1 Introduction

Titanium alloys attract the attention due to their high specific strength, corrosion resistance [1, 2] and biocompatibility [3]. However, their extensive use for civilian applications is still hindered by the high costs associated with complicated metallurgical processing [4]. Fortunately, the vacuum induction melting method can help to reduce the production costs of titanium alloys [5]. Additionally, the key challenge is to select a suitable refractory to contain the molten titanium alloys [6].

Common oxides, such as Al2O3 [7, 8], ZrO2 [9], CaO [10] and Y2O3 [11, 12], have demonstrated higher corrosion resistance in contact with titanium alloys than carbide, nitride and boride ceramics [13]. However, none of these materials proves to be totally inert against titanium alloy melts. Recently, a perovskite structure material, CaZrO3 [14, 15] or BaZrO3, shows good potential as an alternative refractory [16].

Functional thin CaZrO3 coatings obtained by centrifuging and spraying according to the replica technique improve the properties of titanium alloy casting molds [17]. Schafföner et al. [18] demonstrated that CaZrO3 exhibits good corrosion resistance in contact with titanium alloy melts. In addition, self-prepared shell molds based on CaZrO3 ceramic are better than commercial shell systems [19].

A study by Gao et al. [20] revealed that the shape-memory TiNi alloy prepared using the BaZrO3 crucible could achieve a higher shape-memory rate and fatigue life than that prepared using the graphite crucible. In addition, the TiAl alloy prepared using the BaZrO3 crucible exhibited a higher purity than that using the MgO and Al2O3 crucibles on the laboratory scale [21]. Furthermore, a pilot-scale preparation of TiAl alloy was also achieved in the BaZrO3 crucible [22].

In fact, the difficulties of the metallurgy of titanium alloys with a refractory crucible lie in the contamination of alloys and the erosion of the refractory crucible. These phenomena will get worse with the increase in titanium content in the alloys due to the increased chemical activity. The researchers found that the BaZrO3 refractory exhibited insufficient stability during melting the alloys with high titanium content, confirming serious oxygen contamination in the alloy and damage to the crucible [23]. It limits the further application of the BaZrO3 refractory. Thus, to improve the stability of BaZrO3 refractories, it is necessary to broaden its extent of melting titanium alloys.

Several composite molds, including Y2O3–Al2O3–MgO–CaO [24], BN–ZrO2 [25] and ZrO2–MxOy (M = La, Ce, and Nb) [26], have been proposed. In our previous study, we proposed an effective strategy for designing a dual-phase material using Y2O3 and BaZrO3 [27]. This material was able to maintain the stability of both BaZrO3 and Y2O3, and the dissolution of Y2O3 in BaZrO3 leads to the formation of a solid-solution phase BaZr1−xYxO3−δ, which improved the stability of the refractory. Kuang et al. [28] demonstrated that the penetration of alloy melts into the refractory is related to the wettability, and a theoretical calculation by Zheng et al. [29] indicated that the non-wettability performance of BaZrO3 refractory was better than that of Y2O3 refractory when in contact with TiAl alloy melts.

Although the previous study also revealed that the BaZrO3/Y2O3 refractory exhibited good performance in the process of melting the TiAl alloys [27]. However, the effect of Y2O3 on the phase composition and microstructure of BaZrO3 crucibles has not been thoroughly studied, and the activity of Ti in the TiAl alloy melt is insufficient to obtain a significant difference in the interaction between the refractory and TiAl alloys. In this study, the phase composition and microstructure of the surface and the cross-section of BaZrO3/Y2O3 crucible were carefully investigated, and the interaction mechanism between the BaZrO3/Y2O3 crucible and a high titanium content Ti2Ni alloy was further clarified in comparison with the BaZrO3 crucible. The basic research focused on the dual-phase refractory may help meet future demands for melting titanium alloys.

2 Experimental

2.1 Preparation of crucibles

In this study, two different types of crucibles made of fused BaZrO3 powder (size of  13 μm, Shandong Haoyao New Material Co., Ltd., Shandong, China) and commercial Y2O3 powder (purity of 99.9%, size of 5 μm, Sinopharm Chemical Regent Co., Ltd., Shanghai, China) were designed and produced. To prepare the BaZrO3/Y2O3 dual-phase crucible, according to previous research results [27], the ingredients were weighed carefully according to the molar ratio of BaZrO3 toY2O3 of 2:1. In order to mix the ingredients well, the raw materials were first mixed in alcohol, ball-milled on a ball mill for 8 h, and finally dried at 120 °C for 12 h. As shown in Fig. 1, the dried raw material was isostatically cold pressed for 3 min at a pressure of 150 MPa in a U-shaped crucible mold and pressed into a green body. Finally, the BaZrO3/Y2O3 dual-phase crucibles were sintered at 1750 °C for 6 h in a Si–Mo rod sintering furnace. In addition, the preparation conditions of the BaZrO3 crucible are the same as those of the BaZrO3/Y2O3 dual-phase crucible. The BaZrO3 crucible after sintering was 5.5 cm in width and 7 cm in height, and the Y2O3 crucible after sintering was 4.5 cm in width and 6 cm in height.

Fig. 1
figure 1

Schematic diagram of preparation route of crucibles and melting experiment

2.2 Erosion resistance test

Ti2Ni alloys were fabricated in a water-cooled copper induction furnace using a 2:1 molar ratio of titanium sponge (99.9%) to nickel plate (99.9%). The alloys were melted in a medium frequency vacuum induction furnace with the crucibles. Before the melting, the furnace chamber was evacuated down to 10−3 Pa and then backfilled with pure argon up to 0.06 MPa for three times. The alloys were melted in the crucible at 1600 °C for 3 min and then poured into a graphite crucible to obtain the alloy ingots. This melting process was repeated for three times, and each crucible is used three times with each time using a different master alloy prepared from the same batch. The detailed flowchart including the preparation of refractory crucibles and the master alloys is also shown in Fig. 1. The phase composition was analyzed by X-ray diffraction (XRD) (D8 Advance, Bruker) with nickel-filtered Cu Kα radiation. The microstructure analysis was examined using a scanning electron microscope (JSM-6700F) combined with the energy-dispersive spectrometry (EDS) equipment. Inductively coupled plasma atomic emission spectrometry (ICP) and LECO TC600 O/N analyzer were utilized to examine the alloy impurity element caused by the crucible refractory. The erosion layer was measured with the Ruler software, and an average was obtained after multiple measurements.

3 Results and discussion

3.1 Phase composition and microstructure of crucibles

Figure 2 shows the scanning electron microscopy (SEM) images of the surface and inner structure of the BaZrO3 and BaZrO3/Y2O3 crucibles after sintering at 1750 °C for 6 h. Combining the EDS results in Table 1 and Fig. 2a, the surface of the BaZrO3 crucible was composed of BaZrO3 and ZrO2. Our previous study revealed that ZrO2 could be formed as a second phase during the electric arc furnace process of BaZrO3 raw material due to the evaporation of BaO [27]. The backscattered electron (BSE) image shows that the amount of ZrO2 in the surface was little. After the grinding and polishing, the interior of the BaZrO3 crucible consisted of ZrO2 (grain C) and BaZrO3 (grain D), as shown in Fig. 2b, and this was consistent with the phase composition of the crucible surface. In addition, the number of pores in the surface was less than that in the interior. A study by Ahmed et al. [30] revealed that the growth of grains on the surface had more advantages due to easier elimination of pores and lower stress.

Fig. 2
figure 2

SEM image of surface of BaZrO3 crucible and EDS element mapping image for O, Zr, and Ba elements (a), interior of BaZrO3 crucible after grinding and polishing (b), SEM image of surface of BaZrO3/Y2O3 crucible and EDS element mapping image for O, Y, Zr, and Ba elements (c) and interior of BaZrO3/Y2O3 crucible after grinding and polishing (d)

Table 1 EDS results of grains A–H in Fig. 2

The surface of the BaZrO3/Y2O3 crucible in Fig. 2c showed that it consisted of some grains of different shapes, marked as grain E and grain F, respectively. EDS results in Table 1 revealed that these two grains are Ba1−xZr1−yYyO3−δ and Y2O3(ZrO2). To confirm the distribution of Ba element in the surface of the BaZrO3/Y2O3 crucible, an additional element mapping analysis is shown in the inset in Fig. 2c, which indicated that only 3 at.% Ba was detected, confirming the occurrence of evaporation of BaO on the surface during the sintering. As shown in Fig. 2c, the main phase of the surface of the BaZrO3/Y2O3 crucible is Y2O3(ZrO2), and only a small number of Ba1−xZr1−yYyO3−δ phase distributed in it. Additionally, no pores were observed on the surface, indicating that the densification of the surface in the BaZrO3/Y2O3 crucible was better than that in the BaZrO3 crucible. Figure 2d shows the interior of the BaZrO3/Y2O3 crucible, presenting that it also consisted of two kinds of grains, marked as grain G and grain H. EDS analysis in Table 1 exhibits that they were Y2O3(ZrO2) and BaZr1−xYxO3−δ, resulting from the solution of ZrO2 in Y2O3 and the solution of Y2O3 in BaZrO3, respectively. Although some pores appeared in the interior of the BaZrO3/Y2O3 crucible, the number of pores was significantly lower by comparing with the interior in the BaZrO3 crucible. It confirmed that the Y2O3 addition was beneficial to the densification of the BaZrO3 crucible.

Figure 3 illustrates the XRD patterns of the surface of the BaZrO3 and BaZrO3/Y2O3 crucibles before and after grinding and polishing. In Fig. 3a, it can be seen that the surface of the BaZrO3 crucible without grinding and polishing mainly consisted of BaZrO3. After grinding and polishing, the amount of second phase ZrO2 was significantly increased, confirmed by the analysis in Fig. 2a, b. During the sintering, the low surface stress was conducive to the growth of the main phase BaZrO3 grains on the surface, which also influenced the sintering of second phase ZrO2 [31]. Figure 3b shows that the BaZrO3/Y2O3 crucible consisted of a mixture of BaZrO3 and Y2O3. As shown in Fig. 3c, the magnified XRD pattern of the crucible over the 2\(\theta\) range of 27°–37° exhibited a significant peak shifting of BaZrO3 toward lower angles. It was because the Y3+ ions (0.90 Å) could replace the Zr4+ ions (0.72 Å) in the crystal lattice of BaZrO3, leading to the formation of BaZr1−xYxO3−δ solid solution [32]. According to the Bragg’s law 2dsin\(\theta\)=, the increase in d value would cause a decrease in \(\theta\) value. Therefore, it can be concluded that the Y2O3 peak shifts to higher angles due to the substitution of Zr4+ ions by Y3+ ions. In addition, it can be seen from the phase diagram in Fig. 4 that ZrO2 could dissolve in Y2O3 to form α-Y2O3. Especially for the BaZrO3/Y2O3 crucible, the surface of the crucible before grinding and polishing had a small amount of main phase BaZrO3, and it was different from the interior of the crucible after grinding. Thus, it can be concluded that the solid solution effect between BaZrO3 and Y2O3 was beneficial to the evaporation of BaO [33], which was consistent with the phenomenon in Fig. 2c, d. The theoretical reactions could be used to describe the solid solution between BaZrO3, ZrO2 and Y2O3 below.

Fig. 3
figure 3

XRD patterns of crucibles before and after grinding and polishing. a BaZrO3 crucible; b BaZrO3/Y2O3 crucible; c magnified image over 2θ range of 27°–37° in b

Fig. 4
figure 4

YO1.5–ZrO2 binary diagram. T—Temperature; x(ZrO2)—content of ZrO2; Css—cubic structure; Tss—tetragonal structure

$${\mathrm{BaZrO}}_{3}+{\mathrm{Y}}_{2}{\mathrm{O}}_{3}\xrightarrow{\text{sintering}, {1750{\;}^\circ} \mathrm{C}}{\mathrm{BaZr}}_{1-x}{\mathrm{Y}}_{x}{\mathrm{O}}_{3-\delta }$$
(1)
$${\mathrm{Y}}_{2}{\mathrm{O}}_{3}+{\mathrm{ZrO}}_{2}\xrightarrow{\text{sintering}, 1750\;^\circ \mathrm{C}}{\mathrm{Y}}_{2}{\mathrm{O}}_{3}({\mathrm{ZrO}}_{2})$$
(2)
$${\mathrm{BaZrO}}_{3}+{\mathrm{Y}}_{2}{\mathrm{O}}_{3}\xrightarrow{\text{sintering}, 1750\;^\circ \mathrm{C}}{\mathrm{Ba}}_{1-x}{\mathrm{Zr}}_{1-y}{\mathrm{Y}}_{y}{\mathrm{O}}_{3-\delta }+x\mathrm{BaO}$$
(3)

Figure 5 depicts the cross-section of the BaZrO3 and BaZrO3/Y2O3 crucibles before the melting. From Fig. 5a, the crucible had a two-phase structure, consisting of light gray (spot 1) and black gray (spot 2) substances. EDS results in Table 2 show that they were BaZrO3 (spot 1) and ZrO2 (spot 2), respectively, confirming the analysis in the crucible surface. The element mapping in Fig. 5a shows that the two phases exhibited a good homogeneity. Additionally, some pores were significantly observed in the crucible. Figure 5b shows the cross-section of the BaZrO3/Y2O3 crucible, and the crucible also consisted of black (spot 3) and light gray (spot 4) substances. Being different from the BaZrO3 crucible in Fig. 5a, a large amount of the black gray substance was enriched along the inside wall of the BaZrO3/Y2O3 crucible. The analysis of the element mapping in Fig. 5b exhibited that a large amount of Y and Zr elements were distributed in the black gray substance, and the deficiency of Ba element in this area confirming the occurrence of its evaporation was discussed in Figs. 2c and 3b. In comparing with the BaZrO3 crucible in Fig. 5a, the Y- and Zr-rich layer had a small number of pores, indicating that it had a higher density. Additionally, this layer also had a higher densification than the crucible matrix. Thus, the Y2O3 addition also helped to improve the density of the BaZrO3 crucible due to the effect of solid solution sintering.

Fig. 5
figure 5

SEM images and elemental mapping of cross-section of BaZrO3 crucible (a) and BaZrO3/Y2O3 crucible (b) before melting

Table 2 EDS results of spots 1–4 in Fig. 5

3.2 Interaction between BaZrO3 crucibles and Ti2Ni alloys

Figure 6 illustrates the surface microstructure of the BaZrO3 and BaZrO3/Y2O3 crucibles before and after melting Ti2Ni alloys. Figure 6a displays that the surface of the BaZrO3 crucible before melting exhibited an even surface along with the appearance of some pores. From Fig. 6b, the surface of the BaZrO3 crucible after melting was eroded by the alloy melts, leading to the appearance of the inside grains in the crucible, and a large number of pores were also observed. This was due to the destruction of the grain boundaries on the crucible surface by the molten alloy, which then penetrated inside through the pores of the crucible, further damaging the internal structure. From Fig. 6c, the surface of the BaZrO3/Y2O3 crucible before the melting had a complete and dense structure. From Fig. 6d, a similar erosion phenomenon was also observed in the BaZrO3/Y2O3 crucible after the melting. However, a residual flat surface, which consisted of subsphaeroidal grains, appeared in the BaZrO3/Y2O3 crucible. Apparently, the surface of the BaZrO3/Y2O3 crucible was not totally damaged by the molten alloys, indicating that the Y2O3 addition can improve the erosion resistance of the BaZrO3 crucible in the process of melting titanium alloy.

Fig. 6
figure 6

SEM images of surface of crucibles before and after melting Ti2Ni alloys. a BaZrO3 crucible before melting; b BaZrO3 crucible after melting; c BaZrO3/Y2O3 crucible before melting; d BaZrO3/Y2O3 crucible after melting

Figure 7a displays the magnified images of area I in Fig. 6d, and some subsphaeroidal grains existed on the surface of the crucible. From the surface analysis in Fig. 7b, a uniform distribution of Y and Zr elements in the subsphaeroidal grains was observed, indicating that they were Y2O3(ZrO2) grains. In addition, BaZrO3 grains were mainly distributed inside the crucible from the distribution form of Zr and Ba elements. This suggests that the residual high-stability Y2O3 refractory formed a thin film in the crucible surface, which could effectively prevent the erosion by the alloy melts.

Fig. 7
figure 7

Magnified image of area I in Fig. 6d (a), combination of all elements (b), and EDS element mapping images for Y, Zr, Ba, and O elements (cf)

The XRD patterns of the BaZrO3 and BaZrO3/Y2O3 crucible surface before and after melting are shown in Fig. 8. From Fig. 8a, it can be seen that no other new phases were observed on the inner surface of the BaZrO3 crucible after the melting. Study by Chen et al. [16] indicated that the physical dissolution of the crucible refractory in titanium melts was the main interaction mechanism between the crucible and the alloy melt. However, the peaks of ZrO2 phase at 35.088°, 50.428° and 59.999° disappeared on the BaZrO3 crucible surface after melting. In addition, the intensity of BaZrO3 peaks at 30.238°, 43.264° and 53.687° also exhibited a decreasing trend than that of the peaks before the melting. The detailed peak values are listed in Table 3. It can be concluded that the molten alloys dissolved the second phase of ZrO2 on the crucible surface. Furthermore, the dissolution of BaZrO3 refractory caused the loose and uneven surface structure, leading to the decreased intensity of BaZrO3 peaks. The XRD result shows that no interaction product was generated in the BaZrO3/Y2O3 crucible surface after the melting. From Table 3, it can be seen that the intensity of BaZr1−xYxO3−δ peaks at 30.357°, 43.403° and 53.889° before the melting exhibited a decreasing trend than that of the peaks before the melting. However, the intensity of Y2O3 peaks at 29.599°, 34.269° and 49.157° exhibited an increasing trend after the melting. It appears that the molten alloy dissolved a significant amount of the BaZrO3 refractory in the surface, leaving behind residual Y2O3 refractory, as observed in Fig. 6. Hence, the residual Y2O3 refractory exhibited a larger XRD intensity than BaZrO3 refractory. According to the comprehensive analysis in Figs. 6 and 7, it can be concluded that the BaZrO3 refractory exhibited a priority erosion phenomenon than the Y2O3 refractory in the surface of the BaZrO3/Y2O3 crucible. The study by Oyama et al. [34] revealed that the thermodynamic stability of Y2O3 was higher than that of BaZrO3. It should be noted that this interesting phenomenon might be decided by the competition erosion behavior of refractory in molten titanium alloy. It meant that the refractory with lower stability would exhibit a priority erosion phenomenon. And the relative erosion degree of the refractory in the molten titanium alloys should be declined in the following sequence: ZrO2 > BaZrO3 > Y2O3.

Fig. 8
figure 8

XRD patterns of surface of crucibles before and after melting Ti2Ni alloys

Table 3 Intensity of characteristic peaks of different phases in surface of crucibles

To further investigate the performance of the BaZrO3 and BaZrO3/Y2O3 crucibles for melting Ti2Ni alloys, the microstructure images of the cross-section of the crucibles before and after melting are shown in Fig. 9. From Fig. 9a, the BaZrO3 crucible before melting is observed to have a complete structure with a light-yellow color. After melting, a contact layer with a gray color, with 3190 μm in thickness, was observed in the BaZrO3 crucible, as shown in Fig. 9b. The magnified SEM image of area II in Fig. 9b is shown in Fig. 9c. It can be seen that the crucible had a 485 μm thick erosion layer, confirmed by the loose and lamellar structure. During the melting, the alloy melt would erode the surface of the crucible and further permeate along the porosity in the crucible, resulting in the formation of the erosion layer. In addition, due to the capillary force, the alloy melt would further permeate into the crucible, resulting in a thicker contact layer. Figure 9d shows that the cross-section of the BaZrO3/Y2O3 crucible before the melting also presented a complete structure. After the melting, it can be seen from Fig. 9e that a contact layer with a thickness of 3750 μm appeared in the crucible. The enlarged view of area III in Fig. 9f shows that the thickness of the erosion layer in the crucible was only 63 μm with no appearance of cracks. Hence, it indicates that the BaZrO3/Y2O3 crucible exhibited a better erosion resistance to Ti2Ni alloy melts than the BaZrO3 crucible.

Fig. 9
figure 9

Images of cross-section of crucibles. a BaZrO3 crucible before melting; b BaZrO3 crucible after melting; c magnified picture of area II in b; d BaZrO3/Y2O3 crucible before melting; e BaZrO3/Y2O3 crucible after melting; f magnified picture of area III in e

Figure 10 depicts the SEM images of the cross-section of the erosion layer in the BaZrO3 and BaZrO3/Y2O3 crucibles. In Fig. 10a, it can be observed that the structure of the erosion layer in the BaZrO3 crucible was damaged by the alloy melts, resulting in dispersed grains. In addition, from Fig. 10a, there was no obvious ZrO2 phase in the erosion layer, indicating that the molten alloys dissolved a large amount of phase ZrO2, and it was in agreement with the analysis in Fig. 8. Figure 10b shows the microstructure of the erosion layer in the BaZrO3/Y2O3 crucible. Although the Y and Zr-rich layer in the crucible was damaged by the alloy melts, this erosion layer still exhibited a relatively dense structure in comparison with that in the BaZrO3 crucible. Furthermore, the BaZrO3/Y2O3 crucible showed excellent erosion resistance to the alloy melts due to the uniform distribution of the Y, Zr, and Ba elements in the erosion layer.

Fig. 10
figure 10

SEM images and element mapping of cross-section of BaZrO3 crucible (a) and BaZrO3/Y2O3 crucible (b) after melting

Figure 11 displays the measured concentration of O, Zr and Y elements in the alloys obtained from the BaZrO3 and BaZrO3/Y2O3 crucibles. From Fig. 11a, it can be seen that the oxygen concentration in the alloys melted in the BaZrO3 crucible decreased from 0.50 to 0.39 wt.%. In addition, the alloys melted in the BaZrO3/Y2O3 crucible also exhibited a decreased oxygen concentration (from 0.33 to 0.11 wt.%). Furthermore, as demonstrated in Fig. 11b, the Zr concentration in the alloys decreased from 0.640 to 0.635 wt.% after melting in the BaZrO3 crucible. However, the Zr concentration in the alloys melted in the BaZrO3/Y2O3 crucible also had a significant downward trend (from 0.350 to 0.075 wt.%). Compared to the BaZrO3 crucible, the BaZrO3/Y2O3 crucible showed less O and Zr contaminations in the alloys, indicating its superior erosion resistance to the Ti2Ni melts. The analysis in Fig. 6 supports this conclusion by suggesting that the remaining Y2O3 refractory in the crucible surface had excellent erosion resistance to the Ti2Ni melts, which prevented the further dissolution of the crucible refractory. Figure 11b shows that the Y concentration in the alloys obtained from the BaZrO3/Y2O3 crucible was 0.078, 0.130 and 0.130 wt.%, respectively. It was because there would be some Y contamination in the alloy melts due to the presence of Y2O3. Apparently, the low refractory contamination of the BaZrO3/Y2O3 crucible could reflect its good erosion resistance to the alloys.

Fig. 11
figure 11

Concentration of O element (a) and Zr and Y elements (b) in alloys melted in BaZrO3 and BaZrO3/Y2O3 crucibles

3.3 Improved mechanism of stability of BaZrO3/Y2O3 crucible

In our previous study [27], although the dissolution of the BaZrO3/Y2O3 refractory was mentioned in the melting of titanium alloys, the mechanism of its improved stability was not clarified. Generally, the interaction extent between the crucible and the titanium alloy was influenced by the thermodynamic stability of the crucible refractory and the interaction kinetics [35]. Specifically, the lower Gibbs free energy of the refractory and the higher density of the crucible meant better erosion resistance.

Figure 12a shows the dependence of the change in Gibbs free energy ∆G versus the temperature for the formation of Y2O3, BaZrO3, ZrO2, Ti and TiO2, which was obtained from the HSC software. Meanwhile, the Gibbs free energy of Y2O3 was more negative than that of BaZrO3, indicating that Y2O3 was more stable than BaZrO3. Figure 12b is about the Y2O3–ZrO2 phase diagram system, which shows the change of Gibbs free energy of α-Y2O3 phase with the amount of ZrO2 solid solution at 1750 °C. With the increase in ZrO2 solid solution, the energy of the whole system decreases, and the α-Y2O3 phase is more stable. Thus, it can be confirmed that the stability of the BaZrO3/Y2O3 crucible refractory was superior to that of the BaZrO3 crucible refractory. Furthermore, due to the effect of the solid solution sintering, the BaZrO3/Y2O3 crucible exhibited a high density than that of the BaZrO3 crucible, confirmed by the analysis in Figs. 2 and 5. Thus, the low porosity of the BaZrO3/Y2O3 crucible meant that it had a lower interaction rate. In summary, the BaZrO3/Y2O3 crucible had a thinner erosion layer and a lower contamination extent to the alloy melts in comparison with the BaZrO3 crucible.

Fig. 12
figure 12

Ellingham diagram of Y2O3, BaZrO3, ZrO2, Ti and TiO2 (a) and change in Gibbs free energy (G) in solid solution region of ZrO2–YO1.5 binary phase diagram (b)

Figure 13 presents a schematic of interactions between the Ti2Ni melt and the crucible during the melting process. The dissolution of the crucible refractory damaged the structure of the crucible, leading to crucible erosion. As shown in Fig. 13a, ZrO2 grains, which were on the surface of the BaZrO3 crucible, would be totally dissolved by the alloy melts. From Fig. 13b, BaZrO3 grains would be firstly dissolved, resulting in the residual Y2O3 grains on the surface of BaZrO3/Y2O3 dual-phase refractory crucible. In addition, Y2O3 grains with higher stability could effectively resist melt erosion. Thus, the BaZrO3/Y2O3 dual-phase refractory crucible demonstrated superior erosion resistance to Ti2Ni alloy melts than the BaZrO3 crucible.

Fig. 13
figure 13

Schematic diagram of interaction between crucibles and alloy melts. a BaZrO3 crucible; b BaZrO3/Y2O3 crucible

4 Conclusions

  1. 1.

    The BaZrO3/Y2O3 crucible demonstrated excellent resistance to the Ti2Ni alloy, with only a 63 μm thick erosion layer compared to the 485 μm thick layer in the fused BaZrO3 crucible.

  2. 2.

    After melting the alloys in the BaZrO3 crucible, the oxygen concentration in the alloys varied from 0.50 to 0.39 wt.%. However, the BaZrO3/Y2O3 crucible exhibited lower refractory contamination to the alloys, and the oxygen concentration varied from 0.33 to 0.11 wt.%.

  3. 3.

    The improved stability of BaZrO3/Y2O3 was attributed to the densification of the BaZrO3 crucible with the Y2O3 addition and the generation of phase Y2O3(ZrO2) with higher stability. It could reduce the interaction extent of the BaZrO3/Y2O3 crucible with Ti2Ni alloy melts.

  4. 4.

    In the melting process, BaZrO3 in the dual-phase refractory crucible would be firstly dissolved into the alloy melts, leading to the formation of a complete Y2O3 layer, which acted as a barrier to resist the erosion of the alloy melts.