Next Article in Journal
The Effect of Thickness on Strength of Single Lap Orbital Riveted Aluminum/Composite Joints Used in Marine Environments
Next Article in Special Issue
Self-Diffusion Coefficients of Components in Liquid Binary Alloys of Noble Metals
Previous Article in Journal
Preparation and Heat Dissipation Properties Comparison of Al and Cu Foam
Previous Article in Special Issue
Enhanced Absorption Performance of Dye-Sensitized Solar Cell with Composite Materials and Bilayer Structure of Nanorods and Nanospheres
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Copper on the Formation of L12 Intermetallic Phases in Al–Cu–X (X = Ti, Zr, Hf) Alloys

1
Laboratory of Metallurgical Melts, Institute of Metallurgy of the Ural Branch of the Russian Academy of Sciences, 620016 Yekaterinburg, Russia
2
Department of Foundry Engineering and Strengthening Technologies, Faculty of Metallurgy and Metallurgical Science, Ural Federal University, 620002 Yekaterinburg, Russia
*
Author to whom correspondence should be addressed.
Metals 2022, 12(12), 2067; https://doi.org/10.3390/met12122067
Submission received: 3 November 2022 / Revised: 22 November 2022 / Accepted: 25 November 2022 / Published: 30 November 2022

Abstract

:
Transition metal trialuminides of the Al3X type of groups 4 and 5 of the periodic system have reduced density, high melting points, and corrosion resistance. Aluminides with a cubic lattice of the Al3Sc type can be used as a nucleating phase for aluminum alloys. However, low plasticity and a tetragonal lattice limit their application. In this work, we stabilized the metastable cubic lattice of Al3X-type aluminides by replacing aluminum with copper. The conditions for the formation of L12 metastable aluminides in the Al–Cu–TM (TM: Ti, Zr, Hf) alloys were studied using a wide range of copper concentrations. A high concentration of copper (hypereutectic alloys) is the one of the necessary conditions for the formation of (Al1−xCux)3Ti, (Al1−xCux)3Zr, (Al1−xCux)3Hf aluminides. With an increase in the copper concentration, the number of metastable aluminides sharply increased. The process of their formation strongly depended on the sequence of dissolution of the corresponding components in the melts. The low volume fraction of precipitated titanium aluminides was the result of insufficient supersaturation of α-Al with titanium (at the peritectic temperature) compared to that for alloys with zirconium and hafnium. Under identical synthesis conditions in the crystal lattice of metastable aluminides formed in experimental Al–Cu–Ti, Al–Cu–Zr, Al–Cu–Hf alloys, copper was found to substitute up to 8, 10, and 13 at.% of aluminum, respectively. The crystallographic and dimensional similarities of the lattices in metastable transition metal aluminides and in α-Al suggest their usefulness as modifying additions in aluminum-based alloys.

1. Introduction

High-tech industries require the search for new materials with improved technological properties. Group 4 and 5 transition metal (TM) trialuminides have attracted much attention due to their low density, high melting temperatures and oxidation resistance [1,2,3,4,5]. Aluminides of the Al3X type are considered as independent materials [3,5,6,7,8] or as reinforcing additions in aluminum alloys [4,9,10,11,12,13,14]. The authors of [15,16,17,18] continue to study the synthesis conditions and properties of metal-intermetallic laminate composites, where Al3X intermetallic compounds serve as one of the layers. Another practical application of Al3X aluminides is their use as modifying additions for grain refinement of aluminum alloys [19,20,21,22]. The Al3X intermetallic compounds obtained under equilibrium conditions have a tetragonal structure (D022 or D023) and low plasticity, which limits their application. One of the ways to increase the plasticity of aluminides is to transform the tetragonal lattice into a cubic one.
It is known [3,5,7,23,24,25,26] that the cubic lattice of aluminides can be stabilized by adding period 4 elements (Cr, Mn, Fe, Co, Ni, Cu, Zn.). This increases the plasticity of aluminides and, accordingly, improves the elastic properties of aluminide alloys. Such alloys are usually produced either by mechanical alloying or by melting in an electric arc furnace under vacuum with a non-consumable electrode and crystallization on a water-cooled copper hearth [3,23,24,25,26,27]. Cast alloys containing L12 type TM aluminides are produced by high-temperature overheating followed by fast melt cooling [22,28,29,30]. Such alloys can be used as modifying master alloys to refine the grain structure of other alloys [31,32], since aluminides with the fcc lattice (L12-type) are especially attractive due to their high degree of similarity with the α-Al lattice. In addition, Al3X aluminides have a low density, a high elastic modulus and high melting temperatures. They are ideal strengthening phases in high-strength, heat-resistant aluminum-based alloys.
The effect of copper addition on the formation of L12 aluminides in ternary Al–Cu–X (X: Ti, Zr, Hf) alloys was studied. When analyzing the energy dispersive X-ray microanalysis (EDX) data, for the copper-containing TM aluminides formed in the alloys, the authors relied on calculated and experimental data [8]. In [8], it was shown that the substitution of copper instead of aluminum in the crystal lattice of aluminides is energetically more favorable than the substitution of TM with copper. The EDX analysis results were compared with the calculated and experimental data [8] on the smelting of ternary (Al,Cu)0.75Zr0.25 and (Al,Cu)0.75Ti0.25 aluminide alloys.

2. Materials and Methods

Smelting of alloys based on aluminum and copper TM additions (Ti, Zr, Hf) was carried out in a graphite crucible in the ERF-93 electric resistance furnace (Ural Scientific Center, Yekaterinburg, Russia) in an argon atmosphere. Synthesis was carried out at temperatures from 1200 to 1300 °C to ensure sufficient overheating of the melt above the liquidus. Commercial purity aluminum (99.7% purity), high purity titanium (99.95%) and zirconium (99.99%) crystal bars grown by the iodide process, as well as electrolytic hafnium (99.68%) and M00 oxygen-free copper (99.96%) were used for synthesis. All the reagents used in the work were produced by REAHIM Trading House LLC, Yekaterinburg, Russia. The smelting process was carried out in two ways.
The first one consisted of the simultaneous addition of M00 copper rods and a pre-synthesized master-alloy (containing TM) to the aluminum melt. In the second case, the copper was dissolved in an aluminum melt and then a master-alloy (containing TM) was added. In both cases, after the dissolution of all components, the melts were stirred for 30 min and poured into a bronze mold. The crystallization rate in the mold (dimensions: 100 × 80 × 10 mm) was ≈200 °C/s. The weight of the resulting ingots was ≈200 g. Samples for chemical and metallographic analysis were taken from the lower part of the ingot.
Metallographic analyses of the samples were performed using an GX-57 optical inverted microscope (OLYMPUS Corporation, Tokyo, Japan) (magnifications from 50× to 1500×) and EVO 40 scanning electron microscopes (Carl Zeiss AG, Oberkochen, Germany). All electron images presented in this work were obtained using a back-scattering electron (BSE) detector (Carl Zeiss AG, Oberkochen, Germany) (with high sensitivity to differences in atomic number). To determine the chemical composition of the matrix and TM aluminides, an energy dispersive X-ray microanalysis (EDX) was carried out using an INCA X-Act system (Oxford Instruments plc, Abingdon, Great Britain).
The chemical composition of the synthesized alloys was determined by inductively coupled plasma optical emission spectrometry using the Optima 2100DV. Diffraction studies were performed using an XRD-7000 X-ray diffractometer (Shimadzu Corporation, Kyoto, Japan) (CuKα-radiation, graphite monochromator, the accumulation of data during 1.8 s in increments of 0.03 degrees in the range of 20−80 degrees). Unit cell parameters were calculated with the east-square method using the positions (in 2θ) of 10 peaks.

3. Results and Discussion

The compositions of the studied Al–Cu–Ti, Al–Cu–Zr and Al–Cu–Hf ternary alloys varied with a range of copper concentrations, from hypoeutectic to hypereutectic. The main structural components of these alloys were the θ-phase Al2Cu (indicated by the green arrows in the Figures), the α-Al + θ eutectic (indicated by the blue arrows) and TM trialuminides with a cubic lattice (red arrows). The melting temperature was determined by the at.% of the transition metal in the corresponding Al-TM system and its melting temperature. Depending on the degree of overheating of the melts above the liquidus, either stable aluminides with a tetragonal (D022 or D023) lattice structure, metastable with a cubic (L12) structure, or both, were formed in the alloys. Since the aim of the study was to obtain metastable aluminides with a cubic lattice having the L12 structure in alloys, the table below lists the compositions of only those alloys in which aluminides of the L12 structural type were formed. The Table 1 lists the compositions of the alloys containing L12 aluminides. The composition of the aluminides was estimated from the results of 5–10 EDX spectra analyses. The degree of supersaturation of the TM alloys was determined from the ratio of the atomic fraction of the calculated critical C0 value (TM concentration at the temperature of peritectic transformation) and the TM fraction in the alloy.

3.1. Al–Cu–Ti System

Stable D022 intermetallic compounds were formed in the alloys synthesized at 1250 and 1300 °C, with a wide range of copper concentrations (from 9.98 to 32.24 wt.%—Hereinafter, unless otherwise indicated, wt.%—) and with a low titanium content. Only in the Al–41.91Cu–1.64Ti and Al–40.33Cu–2.75Ti alloys (T1 and T2), along with stable aluminides (needles and plates), single metastable dendrites and L12 cuboids (size is less than 7 μm) were formed; they were visible at high magnifications (see Figure 1). Their average composition (at.%) was 70.68Al–6.36Cu–22.96Ti. The stoichiometric composition of the aluminides can be written in the following form: (Al0.92Cu0.08)nTi. It can be seen that copper substituted 8 at.% of the aluminum in the L12 crystal lattice of titanium aluminide.
The compositions of the aluminides formed in the T1 and T2 alloys were compared with that of the L12 cast aluminide alloy (70.1Al–6.9Cu–23.0Ti) synthesized in [8], (at.%). It was shown that the differences in their compositions were in the tenths and hundredths of a percent. The stoichiometric composition of the aluminides can be written in the following form: (Al0.91Cu0.09)3Ti. It can be seen that copper substituted 9 at.% of the aluminum.
Note that in [8], two series of ternary aluminide alloys, i.e., Al–Cu–Ti and Al–Cu–Zr, with copper contents of 8 and 12 at.% were synthesized. Melting was carried out in order to determine the copper concentration, which stabilizes the metastable L12 lattice of titanium and/or zirconium trialuminides. In cast alloys, the formation of one or two additional phases was identified. Such phases disappeared after homogenizing annealing at 1000 °C for 72 h. In addition, after annealing, the scatter of data on the composition of the aluminide alloy decreased, approaching a stoichiometric one.

3.2. Al–Cu–Zr System

Two hypereutectic alloys, i.e., Al–38.78Cu–4.56Zr and Al–36.28Cu–9.02Zr (Z1 and Z2), differed in their zirconium content and were melted at 1200 °C. In the first alloy (Figure 2a), as a result of significant overheating of the melt, only metastable trialuminides were formed. They had a cubic (L12) structure and the following composition: 68.34Al–6.27Cu–25.39Zr. These intermetallics formed large (50 to 150 µm) clusters of fragmented cuboids and dendrites of various growth shapes. Note that the light inclusions visible on the intermetallic compounds were undissolved zirconium particles. In the second alloy, the Zr content was twice as high and, accordingly, the overheating was less. In this case, both structural types of aluminides were formed. Intermetallic particles having the L12 lattice and a composition of 65.39Al–9.05Cu–25.56Zr (Z2) formed clusters. The size of the clusters, like that of the aluminides themselves, was an order of magnitude smaller than in the Z1 alloy (see Figure 2).
It can be seen that the higher the atomic fraction of Zr in the alloy, the greater the number of aluminides and the smaller their size. The composition of intermetallic particles formed in two alloys of the Al–Cu–Zr system (Z1 and Z3) was compared. It was shown that with an increase in the concentration of Zr in the alloys from 1.81 to 3.67 at.%, the atomic fraction of Cu in the formed aluminides increased from 6.27 up to 9.05 at.%. The stoichiometric composition of these aluminides (Al0.92Cu0.08)3Zr and (Al0.88Cu0.12)3Zr was also analyzed. It can be concluded that with an increase in the atomic percentage of Zr, the substitution of aluminum with copper in the aluminides crystal lattice increased from 8 to 12 at.%, although the copper content in the alloy (Z3) was less than in (Z1). This once again confirmed that the composition and size of the formed L12 aluminides depended mainly on the atomic percentage of TM, in this case, zirconium.
The next two alloys, Al–31.47Cu–8.63Zr and Al–31.25Cu–8.37Zr (Z2 and Z4), having practically the same copper and zirconium compositions, were melted at 1300°C, but with a different sequence of copper and zirconium dissolution in aluminum melt. In the first alloy (the dissolution of copper in an aluminum melt, and addition of master-alloy containing zirconium after that), only metastable L12 trialuminides comprising 67.08Al–7.27Cu–25.65Zr or (Al0.90Cu0.10)3Zr were formed. The aluminides formed clusters (up to 100–150 µm in size) of particles of different shapes and sizes (up to 10–20 µm). In the second alloy (Figure 3b), along with a small amount of metastable L12 trialuminides, a lot of needles and plates with the D023 lattice structure and composition (Al0.97Cu0.03)3Zr were formed. In such particles, compared with L12 aluminides, copper substituted three times less aluminum. Such a difference in the structure of these alloys indicated a significant effect of the sequence of dissolution of copper and TM in the aluminum melt.
Formed in three cast alloys of the Al–Cu–Zr system, zirconium aluminides with L12 structure had an average composition of 66.94Al–7.53Cu–25.53Zr. This was close to the composition of the 67.90Al–7.50Cu–24.70Zr cast aluminide alloy, synthesized in [8]. In both cases, copper substituted up to 10 at.% of the aluminum in the crystal lattice of the aluminides. According to the data presented in [8], the cubic (L12) lattice parameter of this phase in the cast state was a = 0.4070 nm. This value is close to that measured in our work, a = 0.40662(7) nm for the 65.39Al–9.05Cu–25.56Zr phase formed in the Z3 alloy.

3.3. Al–Cu–Hf Yystem

Among the five alloys of this system, Al–29.73Cu–13.24Hf and Al–29.47Cu–13.69Hf (H1 and H2) had similar compositions, both in terms of their copper and hafnium concentrations; however, they differed in melting temperature and in the sequence of Cu and Hf dissolution in molten Al. The H1 alloy was synthesized at 1250 °C using the second method (first, copper was dissolved in the aluminum melt, then hafnium), and alloy H2 using the first method (conventional scheme), but at a temperature of 1300 °C (i.e., with greater overheating of the melt). In both alloys, as can be seen in Figure 4, mainly stable (plates and needles) hafnium aluminides and only a small number of metastable ones (clusters of cuboid particles) were formed.
Subsequently, 5 wt.% of Cu was added to the alloys, and melting was carried out at the same temperatures, i.e., 1250 °C and 1300 °C. As a result, stable and metastable hafnium trialuminides were formed in the alloys, but with a different ratio (see Figure 5). An increase in the copper concentration in the alloys changed the ratio of stable and metastable trialuminides in favor of the formation of a larger amount of the latter. This is clearly seen from a comparison of the particle growth shapes and sizes and the volume fraction of intermetallic particles precipitated in the alloys (Figure 4 and Figure 5).
An EDX analysis of the aluminides formed in the H4 and H3 alloys showed that the copper concentration in L12 intermetallic compounds was three times higher than that in the aluminides with the D022 structure. The chemical composition of the metastable aluminides in the investigated alloys differed insignificantly, amounting to 69.90Al–7.39Cu–22.71Hf and 69.03Al–7.67Cu–23.30Hf, respectively. The average stoichiometric composition can be written as (Al0.90Cu0.10)nHf. It can be seen that in the crystal lattice of the aluminides, copper substituted about 10 at.% of aluminum (for both alloys).
Finally, the hypereutectic H5 alloy with a high (3.50 at.%) hafnium concentration was synthesized at 1200 °C. Only metastable aluminides of composition (at.%) 67.09Al–10.03Cu–22.88Hf or (Al0.87Cu0.13)nHf were formed in this alloy. It can be seen that copper substituted 13 at.% of the aluminum. The distribution and growth patterns of L12 aluminides are clearly seen in Figure 6. According to the EDX data, Al2.5Cu0.5Hf aluminides with a cubic lattice (L12) a = 0.40441(5) nm formed in the H5 alloy. The XRD analysis results for the Al–36.28Cu–9.02Zr (Z3), Al–35.66Cu–15.36Hf (H5) alloys are shown in Figure 7.
Thus, in two alloys (Z1, Z2) of the Al–Cu–Zr system and in one (H5) of the Al–Cu–Hf system, only aluminides having the L12 structural structure were formed. The aluminides had compositions of (Al0.92Cu0.08)3Zr, (Al0.90Cu0.10)3Zr and (Al0.87Cu0.13)nHf, in which copper substituted 8, 10, and 13 at.% of aluminum, respectively. The composition of the aluminides was determined by the atomic fraction of TM in the alloys.

3.4. Formation of Trialuminides with a Cubic Lattice in Alloys of the Al–Cu–X (X: Ti, Zr, Hf) System

The aluminide phases having the L12 structure formed in the Al–Cu–X (X: Ti, Zr) alloys smelted in this work were compared with the L12 aluminide alloys (in the cast state) synthesized in [8]. It was shown that they were almost completely identical both in composition and in the percentage of substitution of copper in the aluminum crystal lattice. However, the volume fraction of Al3Ti particles precipitated from the α-Al solid solution (in two alloys of the Al–Cu–Ti system) turned out to be very low. This was the result of insufficient supersaturation with titanium, according to the classical theory of nucleation. To verify this fact, the following calculations were carried out:
1.
the nucleation rate per unit volume [33,34], which is a function of the diffusion activation energy (Q);
2.
the chemical driving force ( Δ F ch );
3.
the elastic strain energy ( Δ F el ), which reduces the chemical driving force ( Δ F ch ), preventing nucleation.
According to [35], the chemical driving force Δ F ch for nucleation of Al3X aluminides can be found from the following equation:
Δ F ch = RT V Al 3 X C Al 3 X C α 1 C α ln C 0 C α
where R = 8.314 (J/mol∙K) is the universal gas constant; C Al 3 X is the atomic fraction of TM in the Al3X intermetallic compound equal to 0.25; C 0 is the atomic fraction of TM in the α-Al solid solution; C α is the solubility of TM in the solid α-Al; and V Al 3 X is the molar volume of the Al3X phase, determined by the following equation:
V Al 3 X = N a a Al 3 X 3 4
where N a = 6.022 × 10 23 mol 1 is Avogadro constant and a Al 3 X is the lattice parameter of the Al3X phase.
The values of the cubic lattice parameter of the Al3Ti, Al3Zr, and Al3Hf trialuminides under consideration, obtained experimentally in [36], were 0.3967, 0.4080, and 0.4048 nm, respectively.
The solubility limit of TM in Al could be found from the following equation (according to [37]):
C α Al 3 X = exp 4 Δ G Al 3 X Δ G X kT = exp Δ S vib X 4 Δ S vib Al 3 X k × exp 4 Δ H Al 3 X Δ H X kT
where k = 1.38 × 10 23 J · K 1 is the Boltzmann constant; Δ G Al 3 X is the free energy of formation for Al3X aluminides; Δ G X is the free energy of formation of a single TM in Al (fcc); Δ S vib X is the vibrational entropy of 1 atom of TM; Δ S vib Al 3 X —vibrational entropy of Al3X (per 1 atom); Δ H Al 3 X is the formation enthalpy for Al 3 X per 1 atom; and Δ H X is the enthalpy of formation (per 1 atom) for TM impurity in Al.
Table 2 and Table 3 present literature data on the enthalpies and vibrational entropies included in Equation (3) for Al3X (X = Ti, Zr, Hf) compounds and TM atoms, as determined in [37] based on first-principles calculations.
The exponential dependences of C α Al 3 X on temperature for L12 trialuminides were obtained by substituting the tabular values into Equation (3).
C α Al 3 Ti = 2.334 × exp 4398.1 T
C α Al 3 Zr = 10.528 × exp 8285.6 T
C α Al 3 Hf = 8.602 × exp 7600.9 T
The values of chemical driving force Δ F ch were calculated by substituting the values determined from Equations (4)–(6) into Equation (1) (for a given T). The results are given in Table 4.
In reactions involving a solid solution, it is necessary to take into account the elastic strain energy because of the formation of a second phase, which is fully or partially coherent with the matrix. This causes stress on the lattices of the matrix and the precipitating intermetallic phase. The elastic strain energy, Δ F el , reduces the chemical driving force, and hence, the nucleation rate. Considering only volumetric strains (for isotropic materials), the strain energy per unit volume for a coherent particle can be calculated using Equation (7) [33,34].
Δ F el = 2 μ 1 + ν 1 ν δ 2
where for Al at ambient temperature the shear modulus µ = 25.4 GPa [38] and Poisson’s ratio ν = 0.345 [39].
The lattice parameters mismatch (δ) of L12 intermetallic compounds with the α-Al lattice (absolute value) was determined using Equation (8).
δ = 1 a A l 3 X a α Al × 100 %
where a α A l   = 0.40496 nm [36] is the lattice parameter of α−Al.
The value of δ was much larger for Al3Ti (δ = 2.04%) intermetallic compounds. For Al3Zr, it was significantly smaller (δ = 0.75%). The value of the lattice parameters closest to α-Al was characteristic of Al3Hf (δ = 0.04%) intermetallic compounds [36]. Using the known values of µ, ν and δ in Equation (7), for each of the systems, the values of Δ F el were determined to be 43.4, 5.9 and 0.017 MJ · m 3 for Al3Ti, Al3Zr and Al3Hf, respectively. Thus, the elastic strain energy preventing the nucleation for Al3Ti (L12) was 7.4 times higher compared to that for Al3Zr (L12) and was incomparably higher than that for Al3Hf (L12).
The critical value of C0 from Equation (1) is necessary to overcome the elastic strain barrier. This value can be calculated by equalizing the values of Δ F ch and Δ F el (at peritectic temperature). For the Al–Ti alloys, the C0 value was equal to 2.7 at.% Ti; it exceeded the C0 values for Zr and Hf, equal to 0.15 at.% and 0.23 at.%, by factors of 18 and 11.7, respectively. Thus, the Al–Ti system at the peritectic temperature required a significantly (at least by one order of magnitude) greater supersaturation with TM compared to the Al–Zr and Al–Hf systems.
Previously [40,41], it was shown that higher cooling rates are required to achieve a greater driving force of nucleation, which depends on the value of C0. Therefore, to achieve a high volume fraction of primary aluminides in the Al-Ti system, higher crystallization rates are required. As noted above, the rate of nucleation, in addition to Δ F ch and Δ F el , strongly depend on the Q value (the activation energy of the TM diffusion process in Al). According to [42], the Q value for Ti in α-Al is 260 kJ/mol, i.e., higher than for Zr (242 kJ/mol) and Hf (241 kJ/mol). Hence, low values of the chemical driving force Δ F ch , as well as high values of the elastic strain energy Δ F el and diffusion activation energy Q inhibit the nucleation process in the Al–Ti system. This fact explains the isolated cases (practically an absence) of the formation of Al3Ti aluminides having a cubic lattice in the Al–Cu–Ti alloys synthesized in this work in comparison with the formation of Al3Zr and Al3Hf in the Al–Cu–Zr and Al–Cu–Hf systems. It is noteworthy that the Al-Zr and Al-Hf systems had similar properties (to each other) in terms of diffusion. However, Al-Zr required almost two times less supersaturation, and hence, lower cooling rates to overcome the energy barrier associated with Δ F el .

4. Conclusions

Our investigation of the conditions for the formation of metastable aluminides in Al–Cu–Ti, Al–Cu–Zr and Al–Cu–Hf alloys with a wide range of copper concentrations made it possible to draw the following conclusions:
(1)
One of the necessary conditions for the formation of (Al1-xCux)3Ti, (Al1-xCux)3Zr, (Al1-xCux)3Hf aluminides having L12 structure is the hypereutectic concentration of copper in the alloys. Ceteris paribus, an increase in the copper concentration (above the eutectic) in the alloys rapidly changed the ratio of stable and metastable trialuminides in favor of the formation of a larger amount of the latter. The sequence of dissolution in liquid aluminum of copper and then of TM (Zr or Hf) significantly accelerated the process of formation of metastable aluminides in the alloys.
(2)
Under identical synthesis conditions in the crystal lattice of metastable aluminides formed in experimental Al–Cu–Ti, Al–Cu–Zr and Al–Cu–Hf alloys, copper substituted up to 8, 10, and 13 at.% of the aluminum, respectively. The amount of aluminum substituted with copper in the crystal lattice of metastable aluminides was three times higher than in stable ones.
(3)
In the studied Al–Cu–Ti alloys, the low volume fraction of precipitating aluminides was the result of insufficient supersaturation of α-Al with titanium at the peritectic temperature (according to calculations based on the classical nucleation theory). The degree of supersaturation of the investigated alloys with titanium was 2 and 3.5 times, with zirconium being in the range of 26 and 53 times, and hafnium 174 to 219 times. As such, the difference in the degree of supersaturation of the alloys was 2–3 orders of magnitude.
(4)
Al2.5Cu0.5Zr and Al2.5Cu0.5Hf aluminides having the L12 structure with lattice parameters a = 0.40662(7) nm (Z3) and a = 0.40441(5) nm (H5), respectively, were synthesized. The high degrees of crystallographic and dimensional similarity between the lattices of aluminides and α-Al (0.4050 nm) suggest the applicability of these intermetallic compounds as modifiers of aluminum alloys.

Author Contributions

Conceptualization, P.K.; methodology, E.P.; synthesis and investigations, P.K., S.P., A.S. and I.G.; theoretical analysis, I.G. and E.P.; resources, P.K., S.P. and A.S.; writing—original draft preparation, P.K. and E.P.; writing—review and editing, E.P. and I.G.; project administration, P.K. All authors have read and agreed to the published version of the manuscript.

Funding

This study was performed in terms of a state task of the Institute of Metallurgy of the Ural Branch of the Russian Academy of Sciences, using equipment of the Center for Collective Use «Ural-M».

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors have no conflict of interest to declare that are relevant to the content of this article.

References

  1. Gao, T.; Liu, X. Replacement with Each Other of Ti and Zr in the Intermetallics of Al–(Si–)Ti–Zr Alloys. J. Mater. Sci. Technol. 2013, 29, 291–296. [Google Scholar] [CrossRef]
  2. Karpets, M.V.; Milman, Y.V.; Barabash, O.M.; Korzhova, N.P.; Senkov, O.N.; Miracle, D.B.; Legkaya, T.N.; Voskoboynik, I.V. The influence of Zr alloying on the structure and properties of Al3Ti. Intermetallics 2003, 11, 241–249. [Google Scholar] [CrossRef]
  3. Nayak, S.S.; Murty, B.S. Synthesis and stability of L12–Al3Ti by mechanical alloying. Mater. Sci. Eng. A 2004, 367, 218–224. [Google Scholar] [CrossRef]
  4. Ferreira, S.C.; Rocha, L.A.; Ariza, E.; Sequeira, P.D.; Watanabe, Y.; Fernandes, J.C.S. Corrosion behaviour of Al/Al3Ti and Al/Al3Zr functionally graded materials produced by centrifugal solid-particle method: Influence of the intermetallics volume fraction. Corros. Sci. 2011, 53, 2058–2065. [Google Scholar] [CrossRef] [Green Version]
  5. Kogachi, M.; Minamigawa, S.; Nakahigashi, K. Long-range order in L12 ternary intermetallic compounds Al3Ti-X (X = Fe,Ni,Cu,Ag). Scr. Metall. Mater. 1992, 27, 407–412. [Google Scholar] [CrossRef]
  6. Milman, Y.V.; Miracle, D.B.; Chugunova, S.I.; Voskoboinik, I.V.; Korzhova, N.P.; Legkaya, T.N.; Podrezov, Y.N. Mechanical behaviour of Al3Ti intermetallic and L12 phases on its basis. Intermetallics 2001, 9, 839–845. [Google Scholar] [CrossRef]
  7. Mabuchi, H.; Kito, A.; Nakamoto, M.; Tsuda, H.; Nakayama, Y. Effects of manganese on the L12 compound formation in Al3Ti-based alloys. Intermetallics 1996, 4, S193–S199. [Google Scholar] [CrossRef]
  8. Ghosh, G.; Vaynman, S.; Asta, M.; Fine, M.E. Stability and elastic properties of L12-(Al,Cu)3(Ti,Zr) phases: Ab initio calculations and experiments. Intermetallics 2007, 15, 44–54. [Google Scholar] [CrossRef]
  9. Wang, X.; Ma, S. Mechanical properties and fracture of in-situ Al3Ti particulate reinforced A356 composites. Mater. Sci. Eng. A 2019, 754, 46–56. [Google Scholar] [CrossRef]
  10. Prusov, E.; Deev, V.; Rakhuba, E. Aluminum Matrix In-Situ Composites Reinforced with Mg2Si and Al3Ti. Mater. Today Proc. 2019, 11, 386–391. [Google Scholar] [CrossRef]
  11. Suzuki, A.; Kosugi, N.; Takata, N.; Kobashi, M. Microstructure and compressive properties of porous hybrid materials consisting of ductile Al/Ti and brittle Al3Ti phases fabricated by reaction sintering with space holder. Mater. Sci. Eng. A 2020, 776, 139000. [Google Scholar] [CrossRef]
  12. Himmler, D.; Randelzhofer, P.; Körner, C. In-situ Al3Ti particle reinforcement for stiff aluminum die castings. J. Alloys Compd. 2022, 904, 163984. [Google Scholar] [CrossRef]
  13. Gong, Y.P.; Ma, S.M.; Hei, H.J.; Ma, Y.; Zhou, B.; Wang, Y.S.; Yu, S.W.; Wang, X.; Wu, Y.C. Tailoring microstructure and its effect on wear behavior of an Al–7Si alloy reinforced with in situ formed Al3Ti particulates. J. Mater. Res. Technol. 2020, 9, 7136–7148. [Google Scholar] [CrossRef]
  14. Dinaharan, I.; Ashok Kumar, G.; Vijay, S.J.; Murugan, N. Development of Al3Ti and Al3Zr intermetallic particulate reinforced aluminum alloy AA6061 in situ composites using friction stir processing. Mater. Des. 2014, 63, 213–222. [Google Scholar] [CrossRef]
  15. Vecchio, K.S.; Jiang, F.C. Fracture toughness of Ceramic-Fiber-Reinforced Metallic-Intermetallic-Laminate (CFR-MIL) composites. Mater. Sci. Eng. A 2016, 649, 407–416. [Google Scholar] [CrossRef]
  16. Lazurenko, D.V.; Petrov, I.Y.; Mali, V.I.; Esikov, M.A.; Kuzmin, R.I.; Lozanov, V.V.; Pyczak, F.; Stark, A.; Dovzhenko, G.D.; Bataev, I.A.; et al. Ti-Al3Ti metal-intermetallic laminate (MIL) composite with a cubic titanium trialuminide stabilized with silver: Selection of fabrication regimes, structure, and properties. J. Alloys Compd. 2022, 916, 165480. [Google Scholar] [CrossRef]
  17. Vecchio, K.S. Synthetic multifunctional metallic-intermetallic laminate composites. JOM 2005, 57, 25–31. [Google Scholar] [CrossRef]
  18. Zhang, X.; Huang, Y.; Liu, Y.; Ren, X. A comprehensive DFT study on the thermodynamic and mechanical properties of L12-Al3Ti/Al interface. Vacuum 2021, 183, 109858. [Google Scholar] [CrossRef]
  19. Wang, F.; Qiu, D.; Liu, Z.; Taylor, J.; Easton, M.; Zhang, M.X. Crystallographic study of Al3Zr and Al3Nb as grain refiners for Al alloys. Trans. Nonferrous Met. Soc. China 2014, 24, 2034–2040. [Google Scholar] [CrossRef]
  20. Yang, T.; Wei, M.; Ding, Z.; Han, X.; Li, J. First-principle calculations on the Al/L12-Al3Zr heterogeneous nucleation interface. Calphad 2020, 69, 101768. [Google Scholar] [CrossRef]
  21. Cui, Y.; King, D.J.M.; Horsfield, A.P.; Gourlay, C.M. Solidification orientation relationships between Al3Ti and TiB2. Acta Mater. 2020, 186, 149–161. [Google Scholar] [CrossRef]
  22. Popova, E.A.; Kotenkov, P.V.; Pastukhov, E.A.; Shubin, A.B. Master alloys Al-Sc-Zr, Al-Sc-Ti, and Al-Ti-Zr: Their manufacture, composition, and structure. Russ. Metall. 2013, 2013, 590–594. [Google Scholar] [CrossRef]
  23. Moon, K.I.; Chang, K.Y.; Lee, K.S. The effect of ternary addition on the formation and the thermal stability of L12 Al-Zr alloy with nanocrystalline structure by mechanical alloying. J. Alloys Compd. 2000, 312, 273–283. [Google Scholar] [CrossRef]
  24. Moon, K.I.; Lee, S.H.; Kim, S.J. The effect of Cu and Zn on the phase stability of L12 Al3Hf intermetallic compound synthesized by mechanical alloying. Intermetallics 2002, 10, 793–800. [Google Scholar] [CrossRef]
  25. Desch, P.B.; Schwarz, R.B.; Nash, P. Formation of metastable L12 phases in Al3Zr and Al-12.5%X-25%Zr (X = Li, Cr, Fe, Ni, Cu). J. Less Common Met. 1991, 168, 69–80. [Google Scholar] [CrossRef]
  26. Lee, S.H.; Moon, K.I.; Lee, K.S. Enhancement of the fracture toughness of bulk L12-based (Al+12.5at.%M)3Zr(M=Cu,Mn) intermetallics synthesized by mechanical alloying. Intermetallics 2006, 14, 1–8. [Google Scholar] [CrossRef]
  27. Srinivasan, S.; Desch, P.B.; Schwarz, R.B. Metastable phases in the Al3X (X = Ti, Zr, and Hf) intermetallic system. Scr. Metall. Mater. 1991, 25, 2513–2516. [Google Scholar] [CrossRef]
  28. Pandey, S.K.; Suryanarayana, C. Structure and transformation behavior of a rapidly solidified Al-6.4wt.%Hf alloy. Mater. Sci. Eng. A 1989, 111, 181–187. [Google Scholar] [CrossRef]
  29. Popova, E.; Kotenkov, P.; Shubin, A.; Gilev, I. Formation of Metastable Aluminides in Al–Sc–Ti (Zr, Hf) Cast Alloys. Met. Mater. Int. 2020, 26, 1515–1523. [Google Scholar] [CrossRef]
  30. Popova, E.A.; Kotenkov, P.V.; Gilev, I.O.; Melchakov, S.Y.; Shubin, A.B. Effect of Temperature on the Formation of Stable and Metastable Aluminide Phases in Al–Zr–Nb Alloys. Russ. J. Non-Ferr. Met. 2020, 61, 319–324. [Google Scholar] [CrossRef]
  31. Watanabe, Y.; Yamazaki, R.; Yamanaka, K.; Sato, H. Grain refinement of pure Al using Al2.5Cu0.5Ti particles with an L12 structure. J. Mater. Process. Technol. 2018, 255, 400–410. [Google Scholar] [CrossRef]
  32. Popova, E.A.; Shubin, A.B.; Kotenkov, P.V.; Bodrova, L.E.; Dolmatov, A.V.; Pastukhov, E.A.; Vatolin, N.A. Al-Sc-Zr master alloy and estimation of its modifying capacity. Russ. Metall. 2011, 8, 715–718. [Google Scholar] [CrossRef]
  33. Russell, K.C. Ch. 6: Nucleation in Solids. In Phase Transformations; Aaronson, H.I., Ed.; American Society for Metals: Metals Park, OH, USA, 1970. [Google Scholar]
  34. Wagner, R.; Kampmann, R.; Voorhees, P.W. Ch. 5: Homogeneous Second-Phase Precipitation. In Phase Transformations in Materials; Kostorz, G., Ed.; Wiley-VCH: New York, NY, USA, 2001; pp. 309–407. [Google Scholar]
  35. Doherty, D. Ch. 14: Diffusive Phase Tranformations in the Solid State. In Physical Metallurgy, 3rd ed.; Cahn, R.W., Haasen, P., Eds.; Elsevier: Amsterdam, The Netherlands, 1983; pp. 933–1030. [Google Scholar]
  36. Knipling, K.E.; Dunand, D.C.; Seidman, D.N. Criteria for developing castable, creep-resistant aluminum-based alloys—A review. Mat. Res. Adv. Tech. 2006, 97, 246–265. [Google Scholar] [CrossRef] [Green Version]
  37. Liu, Z. Thermodynamics of Nano-Scale Precipitate-Strengthened Fe-Cu and Al-Transition-Metal Systems from First-Principles Calculations. Ph.D. Thesis, Northwestern University, Evanston, IL, USA, 2006. [Google Scholar]
  38. Frost, H.J.; Ashby, M.F. Deformation-Mechanism Maps: The Plasticity and Creep of Metals and Ceramics; Pergamon Press: New York, NY, USA, 1982. [Google Scholar]
  39. Meyers, M.A.; Chawla, K.K. Mechanical Metallurgy: Principles and Applications; Prentice-Hall: Hoboken, NJ, USA, 1984. [Google Scholar]
  40. Ohashi, T.; Ichikawa, R. Grain Refinement in Aluminium-Zirconium and Aluminium-Titanium Alloys by Metastable Phases. Mat. Res. Adv. Tech. 1973, 64, 517–521. [Google Scholar]
  41. Kerr, H.W.; Cisse, J.; Bolling, G.F. Equilibrium and Nonequilibrium Peritectic Transformations. Acta Metall. 1974, 22, 677–686. [Google Scholar] [CrossRef]
  42. Fujikawa, S.I. Solid State Diffusion in Light Metals. J. Jpn. Inst. Light Met. 1996, 46, 202–215. [Google Scholar] [CrossRef]
Figure 1. Distribution of structural components in the Al–40.33Cu–2.75Ti (T2) alloy: (a) magnification × 500; (b) magnification × 1500. The arrows point to main structural components of the alloy: the θ-phase Al2Cu (green arrows), the α-Al + θ eutectic (blue arrows) and L12 trialuminides (red arrows).
Figure 1. Distribution of structural components in the Al–40.33Cu–2.75Ti (T2) alloy: (a) magnification × 500; (b) magnification × 1500. The arrows point to main structural components of the alloy: the θ-phase Al2Cu (green arrows), the α-Al + θ eutectic (blue arrows) and L12 trialuminides (red arrows).
Metals 12 02067 g001
Figure 2. Distribution and growth forms of zirconium aluminides in alloys: (a) Al–38.78Cu–4.56Zr (Z1) and (b) Al–36.28Cu–9.02Zr (Z2) (magnification × 1000). The arrows point to main structural components of these alloys: the θ-phase Al2Cu (green arrows), the α-Al + θ eutectic (blue arrows) and L12 trialuminides (red arrows).
Figure 2. Distribution and growth forms of zirconium aluminides in alloys: (a) Al–38.78Cu–4.56Zr (Z1) and (b) Al–36.28Cu–9.02Zr (Z2) (magnification × 1000). The arrows point to main structural components of these alloys: the θ-phase Al2Cu (green arrows), the α-Al + θ eutectic (blue arrows) and L12 trialuminides (red arrows).
Metals 12 02067 g002
Figure 3. Distribution, sizes and forms of growth of aluminides in alloys: (a) Al–31.47Cu–8.63Zr (Z2); (b) Al–31.25Cu–8.37Zr (Z4) (magnification × 1000). The arrows point to main structural components of these alloys: the θ-phase Al2Cu (green arrows), the α-Al + θ eutectic (blue arrows) and L12 trialuminides (red arrows).
Figure 3. Distribution, sizes and forms of growth of aluminides in alloys: (a) Al–31.47Cu–8.63Zr (Z2); (b) Al–31.25Cu–8.37Zr (Z4) (magnification × 1000). The arrows point to main structural components of these alloys: the θ-phase Al2Cu (green arrows), the α-Al + θ eutectic (blue arrows) and L12 trialuminides (red arrows).
Metals 12 02067 g003
Figure 4. Distribution and forms of growth of stable and metastable aluminides in alloys: (a) Al – 29.73Cu – 13.24Hf (H1); (b) Al–29.47Cu–13.69Hf (H2) (magnification × 1000). The arrows point to main structural components of these alloys: the θ-phase Al2Cu (green arrows), the α-Al + θ eutectic (blue arrows) and L12 trialuminides (red arrows).
Figure 4. Distribution and forms of growth of stable and metastable aluminides in alloys: (a) Al – 29.73Cu – 13.24Hf (H1); (b) Al–29.47Cu–13.69Hf (H2) (magnification × 1000). The arrows point to main structural components of these alloys: the θ-phase Al2Cu (green arrows), the α-Al + θ eutectic (blue arrows) and L12 trialuminides (red arrows).
Metals 12 02067 g004
Figure 5. Distribution and forms of growth of stable and metastable aluminides in alloys: (a) Al–34.09Cu–13.04Hf (H4); (b) Al–32.83Cu–13.04Hf (H3) (magnification × 500). The arrows point to main structural components of these alloys: the θ-phase Al2Cu (green arrows), the α-Al + θ eutectic (blue arrows) and L12 trialuminides (red arrows).
Figure 5. Distribution and forms of growth of stable and metastable aluminides in alloys: (a) Al–34.09Cu–13.04Hf (H4); (b) Al–32.83Cu–13.04Hf (H3) (magnification × 500). The arrows point to main structural components of these alloys: the θ-phase Al2Cu (green arrows), the α-Al + θ eutectic (blue arrows) and L12 trialuminides (red arrows).
Metals 12 02067 g005
Figure 6. Distribution and growth patterns of metastable aluminides in Al–35.66Cu–15.36Hf (H5) alloy: (a) magnification × 250; (b) magnification × 1000. The arrows point to main structural components of the alloy: the θ-phase Al2Cu (green arrows), the α-Al + θ eutectic (blue arrows) and L12 trialuminides (red arrows).
Figure 6. Distribution and growth patterns of metastable aluminides in Al–35.66Cu–15.36Hf (H5) alloy: (a) magnification × 250; (b) magnification × 1000. The arrows point to main structural components of the alloy: the θ-phase Al2Cu (green arrows), the α-Al + θ eutectic (blue arrows) and L12 trialuminides (red arrows).
Metals 12 02067 g006
Figure 7. XRD pattern for the alloys: Al–36.28Cu–9.02Zr (Z3—blue curve), Al–35.66Cu–15.36Hf (H5—red curve), green markers indicate the aluminide phase with a cubic L12 lattice.
Figure 7. XRD pattern for the alloys: Al–36.28Cu–9.02Zr (Z3—blue curve), Al–35.66Cu–15.36Hf (H5—red curve), green markers indicate the aluminide phase with a cubic L12 lattice.
Metals 12 02067 g007
Table 1. Compositions of Al–Cu–X (X=Ti, Zr, Hf) alloys and the L12 aluminides formed in them.
Table 1. Compositions of Al–Cu–X (X=Ti, Zr, Hf) alloys and the L12 aluminides formed in them.
Alloy№Alloys Composition, wt.% (at.%),
Degree of Supersaturation with TM.
Synthesis Temperature, °CComposition of Sluminides Having the L12 Crystal StructureStable Aluminides Having the
D023 and D022 Crystal Structure
T1Al–41.91Cu–1.64Ti
(Al–23.70Cu–1.23Ti),
2 times
1250Cuboids and dendrites
70.68Al–6.36Cu–22.96Ti
or (Al0.92Cu0.08)nTi *
Needles and plates
containing 5.71 at.%Cu
T2Al–40.33Cu–2.75Ti
(Al–22.68Cu–2.05Ti),
3.4 times
1300Cuboids and dendrites
70.68Al–6.36Cu–22.96Ti
or (Al0.92Cu0.08)nTi *
Needles and plates
containing 5.71 at.%Cu
Z1Al–38.78Cu–4.56Zr
(Al–22.13Cu–1.81Zr),
26 times
1200Cuboids and dendrites
68.34Al–6.27Cu–25.39Zr
or (Al0.92Cu0.08)3Zr
There are no particles having this crystal structure
Z2Al–31.47Cu–8.63Zr
(Al–17.64Cu–3.37Zr)
48 times
1300Cuboids and dendrites
67.08Al–7.27Cu–25.65Zr
or (Al0.90Cu0.10)3Zr
There are no particles having this crystal structure
Z3Al–36.28Cu–9.02Zr
(Al–21.19Cu–3.67Zr)
52 times
1200Cuboids and dendrites
65.39Al–9.05Cu–25.56Zr
or (Al0.88Cu0.12)3Zr
Needles and plates
Z4Al–31.25Cu–8.37Zr
(Al–17.45Cu–3.26Zr)
47 times
1300Single cuboids,
not analyzed
Plates
(Al0.97Cu0.03)3Zr
H1Al–29.73Cu–13.24Hf)
(Al–17.63Cu–2.79Hf
174 times
1250Single cuboids,
not analyzed

Plates
containing 3.80 at.%Cu
H2Al–29.47Cu–13.69Hf
(Al–17.54Cu–2.91Hf)
182 times
1300Single cuboids,
not analyzed
Plates
containing 3.80 at.%Cu
H3Al–32.83Cu–13.04Hf
(Al–19.92Cu–2.82Hf)
176 times
130069.03Al–7.67Cu–23.30Hf
or (Al0.90Cu0.10)nHf *
Single plates and needles
Al3Hf
H4Al–34.09Cu–13.04Hf
(Al–20.91Cu–2.84Hf)
178 times
125069.90Al–7.39Cu–22.71Hf
or (Al0.90Cu0.10)nHf *
Plates
containing 3.55 at.%Cu
and needles Al3Hf
H5Al–35.66Cu–15.36Hf
(Al–22.81Cu–3.50Hf)
219 times
120067.09Al–10.03Cu–22.88Hf
or (Al0.87Cu0.13)nHf *
There are no particles having this crystal structure
* According to EDX analysis, value of n is in the range from 3 to 4.
Table 2. Mixing enthalpy of Al3X (X = Ti, Zr, Hf) compounds and TM atoms.
Table 2. Mixing enthalpy of Al3X (X = Ti, Zr, Hf) compounds and TM atoms.
System Δ H = 4 Δ H A l 3 X ,
eV/atom
Δ H = Δ H X ,
eV/atom
Δ H s o l = 4 Δ H A l 3 X Δ H X ,
eV/atom
L12ImpurityL12
Al-Ti−1.576−1.197−0.379
Al-Zr−1.964−1.250−0.714
Al-Hf−1.609−0.954−0.655
Table 3. Vibrational entropy of Al3X (X = Ti, Zr, Hf) compounds and TM atoms.
Table 3. Vibrational entropy of Al3X (X = Ti, Zr, Hf) compounds and TM atoms.
System Δ S v i b = 4 Δ S v i b A l 3 X ,
k/atom
Δ S v i b = Δ S v i b X ,
k/atom
Δ S v i b X 4 Δ S v i b A l 3 X ,
k/atom
L12ImpurityL12
Al-Ti−3.365−2.5180.847
Al-Zr−3.354−1.0002.354
Al-Hf−2.989−0.8372.152
Table 4. Calculation of the chemical driving force Δ F ch , which determines the precipitation of the Al3X phase.
Table 4. Calculation of the chemical driving force Δ F ch , which determines the precipitation of the Al3X phase.
Al3X (L12) V A l 3 X ,   m 3 · m o l 1 C α A l 3 X ,  
Atomic Fraction
(T = 698 K)
C A l 3 X ,  
Atomic Fraction
Δ F c h ,   M J · m 3
(Peritectic Temperature)
Al3Ti0.94 × 10–50.021510.25193.7 × ln(46.5 × C0)
Al3Zr1.02 × 10–50.001480.25189.3 × ln(675.7 × C0)
Al3Hf0.99 × 10–50.002280.25192.4 × ln(438.6 × C0)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Popova, E.; Kotenkov, P.; Gilev, I.; Pryanichnikov, S.; Shubin, A. Effect of Copper on the Formation of L12 Intermetallic Phases in Al–Cu–X (X = Ti, Zr, Hf) Alloys. Metals 2022, 12, 2067. https://doi.org/10.3390/met12122067

AMA Style

Popova E, Kotenkov P, Gilev I, Pryanichnikov S, Shubin A. Effect of Copper on the Formation of L12 Intermetallic Phases in Al–Cu–X (X = Ti, Zr, Hf) Alloys. Metals. 2022; 12(12):2067. https://doi.org/10.3390/met12122067

Chicago/Turabian Style

Popova, Elvira, Pavel Kotenkov, Ivan Gilev, Stepan Pryanichnikov, and Alexey Shubin. 2022. "Effect of Copper on the Formation of L12 Intermetallic Phases in Al–Cu–X (X = Ti, Zr, Hf) Alloys" Metals 12, no. 12: 2067. https://doi.org/10.3390/met12122067

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop