Acessibilidade / Reportar erro

Glass Formation of Null-Matrix Cu-Hf-(Zr, Ti) Alloys

Abstract

An approach following the null-matrix, Ti/Zr ratio of 2.08, for neutron diffraction was used to calculate the glass formation in Cu-Ti-(Hf,Zr) alloys. The Cu55Zr14.6Ti30.4, Cu61.48Hf18.52Zr6.48Ti13.52, Cu58.1Hf16.9Zr8.1Ti16.9 and Cu52.46Hf14.2Zr10.8Ti22.54 compositions (at. %), were calculated and prepared by argon arc melting. Copper die suction casting was employed to produce conical shaped samples with diameters decreasing from 8 mm to 1 mm. X-ray diffraction and Neutron Diffraction were used to characterize the alloys with the aim of obtaining the critical glassy diameter, dc, of the alloys. The results showed the composition with the biggest dc was the Cu61.48Hf18.52Zr6.48Ti13.52 alloy, with dc = 6 mm. Thermal parameters were obtained by differential scanning calorimetry and the maximum values for glass transition temperature, Tg, (747 K), crystallization temperature, Tx, (772 K), solidus temperature, Tm, (1152 K) and liquidus temperature, Tl, (1230 K) corresponded to the Cu61.48Hf18.52Zr6.48Ti13.52 composition. The results also showed that the parameters obtained from thermal analysis did not correlate with the GFA obtained. However, a good correlation of GFA and dc was found by means of the topological model, where the highest % packing efficiency (53.57 %) was found for the Cu61.48Hf18.52Zr6.48Ti13.52 alloy, which also showed the highest dc value with 6 mm.

Keywords:
metallic glasses; rapid-solidification; calorimetry; neutron diffraction; X-ray diffraction; topological model


1. Introduction

In recent years, the research in metallic glass formation has been focused on Cu-Hf and Cu-Zr based bulk metallic glasses (BMG, typically referred to a critical casting thickness larger than 1 mm) due to its critical glassy diameter, dc, and the low critical cooling rate, Rc, required for vitrification. The glass forming ability, GFA, of binary alloys Cu100-xHfx (x=50, 45, 40, 35, 30 24 at.%) has been investigated and the maximum reliable, dc, obtained was 1 mm for the Cu65Hf35 alloy 11 Figueroa IA, Plummer JD, Lara-Rodriguez GA, Novelo-Peralta O, Todd I. Metallic glass formation in the binary Cu-Hf system. Journal of Materials Science. 2013;48:1819-1825.. The bulk glass formation in binary Cu-Zr system was studied for Cu100−xZrx (x=34, 36, 38.2, 40 at.%) by the copper mold casting method and the conclusion was that the GFA had a strong compositional dependence 22 Xu D, Lohwongwatana B, Duan G, Johnson WL, Garland C. Bulk metallic glass formation in binary Cu-rich alloy series - Cu100-xZrx (x = 34, 36, 38.2, 40 at.%) and mechanical properties of bulk Cu64Zr36 glass. Acta Materialia. 2004;52(9):2621-2624.. Besides, in case of the ternary Cu-Hf-Ti alloys, an exhaustive study has been carried out and the results showed an improvement in the critical diameter, dc. The highest dc obtained was 5 mm for the Cu56Hf25Ti19 and Cu55Hf26Ti19 alloy compositions 33 Figueroa IA, Báez-Pimiento S, Plummer JD, Novelo-Peralta O, Davies HA, Todd I. A detailed study of metallic glass formation in copper-hafnium-titanium alloys. Acta Metallurgica Sinica. 2012;25(6):409-419.. In addition, the glass formation and mechanical properties in the Cu-Zr-Hf-Ti, Cu-Zr-Ti and Cu-Hf-Ti systems for certain compositions were studied in detail 44 Inoue A, Zhang W, Zhang T, Kurosaka K. High-strengh Cu-based bulk glassy alloys in Cu-Zr-Ti and Cu-Hf-TI ternary systems. Acta Materialia. 2001;49(14):2645-2652.,55 Inoue A, Zhang W, Zhang T, Kurosaka K. Cu-based bulk glassy alloys with good mechanical properties in Cu-Zr-Hf-Ti system. Materials Transactions. 2001;42(8):1805-1812. On the other hand, it has been reported that additions of small amounts of B, Y and Si increased the GFA in these alloys 66 Figueroa IA, Rawal R, Stewart P, Carrol PA, Davies HA, Jones H, et al. Bulk glass formation and mechanical properties for Cu-Hf-Ti-M (M = B,Y) alloys. Journal of Non-Crystalline Solids. 2007;353(8-10):839-841.,77 Figueroa IA, Davies HA, Todd I. High glass formability for Cu-Hf-Ti alloys with small additions of Y and Si. Philosophical Magazine: A. 2009;89(27):2355-2368. Certainly, there is a compositional range in the Cu-Hf-Ti system in which BMG can be produced, thus, it is very important to propose a method for predicting the formation of BMG in specific compositions to facilitate the study of these materials. A number of thermal parameters have been proposed to determine the origin of GFA, i.e. the reduced glass transition temperature Trg = Tg/Tl (where Tg and Tl are glass transition and liquidus temperatures, respectively) 88 Turnbull D. Under what conditions can a glass be formed. Contemporary Physics. 1969;10(5):473-488., the supercooled liquid region, ΔTx = Tx- Tg (where Tx = crystallization temperature) 99 Inoue A. High strength bulk amorphous alloys with low critical cooling rates. Materials Transactions, JIM. 1995;36(7):866-875., and the parameters γ = Tx/(Tg+Tl) 1010 Lu ZP, Liu CT. A new glass-forming ability criterion for bulk metallic glasses. Acta Materialia. 2002;50(13):3501-3512., γm = (2Tx-Tg)/Tl1111 Du XH, Huang JC, Liu CT, Lu ZP. New criterion of glass forming ability for bulk metallic glasses. Journal of Applied Physics. 2007;101(8):086108. and β = (Tx*Tg)/(Tl-Tx)^2 1212 Yuan ZZ, Bao LB, Lu Y, Zhang CP, Yao L. A new criterion for evaluating the glass-forming ability of bulk glass forming alloys. Journal of Alloys and Compounds. 2008;459(1-2):251-260.. It has been proven that these parameters could help to predict the GFA for some alloys; however, they present limitations that do not make it suitable to be applied in all systems. Additionally, some structural parameters and topological criteria have been used to explain the high GFA of bulk glass-forming alloys. Egami and Waseda proposed the atomic strain criterion to predict the GFA for binary systems1313 Egami T, Waseda Y. Atomic size effect on the formability of metallic glasses. Journal of Non-Crystalline Solids. 1984;64(1-2):113-134. and Miracle and co-workers proposed another topological criterion that considers only the geometry of the packing of spheres of different diameters 1414 Miracle DB. The efficient cluster packing model - An atomic structural model for metallic glasses. Acta Materialia. 2006;54(16):4317-4336.,1515 Miracle DB. A structural model for metallic glasses. Nature Materials. 2004;3(10):697-702. to predict the best composition in a metallic system, in order to produce bulk metallic glasses. In the present work, following the null-matrix approach of Ti/Zr ratio = 2.08, for neutron diffraction, four new Cu-Ti-(Hf,Zr) BMG were calculated and experimentally produced. X-ray diffraction and Neutron Diffraction techniques were used to characterize the alloys in order to obtain the dc, of such alloys. Thermal parameters were obtained by deferential scanning calorimetry (DSC) and the % of Packing Efficiency was also calculated by means of a modified topological model 1515 Miracle DB. A structural model for metallic glasses. Nature Materials. 2004;3(10):697-702..

2. Experimental Procedures

The compositions of the alloys proposed were calculated, as mentioned above, keeping the Ti/Zr ratio equal to 2.08. The alloy compositions calculated are shown in Table 1.

Table 1
Alloys composition studied

The alloy ingots were prepared by melting the pure elements, i.e. Hf (99.8% pure), Ti (99.6 % pure), Cu (99.99 % pure) and Zr (99.2 % pure) with an atmosphere controlled arc furnace. In order to reduce the oxidation of pure elements, a Ti-gettered and high purity argon were used. Each ingot was re-melted at least five times with the aim of obtaining good chemical homogeneity. The alloy compositions represent the nominal values but the weight losses in melting were <0.1%. Copper die suction casting was employed to produce conical shaped samples with diameters decreasing from 8 mm to 1 mm. Ribbon glassy samples were produced by chill-block melt spinning in a sealed helium atmosphere at a roll speed of 25 m/s with an injection pressure of 0.4 bar and a nozzle orifice diameter of 0.8 mm. The gap between the crucible and the copper wheel was approximately 5 mm. The critical glassy diameter, dc, of the cast alloys samples was obtained by X-ray diffraction (XRD) using Co radiation (λ=1.78897 Å) in a diffractometter D5000, SIEMENS. Additionally, neutron diffraction scans were performed at 298 K up to 3 mm for the suction cast samples, with an incident neutron wavelength of ≈ 0.7 Å. The neutron diffraction patterns were taken with the samples in an evacuated bell jar (to avoid air scattering); plus a pattern of the empty bell jar to give the background scattering, and a cylindrical vanadium rod of 6 mm of diameter for normalization purposes. The glass transition (Tg), crystallization (Tx), solidus (Tm) and liquidus (Tl) temperatures were determined by differential scanning calorimetry (DSC) using a TA SDT Q600 calorimeter at a heating rate of 0.33 K/s in alumina crucibles. The DSC measurements were calibrated using a fresh zinc standard, giving an accuracy of ±0.2 K and ±0.02 mW.

3. Results and Discussion

The composition of the ternary Cu-Zr-Ti alloy here proposed has almost the same composition of the ternary Cu-Hf-Ti alloy reported in the literature, this is because the lattices parameter of Hf and Zr elements are approximately the same (a0 = 3.1946 Å, b0 = 5.0511 Å and c0 = 1.5811 Å for Hf; a0 = 3.2312 Å, b0 = 5.1477 Å and c0 = 1.5931 Å for Zr 1616 Pearson WB. A handbook of lattice spacings and structures of metals and alloys. Oxford: Pergamon Press; 1958., this should result in a closely isomorphous replacement. As mentioned in the experimental section, the ternary Cu-Zr-Ti and quaternary Cu-Hf-Zr-Ti alloys were calculated by keeping the Ti/Zr ratio = 2.08. Based on the above, and according to the reported alloys with high GFA 1, the resulting ternary composition with 55 at.% of Cu was the Cu55Zr14.6Ti30.4 alloy. For quaternary alloys, the same procedure was used to propose three compositions that might show a fully glassy phase, i.e. above 1 mm in diameter. A comparison of the curves obtained by means of neutron diffraction for all alloys is shown in Figure 1. The four alloys showed the characteristic features commonly observed in most metallic glasses, namely a sharp first peak, a second peak with a shoulder on the high Q side and well developed oscillations out to Qmax. This proved that the proposed alloys did form BMGs with dc ≥ 2 mm.

Figure 1
Neutron diffraction pattern for the alloys analyzed by neutron diffraction with dc ≥ 2 mm.

As the samples that were analyzed by neutron diffraction had diameters of 1, 2 and 3 mm, it was suspected that dc was larger. Therefore, X-Ray analysis was performed for samples of up to 8 mm in diameter, in order to obtain the dc. These results are shown in Figure 2. The proposed ternary Cu55Zr14.6Ti30.4 alloy composition is very similar to Cu55Hf15Ti30 3, both had a dc = 2 mm. In the case of quaternary alloys, the Cu61.48Hf18.52Zr6.48Ti13.52 alloy displayed a dc = 6 mm, which is the maximum value obtained for the investigated alloys. By increasing the amount of titanium (keeping the Ti2.08Zr1 ratio), it was observed that dc tended to decrease. This was observed for the Cu58.1Hf16.9Zr8.1Ti16.9 alloy with a dc = 5 mm and finally the Cu52.46Hf14.2Zr10.8Ti22.54 alloy with dc = 2mm (lowest dc value obtained).

Figure 2
XRD diffraction pattern from the cross section withs dc of the proposed alloy

The chemical mixing enthalpy (ΔHmix) was calculated based on the extended regular solution model using the reported values in reference 1717 Takeuchi A, Inoue A. Calculations of mixing enthalpy and mismatch entropy for ternary amorphous alloys. Materials Transactions, JIM. 2000;41(11):1372-1378.:

(1) H mix = i = 1 i j 3 Ω ij c i c j

Where Ωij is the regular solution interaction parameter between i-th and j-th elements, and is assumed to be constant in this case. ci,j is the composition of i and j elements. According to Miedema’s macroscopic model for binary liquid alloys 1818 Boer FR, Mattens WCM, Boom R, Miedema AR, Niessen AK. Cohesion in metals: Transition Metal Alloys. Netherlands: North-Holland; 1988., Ωij = 4(ΔHmix). The coefficient “4” is due to the definition at the equiatomic composition in a binary A-B system. Table 2 shows the ΔHmix for the proposed alloys and their respective dc. The values obtained do not follow a trend over dc, since the alloy with highest dc (6 mm), of the quaternary alloys, did not show the most negative value of ΔHmix, which corresponds to dc = 5 mm for the Cu58.1Hf16.9Zr8.1Ti16.9 allow with ΔHmix = -14.54 kJ mol-1. However, it can be considered that the ΔHmix values are very similar between the three alloys, despite the wide variation in dc values.

Table 2
Analyzed alloys with their dc and ΔHmix values

Since no correlation was found between the obtained dc and the chemical mixing enthalpy values, and in order to attempt to understand the GFA of the studied alloys, several parameters were calculated from the values obtained by the DSC. Figure 3, shows the DSC curves obtained for the investigated alloys. The magnitude of Tg, Tx, Tm and Tl parameters and some GFA indicators like ΔTx99 Inoue A. High strength bulk amorphous alloys with low critical cooling rates. Materials Transactions, JIM. 1995;36(7):866-875., Trg88 Turnbull D. Under what conditions can a glass be formed. Contemporary Physics. 1969;10(5):473-488., γ 1010 Lu ZP, Liu CT. A new glass-forming ability criterion for bulk metallic glasses. Acta Materialia. 2002;50(13):3501-3512., γm1111 Du XH, Huang JC, Liu CT, Lu ZP. New criterion of glass forming ability for bulk metallic glasses. Journal of Applied Physics. 2007;101(8):086108. and β 1212 Yuan ZZ, Bao LB, Lu Y, Zhang CP, Yao L. A new criterion for evaluating the glass-forming ability of bulk glass forming alloys. Journal of Alloys and Compounds. 2008;459(1-2):251-260. are shown in Table 3.

Figure 3
DSC curves of the studied alloys at a continuous heating rate of 0.33 K/min

Table 3
Critical diameters and thermal parameters of the studied alloys

Figure 4 shows the changes in ΔTx , Trg, γ, γm , β parameters and dc as a function of Ti. Generally, the ΔTx parameter has been related to GFA 11 and can be expressed as GFA ~ ΔTx. However, for these alloys, at dc = 6 mm (the highest obtained value), the magnitude of ΔTx was the lowest. It was also observed that as dc values increased, ΔTx decreased. For that, the GFA of the Cu-Hf-Zr-Ti proposed alloys is not reflected by the ΔTx parameter. On the other hand, the minimum values for the γ, γm, β parameters corresponded to a dc = 6 mm. The highest values of these parameters were obtained for the alloy with dc = 5 mm and finally, these values decreased down to dc = 2 mm. Therefore, the γ, γm, β parameters did not show a significant correlation with GFA. On the other hand, the Trg parameter did show the best correlation with the GFA, because, unlike the previous ones, the lowest value obtained was closely related to the lowest dc value (2 mm) and at dc = 5 mm and dc = 6 mm the values obtained for Trg are almost the same (0.610 and 0.607, respectively). In summary, the GFA of the proposed alloys is not well described by ΔTx, Trg, γ, γm and β parameters.

Figure 4
Changes in ΔTx, Trg, γ, γm, β parameters and dc as a function of Ti (at. %)

As the physical concept of null-matrix was used, an attempt to explain such correlating was also explored. Since positive neutron scattering length of a nucleus is the result of the repulsive potential of the neutron as it approaches the nucleus and the negative scattering length means the neutron is subjected to an attractive potential (a negative scattering length indicates that the Fermi pseudo-potential for the neutron scattering is attractive). The existence of positive and negative scattering lengths for atomic nuclei is somehow helping to improve the GFA of the alloys investigated. With these results, it was not possible to explain the GFA in terms of negative or positive scattering lengths. This suggests that the topological and particularly the chemical order in these BMGs will be a major influence on the viscosity and diffusivity in the melt, as well as the crystal nucleation and growth characteristics. If the partial pair distributions of these BMGs are obtained, these conjectures will be clarified (this results will be reported elsewhere in detail).

Using the topological model 1515 Miracle DB. A structural model for metallic glasses. Nature Materials. 2004;3(10):697-702., in order to understand the GFA, the preferred values R* were calculated by means of the ratio (R) of the solute atom radius (x) to the solvent atom radius (y), R=x/y. In the quaternary alloys, the atomic radii of the component atoms are: Cu = 0.127 nm, Ti = 0.146, Zr = 0.158 nm and Hf = 0.167 nm 1414 Miracle DB. The efficient cluster packing model - An atomic structural model for metallic glasses. Acta Materialia. 2006;54(16):4317-4336.. Therefore RZr/Cu = 1.24, RTi/Cu = 1.15 and RHf/Cu = 1.31. With these values, the dimensions of the cluster lattice unit cell length (Ao) were calculated. Since the chemical composition of the alloys is already defined, and using Zr as the basis for calculation, the numbers of atoms per cluster lattice unit cell and the cluster-packing factor were calculated. Table 4 shows the obtained values for the proposed alloys and the Cu70.99Zr14.51Hf7.25Ti7.25 alloy 1919 Soto CEN, Vargas IAF, Velázquez JRF, Rodriguéz GAL, Matínez JAV. Composition, elastic property and packing efficiency predictions for bulk metallic glasses in binary, ternary and quaternary systems. Materials Research. 2016;19(2):285-294. calculated by means of the topological model, which according to the same, it would be the alloy with the maximum packing factor for the Cu-Hf-Zr-Ti alloy family. The criterion of having Zr as a base element is due to the ratio of number of atoms in the alloy was greater than 1. These results could provide a possible explanation of the experimentally obtained dc values, as the highest value of packing efficiency percentage, coincides with the largest dc value. Although the viscosity of the alloys, just above the liquidus (Tl) temperature was not obtained, it is thought that the alloy that showed the highest percentage of packing efficiency would have the highest viscosity value. If the magnitude of viscosity were high enough to difficult the movement and nucleation of atoms, the GFA of any alloy would be greatly enhanced.

Table 4
Number of atoms per cluster lattice cell and packing efficiency % of the calculated alloys

4. Conclusions

By means of the null-matrix method it was possible to calculate the formation of BMG in the Cu-Ti-(Zr-Hf) alloy family. The Cu61.48Hf18.52Zr6.48Ti13.52 alloy showed the largest experimentally, dc, with 6 mm. The Cu58.1Hf16.9Zr8.1Ti16.9 and Cu52.46Hf14.2Zr10.8Ti22.533 alloys also showed a fully glassy phase, dc, of 5 and 2 mm, respectively. The GFA could not be explained in terms of the proposed thermal parameters neither by chemical mixing enthalpy, as they did not follow a trend as a function of dc. The high glass formation of these alloys was attributed not only to their proximity to the quasi-ternary eutectic point, but also to the calculated cluster packing efficiency. The Cu61.48Hf18.52Zr6.48Ti13.52 alloy showed the largest dc (6 mm) and the highest packing efficiency (53.57%). These results showed that the experimentally obtained dc value coincides with the highest calculated packing efficiency percentage.

5. Acknowledgements

The authors would like to acknowledge the financial support from DGAPA-PAPIIT UNAM “IN102319” for funding the project. C. Flores, G. A. Lara-Rodriguez, F. Silvar, A. Tejeda, O. Novelo, C. Ramos, R. Reyes, A. Lopez V., J. M. Garcia and F. Garcia are also acknowledged for their technical support. “Por mi raza hablará el espíritu”.

6. References

  • 1
    Figueroa IA, Plummer JD, Lara-Rodriguez GA, Novelo-Peralta O, Todd I. Metallic glass formation in the binary Cu-Hf system. Journal of Materials Science 2013;48:1819-1825.
  • 2
    Xu D, Lohwongwatana B, Duan G, Johnson WL, Garland C. Bulk metallic glass formation in binary Cu-rich alloy series - Cu100-xZrx (x = 34, 36, 38.2, 40 at.%) and mechanical properties of bulk Cu64Zr36 glass. Acta Materialia 2004;52(9):2621-2624.
  • 3
    Figueroa IA, Báez-Pimiento S, Plummer JD, Novelo-Peralta O, Davies HA, Todd I. A detailed study of metallic glass formation in copper-hafnium-titanium alloys. Acta Metallurgica Sinica 2012;25(6):409-419.
  • 4
    Inoue A, Zhang W, Zhang T, Kurosaka K. High-strengh Cu-based bulk glassy alloys in Cu-Zr-Ti and Cu-Hf-TI ternary systems. Acta Materialia 2001;49(14):2645-2652.
  • 5
    Inoue A, Zhang W, Zhang T, Kurosaka K. Cu-based bulk glassy alloys with good mechanical properties in Cu-Zr-Hf-Ti system. Materials Transactions 2001;42(8):1805-1812.
  • 6
    Figueroa IA, Rawal R, Stewart P, Carrol PA, Davies HA, Jones H, et al. Bulk glass formation and mechanical properties for Cu-Hf-Ti-M (M = B,Y) alloys. Journal of Non-Crystalline Solids 2007;353(8-10):839-841.
  • 7
    Figueroa IA, Davies HA, Todd I. High glass formability for Cu-Hf-Ti alloys with small additions of Y and Si. Philosophical Magazine: A 2009;89(27):2355-2368.
  • 8
    Turnbull D. Under what conditions can a glass be formed. Contemporary Physics 1969;10(5):473-488.
  • 9
    Inoue A. High strength bulk amorphous alloys with low critical cooling rates. Materials Transactions, JIM 1995;36(7):866-875.
  • 10
    Lu ZP, Liu CT. A new glass-forming ability criterion for bulk metallic glasses. Acta Materialia 2002;50(13):3501-3512.
  • 11
    Du XH, Huang JC, Liu CT, Lu ZP. New criterion of glass forming ability for bulk metallic glasses. Journal of Applied Physics 2007;101(8):086108.
  • 12
    Yuan ZZ, Bao LB, Lu Y, Zhang CP, Yao L. A new criterion for evaluating the glass-forming ability of bulk glass forming alloys. Journal of Alloys and Compounds 2008;459(1-2):251-260.
  • 13
    Egami T, Waseda Y. Atomic size effect on the formability of metallic glasses. Journal of Non-Crystalline Solids 1984;64(1-2):113-134.
  • 14
    Miracle DB. The efficient cluster packing model - An atomic structural model for metallic glasses. Acta Materialia 2006;54(16):4317-4336.
  • 15
    Miracle DB. A structural model for metallic glasses. Nature Materials 2004;3(10):697-702.
  • 16
    Pearson WB. A handbook of lattice spacings and structures of metals and alloys Oxford: Pergamon Press; 1958.
  • 17
    Takeuchi A, Inoue A. Calculations of mixing enthalpy and mismatch entropy for ternary amorphous alloys. Materials Transactions, JIM 2000;41(11):1372-1378.
  • 18
    Boer FR, Mattens WCM, Boom R, Miedema AR, Niessen AK. Cohesion in metals: Transition Metal Alloys Netherlands: North-Holland; 1988.
  • 19
    Soto CEN, Vargas IAF, Velázquez JRF, Rodriguéz GAL, Matínez JAV. Composition, elastic property and packing efficiency predictions for bulk metallic glasses in binary, ternary and quaternary systems. Materials Research 2016;19(2):285-294.

Publication Dates

  • Publication in this collection
    13 Jan 2020
  • Date of issue
    2019

History

  • Received
    28 June 2019
  • Reviewed
    29 Aug 2019
  • Accepted
    18 Sept 2019
ABM, ABC, ABPol UFSCar - Dep. de Engenharia de Materiais, Rod. Washington Luiz, km 235, 13565-905 - São Carlos - SP- Brasil. Tel (55 16) 3351-9487 - São Carlos - SP - Brazil
E-mail: pessan@ufscar.br