Вы находитесь на странице: 1из 9

Available online at www.sciencedirect.

com

ScienceDirect
Journal of the European Ceramic Society 34 (2014) 13651373

Role of cement content on the properties of self-owing Al2O3 refractory castables


Cem Gogtas a,1 , Hugo F. Lopez a, , Konstantin Sobolev b
a b

Department of Materials Science and Engineering, University of Wisconsin-Milwaukee, 3200 North Cramer Street, Milwaukee, WI 53211, United States Department of Civil Engineering and Mechanics, University of Wisconsin-Milwaukee, 3200 North Cramer Street, Milwaukee, WI 53211, United States Received 10 June 2013; received in revised form 8 October 2013; accepted 1 November 2013 Available online 12 December 2013

Abstract In this work, Al2 O3 self-owing castables (SFCs) were produced based on various cement contents. The SFCs were sintered at 1273 K, 1573 K and 1773 K and the exhibited properties were experimentally determined. Among the properties determined in this work are bulk density (BD), apparent porosity (AP), water absorption (WA), cold crushing strength (CCS), modulus of rupture (MOR) and fracture toughness (KIC ). It is found that additions of 5% cement lead to SFCs with maximum MOR and KIC values after ring at 1773 K. Firing at 1573 K leads to a reduction in both, MOR and KIC . In SFC containing 3% cement, maximum KIC values of 3.53 MPa m1/2 were achieved after ring at 1573 K. In the low cement SFCs (1 wt%) after ring at 1773 K the exhibited KIC values were below those obtained in either the SFC-3 or SFC-5, but they were signicantly high (3.43 MPa m1/2 ). 2013 Elsevier Ltd. All rights reserved.
Keywords: Self ow castables; Alumina refractories; Cement; Modulus of rupture; Fracture toughness

1. Introduction Refractory castables are classied by The American Society for Testing Materials (ASTM) according to their lime content as (a) conventional castables (CCs) (CaO > 2.5%), (b) low cement castables (LCCs) (2.5% > CaO > 1.0%), (c) ultra low cement castables (ULCCs) (1% > CaO > 0.2%) and (d) zero cement castables (ZCCs) (CaO < 0.2%). CCs contain 1530% calcium aluminate cement which acts as a lubricant in a water suspension and as a binder after hardening for strength at room temperature. High cement CCs are usually porous and open textured, due to the relatively large amounts of water required. They also exhibit a characteristic drop in strength at intermediate temperatures as sluggish sintering does not allow the development of a ceramic bond after breaking down the hydraulic bond. The high lime content of these castables favors the formation of low melting phases

Corresponding author. Tel.: +1 414 229 6005; fax: +1 414 229 6958. E-mail addresses: cgogtas@uwm.edu (C. Gogtas), hlopez@uwm.edu (H.F. Lopez), sobolev@uwm.edu (K. Sobolev). 1 Tel.: +1 414 229 6005; fax: +1 414 229 6958. 0955-2219/$ see front matter 2013 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.jeurceramsoc.2013.11.004

such as anorthite and gehlenite at elevated temperatures. These phases are known to degrade the refractoriness and corrosion resistance of the CCs.1,2 Further improvements in the properties of CCs have been achieved through the development of LCCs and ULLCs.13 In these castables, minimal water contents are needed to achieve the casting consistency. In turn, these castables posses improved cold and hot strength, reduced porosity, increased density and high corrosion resistance. In addition, in LCCs the calcium aluminate cement is replaced by a hydraulic binder such as hydraulic alumina.15 Due to the reduced water content low cement additions in the castables can induce poor owability and vibration is commonly applied in order to promote ow. In turn, this can result in serious space lling limitations in order to avoid any defect formation.69 In the mid-1980s, self-owing castables (SFCs) were developed (LCCs or ULLCs) with a mixing consistency which allows them to ow and degas without the application of vibration.7 In addition, in refractory materials one of the important factors used to control owability and porosity is the particle size distribution (PSD). Accordingly, the modied model of Andreassen has been widely used to establish particle size

1366

C. Gogtas et al. / Journal of the European Ceramic Society 34 (2014) 13651373

distributions.7,10 In the Andreassen model, q is known as the distribution coefcient of the PSD and for q values below 0.25, self ow properties are attained while values of around 0.30 are typical of vibrational castables.10 Alumina based castables have been widely investigated in the literature as alumina posses high strength, high hardness, good thermal shock properties and corrosion resistance.2,514 In the present work, only alumina grades that qualify for structural engineering, particularly with over 90% Al2 O3 and no open porosity are considered. Moreover, the exhibited properties including fracture strength and toughness are investigated as a function of self ow processing and cement additions (15%) including silica. 2. Experimental The starting raw materials and composition of the ceramic refractories produced in this work are given in Table 1. It was found that 5.5% water was enough to obtain a self-owing well dispersed castable system. Flow measurements were carried out by pouring the castable suspension into a truncated ow cone as described in the ASTM standard C-230. All self-ow measurements were based on the ASTM-C71 standard. In addition, the ASTM C-20 standard was used for determinations of apparent porosity (AP), bulk density (BD) and water absorption (WA). The mechanical properties such as cold crushing strength (CCS, ASTM C-133), modulus of rupture (MOR, ASTM C1161) and fracture toughness (KIC , ASTM C 1421) were also measured. The samples were cured for 24 h at room temperature (298 K) prior to testing. After de-molding, the samples were dried at 383 K for 24 h, then red at different temperatures (1273 K, 1573 K and 1773 K) for 3 h and furnace-cooled. Thermo-gravimetric analysis (TGA) and differential thermal analysis (DTA) were employed to disclose possible reactions exhibited by the SFCs as a function of the ring temperatures. Also, after mechanical testing, the fractured samples were cut into 25 mm 10 mm 10 mm bars and phase identication was made using X ray diffraction (XRD). In addition, microstructural
Table 1 Raw materials used in producing self-ow refractories. SFC-5 (wt%) Tabular alumina 36 mm 13 mm 01 mm 325 mesh Secar 71 (CA-14M) Alphabond 300 (hydratable alumina) A 1000SG (calcined alumina) 971U (microsilica) M-ADS (accelerator) M-ADW (retarder) Darvan-811D Citric acid (retarder) Water 17 26 27 13 5 7 5 0.05 0.05 5.5 SFC-3 (wt%) 17 26 27 14 3 1 7 5 0.9 0.1 5.5 SFC-1 (wt%) 17 26 27 14 1 3 7 5 0.9 0.1 5.5

characterization of the fracture surfaces was made using scanning electron microscopy (SEM) including energy dispersive spectroscopy (EDS). 3. Results and discussion 3.1. TGA and DTA Fig. 1 shows the TGA and DTA results of the processed SFCs in terms of weight loss (%) and the DTA (K/mg) parameter, respectively as a function of temperature. As expected, all the SFCs exhibit weight losses with increasing temperatures. In particular, there is a characteristic drop in the weight of the castables in the temperature range of 473573 K. This weight loss is associated with the dehydration of the hydrated cement phases AH3 and C3 AH6 , where C = CaO, A = Al2 O3 , and H = H2 O. According to the literature1,2 the dehydrating temperatures of AH3 and C3 AH6 are between 483573 K and 573633 K respectively. At elevated curing temperatures, the high density stable hydrates (AH3 and C3 AH6 ) produce signicant porosity with large pore sizes as the vapor species are able to escape. In turn, this promotes a sharp decrease in the weight of castables as seen in the above gure. Notice that weight loss (%) is higher in the high cement SFC-5 when compared with either SFC-3 or SFC-1. Since both, SFC-3 and SFC-1 contain hydrated alumina, they do not release chemically bonded water easily at elevated temperatures due to its gel structure when compared with cement.1 The DTA results of the castables show endothermic reactions coupled with the formation of the hydrated phases (AH3 and C3 AH6 ) at around 473573 K. Notice that hydration starts when cement is in contact with water (exothermic reaction) and internally, the castables can reach temperatures of up to 348 K. As the temperature is increased, the process of hydration is followed by the dehydration process and it continues until all the phases lose their water of crystallization. TGA results show a gradual increase in the weight loss (%) for the castables in the temperature range of 5731273 K. In addition, the DTA results show an almost linear reduction in (K/mg) for temperatures above 700 K. 3.2. Processing During sintering, reaction between the calcium aluminate cement, silica and the ne fractions of alumina (reactive and alumina sizes below 20 m) occurs, whereas the coarse alumina grains remain virtually inactive. The reactive alumina is not only a binding agent, but also the main component in achieving good rheological and owing properties. Fig. 2 shows X-ray diffraction intensity peaks (XRD) of the refractory castables dried at 383 K for 24 h. From this gure, it is evident that the refractory structure is predominantly -alumina as the X-ray intensity peaks correspond to the corundum phase. The XRD patterns of the SFCs red at 1773 K for 3 h are shown in Fig. 3. Under these conditions, mullite (3Al2 O3 2SiO2 ) forms in all of the SFCs compositions considered in this work. It is found that the Xray intensity peaks corresponding to the mullite phase decrease

C. Gogtas et al. / Journal of the European Ceramic Society 34 (2014) 13651373

1367

Fig. 1. Exhibited TGA and DTA curves in experimental SFCs.

in intensity as the amount of cement in the SFCs is increased (1%, 3% and 5%). This effect is attributed to the relatively large amounts of CACs (calcium aluminate cement)) which are known to attack mullite and/or inhibit its formation.13 During the ring process it is expected that CaO will react with SiO2 and Al2 O3 to form anorthite (CaOAl2 O3 2SiO2 ).2 In turn, this can lead to avoid or severely restrict the formation of mullite (3Al2 O3 2SiO2 ). The formation of the low melting phase anorthite was observed in both, SFC-5 and SFC-3 refractories but not in the SFC-1 one (see Fig. 3). In particular, it was found that anorthite formed at 1273 K in the SFC-5 and at 1573 K in the SFC-3. Apparently, increasing contents of cement promote the formation of anorthite at relatively low ring temperatures

(1273 K). From the ternary CaOAl2 O3 SiO2 phase diagram,16 it is expected that the transformation sequence followed by the SFC-1 at all the ring temperatures will only give rise to cristobalite formation. The formation of mullite and/or anorthite is effectively suppressed as there is not enough SiO2 or CaO to form either of these phases.

3.3. Physical properties Fig. 4 shows the relationship between cement content and the self ow values exhibited by the experimental castables. As expected, self ow values decrease with increasing hydraulic

Fig. 2. XRD pattern of (a) SFC-5, (b) SFC-3, (c) SFC-1 red at 383 K for 24 h.

1368

C. Gogtas et al. / Journal of the European Ceramic Society 34 (2014) 13651373

Fig. 3. XRD pattern of (a) SFC-5, (b) SFC-3, (c) SFC-1 red at 1773 K for 3 h.

Fig. 4. Self ow values of SFCs as a function of cement content.

alumina contents. This stems from the need for relatively large amounts of water required by hydraulic alumina when compared with cement in order to set the castables. Moreover, in order to control the setting time for the castables small additions (usually around 0.5%) of cement may be necessary,1 otherwise, large amounts of water might be required to maintain ow. The resultant bulk density (BD) and apparent porosity (AP) of the SFCs after drying and ring at elevated temperatures is given

in Fig. 5. Notice from this gure that at 1773 K, the BD values reach a maximum in all the experimental castables. In addition, this corresponds to a reduction in AP to the lowest levels. This is likely to be due to the formation of the high density mullite phase (3.163.22 g/cm3 ) combined with amorphous anorthite (CaOAl2 O3 2SiO2 ) which can easily ll the open aggregate interspaces. The increases in the BD values are found to strongly depend on the ring temperatures and on the cement content. At ring temperatures between 1273 K and 1573 K the SFC-5 yields the highest AP and lowest BD values. Apparently, the lack of enough mullite formation at these temperatures in the SFC-5 is the main cause for the exhibited AP and BD values. Anorthite is found to start forming at 1273 K and 1573 K in SFC-5 and SFC-3, respectively but not in the SFC-1. Hence, this phase is not likely to play a major role in the experimentally determined AP and BD values exhibited in the SFC-1 refractories. The increase in the AP of the castables is attributed to a dominant dehydration process when ring temperatures below or equal to 1273 K were used. In addition, there is a clear

Fig. 5. Apparent porosity, AP and bulk density, BD for the SFCs as a function of the ring temperatures.

C. Gogtas et al. / Journal of the European Ceramic Society 34 (2014) 13651373

1369

Fig. 6. Water absorption, WA (wt%) for the SFCs as a function of the ring temperatures.

Fig. 8. The MOR of SFCs as a function of temperature.

increase in the apparent porosity and a reduction in density as water evaporates and leaves the SFCs. Above 1273 K, closure of the developed porosity starts to occur leading to densication through the formation of new phases such as anorthite, cristobalite and mullite. This is further conrmed by determinations of water absorption in the SFCs (see Fig. 6). Notice from this gure that WA is the highest in the high cement castable, SFC-5 all the way to 1573 K. Yet, ring at 1773 K drastically reduces the amount of WA leading to a maximum BD with minimal AP. In contrast, the amount of WA in the SFC-1 is still relatively high. This can be explained by considering that hydrated alumina does not easily release the chemically bonded water at elevated temperatures due to its gel structure.1 3.4. Mechanical properties Fig. 7 shows the cold crushing strength (CCS) exhibited by the castables at the ring temperatures considered in this work. Notice that up to 1273 K the SFC-5 possesses the highest CCS values when compared with either the SFC-3 or the SFC-1. Apparently, at decreasing cement contents, the development of bonding between aggregates is rather limited at the low ring temperatures. Nevertheless, after heating at 1773 K for 3 h, the SFC-3 exhibits the highest CCS values while the SFC-5 strength values slightly drop down. The apparent reduction in CCS observed in the SFC-5 can be associated with the formation of the low melting anorthite phase. In addition, the

Fig. 7. CCS of the SFCs as a function of the ring temperatures.

CCS exhibited by the SFC-3 does not change much at the ring temperatures employed. Alternatively, ring the SFC-1 at 1573 K leads to a vast improvement in CCS. In turn this can be related to a strong bond formation through the development of the cristobalite phase. The modulus of rupture (MOR) exhibited by SFCs as a function of the ring temperatures is given in Fig. 8. Notice that similar trends are observed in the exhibited MOR values when compared with the ones for the CCS (Fig. 7) (i.e., up to 1273 K the highest MOR values are found in the SFC-5 with a high cement content (5%)). Once again, the presence of cement in the SFC-5 promotes the early development of ceramic bonding with the aggregates when compared with the SFC-3 and SFC-1 refractories. Moreover, ring at 1573 K for 3 h in both, SFC-3 and SFC-1 leads to signicant improvements in the resultant MOR, suggesting the development of ceramic bonding. However, in the SFC-5 the MOR values drop down to 36.34 MPa. This coincides with the formation of the low melting phase anorthite which is expected to reduce the MOR strength of the SFC-5. At 1773 K, all the SFCs show gradual improvements in the MOR values probably due to the formation of mullite. In a related work, Martinovi c et al.15 found that in a low cement castable (1.2% CaO) ring at 1873 K, resulted in MOR, WA, CCS values that are comparable the ones found in this work. In their work, the exhibited property improvements are attributed to the formation of calcium hexaaluminate (CaO6Al2 O3 ) or CA6 phase2 which forms at elevated temperatures. Hence, these castables must be red at temperatures of 1873 for maximum strength as the strength drops by over 50% when ring at lower temperatures (1473 K). Nevertheless, the reported castable properties after ring at temperatures below 1873 K are well below the ones found in this work, particularly MOR and CCS. In their work, ring at 1073 K, gives rise to CCS and MOR values that are not signicantly different from the ones found at 383 K, probably due to the lack of any signicant ceramic reaction. In their work, they use a silica-free composition where the only phases present up to 1673 K are Al2 O3 , CA and CA2 phases. From the published literature,1 the CCS values for CA and CA2 are 60 and 25 MPa, respectively, and they are unable to provide any strength improvements. Alternatively, in the present work the formation of anorthite and cristobalite phases at temperatures between 1273 K and

1370

C. Gogtas et al. / Journal of the European Ceramic Society 34 (2014) 13651373

1573 K. This effect is consistent with the exhibited MOR values found in the high cement castable and it can be attributed to the formation of anorthite. Firing at 1773 K for 3 h improves the fracture toughness of the SFC-5 refractory to maximum KIC values of up to 4.08 MPa m1/2 . The toughness improvements can be related to the formation of mullite concomitant with a reduction in the amount of anorthite and a relatively high densication with reduced porosity. 3.5. Fracture morphologies
Fig. 9. Fracture toughness values of SFCs as a function of temperature.

1573 K promote appreciable enhancements in both, the MOR and CCS values for the investigated SFCs. Hence, the property improvements found in this work are comparable or superior to the ones reported by Martinovi c et al.15 Notice in particular that in the low cement alumina-silicate castables, high strength can be achieved by processing at relatively low ring temperatures (below 1373 K). In addition, ring temperatures as low as 1573 K are sufcient to produce high strength castables in the ultra-low cement compositions (SFC-1). Fig. 9 shows fracture toughness (KIC ) values for the experimental SFCs as a function of the ring temperatures. In all the experimental SFCs the exhibited fracture toughness values were between 3 and 4 MPa m1/2 after ring at or above 1573 K. Typical KIC values for conventional castables are of the order of 0.21.5 MPa m1/2 .2 Hence, it is evident that the SFCs produced in this work can be considered as high toughness refractories. Notice that the KIC values in both, SFC-3 and SFC-1 steadily increase with the ring temperatures. It is found that the KIC tends to increase with the cement content except for the SFC-5 which exhibits a slight reduction in the exhibited KIC values at

Refractory castables are rather brittle at low ring temperatures as the ceramic bond has not yet developed. Fig. 10ac shows the fracture surfaces of a castable after drying at 383 K for 24 h. Notice from this gure that the fracture surfaces consist of aggregates debonded from the refractory castable binder. Thus, it is evident that no strong bond exists between aggregates and no binding system develops at this temperature. Fig. 10b shows a matrix area associated with the debonded aggregates. In addition, porosity of various sizes is observed (see Fig. 10a) and shapes close to spherical morphology (see Fig. 10c). From these gures, it is evident that the fracture surface of the SFC is rather smooth suggesting that it is inherently brittle. Firing at high temperatures promoted the development of aggregate bonding through various reactions. Fig. 9 shows the exhibited KIC for the various SFC refractories after ring at 12731773 K. The exhibited improvements in KIC values are attributed to the formation of ceramic phases such as anorthite, cristobalite and mullite. Fig. 11a shows the fracture surface of a SFC-3 after ring at 1773 K for 3 h. Notice in this gure that most of the alumina aggregates are transgranularly fractured as a strong chemical bonding has developed between the aggregates and the binding phases formed through sintering. For the most part, the number of fractured aggregates increases with

Fig. 10. SEM fractographs of a SFC-3 after drying at 383 K for 24 h. (a) Debonded aggregates and porosity, (b) detail of the debonded area in the surrounding matrix and (c) a pore and its effect on the exhibited fracture path.

C. Gogtas et al. / Journal of the European Ceramic Society 34 (2014) 13651373

1371

Fig. 11. SEM micrographs of SFC-3 after ring at 1773 K for 3 h. (a) Fracture surface, (b) X-ray peak intensity corresponding to Al and (c) a fractured alumina aggregate bonded to surrounding matrix.

the ring temperatures. Fig. 11b and c is SEM micrographs of a fractured aggregate, showing a strong bond with the surrounding matrix. X-ray intensity peaks, (EDS, Fig. 11c) corresponding to the fracture surface indicates that the fractured aggregate is alumina. Apparently, as the aggregates become strongly bonded in the surrounding matrix, the elastic energy needed for debonding surpasses that for fracture of the alumina grains and aggregate fracture becomes prevalent. Fig. 12a and b is SEM micrographs corresponding to aggregate fractures in SFC-3 red at 1273 K and 1773 K, respectively. In both cases, the aggregate fracture is strongly inuenced by the crystal structure of the alumina as multiple ledges or steps develop along the fracture path. In addition, at the highest ring temperatures the fracture path becomes increasingly tortuous as the length of the steps or ledges developed are rather short

compared with the ones found in the SFC-1. Moreover, appreciable closed porosity is also observed within the alumina aggregates. All of these factors can account for the exhibited strength and KIC values, as well as for their relative differences in magnitudes. In particular, they can explain the relatively high KIC values found in the SFC-5 refractories. Fig. 13a and b shows the fracture surfaces of SFC-5 and SFC-1 samples after ring at 1773 K for 3 h. From these gures, it is evident that there is an increasing number of short sized cracks developing in the SFC-5 when compared with the SFC-1 fracture surfaces. The development of multiple cracking in the SFC-5 at 1773 K can be related to the formation of mullite at these temperatures. As mullite develops there is a volume increase associated with the transformation which can promote cracking through the development of residual stresses. Other

Fig. 12. SEM micrograph of SFC-3 after ring (a) at 1273 K and (b) 1773 K for 3 h.

1372

C. Gogtas et al. / Journal of the European Ceramic Society 34 (2014) 13651373

Fig. 13. SEM micrographs of (a) SFC-5 and (b) SFC-1 after ring at 1773 K for 3 h.

factors contributing to the development of residual stresses include elastic anisotropy due to thermal expansion differences among alumina aggregates and also among the new developed phases (anorthite and mullite). Although there are no fracture mechanics models that can quantitatively explain toughening in single phase polycrystalline ceramics, among the suggested toughening mechanisms are,17 (a) microcrack toughening and (b) crack bridging. The mechanism of microcrack toughening assumes that stable grain boundary microcracks are nucleated by the high stresses in the vicintiy of the macroscopic crack tip. These micro-cracks then lower the stress experienced by the tip. In the second approach,17 alteration of the stress intensity of a macro-crack tip is assumed as the crack interacts with a discrete array of micro-cracks. The proposed toughenning mechanisms are in agreement with the observations of an increasing density of short size microcracks found in the SFC-5 castables. In turn, the exhibited improvements in the SFC-5 samples can be attributed to the development of discrete arrays of micro-cracks as a result of relatively high residual stresses. The development of increasing residual stresses can be related to the contribution of increasing phase formation reactions (mullite and anorthite) induced at aggregateaggregate boundaries. Notice that in SFC-1 castables, cristobalite is the dominant phase that develops at the ring temperatures. In turn, the residual stress development associated with the aggregates and the cristobalite phases is not effective enough to promote the development of a high density of short cracks. 4. Summary In this work, low cement self owing castables (SFCs) were processed. A water addition of 5.5% was found to be enough to achieve full self ow properties. It was found that the self ow values decreased as the volume fraction of hydratable alumina was increased. Drying at 383 K, produced a fully corundum matrix structure in all the SFCs. The formation of anorthite was observed in SFC-5 and SFC-3 at 1273 K and 1573 K, respectively. In contrast, cristobalite was found to be present in SFC-1 at 1573 K. At 1773 K, SFCs having 1% and 3% cement content

developed relatively large amounts of mullite as evidenced by X-ray diffraction. The cold crushing strength values were found to reach a maximum at 1573 K in the SFC-3 and SFC-1 castables and at 1273 K for the SFC-5 ones. In addition, in the SFC-3 and SFC-1 castables, KIC and MOR values were found to consistently increase with the ring temperatures as a result of the development of ceramic bonding. In contrast, a slight drop in KIC and MOR values was observed in the SFC-5 at 1573 K, probably due to the formation of anorthite. Yet, the exhibited KIC in the SFC-5 greatly improved after ring at 1773 K. The highest values of BD corresponded to the lowest values of AP and they were attained after ring at 1773 K in all of the castables. Acknowledgements The authors would like to acknowledge the support of Steve Kowalski and Dale Zacherl from Almatis Corporation by providing all of the materials used in this research work. References
1. Lee WE, Vieira W, Zhang S, Ghanbariahari K, Sarpoolaky H, Parr C. Castable refractory concretes. Int Mater Rev 2001;46:14567. 2. Schacht CA. Refractories handbook. New York: Marcel Dekker, Inc.; 2004. 3. Cardoso FA, Innocentini MDM, Miranda MFS, Valenzuela FAO, Pandolfelli VC. Drying behavior of hydratable alumina-bonded refractory castables. J Eur Ceram Soc 2004;24:797802. 4. Gogtas C, Unlu N, Odabasi A, Sezer L, Cnar F, Guner S, et al. Preparation and characterization of self owing refractory materials containing 971U type microsilica. Adv Appl Ceram 2010;109:611. 5. Ye G, Troczynski T. Hydration of hydratable alumina in the presence of various forms of MgO. Ceram Int 2006;32:25762. 6. Gogtas C, Odabasi A, Sezer L, Cinar F, Guner S, Eruslu N, et al. Improvement of properties of self-owing low-cement castables based on brown fused alumina. J Ceram Process Res 2007;8:32430. 7. Evangelista PC, Parr C, Revais C. Control of formulation and optimization of self-slow castables based on pure calcium aluminate. Refract Appl New 2002;7:148. 8. Gogtas C, Unlu N, Odabasi A, Sezer L, Cinar F, Guner S, et al. Characterization of the physical properties of a self-owing low-cement castable based on white-fused alumina. J Ceram Process Res 2008;9:17.

C. Gogtas et al. / Journal of the European Ceramic Society 34 (2014) 13651373 9. Jones V, Cross G. Rheology control of low-cement self-ow castables. Ceram Ind 1998;148:4953 [ABI/INFORM Global]. 10. Myhre B, Hundere AM. The use of particle size distribution in development of refractory castables. In: XXV ALAFAR Congress. 1996. p. 14. 11. Myhre B, Sunde K. Alumina based castables with very low contents hydraulic compound Part 1. In: UNITECR 95. 1995. p. 1922. 12. Cardoso FA, Innocentini MDM, Akiyoshi MM, Pandolfelli VC. Effect of curing time on the properties of CAC bonded refractory castables. J Eur Ceram Soc 2004;24:20738. 13. Zawrah MFM, Khalil NM. Effect of mullite formation on properties of refractory castables. Ceram Int 2001;27:68994.

1373

14. Zhou X, Sankaranarayanane K, Rigaud M. Design of bauxite-based low-cement pumpable castables: a rheological approach. Ceram Int 2004;30:4755. 15. Martinovic S, Vlahovic M, Majstorovic J, Matovic B, Volkov-Hosovic T. Thermal and mechanical properties of high alumina low cement castable. Metall Mater Eng 2012;18:5365. 16. Levin EM. Phase diagrams for ceramists. Columbus, OH: American Ceramic Society; 1964. 17. Hutchinson JW. Mechanisms of toughening in ceramics. In: Germain P, Piau M, Cailleirie D, editors. Theoretical and applied mechanics. North-Holland: Elsevier Science Publishers; 1989. p. 13944.

Вам также может понравиться