SINTERING MECHANISMS OF PORCELAIN STONEWARE TILES

Similar documents
Effect of soda-lime glass on sintering and technological properties of porcelain stoneware tiles

Ceramic Processing Research

STUDY OF POROSITY IN PORCELAIN TILE BODIES

Fabrication and Characterization of Anorthite-Based Porcelain using Malaysian Mineral Resources

Use of glass waste as a raw material in porcelain stoneware tile mixtures

Ceramic Processing Research

MECHANICAL PROPERTIES AND MICROSTRUCTURE OF FAST FIRED CLINKER TILES BASED ON WIERZBKA I RAW MATERIAL

AN INSIGHT INTO THE PYROPLASTICITY OF PORCELAIN STONEWARE TILES

This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and

THE EFFECT OF MAGNESITE ADDITION ON THE PROPERTIES AND PRODUCTION PROCESS OF LOW POROSITY PORCELAIN TYPE CERAMICS

Ceramic Processing Research

EFFECT OF THE Si POWDER ADDITIONS ON THE PROPERTIES OF SiC COMPOSITES

Waste Colored Glasses as Sintering Aid in Ceramic Tiles Production

UTILIZATION OF BASALT FIBERS AS A RAW MATERIAL FOR CLAY CERAMIC PRODUCTION

ScienceDirect. Characterization of sugarcane bagasse ash waste for Its Use in Ceramic Floor Tile

FIXED PROSTHODONTICS Page 1 Lecture: "Dental Porcelains" REVIEW OF CERAMICS AND PORCELAINS: A. Review of Definitions and Terminology:

Design of ceramic microstructures based on waste materials

Characterization of Sintered Ceramic Tiles Produced from Steel Slag

Recycling of Steel Sludge into Red Ceramic

Effect of chemical composition and alumina content on structure and properties of ceramic insulators

Chapter 5: REACTION SINTERING of diamond and silicon powders

Polish Academy of Sciences Institute of Fundamental Technological Research ( Warszawa, Świętokrzyska 21, Poland)

Microstructural Evolution of W-Ni-Fe During Liquid Phase Sintering A Quenching Study

Densification and grain growth of TiO 2 -doped ZnO

PORCELAIN TILES BY THE DRY ROUTE

DEVELOPMENT OF DAN KWIAN CLAY FOR USE IN CERAMIC TILE INDUSTRY

Tribological Behavior of Whiteware with Different Transparent Glazes


MSE 351 Engineering Ceramics I

Mullite Preparation from Kaolin Residue. Química - Aprígio Veloso, 820 CEP Campina Grande, PB, Brasil.

Effect of Materials Design on Properties of Porcelain Insulators

THE ROLE OF CEMENT DUST IN BASALT-DE-ALUMINATED KAOLIN BRICKS

Effects of CaO B 2 O 3 Glass on Sintering and Microwave Properties of Cordierite Ceramics for Low-Temperature Cofired

PHYSICAL PROPERTIES OF AL/CALCINED DOLOMITE METAL MATRIX COMPOSITE BY POWDER METALLURGY ROUTE

Study the Influence of Sintering on the Properties of Porcelain Stoneware Tiles

2. EXPERIMENTAL PROCEDURE, RESULTS AND DISCUSSION

Improvement of Engineering Properties of Fired Clay Bricks Through the Addition of Calcite

Qualification of Micronized Kyanite and Silica Fume

Microstructural evolution of mullite during the sintering of kaolin powder compacts

CERAMIC PROPERTIES OF PUGU KAOLIN CLAYS. PART I: POROSITY AND MODULUS OF RUPTURE. Leonard D. Akwilapo 1* and Kjell Wiik 2.

Microstructure and characteristics of BaTiO 3 ceramic with 4PbO B 2 O 3 (sintering aid)

Processing of Ceramic Materials

A Study on Injection Moulding of Two Different Pottery Bodies

Properties of Fired Bricks by Incorporating TFT-LCD Waste Glass Powder with Reservoir Sediments

Anorthite porcelain: synthesis, phase and microstructural evolution

Silicon carbide particle embedded magnesium oxychloride cement composite bricks for polishing of porcelain stoneware tiles

Effect of Sintering Temperature on the Properties and Microstructure of a Ceramic Product

MSE 351 Engineering Ceramics I

TECHNICAL CERAMICS.

EFFECTS OF LIGHTWEIGHT MULLITE-SILICA RICH GLASS COMPOSITE AGGREGATES ON PROPERTIES OF CASTABLES

Electrical Porcelain Containing Ornamental Rock Waste: Microstructural Development

Ceramic Processing Research

Sintering of Ni-Zn Ferrite Nanopowders By the Constant Heating Rate (Chr) Method

DEVELOPMENT AND PRODUCTION OF CERAMIC TILES FROM WASTE BOTTLE POWDER (MILLED GLASS)

Influence of Raw Materials Ratio and Sintering Temperature on the Properties of the Refractory Mullite-Corundum Ceramics

Low Temperature Sintering Glass Binder for Al 2 O 3 Grinding Wheel

Research on the High Strength Glass Ceramics/Mullite Ceramics Composites

Diffusion Bonding of Boron Carbide to Boron Carbide Using Cu Interlayer

DILATOMETRIC STUDY ON LIQUID PHASE SINTERING OF Al-Ni POWDER COMPACTS

Indentation fatigue in silicon nitride, alumina and silicon carbide ceramics

SYNTHESIS OF FELDSPAR SUBSTITUTES VIA UTILIZATION OF LOCAL MATERIALS

DEGRADATION THE REFRACTORIES OF ROTARY FURNACE LININGS IN THE PRODUCTION OF ZINC OXIDE

Lecture Outline. Mechanical Properties of Ceramics. Mechanical properties of ceramics. Mechanical properties of ceramics

Ceramic Processing Research

Changes in Hot Modulus of Rupture with the Characteristics of Aluminosilicate Refractories

Technical Paper MCC AND HCC: DEFLOCCULATED HIGH PERFORMING CASTABLES RICH IN CALCIUM ALUMINATE BINDER. Christoph Wöhrmeyer, Chris Parr*,

Tio2 Reinforced Al2o3 Composites

Preparation of mullite by the reaction sintering of kaolinite and alumina

Fabrication of slag-glass composite with controlled porosity

Kinetics - Heat Treatment

Mine tailings valorization potential as raw materials for reaction sintered mullite based ceramics

MSE 352 Engineering Ceramics II

Fabrication and thermal properties of Al 2 TiO 5 /Al 2 O 3 composites

Effect of Nano-Crystalline Boehmite Addition on Sintering and Properties of Alumina Ceramics

Water Vapour Permeability of Clay Bricks

Influence of forming method and sintering process on densification and final microstructure of submicrometre alumina ceramics

5.2 Relationships between microstructure and mechanical properties

Synthesis of glass-ceramic glazes in the ZnO Al 2 O 3 SiO 2 ZrO 2 system

Properties of Ceramic Materials

Influence of milling time on the performance of ceramic ball grinding media prepared from refractory waste

Chemistry Department, Faculty of Science, Al-Azhar University [Girls], Nasr City, Cairo, Egypt **

Effects of mechanical activation on the fluxing properties of Gua Musang Feldspar

LSA GLASS-CERAMIC TILES MADE BY POWDER PRESSING

The Glass Ceramics Pyroxene Composition Synthesized on the Solar Furnace

STUDIES ON THE INTERACTION OF FLY ASH WITH LIME IN PRESENCE OF VARYING QUANTITY OF SAND

Ceramic Processing. Engineering Materials. 7/15/2009 Ceramic Processing/S.Rattanachan 1

Soluble Silicates in Refractory and Chemical- Resistant Cements

EFFECT OF COMPACTING PRESSURE ON INTERCONNECTED POROSITY IN IRON PM COMPACTS

PROPERTIES OF 2.45 GHZ MICROWAVE SINTERED SiO2 FROM RICE HUSK ASH AND Al2O3

Haseeb Ullah Khan Jatoi Department of Chemical Engineering UET Lahore

Innovative use of waste CRT glass in wall cladding tiles production

ENHANCEMENT OF MECHANICAL PROPERTIES FOR SELF-REINFORCED HOT-PRESSED SILICON CARBIDE

Production and Characterisation of Glass Foams for Thermal Insulation

Ceramic Processing Research

Powder Technology course Autumn semester Sintering theory. Peter M Derlet Condensed Matter Theory Paul Scherrer Institut.

Experimental O 3. Results and discussion

Use of Factorial Design in the Study of the Influence of Firing Cycle Parameters on the Technological Properties of a Brazilian Kaolinitic Material

Proceedings of XV BALKAN MINERAL PROCESSING CONGRESS Sozopol, Bulgaria, June 12 16, 2013

Preparation of rare earth oxide doped alumina ceramics, their hardness and fracture toughness determinations

Sintering activation of 316L powder using a liquid phase forming powder

Transcription:

SINTERING MECHANISMS OF PORCELAIN STONEWARE TILES M. Dondi*, M. Raimondo, C. Zanelli, CNR-ISTEC, Institute of Science and Technology for Ceramics, Faenza (Italy); P.M. Tenorio Cavalcante, COPPE-UFRJ, Rio de Janeiro (Brasil) ABSTRACT The sintering of porcelain stoneware tiles a glass-bonded material with excellent technical performances produced in slabs up to 1 m 2 is accomplished in roller kilns with fast cycles (<6 min) at maximum temperatures in the 118-124 C range. The main limit in further development of this product is its residual porosity, consisting of 2-8% closed pores, that is detrimental to its mechanical and tribological properties as well as the resistance to stains of polished tiles. Several industrial and experimental body formulations were studied by isothermal and constant rate optical thermodilatometry, RPD, SEM and BET, in order to understand which mechanisms are active during the whole sintering path. In the initial stage, a strong reduction of specific surface occurs with negligible shrinkage, suggesting that surface diffusion is the leading mechanism below C. In the intermediate step, densification is accomplished with increasing rate from 15 to 12 C, being a viscous flow the basic mechanism. However, the maximum shrinkage at each temperature is not related to the amount of viscous phase, so densification seems to be mainly governed by the viscosity of the liquid phase. In the final stage, a more or less conspicuous coarsening is observed, essentially over 1175 C, with development of coarser pores. In this critical step, the amount of residual porosity depends on the contrasting effects of a decreasing densification rate against an increasing coarsening rate. INTRODUCTION Porcelain stoneware is a glass-bonded material with excellent technical performances for ceramic tiles, such as mechanical, wear, frost and chemical resistance. In the last decade, the growth rate of the global production of porcelain stoneware tiles increased more than other ceramic products; in fact, the technical properties of porcelain stoneware, coupled with even more improved aesthetic appearance, gave it a prominent role on the tile market [1]. Porcelain stoneware bodies consist mainly of a mixture of ball clays (3-4%), feldspars (45-5%), quartz-feldspathic sands (1-15%) and several additives such as glass-ceramics frits and pigments. The tiles are obtained by wet grinding, spray-drying, dry pressing, fast drying, and are sintered by fast single firing (<6 minutes cold-to-cold, 118-12 C of maximum temperature and 5 minutes soaking times). Finished products including 1 m 2 slabs have strict requirements in terms of water absorption (<.5%) and planarity. The ceramic material is made up of crystalline phases, both new formed (i.e. mullite) and residual ones (quartz, feldspars) embedded in an abundant glassy matrix [2-3]. The main limit in further development of porcelain stoneware is the presence of a closed porosity (usually 2-8%) which is detrimental to the flexural modulus of rupture (max 8 MPa), Young modulus (max 75 GPa), abrasion resistance (no less than 11 mm 3 of materials removed according to ISO 1545-6) and particularly resistance to stains of polished tiles. The industrial sintering process is not able to remove completely the residual porosity, consisting of both small (<1 µm) spheroidal, gas-filled pores and irregularly shaped, large pores (up to 5 µm) probably resulting from coalescence of smaller ones [4-5]. The industrial attention has been especially paid to the intermediate stage of sintering, where densification occurs with a linear shrinkage around 7-9%; the final stage has been less concerned, notwithstanding the closed pores happened to be to a large extent unsinterable, since prolonged firing resulted in swelling by coarsening. The aim of this study is to understand better the mechanisms active during the whole sintering path of porcelain stoneware tiles, in order to achieve a phenomenological model of the process. In fact, notwithstanding the considerable innovation in the industrial firing technology (e.g. use of roller kilns, accurate control of temperature, great flexibility in time/temperature curves), the sintering process is still developed with an empirical approach (i.e. trial-and-error in the design of both sintering curve and body formulation). A scientific viewpoint is now necessary to accomplish any further material enhancement. MATERIALS AND METHODS Two samples of industrial bodies (ATP and AT) and five simplified compositions were studied: prevalently sodic (Na and NaB) or potassic (K and KB) and mixed sodic-potassic (NaK). Every body consists of a mixture of ball clay, quartz sand and sodic or potassic feldspars (Tab.1). These bodies were experimented at a laboratory scale, simulating the industrial tilemaking process, by mixing the raw materials, wet grinding in porcelain jar with dense alumina media (18h), followed by granulation and humidification with 5-6% water, uniaxial pressing (4 MPa) of 11 55 5 mm 3 tiles, then drying in electric oven at 15±5 C. All samples were characterized by isothermal and constant rate optical thermodilatometric analysis using a hot stage microscope (Expert System Misura). Isothermal tests were carried out at different maximum temperatures (heating rate 8 C/min to 1, 1125, 115, 1175 and 12 C). In order to estimate the actual phase composition at high temperature, the bodies underwent a water quenching: at different temperatures (15-12 C) and times corresponding to the maximum densification (Fig. 1); at different times up to the maximum densification at 12 C (Fig.2).

Table 1 Formulation of industrial and simplified bodies for porcelain stoneware tiles and physical properties of green bodies. Property Unit Na NaK K NaB KB AT ATP Ball clay %wt 4 4 4 35 35 36 37 Na-Feldspar %wt 5 25 41 38 K-Feldspar %wt 25 5 55 55 7 Quartz sand %wt 1 1 1 1 1 23 18 Median particle size µm 5.6 6. 6.6 3.5 4.2 4. 4.9 Bulk density g/cm 3 1.92 2.7 1.95 1.86 1.89 1.91 1.9 Relative density adim..738.7.75.721.735.726.725 12 11 99 97 95 93 1 C 1125 C 12 1175 C 115 C 2 4 6 8 12 14 Time (min) Fig. 1: Isothermal densification curves for the industrial sample AT. The symbols indicate for each temperature the time of maximum densification at which quenching was done. 12 11 99 97 95 12 C 93 2 4 6 8 12 14 Time (min) Fig. 2: Isothermal (12 C) densification curve for the industrial sample AT. The symbols indicate the times of progressive densification at which quenching was done.

Green compacts were characterized by testing: particle size distribution (photosedimentation, ASTM C 958), bulk density (geometrical method) and relative density (specific weight of powders by He pycnometry). The quenched specimens were characterized analysing: phase composition by RIR-RPD method, using CaF 2 as internal standard (Rigaku Miniflex, CuKα radiation); specific surface BET (Flowsorb II 23 Micromeritics); microstructure through SEM observations (Leica Cambridge Stereoscan 36). Some physico-chemical properties of the liquid phase such as viscosity at high temperature were estimated on the basis of its chemical composition [6-7]. 12 Initial stage Intermediate stage 92 Final stage 9 2 3 4 5 6 7 8 9 1 12 13 TEMPERATURE ( C) Fig. 3: Example of constant rate sintering curve of an industrial sample of porcelain stoneware tile. RESULTS AND DISCUSSION The firing behaviour of porcelain stoneware tiles is characterized by three main steps, as can be seen in a typical constant rate curve (Fig. 3): initial stage, corresponding to small size variations, extending usually up to 15-1 C; intermediate stage, accounting for most densification, usually in the 1-12 C range; final stage, where a more or less pronounced expansion occurs, due to a coarsening effect. Initial stage For the study of the early stage of sintering, an industrial body was quenched after firing at 9 C or C from 2 minutes to 24 hours of soaking time. The main sintering mechanism was evaluated by comparing the evolution of both specific surface area (with respect to the value S of the unfired body) and linear shrinkage (Fig. 4). S/S 8 Pure surface diffusion 36 24h 18 2 18 6 24h 6 4 2 2 ATP Pure grain boundary diffusion 9 C C -,5.5 1 1.5 L/L 2 2.5 3 3.5 4 Fig. 4: Relationship between linear shrinkage ( L/L ) and reduction of specific surface area ( S/S ) [8].

Table 2 Phase composition of the industrial body ATP during the initial sintering stage (after quenching at 9 and C). Phase (wt %) 9 C 2 min 9 C 24h 9 C 3h C 2 min C 24h Mullite - - - - 1.8±1.9 Quartz 37.5±1.5 27.1±2.8 27.6±1. 32.8±1.5 23.2±1.1 Cristobalite.8±. 3.3±.4 2.4 ±.2 2.9±.2 2.9 ±.1 Plagioclase 32.±.5 22.8±2.4 23.6±1.5 23.1±.9 13.8±.1 K-Feldspar 4.8±.3 - - - - Amorphous phase 25.±1.8 47.±3. 46.±2.8 41.±3.3 49.±4.1 A strong reduction of specific surface was found already at the lowest temperature for the shortest time (i.e. 9 C 2 min) coupled with a negligible expansion. The specific area reduction increases progressively up to 93%, while the linear shrinkage is always below 1%. This trend of linear shrinkage to specific surface suggests that the leading mechanism for sintering at temperature below C is mainly surface diffusion [8]. For higher temperatures, a viscous mass transport begins with a relevant shrinkage (3.5% at C for 24h) with just a limited further drop of specific surface (%). Moreover, the sintering at 9- C is accompanied by noteworthy phase transformations (Tab. 2). In fact, a considerable amount of amorphous phase is formed, to a large extent after metakaolinite (mullite is nor found yet), and to an initial fusion of feldspar-quartz eutectics. At 9 C it is evident a stabilization of the phase composition with increasing time, which represents a possible equilibrium condition. At C, the formation of mullite occurs together with a significant increase of liquid phase, that brings about the formation of about 2% of a feldspar-rich eutectic. Intermediate stage Approaching the intermediate stage of sintering, the seven samples underwent quenching from 1 to 12 C for soaking times corresponding to the maximum densification obtained from the isothermal curves. In all samples, the densification rate attributable to a viscous flow closing the open porosity increases rather regularly up to 12 C, but the shrinkage reaches its maximum value already close to 115 C and it does not vary much at higher temperatures (Fig. 1). However, for each sintering temperature, almost the same amount of vitreous phase was found (Tab. 3) implying that the linear shrinkage is not proportional to the quantity of liquid phase produced. As a matter of fact, the lower the viscosity of the liquid phase, the larger the shrinkage (Fig. 5). This result suggests that the densification rate by viscous flow depends more on the viscosity values than on the amount of liquid phase made available during the process. Probably, the control of the densification rate exerted by the solubility of the solid in the liquid phase is not so strict as expected. A prominent role seems to be played by the strong dependence of viscosity on temperature. 18 16 VISCOSITY (MPas) 14 12 1 K 8 6 ATP 4 AT 2 LINEAR SHRINKAGE ( L/L),2,3,4,5,6,7,8,9 Fig. 5: Relationship between linear shrinkage and viscosity of the liquid phase for samples ATP, AT and K.

Table 3 Amount of vitreous phase at the maximum densification for different temperatures C AT ATP NaB Na NaK K KB 1 63.8 69.5 77.8 75.3 71. 69.4 62.1 1125 74.2 7.8 74.6 73.8 74.6 68.5 68.2 115 77. 68.8 72.5 68.7 73.7 74.3 69.8 1175 74.5 71.2 76.3 73.2 65.9 76.6 69.4 12 78.3 69.5 73.8 69.5 72.6 75.5 71.6 Moreover, the different samples were isothermally treated at 12 C and were quenched for times shorter than those of maximum densification. There are relevant differences among the various body compositions, both in tems of maximum shrinkage and densification rate (Fig. 6). At the intermediate step, the sintering rates range between 1.5 and 4.6 %/min ( V/V ), with the industrial compositions containing also alkaline-earth oxides that exhibit the faster ones (Tab. 4). Overall, the sodium-rich bodies are characterised by higher apparent energies of activation of the viscous flow, as these energy values seem to depend to a large extent on the Na/K ratio. The relationship with the kinetic parameter k is much less significant, probably because other variables for instance the particle size distribution play a critical role (Fig. 7). 12 Na ATP NaK K AT 92 9 KB NaB 2 4 6 8 12 14 Fig. 6: Isothermal curves at 12 C for the different porcelain stoneware bodies. Table 4 Sintering rates and apparent energies of activation of initial densification Parameter Unit AT ATP NaB Na NaK K KB Isothermal sintering rate (12 C) Apparent energy of activation of initial densification (E a ) ( V/V ) min -1 2.38 4.36 4.63 1.52 2. 2.4 2.49 kcal/mol 158.3 225.6 156.9 195.3 163.6 71.3.4 Avrami parameter (k) adim..29.25.291.72.377.69.137

25.4.35 2.3 E a (kcal/mol) 15.25.2.15 k (adim.) 5 Ea k.1.5...5 1. 1.5 2. 2.5 3. 3.5 4. 4.5 5. Na 2 O/K 2 O Fig. 7: Correlation between the Na/K ratio of porcelain stoneware bodies with the apparent energy of activation of the viscous flow (E a ) and the Avrami kinetic parameter k. The phase composition exhibits an evolution similar to that presented in Fig. 8 in all samples. The partial melting of feldspars-quartz eutectics is very quick, as quartz and particularly feldspar amounts are greatly reduced after 5 minutes, and both plagioclase and K-feldspar are not found no longer after ten minutes. In contrast, the percentages of quartz and mullite are quite stable with time (Fig. 8a). On the other hand, the vitreous phase is already predominant after five minutes of treatment and its amount appears to be more or less constant with time (Fig. 8b). Overall, the system seems to reach rapidly a sort of equilibrium, at which the liquid phase does not change significantly its composition and thus its viscosity remains more or less constant during densification. 12 1 8 % 6 4 2 a Quartz Na-Feldspar Mullite 25 2 15 1 5 2.6 2.4 2.2 2 MPa s Viscosity Amorphous phase b 8 75 7 65 6 5 1 15 2 25 3 1.8 5 1 15 2 25 3 55 Fig. 8: Evolution of the phase composition and the estimated viscosity of the liquid phase of sample ATP at 12 C for increasing time. Final stage The final stage takes place over 1175 C and appears to be mainly characterized by constraints in pore removal. Coarsening and solubility of gases filling the closed pores become the most important phenomena affecting the tile microstructure. The porcelain stoneware bodies here studied, once the maximum density is achieved, tend more or less rapidly to expand, but this phenomenon occurs with a different rate for the various bodies (Fig. 6). In some compositions, the coarsening of closed pores is well appreaciable just after the maximum densification was achieved (Fig. 9). Therefore, the amount of residual porosity in the final step depends on the contrasting effects of two competing mechanisms: a decreasing densification rate against an increasing coarsening rate.

11 Na 12 C 5 TILE SIZE (%) 99 2 µm Na 12 C 97 2 µm 2 4 6 8 1 Fig. 9: Example of coarsening during the isothermal sintering (12 C) of porcelain stoneware (body Na). CONCLUSIONS Different mechanisms are active during the sintering of porcelain stoneware tiles. The most important are: surface diffusion, accounting for a strong reduction of specific surface with negligible shrinkage in the early stage (below C); viscous flow, responsible of densification occurring mainly in the 1-12 C range; pore coarsening, contrasting the decreasing sintering rate in the final stage (>1175 C) and resulting in swelling/bloating for prolonged firing. The maximum densification achievable increases progressively in the 15-12 C range, while the melting of feldspars-quartz eutectics occurs already at 9 C and a large mass of viscous phase is present over C. However, the firing shrinakge is not proportional to the amount of liquid phase. This picture suggests that densification depends essentially on the viscosity of the liquid phase, which exhibiting limited changes in composition with time, seems to be mainly affected by temperature. The sintering rate is very fast up to 7-8% of pore removal, accounting for 2-5% volumetric shrinkage per minute at 12 C. The apparent energy of activation for viscous flow is higher in sodic bodies (157-226 kcal/mol) than in potassic ones (71- kcal/mol). Coarsening and solubility of gases filling the closed pores are the prevalent mechanisms affecting the tile microstructure. Porcelain stoneware bodies tend to expand in the final stage, though this phenomenon occurs in the different bodies with a variable rate. In some compositions, a noteworthy pore coarsening appears just after the maximum densification was achieved. In this critical step, the amount of residual porosity seems to depend on the competing effects of decreasing densification rate against increasing coarsening rate. REFERENCES [1] T. Manfredini, G.C. Pellacani, M. Romagnoli, Porcelainized stoneware tiles, Am. Ceram. Soc. Bull 74 (1995) 76-79 [2] G. Biffi, Porcelainized Stoneware, Gruppo Editoriale Faenza, Italy, 19 [3] M. Dondi, G. Ercolani, M. Marsigli, C. Melandri and C. Mingazzini, The chemical composition of porcelain stoneware tiles and its influence on microstructure and mechanical properties. Interceram, 48 [2] (1999) 75-83. [4] E. Sanchez, M. J Orts, K. garcia-ten, V. Cantavella, Porcelain tile composition effect on phase formation and end product, Am. Ceram. Soc. Bull. 8 (6) (21) 43-49 [5] C. Leonelli, F. Bondioli, P.Veronesi, M. Romagnoli, T. Manfredini, G.C. Pellacani, V. Cannillo: Enhancing the mechanical properties of porcelain stoneware tiles: a microstructure approach, J. Eur. Ceram. Soc., 21 (21), 785-793 [6] T. Lakatos, L-G. Johansson, B. Simmingsköld, Viscosity temperature relations in the glass system SiO2-Al2O3- Na2O-K2O-CaO-MgO in the composition range of technical glasses, Glass Technology 13 [3] (1972) 88-95. [7] F. Cambier, A. Leriche, Vitrification, Materials Science and Technology A Comprehensive Treatment, Vol 17B Processing of Ceramics Part II (19) VCH [8] R.M. German, Overview of Key Directions and Problems in Computational and Numerical Techniques in Powder Metallurgy, (1993)