Next Article in Journal
Microstructure and Residual Stresses of AA2519 Friction Stir Welded Joints under Different Heat Treatment Conditions
Next Article in Special Issue
Effect of Co Substitution on Crystallization and Magnetic Behavior of Fe85.45−xCoxCu0.55B14 Metallic Glass
Previous Article in Journal
Biodegradable Blends of Grafted Dextrin with PLGA-block-PEG Copolymer as a Carrier for Controlled Release of Herbicides into Soil
Previous Article in Special Issue
Study of the Morphology and Properties of Biocompatible Ca-P Coatings on Mg Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Correlation between High Temperature Deformation and β Relaxation in LaCe-Based Metallic Glass

School of Mechanics, Civil Engineering and Architecture, Northwestern Polytechnical University, Xi’an 710072, China
*
Author to whom correspondence should be addressed.
Materials 2020, 13(4), 833; https://doi.org/10.3390/ma13040833
Submission received: 14 December 2019 / Revised: 31 January 2020 / Accepted: 5 February 2020 / Published: 12 February 2020
(This article belongs to the Special Issue Structure and Properties of Crystalline and Amorphous Alloys)

Abstract

:
High-temperature deformation around the glass transition temperature Tg and the dynamic mechanical behavior of La30Ce30Al15Co25 metallic glass were investigated. According to dynamic mechanical analysis (DMA) results, La30Ce30Al15Co25 metallic glass exhibits a pronounced slow β relaxation process. In parallel, strain-rate jump experiments around the glass transition temperature were performed in a wide range of strain rate ranges. The apparent viscosity shows a strong dependence on temperature and strain rate, which reflects the transition from non-Newtonian to Newtonian flow. At low strain or high temperature, a transition was observed from a non-Newtonian viscous flow to Newtonian viscous flow. It was found that the activation volume during plastic deformation of La30Ce30Al15Co25 metallic glass is higher than that of other metallic glasses. Higher values of activation volume in La30Ce30Al15Co25 metallic glass may be attributed to existence of a pronounced slow β relaxation. It is reasonable to conclude that slow β relaxation in La30Ce30Al15Co25 metallic glass corresponds to the “soft” regions (structural heterogeneities) in metallic glass.

Graphical Abstract

1. Introduction

Over the past few decades, metallic glasses have attracted a great deal of interest in researchers due to their unique disordered structure. Unlike their crystalline counterparts, metallic glasses do not possess defects like grain boundaries and dislocations [1,2,3,4,5,6,7]. Therefore, they show specific mechanical, chemical and physical properties, such as high strength, high hardness and good corrosion resistance [1,4,8,9,10,11]. As a consequence, metallic glasses are great functional and structural materials from a point of practical application. However, brittle fractures at room temperature severely restrict their application as an engineering material [12]. It is important to tune the plasticity of metallic glasses. Interestingly, when the deformation temperature approaches the glass transition temperature Tg, metallic glasses usually show excellent thermoplastic capacity [13,14].
It is well documented that amorphous materials stay in a metastable state of energy. Structural relaxation occurs in amorphous materials over time, which is an essential and important feature of these materials. It is believed that there are two typical relaxation processes, i.e., the main α relaxation and the secondary β relaxation, in metallic glasses [8,15,16]. On the one hand, investigations have proven that β relaxation is a reversible process, which is linked to local atomic or molecular movement in amorphous materials [17], while on the other hand β relaxation is an irreversible relaxation, which is related to the dynamic glass transition phenomenon of amorphous materials [18]. Importantly, recent studies have shown that β relaxation may be associated with the plasticity of metallic glasses [17,19].
The plastic deformation of metallic glasses can be divided into two types: homogeneous deformation and inhomogeneous deformation. It has been reported that the mechanical deformation behavior of metallic glasses is sensitive to the temperature and strain rate [20,21]. Similarly, the effects of strain on the plastic deformation of polymers have also been studied, and related mechanisms have been discussed in previous studies [22,23]. Deformation is heterogeneous at low temperatures, which is accompanied by the nucleation and quick propagation of a single shear band, with a brittle fracture then occurring in metallic glasses [24]. At high temperatures (T > 0.8 Tg), the deformation is homogeneous, and metallic glasses show excellent plasticity [21,25,26]. Previous investigation suggested that the correlation between strength and temperature can be explained by the relaxation activation mechanism [27,28]. Some literature has reported that the transition from Newtonian to non-Newtonian rheology will occur with changes in the temperature and strain rate [21,26,29]. The variation of apparent viscosity is often used as an important indicator of the plastic deformation of metallic glasses. Many theories or models have been proposed to explain the deformation behavior of metallic glasses. The shear transformation zones (STZs) model presumes that atomic clusters undergo torsional deformation under external shear stress [30], whereas the free volume model assumes that plastic deformation is related to the generation and disappearance of free volumes [31].
In the current study, we chose La30Ce30Al15Co25 (at. %) metallic glass as a model alloy, due to the LaCe-based metallic glass showing a pronounced slow β relaxation based on the dynamic mechanical analysis [32]. Dynamic mechanical behavior and deformation at high temperatures in LaCe-based metallic glasses have been investigated. The deformation experimental data were analyzed in the framework of the physical models.

2. Experimental Procedure

A master alloy of a typical La30Ce30Al15Co25 (at. %) metallic glass was prepared by the arc melting method. In order to ensure the homogeneity of the chemical composition of the model alloy, the master alloy was melted at least six times. The metallic glass ribbon used in the current research was prepared by a single roll melt spinning technique in an argon atmosphere. The amorphous nature of the model alloy was determined by X-ray diffraction (XRD, D8 ADVANCE Bruker AXS Gmbh, Karlsruhe, Germany) using Cu Kα radiation at room temperature. The glass transition temperature and onset temperature of crystallization are determined by the differential scanning calorimetry (DSC) technique (commercial instrument NETZCH 404, Bavaria, Germany) with a heating rate of 20 K/min in a nitrogen atmosphere.
The dynamic mechanical behavior of La30Ce30Al15Co25 metallic glass was carried out in a commercial dynamic mechanical analysis (DMA, TA Q800, USA) in a tensile mode. The DMA experiment was performed at a fixed frequency of 1 Hz, with a constant heating rate of 2 K/min. The sample was subjected to a periodic sinusoidal strain, and its response stress was simultaneously measured. The complex form of modulus E * = E + i E can be obtained, where E is the loss modulus and E is the storage modulus.
High-temperature deformation experiments on La30Ce30Al15Co25 metallic glass ribbon were conducted in a film tensile mode in DMA Q800 at 435, 440, 445 and 450 K, respectively. Initially, La30Ce30Al15Co25 metallic glass ribbon was heated to the target temperature, at a rate of 20 K/min. The sample was then kept at the testing temperature for 5 min to ensure thermal equilibration prior to the tensile experiments. The strain rate ranged from 5 × 10−5 s−1 to 2.5 × 10−4 s−1.

3. Results and Discussion

The glass nature of the model alloy was verified prior to the experiment. Figure 1 displays the XRD pattern of as-cast La30Ce30Al15Co25 metallic glass ribbon. No crystalline peak was detected, which confirms the amorphous nature of the model alloy.
In order to explore the thermal characteristics of La30Ce30Al15Co25 metallic glass, the DSC measurement was performed at a heating rate of 20 K/min. The glass transition temperature Tg and onset temperature of crystallization Tx were determined by ASTM-E1356. As shown in Figure 2, the glass transition temperature Tg is 430 ± 2 K, and the onset temperature of crystallization Tx is 460 ± 2 K. Figure 3 shows the temperature dependence of the normalized storage modulus and loss modulus of La30Ce30Al15Co25 metallic glass. Three different regions can be observed: (1) At a low temperature region (from room temperature to 400 K), a peak corresponding to slow β relaxation around 348 K can be observed, which is considered to be related to the structure inhomogeneity of metallic glass in previous research [19]. However, many metallic glasses show only an “excess wings” or “shoulders” based on the loss modulus [15]; (2) As the temperature increases from 400 K and 480 K, which mainly corresponds to the supercooled liquid region of La30Ce30Al15Co25 metallic glasses. Clearly, the storage modulus drops drastically while the loss modulus increases rapidly up to its maximum, which corresponds to the main α relaxation. The magnitude of α relaxation is much higher than β relaxation, which can be attributed to the large-scale movement of atoms [1,16]; (3) When the temperature surpasses 480 K, wherein the sample undergoes an irreversible structural transformation, with the storage modulus again increasing due to the introduction of the crystalline phase with the temperature. These results are in good agreement with the previous study [16,32].
In order to further study the mechanical behavior of LaCe-based metallic glass at high temperatures, tensile strain-rate jump tests were carried out, with the strain rates ranging from 5 × 10−5 s−1 to 2.5 × 10−4 s−1. Typically, the deformation behaviors of metallic glasses depend on both the temperatures and applied strain rates [20,21,33]. A typical tensile strain-stress curve of La30Ce30Al15Co25 metallic glass ribbon during a strain-rate jump test at 450 K was shown in Figure 4.
It is clear to see that the steady flow stress increases by increasing the strain rate. At a given strain rate, the steady flow stress will drop by increasing the measurement of the temperature. According to the strain-rate jump tests, the apparent viscosity can be calculated as [34]
η = σ f 3 ε ˙
where σ f is the steady state flow stress and ε ˙ is the applied strain rate. Figure 5 shows the effect of the temperature and strain rate on the apparent viscosity of La30Ce30Al15Co25 metallic glass. It can be found that the apparent viscosity gradually decreases by increasing the strain rate at a lower temperature. When the temperature is below 450 K, the viscosity changes significantly with the strain rate. The current research is in accordance with the non-Newtonian behavior rheological properties found in previous studies [21,35]. On the other hand, when the temperature is above 450 K, the deformation turns into Newtonian behavior. Obviously, the apparent viscosity is not sensitive to the strain rate [33,36].
The high-temperature tensile deformation of metallic glasses is usually described as [31]
ε ˙ = 2 c f v D exp ( Δ G m KT ) sin h ( σ V 2 3 KT )
where K is the Boltzmann constant, T is the temperature, v D is the Debby frequency, Δ G m is the activation energy for defect migration and c f is the concentration of flow defect. The hyperbolic sine term takes the effect of the applied stress on the energy barrier into account. V is the activation volume, defined as the volume corresponding to each elementary defect jump. The energy barrier will be affected by the activation volume when the flow stress increases. Up to now, there is no perfect explanation for the defects in high-temperature deformation. It should be noted that the defect concentration is closely related to the change of free volume [31]. Equation (2) can also be simplified to the following formula [37]
ε ˙ = ε ˙ 0 sin h ( σ V 2 3 KT )
with
ε ˙ 0 ( c f ) = 2 c f v D exp ( Δ G m KT )
Figure 6 shows the correlation between the flow stress and applied strain rate at various temperatures. The activation volume V and frequency factor ε ˙ 0 can be determined (as shown in Table 1). Based on Equation (2), the values of activation volume V and the frequency factor ε ˙ 0 can be derived for La30Ce30Al15Co25 metallic glass. It can be clearly seen from Table 1 that the activation volume increases with the temperature. The activation can be defined as V = γ 0 Ω f , where γ 0 represents the local shear strain and Ω f is the volume of the involved defect. γ 0 is often taken as 0.125 in previous studies [37]. Taking into account the values of γ 0 , the volume of the LaCe-base metallic glass flow defect volume Ω f can be determined to be about 4.1 nm3. In addition, according to the activation volume data shown in Figure 7, the value Ω f of several typical metallic glasses ranges from 0.8 nm3 to 2.24 nm3. LaCe-based metallic glass involves greater flow defects in high-temperature deformation than that of other typical metallic glasses. This result is consistent with the previous report on La65Ni15Al25 metallic glass [38].
In Figure 7, it is clearly observed that the activation volume of La30Ce30Al15Co25 metallic glass at high-temperature deformation is more than 0.4 nm3. The average volume of atoms values Ω is considered to be 0.013 nm3 [39]. It is reasonable to conclude that the activation volume for La30Ce30Al15Co25 metallic glass corresponds to about 40 atomic volumes. Interestingly, Figure 7 can be divided into two regions: Region (I) is the activation volume of La30Ce30Al15Co25 and La65Ni15Al25 [38] metallic glasses, which, according to the DMA results, exhibit pronounced slow β relaxation; and Region (II), in which metallic glasses do not exhibit distinct slow β relaxation with a smaller activation volume, such as Cu-based, Zr-based and Ti-based metallic glasses, whose activation volume involved in the local motion of the atom is approximately 20 atomic volumes. A larger activation volume is linked to more atoms being activated during the plastic deformation. In general, a larger activation volume perhaps shows a high concentration of defects to some extent.
The microstructure heterogeneity of amorphous alloys has been confirmed by electron microscopy or simulations. It is believed that β relaxation is related to structural heterogeneity. In parallel, the activation of β relaxation is considered to be an indicator of achieving plastic properties [40,41]. In addition, β relaxation is usually considered to be a partial migration of local atoms and the plastic units, i.e., STZs, which are involved in the motions of dozens of atoms. The activation energy of β relaxation and STZs have been confirmed to be almost the same [42]. Several models have been proposed to explain the structural heterogeneity of metallic glasses with the help of experiment or theory. The flow units model suspects that deformation in amorphous alloys is borne by the flow units [43]. Perez proposed the “quasi-point defects” (QPD) model, which assumes that there are fluctuations of entropy or enthalpy in amorphous solids [44]. It is well accepted that β relaxation is associated with structural heterogeneity in metallic glasses [8,16]. Generally, it is believed that the structural heterogeneity of metallic glasses includes a soft “liquid-like” region and a hard “solid-like” region [8,42]. β relaxation probably originated from the “soft” region [38,41], where the atoms are more loosely arranged and easier to move. Therefore, more atoms could be activated under mechanical or thermal stimulation in LaCe-based metallic glasses. Figure 8a,b show the activation volume diagram for the metallic glasses which have pronounced β relaxation or not, respectively. From Figure 8, it can be seen that more atoms could be activated in metallic glasses with a pronounced β relaxation process. As evidenced by Figure 7, two La-based metallic glasses with pronounced β relaxation in area (I) possess more activation volumes than other metallic glasses without obvious β relaxation in area (II). This indicates that metallic glasses with pronounced β relaxation have more atoms involved in their motion during a steady flow at a high temperature, which is in agreement with the above analysis.
In the Newtonian flow region, it is assumed that the flow defects concentration c f is related to the equilibrium volume of a Newtonian state. Therefore, Newtonian equilibrium viscosity η N can be determined by:
η N = 2 3 KT 3 ε 0 ˙ V
Newtonian equilibrium viscosity ηN at different temperatures can be obtained from the experimental data. The values are 1.18 × 1012 Pa s, 3.84 × 1012 Pa s, 3.04 × 1012 Pa s and 6.63 × 1010 Pa s for T = 435 K, 440 K, 445 K and 450 K, respectively. In addition, the viscosity master curve is an approach to discuss the deformation of metallic glass. Figure 9 shows a master curve between the normalized viscosity η / η N and ε ˙ η N. Viscosity at different temperatures was normalized with the corresponding Newtonian viscosity, and the main curve was obtained by shifting with 440 K as the reference temperature. The master curve can be well described by a stretched exponential function, the Kohlraush–Williams–Watts (KWW) model, based on a previous investigation [29]
η η N = 1 exp [ ( σ tc ε ˙ η N ) β kww ]
where σ tc and β kww are fitting parameters. σ tc can be regarded as the characteristic stress of non-Newtonian rheology to Newtonian rheological transformation. Usually, β kww ranges from 0 to 1. It can be seen from Figure 9, and experimental data can be well described by Equation (6). The value of characteristic stress σ tc = 32.6   MPa is related to a transformation from non-Newtonian to Newtonian behavior. β kww = 0.7 is in good accordance with the values of β kww in previous research results: Zr41.2Ti13.8Cu12.5Ni10Be22.5 [46], La55Al25Ni20 [47] and Zr65Cu18Ni7Al10 [29]. It could be concluded that a smaller β kww value means a larger deviation from Newtonian flow.
According to previous investigations, the master curve of viscosity can be described as [48]
η η N = σ V / ( 2 3 KT ) sin h [ σ V / ( 2 3 KT ) ] = X sin hX
where X = σ V / ( 2 3 KT ) . Figure 10 shows the master curve between the normalized viscosity η / η N and X at different temperatures. It should be noted that Newtonian flow occurs when the normalized viscosity tends to a constant. With the variation of temperature and strain rate, there will be a transition between Newtonian and non-Newtonian flow.

4. Conclusions

High-temperature deformation behavior and the dynamic mechanical relaxation process of La30Ce30Al15Co25 metallic glass were investigated in the current work. Compared with other metallic glasses, La30Ce30Al15Co25 metallic glass shows a pronounced slow β relaxation, based on the DMA results. We found that the activation volumes during the steady flow of La-based and La30Ce30Al15Co25 metallic glasses with a pronounced β relaxation are much higher than that of other typical metallic glasses without obvious β relaxation. This phenomenon can be attributed to the same origin of the β relaxation of metallic glasses, i.e., the structural heterogeneity.

Author Contributions

Conceptualization, J.Q.; methodology, Y.C. and J.Q.; formal analysis, Y.C. and J.Q.; investigation, Y.C. and J.Q.; resources, J.Q.; writing—original draft preparation, Y.C.; writing—review and editing, J.Q.; supervision, J.Q.; project administration, J.Q.; funding acquisition, Y.C. and J.Q. All authors have read and agreed to the published version of the manuscript.

Funding

The research of JCQ was supported by the Fundamental Research Funds for the Central Universities (Nos. 3102019ghxm007 and 3102017JC01003), Astronautics Supporting Technology Foundation of China (2019-HT-XG) and the Natural Science Foundation of Shaanxi Province (No. 2019JM-344). YHC acknowledges financial support from the Seed Foundation of Innovation and Creation for Graduate Students in Northwestern Polytechnical University (Grant No. ZZ2019014).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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]
  2. 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]
  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. Wang, W.H.; Dong, C.; Shek, C.H. Bulk Metallic Glasses. Mater. Sci. Eng. R 2008, 321, 502–503. [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. Pang, S.J.; Zhang, T.; Asami, K.; Inoue, A. Synthesis of Fe-Cr-Mo-C-B-P bulk metallic glasses with high corrosion resistance. Acta Mater. 2002, 50, 489–497. [Google Scholar] [CrossRef]
  7. 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] [Green Version]
  8. 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]
  9. Jia, Z.; Wang, Q.; Sun, L.; Wang, Q.; Zhang, L.C.; Wu, G.; Luan, J.H.; Jiao, Z.B.; Wang, A.; Liang, S.X. 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]
  10. Sha, Z.; She, C.; Xu, G.; Pei, Q.; Liu, Z.; Wang, T.; Gao, H. Metallic glass-based chiral nanolattice: Light weight, auxeticity, and superior mechanical properties. Mater. Today 2017, 20, 569–576. [Google Scholar] [CrossRef]
  11. Inoue, A.; Takeuchi, A. Recent development and application products of bulk glassy alloys. Acta Mater. 2011, 59, 2243–2267. [Google Scholar] [CrossRef]
  12. Maass, R.; Derlet, P. Micro-plasticity and recent insights from intermittent and small-scale plasticity. Acta Mater. 2018, 143, 338–363. [Google Scholar] [CrossRef] [Green Version]
  13. Li, N.; Chen, Y.; Jiang, M.; Li, D.; He, J.; Wu, Y.; Liu, L. A thermoplastic forming map of a Zr-based bulk metallic glass. Acta Mater. 2013, 61, 1921–1931. [Google Scholar] [CrossRef] [Green Version]
  14. Lu, J.; Ravichandran, G.; Johnson, W.L. Deformation behavior of the Zr41.2Ti13.8Cu12.5Ni10Be22.5 bulk metallic glass over a wide range of strain-rates and temperatures. Acta Mater. 2003, 51, 3429–3443. [Google Scholar] [CrossRef]
  15. Qiao, J.; Wang, Q.; Crespo, D.; Yang, Y.; Pelletier, J.M. Amorphous physics and materials: Secondary relaxation and dynamic heterogeneity in metallic glasses: A brief review. Chinese Phys. B 2017, 26, 016402. [Google Scholar] [CrossRef] [Green Version]
  16. Wang, W.H. Dynamic relaxations and relaxation-property relationships in metallic glasses. Prog. Mater. Sci. 2019, 106, 100561. [Google Scholar] [CrossRef]
  17. Wang, Q.; Liu, J.; Ye, Y.; Liu, T.; Wang, S.; Liu, C.; 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]
  18. Qiao, J.C.; 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] [Green Version]
  19. Yu, H.B.; Wang, W.H.; Samwer, K. The β relaxation in metallic glasses: An overview. Mater. Today 2013, 16, 183–191. [Google Scholar] [CrossRef]
  20. Ge, Y.; Song, W.; Wang, X.; Luo, Z.; Li, W.; Lin, J. Temperature and strain rate dependence of deformation behavior of Zr65Al7.5Ni10Cu17.5. Mater. Chem. Phys. 2010, 124, 25–28. [Google Scholar] [CrossRef]
  21. Huang, Y.; Shen, J.; Sun, Y.; Sun, J.; Chen, J.J.J. High temperature deformation behaviors of Ti40Zr25Ni3Cu12Be20 bulk metallic glass. J. Alloys Compd. 2010, 504, S82–S85. [Google Scholar] [CrossRef]
  22. Hiss, R.; Hobeika, S.; Lynn, C.; Strobl, G. Network stretching, slip processes, and fragmentation of crystallites during uniaxial drawing of polyethylene and related copolymers. A comparative study. Macromolecules 1999, 32, 4390–4403. [Google Scholar] [CrossRef]
  23. G’sell, C.; Hiver, J.; Dahoun, A. Experimental characterization of deformation damage in solid polymers under tension, and its interrelation with necking. Int. J. Solids Struct. 2002, 39, 3857–3872. [Google Scholar] [CrossRef]
  24. Greer, A.; Cheng, Y.; Ma, E. Shear bands in metallic glasses. Mater. Sci. Eng. R 2013, 74, 71–132. [Google Scholar] [CrossRef]
  25. Tian, Y.B.; Lin, J.G.; Li, W.; Ma, M.; Luo, Z.C.; Jiang, W.J. Deformation behavior of a Cu-based amorphous alloy under different strain rates. J. Appl. Phys. 2011, 109, 083508. [Google Scholar] [CrossRef]
  26. Zhang, C.; Qiao, J.C.; Pelletier, J.M.; Yao, Y. Thermal activation in the Zr65Cu18Ni7 Al10 metallic glass by creep deformation and stress relaxation. Scr. Mater. 2016, 113, 180–184. [Google Scholar] [CrossRef]
  27. Spieckermann, F.; Wilhelm, H.; Polt, G.; Ahzi, S.; Zehetbauer, M. Rate Mechanism and Dislocation Generation in High Density Polyethylene and Other Semicrystalline Polymers. Polymer 2014, 55, 1217–1222. [Google Scholar] [CrossRef]
  28. Richeton, J.; Ahzi, S.; Daridon, L.; Remond, Y. A Formulation of the Cooperative Model for the Yield Stress of Amorphous Polymers for a Wide Range of Strain Rates and Temperatures. Polymer 2005, 46, 6035–6043. [Google Scholar] [CrossRef]
  29. Zhang, C.; Qiao, J.; Pelletier, J.; Yao, Y. Arrhenius Activation of Zr65Cu18Ni7Al10 Bulk Metallic Glass in the Supercooled Liquid Region. Intermetallics 2017, 86, 88–93. [Google Scholar] [CrossRef] [Green Version]
  30. Argon, A.S. Plastic deformation in metallic glasses. Acta Metall. 1979, 27, 7–58. [Google Scholar] [CrossRef]
  31. Spaepen, F. A Microscopic Mechanism for Steady State Inhomogeneous Flow in Metallic Glasses. Acta Metall. 1976, 25, 407–415. [Google Scholar] [CrossRef]
  32. Qiao, J.C.; Chen, Y.H.; Casalini, R.; Pelletier, J.M.; Yao, Y. Main α relaxation and slow β relaxation processes in a La30Ce30Al15Co25 metallic glass. J. Mater Sci. Technol. 2019, 35, 982–986. [Google Scholar] [CrossRef]
  33. Zhang, X.Y.; Yuan, Z.Z.; Feng, X.L.; Cui, L.Z.; Li, D.X. Homogeneous viscous flow behavior of a Cu–Zr based bulk metallic glass composites. Mater. Sci. Eng. A 2015, 620, 352–358. [Google Scholar] [CrossRef]
  34. Chan, K.C.; Chen, Q.; Liu, L. Deformation behavior of Zr55.9Cu18.6Ta8Al7.5Ni10 bulk metallic glass matrix composite in the supercooled liquid region. Intermetallics 2007, 15, 500–505. [Google Scholar] [CrossRef]
  35. Lee, K.S.; Chang, Y.W. Deformation behavior of Zr-based bulk metallic glass in an undercooled liquid state under compressive loading. Met. Mater. Int. 2005, 11, 53. [Google Scholar] [CrossRef]
  36. Liu, Y.; Blandin, J.J.; Kapelski, G.; Suéry, M. High temperature deformation of a Cu40Zr44Ag8Al8 bulk metallic glass. Mater. Sci. Eng. A 2011, 528, 3748–3753. [Google Scholar] [CrossRef]
  37. Mei, J.N.; Soubeyroux, J.L.; Blandin, J.J.; Li, J.S.; Kou, H.C.; Fu, H.Z.; Zhou, L. Homogeneous deformation of Ti41.5Cu37.5Ni7.5Zr2.5Hf5Sn5Si1 bulk metallic glass in the supercooled liquid region. Intermetallics 2011, 19, 48–53. [Google Scholar] [CrossRef]
  38. Qiao, J.C.; Pelletier, J.M.; Blandin, J.J.; Gravier, S. High temperature deformation in a lanthanum based bulk metallic glass showing a pronounced secondary relaxation. Mater. Sci. Eng. A 2013, 586, 57–61. [Google Scholar] [CrossRef]
  39. Yao, Z.F.; Qiao, J.C.; Pelletier, J.M.; Yao, Y. High temperature deformation behaviors of the Zr63.36Cu14.52Ni10.12Al12 bulk metallic glass. J. Mater. Sci. 2016, 51, 4079–4087. [Google Scholar] [CrossRef]
  40. Yu, H.B.; Samwer, K.; Wu, Y.; Wang, W.H. Correlation between beta relaxation and self-diffusion of the smallest constituting atoms in metallic glasses. Phys. Rev. Lett. 2012, 109, 095508. [Google Scholar] [CrossRef] [Green Version]
  41. 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] [PubMed]
  42. Wang, W.H. Correlation between relaxations and plastic deformation, and elastic model of flow in metallic glasses and glass-forming liquids. J. Appl. Phys. 2011, 110, 053521. [Google Scholar]
  43. 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]
  44. Perez, J. Defect diffusion model for volume and enthalpy recovery in amorphous polymers. Polymer 1988, 29, 483–489. [Google Scholar] [CrossRef]
  45. Lee, K.S.; Lee, Y.-S. High-temperature deformation and crystallization behavior of a Cu36Zr48Al8Ag8 bulk metallic glass. J. Mater. Sci. 2012, 47, 2472–2478. [Google Scholar] [CrossRef]
  46. Wang, Q.; Pelletier, J.M.; Blandin, J.J.; Suéry, M. Mechanical properties over the glass transition of Zr41.2Ti13.8Cu12.5Ni10Be22.5 bulk metallic glass. J. Non-Cryst. Solids 2005, 351, 2224–2231. [Google Scholar] [CrossRef]
  47. Aken, B.v.; Hey, P.d.; Sietsma, J. Structural relaxation and plastic flow in amorphous La50Al25Ni25. Mater. Sci. Eng. A 2000, 278, 247–254. [Google Scholar] [CrossRef]
  48. Bletry, M.; Guyot, P.; Blandin, J.J.; Soubeyroux, J.L. Free volume model: High-temperature deformation of a Zr-based bulk metallic glass. Acta Mater. 2006, 54, 1257–1263. [Google Scholar] [CrossRef] [Green Version]
Figure 1. XRD pattern of the as-cast La30Ce30Al15Co25 metallic glass.
Figure 1. XRD pattern of the as-cast La30Ce30Al15Co25 metallic glass.
Materials 13 00833 g001
Figure 2. DSC curve of La30Ce30Al15Co25 metallic glass (heating rate: 20 K/min). Inset is the definition of the glass transition temperature Tg and the onset temperature of crystallization Tx.
Figure 2. DSC curve of La30Ce30Al15Co25 metallic glass (heating rate: 20 K/min). Inset is the definition of the glass transition temperature Tg and the onset temperature of crystallization Tx.
Materials 13 00833 g002
Figure 3. Evolution of the normalized storage modulus and loss modulus with temperature of La30Ce30Al15Co25 metallic glass (the driving frequency is 1 Hz and the heating rate is 2 K/min). Eu is the unrelaxed modulus, which is assumed as the storage modulus at an ambient temperature.
Figure 3. Evolution of the normalized storage modulus and loss modulus with temperature of La30Ce30Al15Co25 metallic glass (the driving frequency is 1 Hz and the heating rate is 2 K/min). Eu is the unrelaxed modulus, which is assumed as the storage modulus at an ambient temperature.
Materials 13 00833 g003
Figure 4. A typical tensile strain-stress curve of La30Ce30Al15Co25 metallic glass ribbon during strain-rate jump experiments at a given temperature of 450 K.
Figure 4. A typical tensile strain-stress curve of La30Ce30Al15Co25 metallic glass ribbon during strain-rate jump experiments at a given temperature of 450 K.
Materials 13 00833 g004
Figure 5. The apparent viscosities of La30Ce30Al15Co25 metallic glass ribbon as functions of the strain rates at different temperatures.
Figure 5. The apparent viscosities of La30Ce30Al15Co25 metallic glass ribbon as functions of the strain rates at different temperatures.
Materials 13 00833 g005
Figure 6. Correlation between the flow stress and strain rate of La30Ce30Al15Co25 metallic glass at different temperatures. The solid line is the fitting curve.
Figure 6. Correlation between the flow stress and strain rate of La30Ce30Al15Co25 metallic glass at different temperatures. The solid line is the fitting curve.
Materials 13 00833 g006
Figure 7. The activation volume varies with testing temperature for typical metallic glasses in high-temperature deformation (testing temperature around the glass transition temperature Tg). (I) La-based metallic glasses with pronounced β relaxation (II) other metallic glasses without obvious β relaxation. The activation volume data obtained in the present research and the previous literatures [37,38,45].
Figure 7. The activation volume varies with testing temperature for typical metallic glasses in high-temperature deformation (testing temperature around the glass transition temperature Tg). (I) La-based metallic glasses with pronounced β relaxation (II) other metallic glasses without obvious β relaxation. The activation volume data obtained in the present research and the previous literatures [37,38,45].
Materials 13 00833 g007
Figure 8. The schematic diagram of activation volumes in metallic glasses at the high-temperature deformation: (a) Metallic glasses show a pronounced β relaxation; and (b) β relaxation is not evident (i.e., excess wing or shoulder) in metallic glasses.
Figure 8. The schematic diagram of activation volumes in metallic glasses at the high-temperature deformation: (a) Metallic glasses show a pronounced β relaxation; and (b) β relaxation is not evident (i.e., excess wing or shoulder) in metallic glasses.
Materials 13 00833 g008
Figure 9. A master curve of La30Ce30Al15Co25 metallic glass at high temperature tensile deformation between normalized viscosity η / η N and ε ˙ η N. The solid line is fitted by Equation (6).
Figure 9. A master curve of La30Ce30Al15Co25 metallic glass at high temperature tensile deformation between normalized viscosity η / η N and ε ˙ η N. The solid line is fitted by Equation (6).
Materials 13 00833 g009
Figure 10. Master curve between normalized viscosity η / η N and X at different temperatures, where X = σ V / ( 2 3 KT ) . The solid line is fitted by Equation (7).
Figure 10. Master curve between normalized viscosity η / η N and X at different temperatures, where X = σ V / ( 2 3 KT ) . The solid line is fitted by Equation (7).
Materials 13 00833 g010
Table 1. The activation volume and frequency factor ε ˙ 0 of La30Ce30Al15Co25 metallic glass at different temperatures are calculated from experimental data.
Table 1. The activation volume and frequency factor ε ˙ 0 of La30Ce30Al15Co25 metallic glass at different temperatures are calculated from experimental data.
Temperature (K)Activation Volume (nm3)Frequency Factor ε ˙ 0 (S−1)
4350.4391.20 × 10−5
4400.5073.59 × 10−5
4450.7794.59 × 10−5
4501.2408.70 × 10−5

Share and Cite

MDPI and ACS Style

Chen, Y.; Qiao, J. Correlation between High Temperature Deformation and β Relaxation in LaCe-Based Metallic Glass. Materials 2020, 13, 833. https://doi.org/10.3390/ma13040833

AMA Style

Chen Y, Qiao J. Correlation between High Temperature Deformation and β Relaxation in LaCe-Based Metallic Glass. Materials. 2020; 13(4):833. https://doi.org/10.3390/ma13040833

Chicago/Turabian Style

Chen, Yinghong, and Jichao Qiao. 2020. "Correlation between High Temperature Deformation and β Relaxation in LaCe-Based Metallic Glass" Materials 13, no. 4: 833. https://doi.org/10.3390/ma13040833

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