Next Article in Journal
Compression Strength Mechanisms of Low-Density Fibrous Materials
Next Article in Special Issue
Atomic-Approach to Predict the Energetically Favored Composition Region and to Characterize the Short-, Medium-, and Extended-Range Structures of the Ti-Nb-Al Ternary Metallic Glasses
Previous Article in Journal
Bottom-Up Self-Assembled Supramolecular Structures Built by STM at the Solid/Liquid Interface
Previous Article in Special Issue
Discrete Element Modeling of Intermetallic Matrix Composite Manufacturing by Powder Metallurgy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Crystallization Behavior of Al70Fe12.5V12.5Nb5 Amorphous Alloy Formed by Mechanical Alloying

1
Key Laboratory of Materials Physics, Institute of Solid State Physics, Chinese Academy of Science, Hefei 230031, China
2
University of Science and Technology of China, Hefei 230026, China
3
College of Civil Engineering, Anhui Jianzhu University, Hefei 230061, China
*
Author to whom correspondence should be addressed.
Materials 2019, 12(3), 383; https://doi.org/10.3390/ma12030383
Submission received: 19 December 2018 / Revised: 14 January 2019 / Accepted: 22 January 2019 / Published: 26 January 2019
(This article belongs to the Special Issue Multi-scale Modeling of Materials and Structures)

Abstract

:
In this study, the formation and crystallization of the Al70Fe12.5V12.5Nb5 amorphous alloys has been investigated. The addition of Nb enhances the supercooled liquid region and glass forming ability of the Al-Fe-V amorphous alloys. The Al70Fe12.5V12.5Nb5 amorphous alloy exhibits two distinct crystallization steps and a large supercooled liquid region at more than 100 K. Kissinger and Ozawa analyses showed that the two activation energies for crystallization (Ex) were estimated to be 366.3 ± 23.9 and 380.5 ± 23.9 kJ/mol. Large supercooled liquid regions are expected to gain an application field of Al-based amorphous alloys.

1. Introduction

Currently, the application of amorphous powder in bulk metallic glasses, composite, coating, and 3D printing has widened the application scope of amorphous alloys [1,2,3,4]. In particular, Aluminum-based amorphous alloys have attracted increasing attention due to their exceptional physical properties, low glass forming ability (GFA), and poor thermal stability [5,6,7]. However, Al-based amorphous alloys with low GFA are challenging to be synthesized via the traditional rapid quenching (RQ). The size factor restricts the practical application of Aluminum-based amorphous alloys. Indeed, up to now, the maximum diameter of an Al-based amorphous alloys sample is 2.5 mm by RQ [8]. Mechanical alloying (MA) is an effective process to produce amorphous alloy powders without a high cooling rate [9,10]. Therefore, it is noteworthy to prepare Al-based amorphous powders via MA to widen their application [11,12,13,14,15]. To use amorphous powder, a large supercooled liquid region (ΔTx) is necessary, and enhancement of the GFA and thermal stability is required.
The addition of Nb can improve the GFA, thermal stability, and the corrosion resistance of Zr-based Zr-Ni-Al-Nb bulk metallic glasses with high compressive yield strength [16]. A similar phenomenon has been observed in some Al-based amorphous alloys [17,18]. It has been reported that, in the case of the Nb-Al system, the sequence of phase formation varies widely depending upon Al content: for example, Al80Fe10Nb10 created via an MA process lasted 100 h is a powder mixture [19,20]. However, according to the efficient cluster packing (ECP) model, Nb has the same atomic radius as Al, so they can be mixed with an arbitrary ratio to occupy the Ω position of the solvent atoms without changing the cluster packing structure in the amorphous alloys [21]. Thus, it is interesting to investigate the GFA and thermal stability of in Al-based amorphous alloys in which Nb has replaced Al. Al75V12.5Fe12.5 amorphous alloys with high GFA and thermal stability were prepared via MA and based on the ECP model [22]. In general, a negative heating of the mixing enthalpy (ΔHmix) among elements in the amorphous alloy is beneficial for enhancing the GFA [23]. Nb contribution to the enthalpy leads to a negative ΔHmix with Al (−18 kJ/mol), Fe (−16 kJ/mol), and V (−1 kJ/mol) [24]. Therefore, the GFA and thermal stability expect to be enhanced with the Nb addition in the Al-Fe-V alloys. In this work, we synthesized alloys with 5% of Al replaced by Nb. The GFA and thermal stability were investigated according to the values obtained from differential scanning calorimetry (DSC). The aim of this study is to obtain Al-based amorphous powder with high ΔTx and thermal stability for industrial applications.

2. Experimental Details

The Al70Fe12.5V12.5Nb5 (at %) amorphous alloy powders were prepared via MA of a mixture of elemental Al (99.9%), Fe (99.8%), V (99.9%), and Nb (99.7%) powders. Stearic acid powder (2 wt. %) was used as the process control agent. The MA was conducted in a high-energy planetary ball mill (QM-3SP4, Instrument Factory of Nanjing University, Nanjing, China) using a stainless steel vial and a steel ball bearing (GGr15 with diameters of 20 mm, 10 mm, and 5 mm are mixed) with the rotation speed of 350 rpm and ball-to-powder weight ratio of 20:1 in a highly pure Ar atmosphere. The vial is introduced into the vacuum glove box every 10 h to obtain the powder samples for analysis. The powder samples were characterized via X-ray diffraction (XRD, X’ Pert Pro MPD, Philips) using Cu Ka radiation in the 2θ range of 20° to 90°. The X’Pert HighScore Plus software (Version: 3.0) is used to analyze the XRD curves of the mixed powder after ball milling. The microstructures of the amorphous powders were examined using transmission electron microscopy (TEM, JEOL 2010, JEOL) and selected area electron diffraction (SAED) operating at 200 kV. The thermal properties, such as Tg (the glass transition temperature), Tx (the onset crystallization temperature), Tp1 (the first crystallization peak temperature), Tp2 (the secondary crystallization peak temperature), Tm (the melting temperature), Tl (the liquidus temperature), were measured via the DSC (NETZSCH DSC 404F3) at heating rates of 10, 20, 30, 40 K/min from 303 K to 1473 K in an Ar atmosphere.

3. Results and Discussion

XRD results of Al70Fe12.5V12.5Nb5 alloy powders milled for different periods (10 h, 20 h, 30 h, 40 h, 50 h and 60 h) are shown in Figure 1. It can be seen from Figure 1 that the crystalline phase is formed by ball milling 10 h, X’Pert HighScore Plus software results suggest that the crystalline phase is AlNb2. With the increase of milling time, no new diffraction peaks have been found in the XRD curve, indicating that no new crystalline phase is formed except for AlNb2. From the DSC curve of 10 h to 30 h, the intensity of the diffraction peak corresponding to the initial addition of powder elements is reduced with the strongest diffraction peak position moving towards higher 2θ angles of 0.2°. According to the Scherrer formula, this means that the lattice of Al has infiltrated other atoms with a smaller atomic radius and formed a solid solution. The atomic radius of Fe (0.124 nm) and V (0.132 nm) is shorter than the radius of Al (0.143 nm); therefore, the product is the Al-Fe-V solid solution formed by Fe and V dissolved in Al. When milling to 40 h, the diffraction peak still exists and has broadened, indicating that the amorphous phase is formed and there are crystalline phases and solid solutions at this time. After milling 50 h, the diffraction peaks of all crystals disappear, while different diffuse peaks can be identified at a 2θ angle of 37°. Until the 60 h milling treatment, only the typical broad halo peak showing the amorphous structure exists, indicating that the powders are amorphous.
To verify the amorphous structure, the TEM and selected area electron diffraction (SEAD) images of Al70Fe12.5V12.5Nb5 alloys powder milled for 60 h are shown in Figure 2. There is no visible crystals in the TEM image, and the broad diffuse halo and the absence of rings or diffraction spots in the SEAD pattern further confirms the amorphous structure.
The GFA and thermal stability of the amorphous alloy are related to atomic radii mismatch and thermodynamics of the alloy systems: it can be displayed by the milling time of MA amorphous powders. Compared with the Al75V12.5Fe12.5−xCux amorphous alloys reported in the literature [22], the Al70Fe12.5V12.5Nb5 amorphous alloy milling time is reduced.
The parameter Delta is adopted to describe the effect of atomic radii mismatch of multi-component alloys as follows [25]:
D e l t a = i = 1 n c i ( 1 r i / r ¯ ) 2 .
Here, the n is the number of components in the amorphous alloy system, c i is the atomic percentage of the i component, and r i is the atomic radius, r ¯   ( r ¯ = i = 1 n c i r i ) is the average atomic radius.
By calculating the Delta, the Al70Fe12.5V12.5Nb5 and Al75V12.5Fe12.5−xCux amorphous alloys have the same value. We can then infer that the influence of atomic radii mismatch of the two systems can be neglected.
Another parameter, the equation of the Gibbs free energy ( G m i x ), is used to characterize the thermodynamics of alloy systems.
G m i x = H m i x T S m i x .
According to Equation (2), a low mixing enthalpy ( H m i x ) or a high entropy of mixing ( S m i x ) can decrease the G m i x , thus lowering the crystallization driving force, which consequently renders the formation of the amorphous phase easier. H m i x and S m i x can be calculated by Equations (3) and (4), respectively.
H m i x = i = 1 , i j n Ω i j c i c j ,
S m i x = R i = 1 , i j n c i L n c i ,
where Ω i j = 4 H i j , the H i j is the mixing enthalpy of the i and j elements, and R is the gas constant. It can be concluded that the H m i x increases in the order of the Al70Fe12.5V12.5Nb5 (−11.970 kJ/mol), Al75V12.5Fe12.5 (−10.125 kJ/mol), Al75V12.5Fe6.25Cu6.25 (−8.250 kJ/mol), Al75V12.5Cu12.5 (−6.375 kJ/mol) amorphous alloys. The S m i x decreases in the order of the Al70Fe12.5V12.5Nb5 (0.919R), Al75V12.5Fe12.5 (0.736R), Al75V12.5Cu12.5 (0.736R), Al75V12.5Fe6.25Cu6.25 (0.649R) amorphous alloys. In these amorphous alloys, Al70Fe12.5V12.5Nb5 has the lowest H m i x and the highest S m i x , so the G m i x is the lowest according to Equation (2). This may be the reason why the Al70Fe12.5V12.5Nb5 amorphous alloy milling time is shortest among these amorphous alloys. Thus, we conclude that the GFA and thermal stability of Al70Fe12.5V12.5Nb5 amorphous alloy are higher than Al75V12.5Fe12.5−xCux amorphous alloys.
Figure 3a shows the DSC curves of the Al70Fe12.5V12.5Nb5 amorphous alloy at the heating rates of 10, 20, 30, and 40 K/min. Two exothermic crystallization events can be observed and are associated with the transformations from an amorphous state to the equilibrium phases. The temperature of Tg that indicates glass transition at the different heating rate (see Figure 3b). The thermal properties of Tg, Tx, Tp1, Tp2, Tm, Tl are marked with arrows in Figure 3a and listed in Table 1, which also contains the empirical criteria of GFA [23,26], ΔTx (ΔTx = Tx − Tg), the reduced glass transition temperature Trg(Trg = Tg/Tl), the γ val (γ = Tx/(Tg + Tl)). It is clear that, with the increase of heating rate, Tg, Tx, Tp1, Tp2, turn to a higher temperature, while the ΔTx fluctuates between 106.5 and 117.5 K. This shows that the thermal properties of Al70Fe12.5V12.5Nb5 at different heating rates are influenced by kinetic factors. The values of Tx and ΔTx of Al70Fe12.5V12.5Nb5 amorphous alloy are 854.1 K and 118.6 K at the heating rates of 20 K/min, respectively, whereas the Tm is almost unchanged. It should be noted that the change of ΔTx for Al70Fe12.5V12.5Nb5 is more than 100 K, which is higher than Al75V12.5Fe12.5−xCux amorphous alloys and other Al-based amorphous alloys [27]. The higher ΔTx,Trg, γ indicates that the Al70Fe12.5V12.5Nb5 amorphous alloy has a higher glass forming ability. Therefore, it is a promising candidate for thermoplastic forming BMGs and composite materials via the consolidation of amorphous powders.
The activation energies of Ex, Ep1, Ep2 corresponding to Tx, Tp1, Tp2, respectively, can be estimated via the equations of Kissinger (Equation (5)) and Ozawa (Equation (6)) [28,29]:
Ln   ( β / T 2 ) = ( E / RT ) + Const ,
Ln   ( β ) = ( E / RT ) + Const .
Here, the β is the heating rates, T represents the specific temperature (Tx, Tp1, Tp2), E is the corresponding activation energy (Ex, Ep1, Ep2), and R is the gas constant.
Figure 4 reports the Kissinger and Ozawa plots by linear fitting of ln (β/T2) versus 1/T or ln (β) versus 1/T. The estimated activation energies (Ex, Ep1, Ep2) corresponding to Tx, Tp1, Tp2 were evaluated by the Kissinger and Ozawa methods, and listed in Table 2. The fitted curves are close to the straight lines; the Adjusted R-square is higher than 96, and the results of the two equations have the same tendency, which proves the reliability of the data. It is clear that the effective activation energies evaluated by the Kissinger equation are smaller than Ozawa equation. The Ex values resulting by the application of the Kissinger and Ozawa methods can be calculated: 366.3 ± 23.9 and 380.5 ± 23.9 kJ/mol, respectively.
The DSC curves in Figure 3 have two obvious endothermic peaks, which means that there are two crystallization processes. To further understand the crystallization of the new crystalline phases during the exothermic processes, five heat treatments of the Al70Fe12.5V12.5Nb5 amorphous alloy powders were studied. Figure 5 shows the XRD pattern of the Al70Fe12.5V12.5Nb5 amorphous alloy powders after five different heat treatments. Different heat treatment temperatures have been selected: 773 K (below Tx and above Tg), 973 K (between TP1 and TP2), and 1173 K (above TP2).
The crystal phase is not observed when the sample has been heated to 773 K before rapid cooling (heat treatment a in Figure 5). When amorphous alloy samples are annealed at 773 K for 100 min (heat treatment b), a few Al3Nb phases of the amorphous sample can be observed. It is shown that the short-range atom structure retained in the amorphous alloy has been partly changed after low-temperature annealing, while a certain degree of relaxation has occurred. Although the crystallization temperature has not been reached, the atom diffusion ability of the amorphous alloy has increased after annealing for 100 min with the Nb addition. Al5Fe2 and Al3Nb phases have been mainly observed after initial crystallization and rapid cooling immediately after heating to 973 K (heat treatment c). The Al23V4 phase is also found at 973 K annealing for 100 min (heat treatment d) and heating to 1173 K (heat treatment e). New crystalline phases precipitate when the amorphous sample is annealed, which may be due to the relaxation of the amorphous structure. To reduce the internal free energy of the amorphous alloy, a certain crystallization phase has precipitated in the amorphous alloy.

4. Conclusions

Amorphous alloy Al70Fe12.5V12.5Nb5 (at %) powders were successfully prepared via MA. The replacement of Al with Nb in the Al75V12.5Fe12.5 amorphous alloy effectively enhances its glass forming ability and thermal stability. The Al70Fe12.5V12.5Nb5 amorphous alloy exhibits a larger supercooled liquid region than 100 K, higher GFA with the Trg of 0.522, and γ of 0.398 (20 K/min). The crystallization of the Al70Fe12.5V12.5Nb5 amorphous alloy takes place in two distinct steps, as observed via DSC analyses. Our study provides deeper insights into the development of Al-base amorphous alloys to be used in powder metallurgy industry.

Author Contributions

Conceptualization, X.L. and F.H.; methodology, X.L.; software, X.W.; validation, Y.S. and X.Z.; formal analysis, X.L.; investigation, X.L.; resources, Y.S.; data curation, Y.S.; writing—original draft preparation, X.L.; writing—review and editing, F.H.; visualization, F.H.; supervision, F.H.; project administration, F.H.; funding acquisition, X.L.

Funding

This research was funded by the National Natural Science Foundation of China (51371167). Anhui Provincial Universities Natural Science Research Project (KJ2016JD03).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Maurya, R.S.; Sahu, A.; Laha, T. Effect of sintering temperature on phase transformation during consolidation ofmechanically alloyed Al86Ni6Y6Co2 amorphous powders by spark plasma sintering. J. Non-Cryst. Solids 2016, 453, 1–7. [Google Scholar] [CrossRef]
  2. Zhang, L.; Yang, L.; Leng, J.; Wang, T.; Wang, Y. Alloying Behavior and Properties of Al-Based Composites Reinforced with Al85Fe15 Metallic Glass Particles Fabricated by Mechanical Alloying and Hot Pressing Consolidation. JOM 2017, 69, 748–755. [Google Scholar] [CrossRef]
  3. Gao, M.; Lu, W.; Yang, B.; Zhang, S.; Wang, J. High corrosion and wear resistance of Al-based amorphous metallic coating synthesized by HVAF spraying. J. Alloys Compd. 2018, 735, 1363–1373. [Google Scholar] [CrossRef]
  4. Li, N.; Zhang, J.; Xing, W.; Ouyang, D.; Liu, L. 3D printing of Fe-based bulk metallic glass composites with combined high strength and fracture toughness. Mater. Des. 2018, 143, 285–296. [Google Scholar] [CrossRef]
  5. Liao, J.P.; Yang, B.J.; Zhang, Y.; Lu, W.Y.; Gu, X.J.; Wang, J.Q. Evaluation of glass formation and critical casting diameter in Al-based metallic glasses. Mater. Des. 2015, 88, 222–226. [Google Scholar] [CrossRef]
  6. Yin, J.; Cai, H.; Cheng, X.; Zhang, X. Al-based bulk metallic glass with large plasticity and ultrahigh strength. J. Alloys Compd. 2015, 648, 276–279. [Google Scholar] [CrossRef]
  7. Wu, N.C.; Zuo, L.; Wang, J.Q.; Ma, E. Designing aluminum-rich bulk metallic glasses via electronic-structure-guided microalloying. Acta Mater. 2016, 108, 143–151. [Google Scholar] [CrossRef] [Green Version]
  8. Yang, B.J.; Lu, W.Y.; Zhang, J.L.; Wang, J.Q.; Ma, E. Melt fluxing to elevate the forming ability of Al-based bulk metallic glasses. Sci. Rep. 2017, 7, 1–16. [Google Scholar] [CrossRef] [PubMed]
  9. Koch, C.C. Materials Synthesis by Mechanical Alloying. Annu. Rev. Mater. Res. 1989, 19, 121–143. [Google Scholar] [CrossRef]
  10. Suryanarayana, C. Mechanical Alloying and Milling. Prog. Mater. Sci. 2001, 46, 1–184. [Google Scholar] [CrossRef]
  11. Mula, S.; Mondal, K.; Ghosh, S.; Pabi, S.K. Structure and mechanical properties of Al–Ni–Ti amorphous powder consolidated by pressure-less, pressure-assisted and spark plasma sintering. Mater. Sci. Eng. A 2010, 527, 3757–3763. [Google Scholar] [CrossRef]
  12. Tan, Z.; Wang, L.; Xue, Y.; Cheng, X.; Zhang, L. structural modification of Al65Cu16.5Ti18.5 amorphous powder through annealing and post milling, improving thermal stability. J. Mater. Sci. Technol. 2016, 32, 1326–1331. [Google Scholar] [CrossRef]
  13. Maurya, R.S.; Sahu, A.; Laha, T. Quantitative phase analysis in Al86Ni8Y6 bulk glassy alloy synthesized by consolidating mechanically alloyed amorphous powder via spark plasma sintering. Mater. Des. 2016, 93, 96–103. [Google Scholar] [CrossRef]
  14. Nguyen, V.H.; Nguyen, O.T.; Dudina, D.V.; Le, V.V.; Kim, J.S. Crystallization Kinetics of Al-Fe and Al-Fe-Y Amorphous Alloys Produced by Mechanical Milling. J. Nanomater. 2016, 2, 1–9. [Google Scholar] [CrossRef]
  15. Tan, Z.; Xue, Y.F.; Cheng, X.W.; Zhang, L.; Chen, W.W.; Wang, L.; Zhang, H.F.; Fu, H.M. Effect of element fitting on composition optimization of Al-Cu-Ti amorphous alloy by mechanical alloying. Trans. Nonferrous Met. Soc. China 2015, 25, 3348–3353. [Google Scholar] [CrossRef]
  16. Li, Y.H.; Zhang, W.; Dong, C.; Qiang, J.B.; Xie, G.Q.; Fujita, K.; Inoue, A. Enhancement of glass-forming ability and corrosion resistance of Zr-based Zr-Ni-Al bulk metallic glasses with minor addition of Nb. J. Appl. Phys. 2011, 110, 023513. [Google Scholar] [CrossRef]
  17. Wu, Y.; Wang, X.F.; Han, F.S. Preparation of Al72Ni8Ti8Zr6Nb3Y3 amorphous powders and bulk materials. Int. J. Min. Met. Mater. 2016, 23, 1187–1195. [Google Scholar] [CrossRef]
  18. Mostaan, H.; Karimzadeh, F.; Abbasi, M.H. Thermodynamic analysis of nanocrystalline and amorphous phase formation in Nb–Al system during mechanical alloying. Powder Metall. 2012, 55, 142–147. [Google Scholar] [CrossRef]
  19. Tavoosi, M.; Karimzadeh, F.; Enayati, M.H. Formation and characterization of amorphous–nanocrystalline Al80Fe10M10 [M = Fe, Nb, Ti, Ni, (Ni0.5Ti0.5)] alloys. J. Alloys Compd. 2013, 551, 584–590. [Google Scholar] [CrossRef]
  20. Peng, Z.; Suryanarayana, C.; Froes, F.S. Mechanical alloying of Nb-Al powders. Metall. Mater. Trans. A 1996, 27, 41–48. [Google Scholar] [CrossRef]
  21. Miracle, D.B. The efficient cluster packing model—An atomic structural model for metallic glasses. Acta Mater. 2006, 54, 4317–4336. [Google Scholar] [CrossRef]
  22. Zhu, B.; Wang, X.; Li, X.; Wang, D.; Qin, Y.; Han, F. Bulk amorphous Al75V12.5Fe12.5−xCux alloys fabricated by consolidation of mechanically alloyed amorphous powders. J. Alloys Compd. 2014, 586, 645–649. [Google Scholar] [CrossRef]
  23. Inoue, A. Stabilization of metallic supercooled liquid and bulk amorphous alloys. Acta Mater. 2000, 48, 279–306. [Google Scholar] [CrossRef] [Green Version]
  24. Takeuchi, A.; Inoue, A. Classification of bulk metallic glasses by atomic size difference, heat of mixing and period of constituent elements and its application to characterization of the main alloying element. Mater. Trans. 2005, 46, 2817–2829. [Google Scholar] [CrossRef]
  25. Zhang, Y.; Zhou, Y.J.; Lin, J.P.; Chen, G.L.; Liaw, P.K. Solid-Solution Phase Formation Rules for Multi-component Alloys. Adv. Eng. Mater. 2008, 10, 534–538. [Google Scholar] [CrossRef]
  26. Lu, Z.P.; Liu, C.T. A new glass-forming ability criterion for bulk metallic glasses. Acta Mater. 2002, 50, 3501–3512. [Google Scholar] [CrossRef]
  27. Zhang, W.; Chen, S.; Zhu, Z.; Wang, H.; Li, Y.; Kato, H.; Zhang, H. Effect of substituting elements on thermal stability and glass-forming ability of an Al-based AlNiEr metallic glass. J. Alloys Compd. 2017, 707, 97–101. [Google Scholar] [CrossRef]
  28. Kissinger, H.E. Reaction Kinetics in Differential Thermal Analysis. Anal. Chem. 1957, 29, 1702–1706. [Google Scholar] [CrossRef]
  29. Ozawa, T.A. New method of analyzing thermogravimetric data. Bull. Chem. Soc. Jpn. 1965, 38, 1881–1886. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns (XRD) patterns of Al70Fe12.5V12.5Nb5 powders milled for 10, 20, 30, 40, 50 and 60 h.
Figure 1. X-ray diffraction patterns (XRD) patterns of Al70Fe12.5V12.5Nb5 powders milled for 10, 20, 30, 40, 50 and 60 h.
Materials 12 00383 g001
Figure 2. Transmission electron microscope (TEM) image and selected area electron diffraction (SAED) pattern of Al70Fe12.5V12.5Nb5 alloys powder milled for 60 h.
Figure 2. Transmission electron microscope (TEM) image and selected area electron diffraction (SAED) pattern of Al70Fe12.5V12.5Nb5 alloys powder milled for 60 h.
Materials 12 00383 g002
Figure 3. (a) Differential scanning calorimetry (DSC) curves of the Al70Fe12.5V12.5Nb5 amorphous alloy at the heating rates of 10, 20, 30, and 40 K/min. (b) Enlarged area of the DSC curves shown in (a) of the Tg.
Figure 3. (a) Differential scanning calorimetry (DSC) curves of the Al70Fe12.5V12.5Nb5 amorphous alloy at the heating rates of 10, 20, 30, and 40 K/min. (b) Enlarged area of the DSC curves shown in (a) of the Tg.
Materials 12 00383 g003
Figure 4. The Kissinger and Ozawa plots of Tx, Tp1, Tp2of the Al70Fe12.5V12.5Nb5 amorphous alloy. (a) Kissinger plot, (b) Ozawa plot.
Figure 4. The Kissinger and Ozawa plots of Tx, Tp1, Tp2of the Al70Fe12.5V12.5Nb5 amorphous alloy. (a) Kissinger plot, (b) Ozawa plot.
Materials 12 00383 g004
Figure 5. XRD patterns of the Al70Fe12.5V12.5Nb5 amorphous alloy powders after heat treatment: (a) fast cooling after heating to 773 K; (b) annealing at 773 K for 100 min; (c) fast cooling after heating to 973 K; (d) annealing at 973 K for 100 min; (e) fast cooling after heating to 1173 K.
Figure 5. XRD patterns of the Al70Fe12.5V12.5Nb5 amorphous alloy powders after heat treatment: (a) fast cooling after heating to 773 K; (b) annealing at 773 K for 100 min; (c) fast cooling after heating to 973 K; (d) annealing at 973 K for 100 min; (e) fast cooling after heating to 1173 K.
Materials 12 00383 g005
Table 1. The thermal properties and ΔTx, Trg, γ of Al70Fe12.5V12.5Nb5 measured at 10, 20, 30, and 40 K/min heating rates.
Table 1. The thermal properties and ΔTx, Trg, γ of Al70Fe12.5V12.5Nb5 measured at 10, 20, 30, and 40 K/min heating rates.
β(K/min)Tg(K)Tx (K)Tp1 (K)Tp2 (K)Tm(K)Tl (K)ΔTx(K)Trgγ
10726.7844.2866.21097.71375.21408.7117.50.5160.395
20736.9854.1881.01112.41377.91407.9117.20.5220.398
30749.8862.5888.81120.11377.51416.9112.70.5290.398
40759.0865.5890.91125.71376.51427.3106.50.5320.396
Table 2. Energies of Ex, Ep1, Ep2 estimated with the Kissinger and Ozawa methods.
Table 2. Energies of Ex, Ep1, Ep2 estimated with the Kissinger and Ozawa methods.
EquationsEx (KJ/mol)Ep1 (KJ/mol)Ep2 (KJ/mol)
Kissinger366.3 ± 23.9326.2 ± 35.0489.4 ± 12.2
Ozawa380.5 ± 23.9340.8 ± 35.0507.8 ± 12.2

Share and Cite

MDPI and ACS Style

Liu, X.; Wang, X.; Si, Y.; Zhong, X.; Han, F. Crystallization Behavior of Al70Fe12.5V12.5Nb5 Amorphous Alloy Formed by Mechanical Alloying. Materials 2019, 12, 383. https://doi.org/10.3390/ma12030383

AMA Style

Liu X, Wang X, Si Y, Zhong X, Han F. Crystallization Behavior of Al70Fe12.5V12.5Nb5 Amorphous Alloy Formed by Mechanical Alloying. Materials. 2019; 12(3):383. https://doi.org/10.3390/ma12030383

Chicago/Turabian Style

Liu, Xuan, Xingfu Wang, Yongli Si, Xiaokang Zhong, and Fusheng Han. 2019. "Crystallization Behavior of Al70Fe12.5V12.5Nb5 Amorphous Alloy Formed by Mechanical Alloying" Materials 12, no. 3: 383. https://doi.org/10.3390/ma12030383

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