clays. Microporous Materials 1993; 1:287-290.
  • 19. Lippens BC, de Boer JH. Studies on pore systems in catalysts. Journal of Catalysis 1965; 4:319-323.
  • 20. Einaga H. Hydrolysis of titanium (IV) in aqueous (Na,H)Cl solution. Journal of the Chemical Society, Dalton Transactions 1979; 12:1917-1919.
  • 21. Onal M. Swelling and cation exchange capacity relationship for the samples obtained from a bentonite by acid activation and heat treatments. Applied Clay Science 2007; 37(1-2):74-80.
  • 22. Volzone C, Porto López JM, Pereira E. Activación ácida de un material esmectítico. Revista Latinoamericana de Ingeniería Química y Química Aplicada 1986; 16:205-215.
  • 23. Grim RE, Kulbicki G. Montmorillonite: high temperature reactions and classification. American Mineralogist 1961; 46:1329-1369.
  • 24. Gregg SJ, Sing KSW. Adsorption, surface area and porosity. 2 ed. London: Academic Press Inc.; 1991.
  • 25. Farmer VC. The infrared spectra of mineral London: Ed. Mineralogical Society; 1974.
  • 26. Volzone C, Cesio AM. Structural modifications of OH-Cr-smectites after thermal treatment up to 1000 °C. Materials Chemistry and Physics 1997; 48(3):216-219.
  • 27. Gamboa JA, Pasquevich DM. Effect of chlorine atmosphere on the anatase-rutile transformation. Journal of the American Ceramic Society, 1992; 75(11):2934-38.
  • 28. Nabivanets BI, Kudritskuga NL. Study on the polymerizatio of titanium (IV) in hydrochloric acid solutions. Russian Journal of Inorganic Chemistry 1967; 12:612-620.
  • *
    e-mail:
  • Publication Dates

    • Publication in this collection
      24 July 2009
    • Date of issue
      June 2009

    History

    • Accepted
      24 Apr 2009
    • Received
      14 Jan 2009
    ABM, ABC, ABPol UFSCar - Dep. de Engenharia de Materiais, Rod. Washington Luiz, km 235, 13565-905 - São Carlos - SP- Brasil. Tel (55 16) 3351-9487 - São Carlos - SP - Brazil
    E-mail: pessan@ufscar.br