Next Article in Journal
Online Tool Wear Monitoring by the Analysis of Cutting Forces in Transient State for Dry Machining of Ti6Al4V Alloy
Next Article in Special Issue
Preparation and Characterization of Mg-RE Alloy Sheets and Formation of Amorphous/Crystalline Composites by Twin Roll Casting for Biomedical Implant Application
Previous Article in Journal
The Microstructure, Mechanical Properties, and Corrosion Resistance of UNS S32707 Hyper-Duplex Stainless Steel Processed by Selective Laser Melting
Previous Article in Special Issue
Metallic Glasses: A New Approach to the Understanding of the Defect Structure and Physical Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dynamic Mechanical Relaxation in LaCe-Based Metallic Glasses: Influence of the Chemical Composition

1
School of Mechanics, Civil Engineering Architecture, Northwestern Polytechnical University, Xi’an 710072, China
2
Departament de Física, Barcelona Research Center in Multiscale Science and Technology & Institut de Tècniques Energètiques, Universitat Politècnica de Catalunya, 08930 Barcelona, Spain
3
MATEIS, UMR CNRS5510, Bat. B. Pascal, INSA-Lyon, F-69621 Villeurbanne CEDEX, France
*
Author to whom correspondence should be addressed.
Metals 2019, 9(9), 1013; https://doi.org/10.3390/met9091013
Submission received: 13 August 2019 / Revised: 3 September 2019 / Accepted: 14 September 2019 / Published: 17 September 2019
(This article belongs to the Special Issue Recent Advancements in Metallic Glasses)

Abstract

:
The mechanical relaxation behavior of the (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, and 0.8) metallic glasses was probed by dynamic mechanical analysis. The intensity of the secondary β relaxation increases along with the Co/Cu ratio, as has been reported in metallic glasses where the enthalpy of mixing for all pairs of atoms is negative. Furthermore, the intensity of the secondary β relaxation decreases after physical aging below the glass transition temperature, which is probably due to the reduction of the atomic mobility induced by physical aging.

1. Introduction

Metallic glasses (MGs) have been extensively studied for several decades because they exhibit unique physical, chemical and mechanical properties and have no crystal defects (i.e., dislocations, grain boundaries, and vacancies) [1,2,3]. Compared to other glassy materials (i.e., amorphous polymers, oxide glasses, and other non-crystalline solids), metallic glasses show high yield strength and resilience, large fracture toughness, and attractive corrosion resistance [4,5,6,7,8]. It is well known that the mechanical and physical properties of metallic glasses, i.e., plasticity, glass transition behavior, and diffusion phenomena, are bound up with their mechanical relaxation modes [1,6,9,10,11]. Nevertheless, below the glass transition temperature, metallic glasses have insufficient ductility due to shear band instability during plastic deformation, which dramatically reduces the use in structural applications [12]. Compared to crystalline metals, metallic glasses are basically characterized by brittleness at room temperature. One way to overcome the macroscopic brittle behavior of metallic glass is to reduce the size. Previous studies have shown that brittle behavior can be mitigated when the sample size is reduced to sub-micron levels, thereby reducing the effects of instability on material behavior [13,14]. Below the glass transition, metallic glasses are thermodynamically in a non-equilibrium state, as there is a large enthalpy difference from the crystallized state [15]. However, the topological structure as well as the physical mechanism of relaxation in glassy materials remain unresolved issues [16,17,18].
In the supercooled liquid phase region of metallic glasses, relaxation processes drive the glass towards more stable states. The primary α relaxation and secondary β relaxation are considered the elementary relaxation processes [19,20,21]. Johari et al. [22] proposed that glasses and glass-forming liquids have two relaxation modes: (i) The primary (α) relaxation, which is a global, structural atomic or molecular rearrangement observed at relatively high temperatures and closely related to the glass transition phenomenon; (ii) The secondary (β) relaxation, a low energy process which is observed under the glass transition temperature Tg. While the α relaxation process shows a complex dependence on temperature, the secondary β relaxation generally submits to an Arrhenius temperature dependence rule. The β relaxation shows up as an over wing in the high-frequency tail or a side shoulder at low temperature of α relaxation [23,24]. Contrary to the main relaxation, β relaxation is associated with the motion of atoms or molecules inside a glass material without topological rearrangement. The study on the connection between the diffusion behavior of the amorphous alloy, plastic deformation, and glass transition on the relaxation process of amorphous alloy is of great significance for the assessment of the potential applications of metallic glasses.
Literature results show that La-based metallic glasses display a conspicuous secondary relaxation [25]. Therefore, La-based metallic glasses are an ideal model system to investigate the relaxation process. In the present study, the dynamic mechanical properties of emblematic LaCe-based metallic glasses were investigated by mechanical spectroscopy. The physical mechanism of mechanical relaxation process was analyzed relied on the Kohlrausch–Williams–Watts (KWW) equation.

2. Experimental Procedure

The (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6 and 0.8) precursor alloy was produced by arc melting in high-purity argon atmosphere, after titanium melting for the removal of residual oxygen. The alloy was re-melted at least four times to ensure chemical homogeneity. Suction casting copper method was eventually used to produce plates of 2 mm thickness.
The structural properties of the samples were checked by X-ray diffraction (XRD, Philips PW 3830, Amsterdam, Netherlands) using monochromatic Cu-Kα radiation. The glass transition Tg and crystallization onset Tx temperatures were determined by differential scanning calorimeter (DSC, NEZTCH 404 C, Bavaria, Germany) at a heating rate of 10 K/min. Dynamical mechanical analysis (DMA, TA Q800, USA) was used to monitor the mechanical relaxation behavior. Samples of 30 mm (length) × 2 mm (width) × 1 mm (thickness) for DMA analysis were produced on a precise water-cooled low-speed cutting machine. DMA measurements were performed on single cantilever, at a 3 K/min heating rate; the complex elastic modulus is denoted as E (storage modulus E′ and the loss modulus E″).

3. Experimental Results and Discussions

3.1. Structural and Thermal Properties

The amorphous nature of (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6 and 0.8) was confirmed by XRD. XRD patterns of (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, and 0.8) bulk metallic glasses, as presented in Figure 1a, exhibit broad diffraction peaks, and no traces of crystalline phases are detected. Therefore, the glassy nature of the (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, and 0.8) bulk metallic glasses (BMG) was verified.
DSC curves of the (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6and 0.8) bulk metallic glasses are presented in Figure 1b. The main thermal events are the glass transition and subsequent crystallization. The glass transition temperatures Tg, indicated by arrows in the figure, increase almost linearly with the substitution of Copper by Cobalt. While the Co-free ((La0.5Ce0.5)65Al10Cu25) alloy has a glass transition temperature of 372 K, the Tg of (La0.5Ce0.5)65Al10(Co0.8Cu0.2)25 alloy increases up to 404 K. Crystallization temperatures Tx show a less predictable behavior. Given that the difference between the crystallization and glass transition temperatures is a parameter largely related to the glass stability, it is observed that (La0.5Ce0.5)65Al10(Co0.2Cu0.8)25 metallic glass is the most stable glass. The XRD patterns and DSC curves corroborate the amorphous properties of the studied alloys.

3.2. Dynamic Mechanical Analysis

3.2.1. Constant Frequency Measurements

Dynamic mechanical analysis is very sensitive to the atomic and molecular mobility of amorphous materials. The (La0.5Ce0.5)65Al10(Co0.4Cu0.6)25 metallic glass dynamic mechanical response was measured from room temperature to 550 K with a heating rate of 3 K/min and a driving frequency of 1 Hz. Figure 2 exhibits the normalized storage modulus E′ and loss modulus E″ as a function of temperature, where Eu represents the unrelaxed modulus at room temperature. The temperature dependence of E′ and E″ is very similar to most other BMGs [9,26,27]. It is worth noticing that when the temperature is below the Tg, loss modulus curve showed no apparent β relaxation [9,28,29]. The storage modulus and loss modulus of the LaCe-based metallic glass vary with temperature, and the process can be segmented into three different regions:
Region (I): In the low temperature region, i.e., beneath 350 K, the normalized storage modulus is large and close to unity. Contrarily, the loss modulus E″ is negligible. Therefore, the glassy material mainly exhibits elastic deformation in this temperature range, and the viscoelastic component can be neglected.
Region (II): The medium temperature region comprises the temperature range of 390 K to 460 K. A large increase of the loss modulus E″ is observed reaching a peak at the temperature Tα, denoting the α relaxation characteristic of amorphous materials. This process is the dynamic glass transition. The storage modulus E′ starts to diminish while the loss modulus E″ boosts. This temperature range falls within the super-cooled liquid region of metallic glasses.
Region (III): The high temperature region starts at 460 K. The storage modulus E′ increases again reaching a value similar to that at room temperature, due to crystallization. The loss modulus E″ falls to a relatively low value.
As shown in Figure 2, the β relaxation is not observed below Tg [22,23]. β relaxation in this glass should be found in the temperature range from 300 to 400 K.
Figure 3 exhibits the normalized dynamic loss modulus E″ as a function of temperature at a constant frequency (1 Hz) in (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, and 0.8) glass alloys. The data are normalized to the values of temperature and loss modulus at the peak of the α relaxation, namely Tα and E″max. Interestingly, the secondary relaxation process depends significantly on the chemical composition of the glass. The intensity of β relaxation of (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, and 0.8) decreases with the increase of the Cu content. The behavior of the loss module reveals a noticeable change in the features of the β relaxation around 0.8 Tg for the different compositions. For (La0.5Ce0.5)65Al10Cu25 metallic glass, the β relaxation merely declares as a weak shoulder. It is significant that in several metallic glasses that exhibit an evident β relaxation, this has been correlated to plasticity [30]. It has also been proven that the mechanical relaxation process, especially the β relaxation, is sensitive to the micro-alloying in metallic glasses.
Previous works indicate that the β relaxation reflects the inherent structural heterogeneities in metallic glasses, described for instance as soft domains, liquid-like regions, local topological structure of loose packing regions, and flow units [1]. In the current study, minor addition of Cobalt in the (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, and 0.8) bulk metallic glasses is very important and reshapes the relaxation mode (i.e., β relaxation).
In order to associate the different behavior of β relaxation and the deformability of metallic glass with its structure, transmission electron microscopy (TEM) was carried out to reveal the microstructural characteristics of metallic glass [31]. The most notable structural feature is that the metallic glass is composed of two types of regions: Light regions with typical sizes ranging from 50 to 200 nm are enveloped by dark boundary regions, which are about 5–20 nm in width. It was further confirmed that both of the two regions are of a glassy nature. It is proposed that the local atomic motions of soft regions are responsible for β relaxations, and the heterogeneous structure improves the plasticity of metallic glasses though the formation of multiple shear bands [32].
The relationship between enthalpy of mixing and β relaxation can be used to qualitatively anticipate the intensity β relaxation of some metallic glasses [33]. An empirical rule on β relaxation has been established [34]: Pronounced β relaxation is associated with alloys where all the atomic pairs have larger and similar negative values of the mixing enthalpy. On the other hand, positive or large fluctuations in the values of mixing enthalpy reduce and even suppress β relaxation [33].
Regarding the β relaxation in (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% metallic glasses, the features of the mixing enthalpy between the constituent atoms correspond to the apparent β relaxation. Figure 4 shows the mixing enthalpy of the constituents of the (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, and 0.8) metallic glasses (The data of mixing enthalpy are derived from the reference [35]). The mixing enthalpy △Hm of the “solvent” atoms, La/Ce, with the “solute” atoms are almost identical: △Hm (La/Ce-Al) = −38 kJ/mol, △Hm (La/Ce-Cu) = −21 kJ/mol, △Hm (La-Co) = −17 kJ/mol and △Hm (La-Cu) = −18 kJ/mol, reflecting the chemical similar chemistry of the rare-earth elements. As for the “solute” atoms, the mixing enthalpy of Cu-Co is positive, 6 kJ/mol, and the main difference appears when comparing the mixing enthalpies of Al-Cu, −1 kJ/mol, to that of Al-Co, −18 kJ/mol. Based on the empirical rules to determine ∆Hmix, the mixing enthalpy of (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, 0.8) metallic glasses is given in Figure 4. We observed an actual decrease of the enthalpy of mixing as the concentration of Co increases. However, due to substitution of Cu by Co, the fluctuation on mixing enthalpies decreases, as the Al-Co mixing enthalpy is substantially more negative than that of Al-Cu and similar to those of La/Ce-Cu. According to the literature, this reduction on the mixing enthalpy fluctuation enhances the β relaxation [27].
Previous works proved that relaxation is connected to dynamic heterogeneity in glasses and related to the local movement of “weak spots” [36]. In particular, the microstructure inhomogeneity of metallic glass has been proven by means of microscopy and simulation [37].

3.2.2. Physical Aging on the Secondary Relaxation of LaCe-Based Metallic Glass

From the thermodynamics point of view, annealing below the Tg drives the glassy state towards a more stable state of lower energy. Figure 5 presents the storage and loss factor of (La0.5Ce0.5)65Al10(Co0.8Cu0.2)25 metallic glass after annealing at 363 K for 24 h, which certainly illustrates that the intensity of the β relaxation reduces by physical aging below Tg.
As proposed in previous works, the β relaxation of metallic glasses is ascribed to the structural heterogeneity or local motion of the “defects” [1,33]. These defects are denominated as flow units [33,38], quasi-point defects (QPDs) [39], liquid-like sites [40], weakly bonded zones or loose packing regions [41]. According to Figure 5, annealing below the glass transition temperature can lead to disappearance of “defects” in metallic glasses. Physical aging causes rearrangement of atoms, resulting in an increase of density and elastic modulus. In the metallic glass, the mobility of atoms is closely related to “defects” concentration. Annealing causes the metallic glass to evolve towards a higher density state with a consequent reduction of the local “free volume” available for atomic rearrangement. Subsequent cooling after the annealing does not alter the glassy state, since the cooling rate is much lower than in the initial production of the glass.
In addition, physical aging below the glass transition temperature Tg leads to enthalpy relaxation of glassy materials. Figure 6 shows the DSC trace of the (La0.5Ce0.5)65Al10(Co0.8Cu0.2)25 metallic glass annealed at 363 K for 24 h. Comparison to the as-produced sample allows the identification of a notable enthalpy recovery, a consequence of the glass relaxation during annealing.
The structural relaxation observed in the DMA test can be analyzed by the growth of the loss factor (tan δ = E″/E′) [42,43]. Figure 7 shows the loss factor (tan δ) evolution versus annealing time in the (La0.5Ce0.5)65Al10(Co0.8Cu0.2)25 bulk metallic glass at 363 K. Previous works have characterized the structural relaxation below Tg in amorphous materials, particularly in bulk metallic glasses, by using the Kohlrausch–Williams–Watts (KWW) equation [9,44].
tan δ ( t a ) tan δ ( t a = 0 ) = A { 1 e [ ( t a τ ) β a g i n g ] }
where A = tan δ ( t a ) tan δ ( t a = 0 ) is the maximum value of the dynamic relaxation. τ is the relaxation time, and βaging is the Kohlrausch exponent with values between 0 and 1. The best fit of Equation (1) to the data on the loss factor of the (La0.5Ce0.5)65Al10(Co0.8Cu0.2)25 metallic glass was obtained with τ = 10,594 s and βaging = 0.4.
The Kohlrausch exponent βKWW reveals the presence of a broad distribution of relaxation times in the glass, with βKWW aging = 1 corresponding to a single Debye relaxation time.
Experimental values of the parameter βKWW in amorphous alloys are in the range of 0.24–1 [45]. For amorphous polymers, values extend from 0.24 (polyvinyl chloride) to 0.55 (polyisobutylene), for alcohols from 0.45 to 0.75, while for orientational glasses and networks values are up to 1. In bulk metallic glasses, it appears that βKWW is related to the fragility of the amorphous materials [42]. Values of βKWW close to 1 indicate that the system is a strong glass former while values less than 0.5 suggest that the glass is a fragile glass [46,47]. According to the available literature, no defined trend characterizes the stretching parameters βKWW in bulk metallic glasses, as it is either temperature dependent or temperature independent [48,49,50]. For example, Pd42.5Ni7.5Cu30P20 bulk metallic glass has very similar fragility parameters, 59 < m < 67, and similar stretched exponents, 0.59 < βKWW < 0.6 [44]. However, experimental data indicate that the Kohlrausch exponent βaging is around 0.4 for temperatures close to Tg [44], as found in the present case.
As proposed by Wang et al. [51], the kinetic parameter βKWW is associated with the dynamic heterogeneity. The βKWW parameter takes low values when the temperature is below the β relaxation peak and increases dramatically when the temperature surpasses the glass transition temperature. The β relaxation in metallic glasses is related to the reversible displacement of the “defects”. When the stress relaxation is performed around the β relaxation temperature, only a small fraction of atoms are allowed to move. Thus, it can be concluded that lower βKWW values around the β relaxation are ascribed to reversible “defects”.

4. Conclusions

The dynamic mechanical properties of LaCe-based bulk metallic glasses were investigated by dynamic mechanical analysis. The experimental results reveal that the mechanical and thermal properties strongly depend on chemical composition. Substitution of Cu by Co enhances β relaxation due to the reduction of the enthalpy of mixing fluctuation. Physical aging below the glass transition temperature Tg reduces the concentration of local “defects” in metallic glasses and causes a decrease of the intensity for the β process. The curves of loss modulus can be well fitted by the KWW model, the fitting parameter βKWW is approximately 0.4, indicating that the plastic deformation of the (La0.5Ce0.5)65Al10(Co0.8Cu0.2)25 metallic glass is related to the microstructural heterogeneity.

Author Contributions

M.L., Q.H. and Y.C. conducted the experiments and analyzed the experimental results. Q.H. and M.L. contributed to the sample preparation. M.L., J.Q., D.C., Y.Y. and J.-M.P. proposed the idea of the work, originated the experiment and write the paper.

Funding

This research was funded by the Fundamental Research Funds for the Central Universities (Nos. 3102018ZY010, 3102019ghxm007 and 3102017JC01003), Astronautics Supporting Technology Foundation of China (2019-HT-XG) and the Natural Science Foundation of Shaanxi Province (No. 2019JM-344). The investigation of Y.H. Chen and M.N. Liu sponsored by the Seed Foundation of Innovation and Creation for Graduate Students in Northwestern Polytechnical University (No. ZZ2019014). D. Crespo acknowledges financial support from MINECO, Spain, grant FIS2017-82625-P, and Generalitat de Catalunya, grant 2017SGR0042.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Qiao, J.C.; Wang, Q.; Pelletier, J.M.; Kato, H.; Casalini, R.; Crespo, D.; Pineda, E.; Yao, Y.; Yang, Y. Structural heterogeneities and mechanical behavior of amorphous alloys. Prog. Mater. Sci. 2019, 104, 250–329. [Google Scholar] [CrossRef]
  2. Hufnagel, T.C.; Schuh, C.A.; Falk, M.L. Deformation of metallic glasses: recent developments in theory, simulations, and experiments. Acta Mater. 2016, 109, 375–393. [Google Scholar] [CrossRef]
  3. Wang, W.H. The elastic properties, elastic models and elastic perspectives of metallic glasses. Prog. Mater. Sci. 2012, 57, 487–656. [Google Scholar] [CrossRef]
  4. Inoue, A. Stabilization of metallic supercooled liquid and bulk amorphous alloys. Acta Mater. 2000, 48, 279–306. [Google Scholar] [CrossRef]
  5. Sun, B.A.; Wang, W.H. The fracture of bulk metallic glasses. Prog. Mater. Sci. 2015, 74, 211–307. [Google Scholar] [CrossRef]
  6. Qiao, J.C.; Chen, Y.H.; Casalini, R.; Pelletier, J.M.; Yao, Y. Main relaxation and slow relaxation processes in a La30Ce30Al15Co25 metallic glass. J. Mater. Sci. Technol. 2019, 35, 982–986. [Google Scholar] [CrossRef]
  7. Zhang, L.-C.; Jia, Z.; Lyu, F.; Liang, S.-X.; Lu, J. A review of catalytic performance of metallic glasses in wastewater treatment: Recent progress and prospects. Prog. Mater. Sci. 2019, 105, 100576. [Google Scholar] [CrossRef]
  8. Jia, Z.; Wang, Q.; Sun, L.; Wang, Q.; Zhang, L.-C.; Wu, G.; Luan, J.-H.; Jiao, Z.-B.; Wang, A.; Liang, S.-X.; et al. Attractive in situ self-reconstructed hierarchical gradient structure of metallic glass for high efficiency and remarkable stability in catalytic performance. Adv. Funct. Mater. 2019, 29, 1807857. [Google Scholar] [CrossRef]
  9. Qiao, J.C.; Pelletier, J.M. Dynamic mechanical relaxation in bulk metallic glasses: A review. J. Mater. Sci. Technol. 2014, 30, 523–545. [Google Scholar] [CrossRef]
  10. Schuh, C.A.; Hufnagel, T.C.; Ramamurty, U. Mechanical behavior of amorphous alloys. Acta Mater. 2007, 55, 4067–4109. [Google Scholar] [CrossRef]
  11. Wang, Q.; Liu, J.J.; Ye, Y.F.; Liu, T.T.; Wang, S.; Liu, C.T.; Lu, J.; Yang, Y. Universal secondary relaxation and unusual brittle-to-ductile transition in metallic glasses. Mater. Today 2017, 20, 293–300. [Google Scholar] [CrossRef]
  12. Ghidelli, M.; Gravier, S.; Blandin, J.J.; Djemia, P.; Mompiou, F.; Abadias, G.; Raskin, J.P.; Pardoen, T. Extrinsic mechanical size effects in thin ZrNi metallic glass films. Acta Mater. 2015, 90, 232–241. [Google Scholar] [CrossRef]
  13. Ghidelli, M.; Idrissi, H.; Gravier, S.; Blandin, J.-J.; Raskin, J.-P.; Schryvers, D.; Pardoen, T. Homogeneous flow and size dependent mechanical behavior in highly ductile Zr65Ni35 metallic glass films. Acta Mater. 2017, 131, 246–259. [Google Scholar] [CrossRef]
  14. Han, Z.; Wu, W.F.; Li, Y.; Wei, Y.J.; Gao, H.J. An instability index of shear band for plasticity in metallic glasses. Acta Mater. 2009, 57, 1367–1372. [Google Scholar] [CrossRef]
  15. Leuzzi, L.; Nieuwenhuizen, T.M. Thermodynamics of the glassy state. J. Stat. Phys. 2008, 133, 1185–1186. [Google Scholar]
  16. Dixon, P.K.; Lei, W.; Nagel, S.R.; Williams, B.D.; Carini, J.P. Scaling in the relaxation of supercooled liquids. Phys. Rev. Lett. 1990, 65, 1108–1111. [Google Scholar] [CrossRef]
  17. Ngai, K.; Wang, Z.; Gao, X.; Yu, H.; Wang, W. A connection between the structural α-relaxation and the β-relaxation found in bulk metallic glass-formers. J. Chem. Phys. 2013, 139, 014502. [Google Scholar] [CrossRef]
  18. Qiao, J.C.; Wang, Q.; Crespo, D.; Yang, Y.; Pelletier, J.M. Amorphous physics and materials: Secondary relaxation and dynamic heterogeneity in metallic glasses: A brief review. Chin. Phys. B 2017, 26, 016402. [Google Scholar] [CrossRef] [Green Version]
  19. Angell, C.A. Formation of glasses from liquids and biopolymers. Science 1995, 267, 1924–1935. [Google Scholar] [CrossRef]
  20. Debenedetti, P.G.; Stillinger, F.H. Supercooled liquids and the glass transition. Nature 2001, 410, 259–267. [Google Scholar] [CrossRef]
  21. Zhao, Z.; Wen, P.; Shek, C.; Wang, W. Measurements of slow β-relaxations in metallic glasses and supercooled liquids. Phys. Rev. B 2007, 75, 174201. [Google Scholar] [CrossRef]
  22. Johari, G.P.; Goldstein, M. Viscous liquids and the glass transition. II. Secondary relaxations in glasses of rigid molecules. J. Chem. Phys. 1970, 53, 2372–2388. [Google Scholar]
  23. Ngai, K.L.; Rendell, R.W. Cooperative dynamics in relaxation: A coupling model perspective. J. Mol. Liq. 1993, 56, 199–214. [Google Scholar] [CrossRef]
  24. Kudlik, A.; Benkhof, S.; Blochowicz, T.; Tschirwitz, C.; Rössler, E. The dielectric response of simple organic glass formers. J. Mol. Struct. 1999, 479, 201–218. [Google Scholar] [CrossRef]
  25. Qiao, J.C.; Wang, Y.-J.; Pelletier, J.M.; Keer, L.M.; Fine, M.E.; Yao, Y. Characteristics of stress relaxation kinetics of La60Ni15Al25 bulk metallic glass. Acta Mater. 2015, 98, 43–50. [Google Scholar] [CrossRef]
  26. Lyu, G.J.; Qiao, J.C.; Pelletier, J.M.; Yao, Y. The dynamic mechanical characteristics of Zr-based bulk metallic glasses and composites. Mater. Sci. Eng., A 2018, 711, 356–363. [Google Scholar] [CrossRef]
  27. Qiao, J.C.; Yao, Y.; Pelletier, J.M.; Keer, L.M. Understanding of micro-alloying on plasticity in Cu46Zr47−xAl7Dyx (0 ≤ x ≤ 8) bulk metallic glasses under compression: Based on mechanical relaxations and theoretical analysis. Int. J. Plast. 2016, 82, 62–75. [Google Scholar] [CrossRef]
  28. Yao, Z.F.; Qiao, J.C.; Pelletier, J.M.; Yao, Y. Characterization and modeling of dynamic relaxation of a Zr-based bulk metallic glass. J. Alloys Compd. 2017, 690, 212–220. [Google Scholar] [CrossRef]
  29. Qiao, J.C.; Casalini, R.; Pelletier, J.M.; Yao, Y. Dynamics of the strong metallic glass Zn38Mg12Ca32Yb18. J. Non-Cryst. Solids 2016, 447, 85–90. [Google Scholar] [CrossRef]
  30. Yu, H.B.; Wang, W.H.; Bai, H.Y.; Samwer, K. The β-relaxation in metallic glasses. Natl. Sci. Rev. 2014, 1, 429–461. [Google Scholar] [CrossRef]
  31. Yu, H.B.; Shen, X.; Wang, Z.; Gu, L.; Wang, W.H.; Bai, H.Y. Tensile plasticity in metallic glasses with pronounced β relaxations. Phys. Rev. Lett. 2012, 108, 015504. [Google Scholar] [CrossRef]
  32. Ichitsubo, T.; Matsubara, E.; Yamamoto, T.; Chen, H.S.; Nishiyama, N.; Saida, J.; Anazawa, K. Microstructure of fragile metallic glasses inferred from ultrasound-accelerated crystallization in Pd-based metallic glasses. Phys. Rev. Lett. 2005, 95, 245501. [Google Scholar] [CrossRef]
  33. Wang, W.H. Dynamic relaxations and relaxation-property relationships in metallic glasses. Prog. Mater. Sci. 2019, 106, 100561. [Google Scholar] [CrossRef]
  34. Yu, H.B.; Samwer, K.; Wang, W.H.; Bai, H.Y. Chemical influence on β-relaxations and the formation of molecule-like metallic glasses. Nat. Commun. 2013, 4, 1345–1346. [Google Scholar]
  35. 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]
  36. Liu, S.T.; Wang, Z.; Peng, H.L.; Yu, H.B.; Wang, W.H. The activation energy and volume of flow units of metallic glasses. Scr. Mater. 2012, 67, 9–12. [Google Scholar] [CrossRef]
  37. Liu, Y.H.; Wang, D.; Nakajima, K.; Zhang, W.; Hirata, A.; Nishi, T.; Inoue, A.; Chen, M.W. Characterization of nanoscale mechanical heterogeneity in a metallic glass by dynamic force microscopy. Phys. Rev. Lett. 2011, 106, 125504. [Google Scholar] [CrossRef]
  38. Liu, S.T.; Jiao, W.; Sun, B.A.; Wang, W.H. A quasi-phase perspective on flow units of glass transition and plastic flow in metallic glasses. J. Non-Cryst. Solids 2013, 376, 76–80. [Google Scholar] [CrossRef]
  39. Perez, J.; Etienne, S.; Tatibouät, J. Determination of glass transition temperature by internal friction measurements. Phys. Status Solidi a 1990, 121, 129–138. [Google Scholar] [CrossRef]
  40. Egami, T. Mechanical failure and glass transition in metallic glasses. J. Alloys Compd. 2011, 509, S82–S86. [Google Scholar] [CrossRef]
  41. Qiao, J.C.; Pelletier, J.M. Kinetics of structural relaxation in bulk metallic glasses by mechanical spectroscopy: Determination of the stretching parameter βKWW. Intermetallics 2012, 28, 40–44. [Google Scholar] [CrossRef]
  42. Qiao, J.; Pelletier, J.M.; Casalini, R. Relaxation of bulk metallic glasses studied by mechanical spectroscopy. J. Phys. Chem. B 2013, 117, 13658–13666. [Google Scholar] [CrossRef]
  43. Pelletier, J.M. Influence of structural relaxation on atomic mobility in a Zr41.2Ti13.8Cu12.5Ni10.0Be22.5 (Vit1) bulk metallic glass. J. Non-Cryst. Solids 2008, 354, 3666–3670. [Google Scholar]
  44. Qiao, J.; Casalini, R.; Pelletier, J.M.; Kato, H. Characteristics of the structural and Johari–Goldstein Relaxations in Pd-based metallic glass-forming liquids. J. Phys. Chem. B 2014, 118, 3720–3730. [Google Scholar] [CrossRef]
  45. Böhmer, R.; Ngai, K.L.; Angell, C.A.; Plazek, D.J. Nonexponential relaxations in strong and fragile glass formers. J. Chem. Phys. 1993, 99, 4201–4209. [Google Scholar] [CrossRef]
  46. Angell, A.C. Relaxation in liquids, polymers and plastic crystals—Strong/fragile patterns and problems. J. Non-Cryst. Solids 1991, 131–133, 13–31. [Google Scholar] [CrossRef]
  47. Raghavan, R.; Murali, P.; Ramamurty, U. Influence of cooling rate on the enthalpy relaxation and fragility of a metallic glass. Metall. Mater. Trans. A 2008, 39, 1573–1577. [Google Scholar] [CrossRef]
  48. Zhang, Y.; Hahn, H. Study of the kinetics of free volume in Zr45.0Cu39.3Al7.0Ag8.7 bulk metallic glasses during isothermal relaxation by enthalpy relaxation experiments. J. Non-Cryst. Solids 2009, 355, 2616–2621. [Google Scholar]
  49. Zhang, T.; Ye, F.; Wang, Y.; Lin, J. Structural Relaxation of La55Al25Ni10Cu10 bulk metallic glass. Metall. Mater. Trans. A 2008, 39, 1953–1957. [Google Scholar] [CrossRef]
  50. Gallino, I.; Shah, M.B.; Busch, R. Enthalpy relaxation and its relation to the thermodynamics and crystallization of the Zr58.5Cu15.6Ni12.8Al10.3Nb2.8 bulk metallic glass-forming alloy. Acta Mater. 2007, 55, 1367–1376. [Google Scholar]
  51. Wang, Z.; Sun, B.A.; Bai, H.Y.; Wang, W.H. Evolution of hidden localized flow during glass-to-liquid transition in metallic glass. Nat. Commun. 2014, 5, 5823. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of the (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, and 0.8), as-cast state. (b) Differential scanning calorimeter (DSC) curves of the (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, and 0.8) at a heating rate of 10 K/min. The glass transition temperature Tg of the different metallic glasses is pointed out in the figure.
Figure 1. (a) XRD patterns of the (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, and 0.8), as-cast state. (b) Differential scanning calorimeter (DSC) curves of the (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, and 0.8) at a heating rate of 10 K/min. The glass transition temperature Tg of the different metallic glasses is pointed out in the figure.
Metals 09 01013 g001
Figure 2. Thermal dependence of the normalized storage E′/Eu and loss modulus E″/Eu of (La0.5Ce0.5)65Al10(Co0.4Cu0.6)25 metallic glass. Measurement was carried out at a fixed frequency of 1 Hz and a heating rate of 3 K/min. Eu is the unrelaxed modulus, which equals the value of E′ at room temperature.
Figure 2. Thermal dependence of the normalized storage E′/Eu and loss modulus E″/Eu of (La0.5Ce0.5)65Al10(Co0.4Cu0.6)25 metallic glass. Measurement was carried out at a fixed frequency of 1 Hz and a heating rate of 3 K/min. Eu is the unrelaxed modulus, which equals the value of E′ at room temperature.
Metals 09 01013 g002
Figure 3. Temperature dependence of the loss modulus E″/E″max in the (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, 0.8) metallic glass (Heating rate: 3 K/min; frequency: 1 Hz).
Figure 3. Temperature dependence of the loss modulus E″/E″max in the (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, 0.8) metallic glass (Heating rate: 3 K/min; frequency: 1 Hz).
Metals 09 01013 g003
Figure 4. Mixing enthalpy of the (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, and 0.8) metallic glasses. The inset displays the mixing enthalpy of constituent atoms (the data are taken from the reference [35]).
Figure 4. Mixing enthalpy of the (La0.5Ce0.5)65Al10(CoxCu1−x)25 at% (x = 0, 0.2, 0.4, 0.6, and 0.8) metallic glasses. The inset displays the mixing enthalpy of constituent atoms (the data are taken from the reference [35]).
Metals 09 01013 g004
Figure 5. Temperature dependence of the normalized storage modulus (a) and loss factor (b) of (La0.5Ce0.5)65Al10(Co0.8Cu0.2)25 metallic glass (heating rate: 3 K/min and frequency: 0.3 Hz) (1) As-cast (2) annealed one (annealing temperature: 363 K and annealing time: 24 h).
Figure 5. Temperature dependence of the normalized storage modulus (a) and loss factor (b) of (La0.5Ce0.5)65Al10(Co0.8Cu0.2)25 metallic glass (heating rate: 3 K/min and frequency: 0.3 Hz) (1) As-cast (2) annealed one (annealing temperature: 363 K and annealing time: 24 h).
Metals 09 01013 g005
Figure 6. Enthalpy relaxation in (La0.5Ce0.5)65Al10(Co0.8Cu0.2)25 metallic glass bulk metallic glasses after aging at 363 K.
Figure 6. Enthalpy relaxation in (La0.5Ce0.5)65Al10(Co0.8Cu0.2)25 metallic glass bulk metallic glasses after aging at 363 K.
Metals 09 01013 g006
Figure 7. Evolution of the storage modulus E′ and loss factor tanδ for (La0.5Ce0.5)65Al10(Co0.8Cu0.2)25 metallic glass with the annealing time. The aging temperature is Tα = 363 K, and the driving frequency is 1 Hz. The red curve is the best fit from Equation (1) obtained for the parameters τ = 10,594 s and βaging = 0.4.
Figure 7. Evolution of the storage modulus E′ and loss factor tanδ for (La0.5Ce0.5)65Al10(Co0.8Cu0.2)25 metallic glass with the annealing time. The aging temperature is Tα = 363 K, and the driving frequency is 1 Hz. The red curve is the best fit from Equation (1) obtained for the parameters τ = 10,594 s and βaging = 0.4.
Metals 09 01013 g007

Share and Cite

MDPI and ACS Style

Liu, M.; Qiao, J.; Hao, Q.; Chen, Y.; Yao, Y.; Crespo, D.; Pelletier, J.-M. Dynamic Mechanical Relaxation in LaCe-Based Metallic Glasses: Influence of the Chemical Composition. Metals 2019, 9, 1013. https://doi.org/10.3390/met9091013

AMA Style

Liu M, Qiao J, Hao Q, Chen Y, Yao Y, Crespo D, Pelletier J-M. Dynamic Mechanical Relaxation in LaCe-Based Metallic Glasses: Influence of the Chemical Composition. Metals. 2019; 9(9):1013. https://doi.org/10.3390/met9091013

Chicago/Turabian Style

Liu, Minna, Jichao Qiao, Qi Hao, Yinghong Chen, Yao Yao, Daniel Crespo, and Jean-Marc Pelletier. 2019. "Dynamic Mechanical Relaxation in LaCe-Based Metallic Glasses: Influence of the Chemical Composition" Metals 9, no. 9: 1013. https://doi.org/10.3390/met9091013

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