Introduction

Calcium monoaluminate (CaAl2O4, CA), calcium dialuminate (CaAl4O7, CA2) and calcium hexa-aluminate (CaAl12O19, CA6) are lime-alumina compounds involved in up-to-date ceramic technology and have attracted considerable attention due to their unique properties. CA (1606 °C) [1, 2] is the main chemically active hydraulic constituent of calcium aluminate cements (CACs) [3]. CA2 is a compound which can be broadly used in modern refractories with a high resistance to thermal shock, especially due to its a high melting temperature (1759 °C) [1, 2] and very low or even negative coefficients of thermal expansion [4,5,6]. CA6 is a non-hydraulic compound and more refractory than the other phases, having a melting point of 1850 °C [1, 2]. Calcium zirconium aluminate (Ca7ZrAl6O18) having incongruent melting point at 1550 °C [7] has recently been analyzed in terms of its hydraulic properties [8,9,10,11]. Considering the above mentioned, the CaO–Al2O3–ZrO2 multiphase composite cements can be recognized as very promising candidates for high-temperature applications.

There are several methods used for production of calcia–alumina–zirconia ceramics. Lee et al. [12] synthesized Al2O3-modified calcia-stabilized zirconia materials from fine powders by sintering at 1450–1600 °C in air atmosphere of green pellets shaped via uniaxial pressing and then isostatically pressed at 200 MPa. Liu et al. [13] prepared porous alumina–zirconia ceramics by infiltrating porous alumina ceramics previously prepared with ultra-fine alumina powder via tert-butyl alcohol (TBA)-based gel-casting method. Ai et al. [14] fabricated zirconia matrix ceramics doped with 7.5 vol% nano-Al2O3 by two-step microwave sintering, two-step conventional sintering and one-step microwave sintering. According to a research paper discussing the formation mechanism of calcium aluminate compounds (Ca3Al2O6, Ca5Al6O14, Ca12Al14O33, CaAl2O4, CaAl4O7 and CaAl12O19) based on high-temperature solid-state reaction within the temperature ranges 1100–1400 °C, CaO gradually diffuses into the Al2O3 core and then a layered distribution of calcium aluminates occurs [15]. Composites of ZrO2–CaAl4O7 (CA2) were synthesized by Bruni et al. [16] using the conventional sintering technique at 1400 °C from the mixture of monoclinic polymorph of zirconia with different amount of calcium aluminate cement (CAC). Calcium hexaluminate (CaAl12O19, CA6, hibonite) in an elongated grain morphology was formed in situ during sintering of zirconia-toughened alumina ceramics with CaO at 1600 °C for 4 h [17]. The CA6–CA2–Al2O3 three-phase composite was also recently synthesized by the solid-state reaction process at an elevated temperature using the previously synthesized CA2 compound and commercial corundum powder as starting materials [18]. They are also known the low thermal expansion ceramics based on calcium dialuminate (CaAl4O7, CA2), calcium dialuminate-calcium zirconate (CaZrO3, CZ) and calcium dialuminate–magnesium aluminate (MgAl2O4, MA) binary phases materials from the literature [4,5,6].

Thus, the present work was aimed at studying of the high-temperature reactions between Ca7ZrAl6O18 and Al2O3 for obtaining of novel composite cements belonging to the system CaO–Al2O3–ZrO2 with favorable hydraulic properties by in situ synthesis method. This approach to the composition and microstructural designs strategies is also a somewhat novel concept for the materials scientist.

Experimental

Materials and cements preparation procedure

The materials used in this study were a commercially available α-Al2O3 (PFR 20 reactive alumina, supplied by Alteo) with the chemical and physical characteristic given in Table 1 and calcium zirconium aluminate, Ca7ZrAl6O18 (C7A3Z; C≡CaO, A≡Al2O3, Z≡ZrO2). High-purity commercial mono-modal reactive alumina powder was used as a raw material for the preparation of binary mixtures of Al2O3 and Ca7ZrAl6O18. For this work, Ca7ZrAl6O18 was obtained by a two-step synthesis procedure. In a first step, the calcium carbonate (99.81% CaCO3, Chempur), α-Al2O3 (99.8% Al2O3, Across Organics) and zirconia (98.08% ZrO2, Merck) powders were mixed with the 7:3:1 molar ratio of CaO, Al2O3 and ZrO2 oxides, and then, the mixture was homogenized for 2 h in a ball mil and pressed into cylinders having a diameter of 2 cm. All green pellets were calcined at 1200 °C for 10 h. In a second step, solid-state sintering of the pellets made from the calcined powder at 1500 °C for 30 h resulted in a single-phase orthorhombic Ca7ZrAl6O18 structure. An intermediate grinding/mixing stage in order to improve homogeneity was necessary. Nevertheless, some trace impurities, in the form of CaZrO3 and Ca12Al14O33 were detected.

Table 1 Chemical and physical characteristics of reactive alumina micropowder [19]

As a next step, novel calcia–zirconia–alumina pellets were prepared by a high-temperature reactive sintering of before synthesized Ca7ZrAl6O18 and reactive Al2O3 micropowder. The initial compositions of the Ca7ZrAl6O18–Al2O3 mixtures were located within the Ca7ZrAl6O18–CaAl2O4–CaZrO3, CaAl2O4–CaAl4O7–CaZrO3 and CaAl4O7–CaAl12O19–ZrO2 composition triangles of a ternary system CaO–Al2O3–ZrO2. They contained 25, 50 or 75% by mass of Al2O3 and 75, 50 or 25% by mass of Ca7ZrAl6O18, respectively. Homogenization of the mixtures proceeded for 2 h. Densification of the mixed powders was conducted by pressureless sintering of pellet samples at temperatures of 1300–1500 °C under air atmosphere (Table 2). The aluminates CaAl2O4 (CA), CaAl4O7 (CA2) and Ca12Al14O33 (C12A7) were also synthesized by two-cycle heating and grinding stoichiometric mixtures of CaCO3 and Al2O3 for use as reference material for calorimetric studies.

Table 2 Percentage composition (% by mass) of ceramic pellet samples and sintering condition

Methods

X-ray diffraction analysis (XRD) was carried out using the X’Pert Pro PANalytical X-ray diffractometer. The XRD patterns of previously synthesized calcium zirconium aluminate phase, reactive alumina, calcium aluminates and calcia–zirconia–alumina composite cements were collected by step scanning with step of 0.02° over the range 2Θ 5°–90° at room temperature.

The phases developed during sintering were compared and confirmed using search–match reference ICDD database. Multiphase Rietveld phase quantification was also done using standard ICSD files supported by the PANalytical HighScore Plus software [20]. Goodness of fit (GOF) values lower than 2 proved a good quality Rietveld refinement for all samples. Various functional groups present in the prepared composite cements were identified by FT-IR. A Bruker Vertex 70v spectrometer was employed for this purpose. The FT-IR spectrometer was operated in the mid-IR region (4000–400 cm−1) with a resolution of 4 cm−1. Each spectrum represents the average of 128 scans. Samples were prepared by the standard KBr pellet method. Moreover, scanning electron microscopy (SEM, Nova NanoSEM 200 FEI) combined with energy-dispersive X-ray spectroscopy (EDS) was used to examine the microstructure development. For this purpose, resin-embedded samples were polished, coated with carbon to be conductive and then examined by SEM.

For microcalorimetric measurements, the pellets were ground and milled to obtain fine powder. Specific surface and grain size distribution of samples were measured by a laser diffraction analyzer (the Master Sizer 2000 Ver. 5.60 apparatus of Malvern (UK)). Heat of hydration of the novel composite cements belonging to the system CaO–Al2O3–ZrO2 (Tables in Fig. 1a–c) and the reference aluminate phases was determined with a TAM Air microcalorimeter (TA Instruments) by integrating the continuous heat flow signal during the 20-h hydration process. Calorimetric measurements of synthetic cements were performed using the in situ mixing treatments of the binder pastes under the following conditions: temperature of 25 °C and a water/solid mass ratio w/s = 1.0. Tests were performed by mixing the pastes inside the calorimeter to allow the study of cement early hydration processes. For this purpose, two grams of dry powder was poured into a glass vial, and water weighed into mounted syringes. Admix ampoule was introduced into the calorimeter prior mixing, thus equilibrating to isothermal environment. Mixing was done after equilibration inside the calorimeter to enable the quantitative access to the heat of the early hydration process. A second glass vial containing sand (providing the same heat capacity that the mass of binder paste) was used as an inert reference in the twin channel of calorimeter to ensure the best stability of the baseline. Heat flow curves were normalized to the mass of total dry binder and reported in terms of power (mW g−1 dry binder) versus time (hours).

Fig. 1
figure 1

The results of quantitative Rietveld analyses made with HighScore Plus (PANalytical) program for the samples 25A–75C7A3Z (A), 50A–50C7A3Z (B) and 75A–25C7A3Z (C) sintered in the temperature range 1300–1500 °C; C≡CaO, A≡Al2O3 and Z≡ZrO2

Results and discussion

Phase transformation

Phase transformation of samples 25A–75C7A3Z, 50A–50C7A3Z and 75A–25C7A3Z sintered in the temperature range 1300–1500 °C was examined by XRD analysis, and the crystalline phase content in the samples as determined by Rietveld refinements of the powder patterns is summarized in the tables presented in Fig. 1a–c.

Analysis of the quantitative phase changes based on the Rietveld XRD quantification shown in Fig. 1a–c is suitable to draw some reliable conclusions related to the mechanism of the reaction within the Ca7ZrAl6O18–Al2O3 system in question. Now let us consider 25A–75C7A3Z sample (Fig. 1a). Calcium zirconium aluminate (Ca7ZrAl6O18) reacted with Al2O3 to form calcium monoaluminate (CaAl2O4) and calcium zirconate (CaZrO3) following sintering at 1300 °C for 1 h. The same applies for the samples 50A–50C7A3Z and 75A–25C7A3Z where the content of CaAl2O4 also increases and the contents Ca7ZrAl6O18 and Al2O3 accordingly decrease (Fig. 1b–c). The reaction proceeds according to Eq. 1.

$${\text{Ca}}_{7} {\text{ZrAl}}_{6} {\text{O}}_{18} + 3{\text{Al}}_{2} {\text{O}}_{3} \to 6{\text{CaAl}}_{2} {\text{O}}_{4} + {\text{CaZrO}}_{3}$$
(1)

On the other hand, in the sample 25A–75C7A3Z calcium monoaluminate (CaAl2O4) reacted simultaneously with Al2O3 to form unstable calcium dialuminate (CaAl4O7) under the same sintering condition. After sintering at 1400 °C, CaAl4O7 disappeared, the amount Ca7ZrAl6O18 decreased and the amount of CaAl2O4 and CaZrO3 increased. A quantitative approach based on XRD results in proposed mechanism of the chemical reaction 2:

$${\text{Ca}}_{7} {\text{ZrAl}}_{6} {\text{O}}_{18} + {\text{CaAl}}_{4} {\text{O}}_{7} + 2{\text{Al}}_{2} {\text{O}}_{3} \to 7{\text{CaAl}}_{2} {\text{O}}_{4} + {\text{CaZrO}}_{3}$$
(2)

Following sintering at 1500 °C for 1 and 3 h, it was found that Ca7ZrAl6O18 completely disappeared due to high-temperature eutectic reactions and formation of Ca12Al14O33. The phase formation at each stage of sintering arising when several different reactions occur at different sintering condition will be discussed in detail later when discussing the microstructure evolution.

In 50A–50C7A3Z composite, a larger content of Al2O3 was enough to remain CaAl4O7 as a stable phase. At the temperature range between 1300 and 1500 °C the increment in CaAl4O7 content and the decrease in amount of both Al2O3 and initially formed CaAl2O4 phase were observed. Thus, the following reaction is proposed:

$${\text{CaAl}}_{2} {\text{O}}_{4} + {\text{Al}}_{2} {\text{O}}_{3} \to {\text{CaAl}}_{4} {\text{O}}_{7}$$
(3)

It was also found that evolution with temperature of the amount of crystalline phases present in the 75A–25C7A3Z material occurred. So, Fig. 1c shows a remarkably strong increase with CaAl12O19 content and the accompanying decrease in content of both CaAl4O7 and Al2O3. The results of quantitative analysis of diffraction data support the proposed mechanism of reaction 4. In the alumina-rich part of the phase diagram calcium zirconate (CaZrO3) not exists in equilibrium and then it must be converted in zirconia according to the chemical reaction 5.

$${\text{CaAl}}_{4} {\text{O}}_{7} + 4{\text{Al}}_{2} {\text{O}}_{3} \to {\text{CaAl}}_{12} {\text{O}}_{19}$$
(4)
$${\text{CaZrO}}_{3} + {\text{Al}}_{2} {\text{O}}_{3} \to {\text{CaAl}}_{2} {\text{O}}_{4} + {\text{ZrO}}_{2}$$
(5)

The final composition of investigated samples is presented as X-ray diffractograms in Figs. 24 with JCPDS standard cards used for phase identification.

Fig. 2
figure 2

X-ray diffraction pattern of the 25Al2O3–75C7A3Z composite sintered at 1500 °C for 3 h

Fig. 3
figure 3

X-ray diffraction pattern of the 50Al2O3–50C7A3Z composite sintered at 1500 °C for 3 h

Fig. 4
figure 4

X-ray diffraction pattern of the 75Al2O3–25C7A3Z composite sintered at 1500 °C for 3 h

Spectroscopic results

Figure 5 shows the IR spectrum of sintering products of Al2O3 and Ca7ZrAl6O18 powders. FT-IR is suitable for analyzing the calcium aluminate phases and final characterization of CaO–Al2O3–ZrO2 materials. According to paper presented by Tarte [21] concerning the infrared spectra of inorganic aluminates and characteristic vibrational frequencies of AlO4 tetrahedra and AlO6 octahedra, two different regions can be identified within the infrared spectra of calcium aluminates: the first one in the region of 700–900 cm−1 that corresponds to the stretching vibrations of a lattice of interlinked AlO4 tetrahedra; and the second one appears in the region of 400–500 cm−1, which can be particularly useful for determining the presence of bending mode of the AlO4 lattice [22, 23]. In the spectrum of the main phase present in considered ceramic composite, CaAl2O4, the relevant signals are centered at 417, 447, 542, 573, 638, 687, 721, 764, 779, 790, 804, 820, 841 and 868 cm−1. The structural arrangement considered as the mentioned groups was also found in a minor phase of the 25Al2O3–75C7A3Z sample, i.e., Ca12Al14O33. Despite the fact that dodecacalcium hepta-aluminate gives the spectrum with a very strong and broad absorption band with a maximum at 835–850 cm−1 and the two strong bands at 573–580 cm−1 and at 457–465 cm−1 [23, 24], the coincidence of positions of some bands in the IR spectra of the two phases occurred. The 50A–50C7A3Z sample gives a spectrum dominated by the absorption bands of calcium dialuminate CaAl4O7 positioned at about 420, 442, 538, 573, 810, 834, 864, 918 and 943 cm−1. The results are in very good agreement with the data presented in Ref. [22, 23]. These bands appear to coincide with some bands seen in the infrared spectrum of CaAl2O4. Similarly, IR-active fundamental band of calcium zirconate [25] is not clearly seen in this spectrum only showing up as a inflection point at 523 cm−1.

Fig. 5
figure 5

FT-IR spectra of the 25Al2O3–75C7A3Z, 50Al2O3–50C7A3Z and 75Al2O3–25C7A3Z composites obtained via sintering at 1500 °C for 3 h

The 75A–25C7A3Z sample gives an extremely complicated spectrum (Fig. 5) due to the coincidence of positions of some bands in the IR spectra of CaAl12O19, α-Al2O3 and CaAl4O7. Thus, the appearance of intense absorption bands at 606 and 638 cm−1 may serve to identify the corundum being the representative compound, since its structure is built up by AlO6 octahedra only [21]. In contrast, structure of calcium hexa-aluminate is based upon the structural arrangements where Al is distributed over three symmetrically independent octahedral sites, one tetrahedral site, and one trigonal bipyramidal site [26]. Thus, by following the approach given in the paper [21] that the characteristic frequency ranges are as follow: “condensed” AlO4 tetrahedra 700–900 cm−1, “isolated” AlO4 tetrahedra 650–800 cm−1, “condensed” AlO6 octahedra, 500–680 cm−1 and “isolated” octahedra AlO6 400–530 cm−1, the specific feature of the IR spectrum of CaAl12O19 is the presence of IR bands at about 400, 467, 530, 553, 619, 656, 702, 741, 788 and 945 cm−1 marked in the 75A–25C spectrum (Fig. 5). The peak positions of these bands are in very good agreement with those presented in Ref. [26]. However, it is important that some peak positions of CaAl12O19 coincide with the peak positions observed for the CaAl4O7. Characteristic IR absorption frequencies of CaAl4O7 were found within a very broad absorption peak located in the region between about 800 and 900 cm−1. IR spectrum recorded for the prepared 75A–25C7A3Z sample is dominated by characteristic infrared absorption bands of functional groups belonging to some calcium aluminates, and it shows no significant band distinctive of cubic, tetragonal and monoclinic polymorphs of zirconia having their strong maxima at 480, 510 and 515 cm−1, respectively [27].

Microstructural evolution

The mechanism of the high-temperature reactions between Ca7ZrAl6O18 and Al2O3 was investigated based on the microstructure observations and chemical analysis with SEM–EDS. The BSE images of the polished cross sections of 25A–75C7A3Z and 75A–25C7A3Z composites sintered at 1300, 1400 and 1500 °C are shown in Figs. 6, 7. From these figures, it can be seen that during the first reaction stage at 1300 °C an attack of the alumina ions along the grain boundaries of calcium zirconium aluminate (Ca7ZrAl6O18) occurred and calcium monoaluminate (CaAl2O4) and calcium zirconate (CaZrO3) (Figs. 6 and 7, point 2) were then formed. This process proceeds according to the general balanced chemical reaction 1. Generally, high-temperature calcium zirconium aluminate (Ca7ZrAl6O18) decomposition in the presence of Al2O3 and the further phase transformations can be explained by the phase diagram of the ternary system CaO–Al2O3–ZrO2 [7, 9]. Let us consider the line joining the points represented by the composition of Ca7ZrAl6O18 (CaO 47.8 mass%, Al2O3 37.2 mass%, ZrO2 15.0 mass%) and Al2O3 100 mass%. Thus, when coarse-grained Ca7ZrAl6O18 is combined and sintered with Al2O3 micropowder, the original Ca7ZrAl6O18 grains will be rimmed by a reaction zone. The composition of this reaction zone could be expected by any point of the line Ca7ZrAl6O18–Al2O3 within the ternary system CaO–Al2O3–ZrO2. This means that the starting powders will go through a series of chemical reactions before the final stable compounds are formed, with the intermediate compounds being non-equilibrium liquid or solid transient phases. This process would start with reaction 1 and proceed according to the general balanced chemical reactions 35.

Fig. 6
figure 6

SEM micrographs of the studied 25A–75C7A3Z composite sintered at 1300 °C (1 h), 1400 °C (1 h) and 1500 °C for 3 h (polished cross sections and fractured cross section). EDS analysis in spots: 1—Ca7ZrAl6O18, 2—CaZrO3 embedded in CaAl2O4 matrix, 3—Al-rich CaAl2O4, 10—Ca12Al14O33, 11—CaAl2O4 and 12—coarse-grained CaZrO3

Fig. 7
figure 7

SEM micrographs of the studied 75A–25C7A3Z composite sintered at 1300 °C (1 h), 1400 °C (1 h) and 1500 °C (1 and 3 h). EDS analysis in spots: 1—Ca7ZrAl6O18, 2—CaZrO3 embedded in CaAl2O4 matrix, 3—Al-rich CaAl2O4, 4—CaAl4O7, 5—Al2O3, 6—CaZrO3 (coarse-grained) or ZrO2 (fine-grained) embedded in CaAl4O7, 7—ZrO2 embedded in CaAl4O7 matrix, 8—CaAl4O7 and 9—CaAl12O19

Due to the diffusion phenomena, the calcium zirconium aluminate grain was altered, having become mainly enriched in zircon and reduced in calcium. At the same time, Zr4+ ions may be considered rather as immobile during this process. Calcium zirconate, CaZrO3, formed during reaction 1 tends to agglomerate in large oval grains embedded in calcium aluminate matrix (xCaO·yAl2O3). Let us consider 25 mass% Al2O3–75 mass% Ca7ZrAl6O18 sample where zirconium migration comes rather from the presence of liquid transient phase that produced the fully reacted ceramic composite. Another is to use starting powders wherein the proportion of ingredients in the starting dry mixture is as follows: 75 mass% Al2O3 and 25 mass% Ca7ZrAl6O18 that will pass through a solid stage, rather than liquid, before the final microstructure is reached (Fig. 7, 1500 °C 3 h). Simultaneously, aluminum ions Al3+ move against a concentration gradient to increase their concentration inside Ca7ZrAl6O18 grain core. As a result of an intense counter-diffusion of calcium ions and aluminum ions through reaction zone, the chemical equilibrium could be reached. Nevertheless, the final composition of the three phases in equilibrium, i.e., Ca7ZrAl6O18–CaZrO3–CaAl2O4 and CaAl4O7–ZrO2–CaAl12O19 was not received as it was expected for both 25A–75C7A3Z and 75A–25C7A3Z, respectively. The investigation of the microstructural evolution within the 75A–25C7A3Z sample confirmed by XRD and FT-IR analysis has shown that the thermodynamic equilibrium was not reached due to the presence of both Al2O3 and CaAl4O7 non-equilibrium phases that could undergo of transition according to reaction 4. Hence, final microstructure of the 75A–25C7A3Z ceramics achieved at 1500 °C and 3 h sintering consist of ZrO2 particles embedded in the CaAl4O7 phase (Fig. 7, point 7) occurring in the initial region of Ca7ZrAl6O18 grain and CaAl12O19 (Fig. 7, point 9). Additionally, the 75A–25C7A3Z composite cement consists of alumina as checked by XRD and FT-IR.

Calorimetric studies

The heat of hydration was measured for the grounded cementitious materials (reference compounds and composite cements) whose particle size distribution known as the median diameter or the medium value of the particle size distribution was about 30 μm. A reference measurement of hydration heat of single component, i.e., CaAl2O4 (CA), CaAl4O7 (CA2), Ca12Al14O33 (C12A7) and Ca7ZrAl6O18 (C7A3Z) (Fig. 8) using isothermal heat flow calorimeter at 25 °C was a useful tool for both studying and assessing the hydration behavior of these cementitious compounds.

Fig. 8
figure 8

Heat flow curves of reference materials, i.e., CaAl2O4-, CaAl4O7-, Ca7ZrAl6O18-, Ca12Al14O33-based cementitious pastes prepared with w/s = 1.0 hydrated at 25 °C. Heat flow curves normalized to mass of binder paste (with in situ mixing)

The heat flow curves for the 25A_75C7A3Z, 50A_50C7A3Z and 75A_25C7A3Z hydrating multiphase cement pastes prepared with a water/solid ratio of 1.0 at 25 °C are shown in Fig. 9a–c, respectively. Looking at the overall results, some relationships have been noticed in observations. Firstly, the heat flow peaks of all composite cements are strongly enhanced by the presence of Ca-containing phases (Fig. 9a–c). It means an increase in the peak intensity of heat flow of hydration with decreasing the Al2O3/Ca7ZrAl6O18 mass ratio of the starting mixtures, as shown in Fig. 9a–c. Interestingly, intensity of the heat flow peak of composite cements belonging to the system CaO–Al2O3–ZrO2 hydration is also strongly reduced with the degree of conversion of Ca7ZrAl6O18 within the temperature range from 1300 to 1500 °C. As may be also seen, the hydration of all multiphase composite cements is accelerated with increasing the Al2O3/Ca7ZrAl6O18 mass ratio of the starting mixtures. The induction period was shortened, and the second exothermic heat effect was accelerated considerably as the Al2O3 amount increased. The influence of alumina on the hydration of Ca-aluminates is a phenomenon well known [28].

Fig. 9
figure 9

Heat flow curves of 25A_75C7A3Z (a), 50A_50C7A3Z (b) and 75A_25C7A3Z (c) pastes hydrated at 25 °C for w/s = 1.0, normalized to mass of binder paste (with in situ mixing). Cements obtained at 1300, 1400 and 1500 °C

Going into details, presented figures (Fig. 9a–c) illustrate the hydraulic activity of each cements, especially, the hydration kinetics of individual constituent phases and the mutual interaction of these constituents. Test results indicated that the phase composition of particular cements expressed in mass percent (Fig. 1a–c) of the possible constituent hydraulic phases, i.e., CaAl2O4 (CA), CaAl4O7 (CA2), Ca7ZrAl6O18 (C7A3Z) and Ca12Al14O33 (C12A7) and/or Al2O3 (A) strongly affects the microcalorimetric curve obtained in the isothermal conditions. Thus, these curves reflect the rate of heat evolution vs. time for each considered composite cements from the system CaO–Al2O3–ZrO2. It is well known that some stages of reaction can be identified on the calorimetric curve, namely: the pre-induction (the heat of wetting effect), induction (II), hydration rate increase (III), hydration rate decrease (IV) and low hydration rate (V), respectively. Thus, calcium monoaluminate hydration (Fig. 8) can be considered as good model of cement reaction with water. It is clearly visible that the shape of the heat flow curves of hydrating cement pastes is characteristic and the observed differences relate mainly to the stages generically, but each one can have different duration depending on the cement mineral composition. The all calorimetric curves shown in Fig. 9a–c exhibit one major feature, an intense first heat peak which occurs immediately after injection and mixing the water with the dry cement powders. This quick release of heat is associated with the wetting and initial dissolution of the cement particles within a few minutes after adding water. Besides, the rate evolution during hydration of the samples heat treated at 1300 °C (samples 75A_25C7A3Z, 50A_50C7A3Z, 25A_75C7A3Z) and 1400 °C (sample 25A_75C7A3Z) for this peak is strongly enhanced by the presence of the highly reactive calcium zirconium aluminate (Ca7ZrAl6O18) phase (Figs. 1a–c). The highly reactive nature of Ca7ZrAl6O18 has been extensively studied in recent years [9, 11].

Microcalorimetric curves of the 25A_75C7A3Z samples (Fig. 9a-a–c) are strongly dominated by the intense exothermic heat associated with CA (CaAl2O4) hydration. It is shown that C7A3Z (1300 °C_25A_75C7A3Z, 1400 °C_25A_75C7A3Z) exhibits a strong accelerating effect, more effective than C12A7 (1500 °C_25A_75C7A3Z), on the hydration of calcium monoaluminate (CA).

The hydration of 50A_50C7A3Z composite cements is characterized by a heat flow peak with dissimilar acceleration rate and intensity which follows the induction period. The hydration of 1300 °C_50A_50C7A3Z cement is characterized by a double heat flow peak with different rate and intensity. Thus, the different behaviors, as regards hydration kinetics, of CA and CA2 are indicated. The hydration reactions of calcium monoaluminate (CA) precede the reaction of CA2 with water which, on the other hand, has the ability to very slow and lengthy reaction. The length of the induction period of 50A_50C7A3Z samples hydration has been found to be related to both CA and CA2 contents and increased with increasing CA2 content within the temperature range from 1300 to 1500 °C. Therefore, it is appropriate to say that CA phase acts as an accelerating agent for the hydration of CA2 phase.

Microcalorimetric curve of the 1500 °C_75A_25C7A3Z sample (Fig. 9c-c.) is strongly dominated by the first stage, i.e., the initial wetting of the cement particles and by the second not intense exothermic heat (a weak diffuse exothermic effect) associated with CA2 (CaAl4O7) hydration. Analyzing the graphs of the composite cements heat treated for 1 h at 1300 and 1400 °C, I observe the well-known accelerating effect of other aluminate compound on the hydration of CA2. Increase in the reactivity of CA2 was accomplished by the presence of reactive phases combinations like CA/C7A3Z (Fig. 9c-a) and CA (Fig. 9c-b). Interestingly, the hydraulic behavior of both CA2 and CA is strongly enhanced by the presence of C7A3Z.

Discussion

In this paper, a new approach for synthesizing composite-designed CaO–Al2O3–ZrO2 cements via the high-temperature reactions within the Ca7ZrAl6O18–Al2O3 system in the temperature range between 1300 and 1500 °C was proposed. The reaction products formed in these samples were well understood by looking at compatibility triangles across the line joining Al2O3 and Ca7ZrAl6O18 of the CaO–Al2O3–ZrO2 phase diagram [7]. The configurations of three-phase regions in equilibrium are the following: I. Ca7ZrAl6O18–CaAl2O4–CaZrO3, II. CaAl2O4–CaZrO3–CaAl4O7, III. CaAl4O7–CaZrO3–ZrO2, IV. CaAl4O7–ZrO2–CaAl12O19 and V. CaAl12O19–ZrO2–Al2O3. The phase transformation in the prepared samples 25 mass% Al2O3–75 mass% Ca7ZrAl6O18, 50 mass% Al2O3–50 mass% Ca7ZrAl6O18 and 75 mass% Al2O3–25 mass% Ca7ZrAl6O18 having the initial phase composition in I, II and IV triangles, respectively, sintered in the temperature range 1300–1500 °C was investigated with X-ray diffraction (XRD), Rietveld XRD Quantification, infrared spectroscopy (IR) and scanning electron microscopy combined with energy-dispersive spectrometry (SEM/EDS) techniques. Final results from the various analytical techniques revealed that phase composition of the calcia–zirconia–alumina composite cements prepared via the high-temperature reactions between Ca7ZrAl6O18 and Al2O3 differs from the strategically designed sample composition due to sintering at different experimental conditions. It was concluded that the 75 mass% Al2O3–25 mass% Ca7ZrAl6O18 specimen was probably not fully equilibrated at 1500 °C due to the predominant diffusion path in solid-state reactions. The 25 mass% Al2O3–75 mass% Ca7ZrAl6O18 and 50 mass% Al2O3–50 mass% Ca7ZrAl6O18 specimens were fully equilibrated at temperatures of 1400 and 1500 °C, respectively.

Calcium zirconium aluminate (Ca7ZrAl6O18) was unstable when it was heated with alumina, and it was subjected to chemical reactions with the formation of calcium aluminates increasingly richer in aluminum oxide (Ca7ZrAl6O18 → CaAl2O4 → CaAl4O7 → CaAl12O19) and Zr-bearing compounds (CaZrO3 or ZrO2). Investigations by SEM showed that calcium zirconate tends to be uniformly distributed and as rounded oval grains within the calcium aluminate matrix with a high CaO/Al2O3 molar ratio corresponding to C12A7 (C≡CaO, A≡Al2O3) in the final microstructure. On the other hand, fine-grained zirconia were concentrated in microareas of the initial Ca7ZrAl6O18 grains, and it was embedded in the Al-richer calcium aluminate matrix. Based on theoretical considerations on the reaction mechanism in the CaO–Al2O3–ZrO2 system and the presented results it may be supposed that the reactions at the single-grain level would proceed from the stage 0 to III, which is graphically presented in Fig. 10.

Fig. 10
figure 10

Schematic diagram of phase distribution sequence in the reaction zone of the binary mixture like Ca7ZrAl6O18–Al2O3

The hydration process of composite cements belonging to the system CaO–Al2O3–ZrO2 was characterized by the different intensity exothermic reactions. The shape of the heat flow curves resulted from calorimetric measurements reflects the cement hydration process. The proportion of the selected CaAl2O4, CaAl4O7, Ca12Al14O33, Ca7ZrAl6O18 or Al2O3 phases changes the shape of the heat flow curve, so the hydraulic reactivity of cement phases and their mutual influence on hydration behavior can be quantified.

Conclusions

Some general conclusions can be drawn based on the results of this work:

  1. 1.

    Reactive sintering process of Ca7ZrAl6O18 and Al2O3 micropowders is an effective way for production of novel composite cements belonging to the system CaO–Al2O3–ZrO2.

  2. 2.

    The designed compositions of these hydraulic binders can be achieved with factors such as: ratio by mass of Ca7ZrAl6O18 to Al2O3, time and sintering temperature.

  3. 3.

    The shrinking core model in which the particle size of Ca7ZrAl6O18 is being consumed by reaction with alumina can be applied to visualization of reactions involving changes in microareas of unreacted core.

  4. 4.

    Phase composition affects the hydration kinetics of hydraulic binders determined using isothermal calorimetry.

  5. 5.

    Due to this unique combination of properties, novel composite cements belonging to the system CaO–Al2O3–ZrO2 can be used in many types of unshaped refractory materials including castables.